Sunteți pe pagina 1din 88

SAND2016-8110J

1
2 Mechanism of the Bauschinger Effect
3 in “Hard Pin” Al-Ge-Si Alloys
4
5
6 Wei Gana, Hyuk Jong Bongb, Hojun Limc, R.K. Bogerd, F. Barlate, and R. H. Wagonerb,1
7
8 a
Medtronic, Mounds View, MN 55112, USA
9 b
The Ohio State University, Dept. Materials Science and Engineering, Columbus, OH 43210, USA
10 c
Sandia National Laboratories, Dept. Comp. Materials and Data Science, Albuquerque, NM 87123, USA
11 d
Simulia Central Region, Cincinnati Office, West Chester, OH 45069, USA
12 e
Pohang University of Science and Technology, GIFT, Pohang, Gyeongbuk, 37673, Rep. of Korea

13 ABSTRACT
14 Wrought Al-Ge-Si alloys were designed and produced to ensure dislocation bypass
15 strengthening (“hard pin” precipitates) without significant precipitate cutting/shearing
16 (“soft pin” precipitates). They were processed from the melt, solution heat treated and
17 aged. Aging curves at temperatures of 393, 433, 473 and 513 K (120, 160, 200 and
18 240 °C) were established and the corresponding precipitate spacings, sizes, and
19 morphologies were measured using TEM. The role of non-shearable precipitates in
20 determining the magnitude of Bauschinger was revealed using large-strain
21 compression/tension (CT) tests. The effect of precipitates on the Bauschinger response
22 was stronger than that of grain boundaries, even for these dilute alloys. The Bauschinger
23 effect increases dramatically from the underaged (UA) to the peak aged (PA) condition
24 and remains constant or decreases slowly through overaging (OA). This is consistent
25 with reported behavior for Al-Cu alloys (maximum effect at PA) and for other Al alloys
26 (increasing effect through OA) such as Al-Cu-Li, Al 6111, Al 2524, and Al 6013. The
27 Al-Ge-Si alloy response was simulated with three microstructural models, including a
28 novel SD (SuperDislocation) model, to reveal the origins of the Bauschinger effect in
29 dilute precipitation-hardened / bypass alloys. The dominant mechanism is related to the
30 elastic interaction of polarized dislocation arrays (generalized pile-up or bow-out model)
31 at precipitate obstacles. Such effects are ignored in continuum and crystal plasticity
32 models.
33
34
35
36
37 Manuscript submitted June 30, 2016.
38

1
Corresponding Author. Tel.: +1 614 292 2079; E-mail address: wagoner.2@osu.edu

1
1
2 I. INTRODUCTION
3
4 Many commercial and model aluminum alloys utilize fine precipitates for strengthening
5 by serving as obstacles to dislocation movement. Such hardening, among other types,
6 can give rise to a transient hardening response upon reverse or non-proportional loading.
7 This transient response is general is referred to as the “Bauschinger effect” (with “effect”
8 uncapitalized) [1], although this term is often used in the more limited sense of simply a
9 reduced yield stress upon a stress reversal (which will be referred to as the “Bauschinger
10 Effect,” capitalized).
11
12 An accurate description of precipitation strengthening and its role in reverse yielding and
13 hardening behavior have importance for, among other things, simulation of sheet forming
14 operations, where stress reversals and non-proportional strain paths are common.
15 Springback is particularly sensitive to the stress under such conditions [2-10]. The
16 generalized Bauschinger effect represents a significant departure from proportional-path
17 stresses for strains typically up to a few pct strain following the reversal [11-15]. At the
18 smallest strains following reversal, there is pronounced nonlinearity that signals the start
19 of such transient behavior [16-18]. The nonlinearity before nominal yield was recently
20 shown to be significant for other than springback applications, in particularly the crush of
21 automotive structural components [16].
22
23 The mechanisms that give rise to the transient response fall into two categories:
24 continuum 2 and dislocation-based models. Continuum models capture stress/strain
25 inhomogeneities but have no intrinsic length scale, such as a Burgers vector, and
26 therefore the scale of the microstructure is irrelevant. That is, coarsening of precipitates
27 of fixed volume fraction would not affect the magnitude of the Bauschinger effect.
28 Examples include phenomenological kinematic hardening models [19,20] following the
29 linear kinematic hardening methods proposed by Prager [21] and Ziegler [22]. Such
30 models fail to capture the Bauschinger effect observed in pure single crystal materials
31 where no second phase or grain texture effect is present [23,24].
32
33 The strength of the Bauschinger effect within the continuum approach depends directly
34 on the volume fraction of second-phase particles. Thus, only composites [25], or
35 smaller-scale alloys with large volume fractions such as dispersion-hardened aluminum
36 [26-28] or dual-phase steels [29] are likely to have Bauschinger behavior dominated by
37 continuum-scale elasto-plasticity 3 . For age-hardened aluminum alloys where

2
Crystal plasticity models that are constructed grain-by-grain (CP-FEM) are a variant of continuum
models in that each grain is treated as a continuum although grain texture (another source of
inhomogeneity) is taken into account. The important distinction for the current work is that neither CP
not continuum treatments take into account the elastic interaction of dislocations and the inherent length
scale of the Burgers vector therein.
3
In fact, recent work by the Wagoner group [120] shows that even in the case of DP 980 steel (UTS
greater than 980 MPa), with large martensite islands of 32 pct volume fraction, the continuum models
capture only a small fraction of the observed Bauschinger effect.

2
1 second-phase particles represent less than 3 pct volume fraction, finite element (FE)
2 simulations show only a minimal Bauschinger effect, disappearing after strains of the
3 order of the elastic strain following a stress reversal (see the Simulations and Discussion
4 section of this paper).
5
6 A basic single-dislocation mechanism for transient yielding and hardening following a
7 stress reversal was proposed by Orowan, who showed that the dislocation looping
8 mechanism can produce significant Bauschinger effects for small volume fractions of
9 closely-spaced particles [30-32]. Elastic and plastic inclusion theories quantify the
10 amount of strengthening and Bauschinger effect from precipitates [33,34].
11
12 Multi-dislocation models conceptually involve the storage and release of
13 elastically-interacting polarized sets of dislocations during stress reversals, producing a
14 transient behavior over a finite strain range (corresponding to the strain needed to propel
15 the dislocations to a new steady-state configuration in the new strain path). These sets can
16 be thought of as generalized dislocation pile-ups or bow-outs [35-38] corresponding to
17 simple models of small numbers of individual dislocation. TEM observations have
18 revealed complex, and difficult to analyze, dislocation interactions during such stress
19 reversal or path changes [39-49].
20
21 Semi-phenomenological models based on dislocation cell structures and slip band
22 evolution [41,44,50-56] are difficult to verify. Some experimental results show that
23 details of the dislocation cell structure do not have a determinant role in the Bauschinger
24 effect [57].
25
26 At the macroscopic level, the Bauschinger effect in binary Al-Cu alloys was originally
27 reported to be strongest at peak aging (i.e. maximum strength) [58,59], consistent with a
28 looping /storage mechanism and a presumption of non-shearable precipitates. However,
29 the original Al-Cu results from the 1980’s are not consistent with more recent results for
30 Al-Cu-Sn [60] and Al-Cu-Li alloys [61,62], which exhibit a Bauschinger effect that
31 continues to increase after peak aging, or at least remains nearly constant or decreases
32 very slowly. Complex commercial alloys 6111 [63], 2524 and 6013 [64,65] commercial
33 Al alloys with an array of particles types, sizes, spacings and coherency exhibit a
34 Bauschinger effect that continues to grow well into the overaging conditions (i.e. for a
35 macroscopically weaker condition featuring large, widely spaced particles) [64].
36
37 In all of these alloy systems, the mechanistic interaction of dislocations with precipitates
38 is complicated by a transition from shear-able, small, coherent, metastable particles at
39 early aging stages, to larger, non-shearable, stable, incoherent, particles in late aging
40 stages [54,55,58,59,61,63,66,67]. Qualitative arguments regarding the role of these
41 complexities in the aging behavior are common but difficult to confirm. In contrast,
42 age-hardened Al-Ge-Si alloys have finely spaced Ge-Si precipitates [68,69] of stable,
43 incoherent, non-shearable [70], fixed-composition diamond-cubic crystal structure [70] at
44 nearly all sizes.
45
46 Thus, the Al-Ge-Si system offers an opportunity to observe the transient behavior at all

3
1 aging stages without these confounding complexities. Significant strengthening occurs
2 during aging for very small volume fractions of precipitates. These properties allow the
3 inherent characteristics of “hard pin4” strengthening in terms of forward and reverse yield
4 and hardening to be revealed experimentally.
5
6 Three Al-Ge-Si systems with various alloy concentrations were produced. The
7 materials were aged for various times at various temperatures, and their precipitate
8 structure (shape, size and spacing) were measured by transmission electron microscope
9 (TEM). Tensile and compression/tension (CT) tests were conducted to obtain the
10 monotonic and reverse hardening behavior of the aged materials. FE simulations were
11 compared with experiments to shed light on the nature of the micro-mechanism
12 producing the Bauschinger effect in “hard pin” strengthened alloys.

13 II. EXPERIMENTAL AND ANALYTICAL PROCEDURES


14
15 Dilute Al-Ge-Si alloys were processed at the Alcoa Technical Center [71]. Three melt
16 chemistries were cast into ingots with nominal dimensions of 51 mm×254 mm×356
17 mm, homogenized for 8 hours (2.9×104 s) at 773 K (500 °C), and forced-air cooled. Hot
18 rolling at 713 K (440 °C) was carried out using 12 passes reducing the initial thickness of
19 51 mm to a final thickness of 5 mm. Cold rolling reduced the strip to its final thickness of
20 2 mm in 3 passes.
21
22
23 A. Compositions
24
25 Tables I and II represent the compositions of the three alloys, along with the expected
26 volume fraction based on the densities of pure metals of Al, Ge, and Si. The volume
27 fractions determined in this way agree with measured volume fractions of particles at
28 peak aging and afterwards. That is, essentially all of the solutes are removed from
29 solution for the peak-aging condition and afterward.
30
31 B. Aging and Characterization
32
33 Test coupons 30 mm×30 mm were sheared from the rolled sheets, solutionized in air at
34 773 K (500 °C) for one hour (3.6×103 s), quenched in water, and blow-dried. These
35 coupons were then aged for various times at temperatures of 393, 433, 473 and 513 K
36 (120, 160, 200 or 240 °C) using a Scientific, Isotemp® Model 725F furnace. A
37 redundant aluminum sheet was preheated in the oven and the samples were placed on top
38 of it in order to increase the temperature rise rate in test coupons for more precise
39 determination of aging times. Micro-hardness data were collected at various aging

4
“Hard pins” are obstacles to dislocation motion that must be bypassed (looped around) during plastic
deformation, as opposed to “soft pins” which may be cut or sheared without significant looping around
them by the impinging dislocations.

4
1 times to identify the mechanical states of the small test coupons.
2
3 The microstructures of the as-received and heat-treated samples were examined using a
4 Philips CM200 transmission electron microscope at 200 kV. To obtain TEM samples,
5 small material coupons were first mechanically thinned to a thickness of about 120 μm,
6 and then electro-polished in 70 pct methanol and 30 pct nitric acid to a final thickness of
7 150 nm at the center of the sample.
8
9 Precipitate size and spacing were collected from TEM images. The maximum
10 magnification was 350,000 times, corresponding to a minimum detectable precipitate size
11 of approximately 1 nm. Most of the precipitates observed were larger than 5 nm.
12 Three to five views were taken for each determination, and about 300 particles were
13 measured manually for each aging condition.
14
15 The precipitate size (d), spacing (λ) and volume fraction (f) were calculated from TEM
16 images. The average precipitate size, d, for the aging condition was obtained by taking
17 the arithmetic mean of all the measured individual particle sizes, which is defined as the
18 square root of the product of the maximum and minimum diameters of the particle
19 ( d ppt = d min d max ). Scatter for the size measurement changes with aging condition, but
20 is about 3~5 nm for most cases (details in Section III.A). The total volume fraction of
21 the precipitates was calculated using the following Equation 1 [72]:
π d 3N
22 f = (Equation 1)
6 A(t + d )
23 where A is the area of micrograph and N is the total number of particles in that area.
24 The foil thickness, t, can be measured using the convergent beam electron diffraction
25 (CBED) method [73,74]. Once the particle size and volume fraction are known, the
26 inter-particle spacing, λ , was calculated using Equation 2 [75]:
27 λ = (1.23 2π / 3 f − 2 2 / 3 )r (Equation 2)
28 where r is the particle radius.
29
30 C. Mechanical Testing
31
32 Vickers micro-hardness tests were conducted with a load of 100 g and 20 s of dwell time
33 on polished sample surfaces. Monotonic tensile tests were conducted at a nominal
34 initial strain rate of 1.7×10-3/s. using standard ASTM E-8 tensile samples [76] .
35
36 Standard tensile properties were correlated to Vickers micro-hardness (HV) for the
37 current Al-Ge-Si alloys and other aluminum alloys as shown in Figures 1 (a) – (d). The
38 other alloys are commercial aluminum alloys (Al 2524 and Al 6013) [64,77] and two
39 friction stir welded alloys (Al 5083-H18 and Al 6111-T4) [78]. The best-fit power-law
40 equations for each tensile stress measure also appear on Figures 1. The correlation
41 between tensile strength and Vickers hardness is nearly linear, while the best correlation
42 with HV is to the “flow stress,” 0.5( σ y + σ UTS ) where σ y and σ UTS are the yield stress
43 and ultimate tensile strength, respectively.

5
1

500
Pure Al
Al-1%Ge-Si
400 Al-0.4%Ge-Si
Al-0.2%Ge-Si
Yield Stress (MPa)

2524, aged
6013, aged
300 5083-H18 & FSW
6111-T4 & FSW
Fitting

200

100 1.7145
y = 0.082x
2
R = 0.9351
0
0 20 40 60 80 100 120 140
Vicker's Hardness (HV)
2
3 (a)

6
500
Pure Al
Al-1%Ge-Si

Ultimate Tensile Stress (MPa)


Al-0.4%Ge-Si
400 Al-0.2%Ge-Si
2524, aged
6013, aged
5083-H18 & FSW
300 6111-T4 & FSW
Fitting

200

1.0875
y = 2.046x
100 2
R = 0.9417

0
0 20 40 60 80 100 120 140
Vicker's Hardness (HV)
1

2 (b)

7
500
Pure Al
Al-1%Ge-Si
Al-0.4%Ge-Si
0.5(YS+UTS) (MPa) 400 Al-0.2%Ge-Si
2524, aged
6013, aged
5083-H18 & FSW
300 6111-T4 & FSW
Fitting

200
1.2847
y = 0.7054x
2
100 R = 0.9535

0
0 20 40 60 80 100 120 140
Vicker's Hardness (HV)
1

2 (c)

8
5
Pure Al
Al-1%Ge-Si
Al-0.4%Ge-Si
4 Al-0.2%Ge-Si
2524, aged
6013, aged
5083-H18 & FSW
3
UTS / YS

6111-T4 & FSW


Fitting

-0.6271
y = 24.957x
1 2
R = 0.6788

0
0 20 40 60 80 100 120 140
Vicker's Hardness (HV)
1

2 (d)

3 Figure 1. Tensile properties and Vickers hardness (HV) correlations for various
4 aluminum alloys: (a) σ y vs. HV, (b) σ UTS vs. HV, (c) 0.5( σ y + σ UTS ) vs. HV, and (d)
5 σ UTS / σ y vs. HV
6
7 The Bauschinger effect was observed using CT tests employing a special sample and a
8 special fixture that applies side forces to prevent buckling at large compressive strains
9 [77-81]. A non-contact EIRTM laser extensometer [82] was used to measure the strain
10 during deformation. Measured stress-strain data were corrected for biaxial stress and
11 friction effects as presented in detail elsewhere [77]. The friction coefficient used in the
12 correction procedure was determined by matching the friction-adjusted tensile curve with
13 standard tension tests.
14
15 D. Bauschinger Effect Characterization
16

9
1 The usual measure of the Bauschinger Effect5 is the “Bauschinger Factor” [83,84], which
2 will be called here as the “Bauschinger Stress Factor” to differentiate it from other
3 measures introduced below. β measures the relative stress magnitude with and without
4 a stress reversal, defined as follows:
σ f +σr
5 β= (Equation 3)
σf
6 where σ f is the signed forward stress before the strain path change and σ r is the
7 signed yield stress (however defined) after the path change. If there is no Bauschinger
8 effect, σ r = - σ f , and β = 0. When yielding occurs at zero stress upon path
9 reversal, β = 1 . The value of β will in general depend on the choice of yield criterion,
10 with β0.2 denoting a standard 0.2 pct offset criterion and β0.4 denoting a 0.4 pct offset
11 yield stress criterion. The latter was adopted based on observations in the literature of
12 excessive scatter of the usual yield criterion for similar CT tests of Mg [85] and Al [64].
13
14 Two additional measures were introduced as follows to extend the characterization of the
15 Bauschinger effect to larger post-reversal strains: 1) the “Bauschinger Strain Factor,” γ ,
16 represents the strain extent of the disturbance (similar to a half-life) while 2) the
17 Bauschinger Energy Factor, ΔE, represents the difference in work dissipated between
18 monotonic and reverse deformation after the reversal strain.
19
20 The quantitative definitions for these new measures are illustrated with the aid of
21 experimental tensile and CT data of Al-1pct Ge-Si, aged for 12 hours (4.3×104 s) as
22 shown in Figure 2 (a). The Bauschinger Strain Factor γ is the increment of strain
23 following reverse yield at which the absolute slope of the reverse loading curve6 equals
24 N (where N = 2 or 4) times the absolute slope of the corresponding monotonic slope (i.e.
25 the slope of a companion tensile test without a stress reversal), or where the absolute
26 stress along the reverse path equals σr, the stress at the reversal. The values
27 corresponding to these values are designated γ2x, γ4x, and γs, respectively.
28

5
“Bauschinger Effect” (both words capitalized) refers the change of yield stress following a stress reversal
compared with the flow stress just prior to reversal. In the current work, the known nonlinearity of
reverse loading [16-18] before yield is ignored. That is, the initial reverse loading is considered linear
up to yielding, as determined conventionally by an offset criterion.
6
All stresses and strains are presented as absolute values for more meaningful, and more simply
interpreted, plots.

10
200
Compression/Tension Tension

σ Δσ=lσ l-lσ l
f r

Absolute True Stress (MPa)


f

ΔE

dσ /dε =2dσ /dε


f f r r

Reverse Yield
100 (0.4 pct offset)

β=(σ +σ )/σ Reverse Yield


f r f (0.2 pct offset)

Load Reversal

0
0.02 0.04 0.06 0.08
Absolute True Strain

1
2 (a)

11
150
γ(dσ /dε =2dσ /dε )
CT CT T T

100
Δσ (MPa)

Reverse Yield Point


(0.2 pct offset)
50 Reverse Yield Point
(0.4 pct offset)

ΔE
0

0
0 0.01 0.02 0.03 0.04
Δε

2 (b)

3 Figure 2. Schematics illustrating the definition of Bauschinger factors, using (a) CT


4 stress-strain plots and (b) Δσ-Δε plots.
5
6 The Bauschinger Energy Factor ΔE is the difference in plastic work done after the
7 reversal or from the yield point, up to an incremental strain of 0.03. That is, it is the
8 integral of (Δσ = |σmonontonic| – |σreverse|) from the strain at zero stress or from the 0.2 pct or
9 0.4 pct offset yield stress. The plastic strains are computed using a Young’s modulus of
10 70 GPa, a typical handbook value for Al. The corresponding energy factors are
11 designated ΔE0, ΔE0.2, and ΔE0.4. The choice of the strain limit of 0.03 past the reversal
12 strain is somewhat arbitrary; it was chosen to capture the bulk of the effect while
13 reducing the error over longer strain ranges in cases where the measured tensile and
14 compressive stresses include a spurious stress offset, whether by mis-calibration of load
15 cells or friction coefficients, or other such problems.
16
17 In order to reveal the transition region with more resolution, a reduced plot following the
18 transition is shown, Figure 2 (b), in terms of Δσ vs. Δε, where Δε is the absolute strain
19 from the zero-stress point following the reversal. Such a plot is convenient for
20 illustrating the definition and computation of the variants of the three Bauschinger
21 factors.

12
1
2
3 III. RESULTS
4
5 Results in two categories are presented. The aging of the model Al-Ge-Si alloys are first
6 established quantitatively, including characterization of hardness and details of the
7 corresponding microstructures. These represent new basic results for these uncommon
8 alloys. Second, the detailed mechanical responses from uniaxial tensile and CT tests are
9 determined and analyzed for several aging states. Together, these are designed to reveal
10 the strengthening mechanisms and origins of the Bauschinger effect for a bypass-only
11 precipitate microstructure.
12
13 A. Aging results
14
15 The precipitate structure of Al-0.2pct Ge-Si, aged at 473 K (200 °C) for 8 hours (2.9×104
16 s) is shown in Figures 3. The precipitate distribution was inhomogeneous in both of the
17 more dilute alloys, Al-0.2pct Ge-Si and Al-0.4pct Ge-Si, making quantification of the
18 microstructures unreliable. Aged at 473 K (200 °C) for several weeks, the maximum
19 increase in tensile strength is about 10 pct for the lower alloy materials. Therefore,
20 complete aging curves were generated only for the richer Al-1pct Ge-Si alloy, which has
21 a more uniform microstructure and more significant precipitation hardening. In the
22 remainder of this paper, “Al-Ge-Si” refers to the 1 pct Ge alloy.
23

2 um 200 nm
24
25 (a) (b)

26 Figure 3. TEM micrographs of Al-0.2pct Ge-Si, aged at 473 K (200 °C) for 8 hours
27 (2.9×104 s): (a) lower magnification, and (b) higher magnification.
28

13
1 The aging curves for Al-Ge-Si at various temperatures are plotted in Figure 4, along with
2 the solution heat treated (SHT) material (HV=47). When aged at 393 and 433 K (120
3 and 160 °C), the peak-aged (PA) and overaged (OA) conditions could not be attained
4 within a realistic time of 1 week (6×105 s). At 513 K (240 oC) the underaged (UA)
5 condition was reached in less than 300 s, making it difficult to control, and the
6 precipitates had an unevenly distributed lath-like nature, Figure 5. Therefore, only the
7 473 K (200 °C) aging of the Al-1pct Ge-Si alloy was selected for detailed TEM
8 characterization. The UA, PA and OA conditions are shown on Figure 4, The UA and
9 OA conditions were selected to have a common hardness of HV=60, while the peak (PA)
10 hardness was HV=667.
11
12 Note: Unless otherwise stated, all remaining data presented in this paper refers
13 to the Al-1pct Ge-Si alloy aged at 473 K (200 °C).
14

70

PA
65
Vicker's Hardness (HV)

UA
60
OA

55 0
433 K(160 C)
513 K 0
0
473 K (200 C)
(240 C)
50 0
393 K (120 C)
SHT
Al-Ge-Si
45
2 3 4 5 6
10 10 10 10 10
Time (s)
15
16 Figure 4. Aging curves of Al-Ge-Si alloy at three temperatures.
17

7
As is show below, corresponding tensile tests for these the UA and OA conditions show similar yield
stresses for the UA and OA conditions as well.

14
1 um
1
2 Figure 5. TEM micrograph of Al-Ge-Si, aged at 513 K (240 °C) for 2 hours (7.2×103 s)
3
4 There was no detectable hardness change for solutionized material over several months at
5 room temperature. Therefore, natural aging was judged to be insignificant. Pure
6 aluminum has a Vickers hardness of 15, corresponding to an ultimate tensile strength
7 (UTS) of 42 MPa. This is less than one-third of the UTS in solution heat treated
8 material (HV=47 and UTS=150 MPa). The difference is attributable to solid solution
9 strengthening; no GP zones or other precipitates were observed in TEM of SHT material,
10 down to a resolution of 1 nm.
11
12 Evenly distributed 0.5~1.0 μm precipitates were apparent in the as-processed alloy (after
13 rolling), Figure 6 (a). Figure 6 (b) reveals dense dislocation networks inside of the
14 material. The as-received material has a hardness of HV=72 (UTS=200 MPa). When
15 solution heat treated at 773 K (500 °C) for 1 hour (3600 s), the dislocation networks
16 disappeared and the grain interiors became free of precipitates, with a few large
17 inclusions observed near the grain boundaries. Because of the very low solubility of Ge
18 or Si in aluminum at room temperature (<0.01 pct) [86-88], diffusion is slow and nearly
19 all of the original solute content at solutionizing temperature is retained during the rapid
20 cooling.
21

15
10 um 100 nm
1
2 (a) (b)
3 Figure 6. Microstructure of Al-Ge-Si, as-received: (a) lower magnification, and (b)
4 higher magnification.
5
6 The effect of low-temperature aging, 393 K (120 °C) for one week (6×105 s), on the
7 microstructure is illustrated in Figures 7. Large cluster-like particles are formed at grain
8 boundaries with an accompanying precipitate-free zone, Figures 7 (a) and (b), both
9 presumably related to enhanced grain boundary diffusion [89]. The grain interiors
10 exhibit homogeneously distributed, equiaxed, fine precipitates, Figures 7 (c) and (d).
11 Quantitative precipitate characteristics are shown on Figures 7 (d).
12

16
2 um 200 nm
1
2 (a) (b)

d=7.1 nm
λ=236 nm
f=0.26 pct

200 nm 50 nm
3

4 (c) (d)

5 Figure 7. TEM micrographs of Al-Ge-Si, aged for one week (6.0×105 s) at 393 K
6 (120 °C): (a) grain junction, (b) grain boundary area, (c) grain interior, and (d) fine
7 precipitates inside a grain.
8
9 Figures 8 show the precipitate structures of the material after aging times of 0.5 hour

17
1 (1.8×103 s), 12 hours (4.3×104 s) and one week at 473 K (200 °C) corresponding
2 respectively to the UA, PA and OA conditions in Figure 4. Virtually all of the
3 precipitates in the OA condition grew into equiaxed shape; no laths or plates were
4 observed. As expected, the average precipitate size for OA is larger than for PA, with a
5 greater dispersion, Figure 9.
6
d=9.4 nm d=15.7 nm
λ=390 nm λ=227 nm
f=0.17 pct f=1.22 pct

100 nm 100 nm
7
8 (a) (b)
9
10
d=25.9 nm
λ=372 nm
f=1.24 pct

100 nm
11
12 (c)

18
1 Figure 8. TEM micrographs of Al-Ge-Si, aged at 473 K (200 0C) for: (a) 0.5 hours
2 (3.6×10 s) (UA), (b) 12 hours (4.3×104 s) (PA), and (c) one week (6.0×105 s) (OA).
3

80

70 UA, d=9 nm <S.D.>=3 nm

60 PA, d=16 nm, <S.D.>=4 nm


Percentage (pct)

50

40 OA, d=26 nm,


<S.D.>=7 nm
30

20

10

0
5 10 15 20 25 30 35 40 45 50
Diameter (nm)
4
5 Figure 9. Precipitate size distribution of Al-Ge-Si aged at 473 K (200 °C). The
6 average diameter (d) and standard deviation are included.
7
8
9 All of the quantitative precipitate measurement results for Al-Ge-Si are summarized in
10 Figure 10. Each label contains three numbers, d (average precipitate diameter), λ
11 (average precipitate spacing) and f (precipitate volume fraction). The PA materials at
12 various temperatures have similar precipitate characteristics and the volume fractions
13 approach the limit of 1.25 pct. That is, all Si and Ge are out of solution in the PA
14 condition (and of course thereafter as well).

19
70
d,λ,f= 18.4,264, 15.7, 227 15.2, 229
1.24 pct 1.22 pct 1.14 pct

65
Vicker's Hardness (HV) 10,262,
0.41 pct
7.9,268,
0.25 pct 7.1,236,
0.26 pct
60 9.4,390 26.1,373,
0.17 pct 1.25 pct
25.9, 372,
1.24 pct

55 0
433 K(160 C)
513 K 0
0
473 K (200 C)
(240 C)
50 0
393 K (120 C)
SHT
Al-Ge-Si
45
2 3 4 5 6
10 10 10 10 10
Time (s)
1
2 Figure 10. Aging curves for Al-Ge-Si with quantitative precipitate measurement results.
3 The precipitate structure parameters are plotted in Figures 11 (a) - (d) with respect to θ, a
4 time quantity adjusted for aging temperature that is defined as follows:
ΔG
5 θ = t * exp(− ) (Equation 4)
RT
6 where t is the aging time, T is the aging temperature, R is the gas constant and ΔG refers
7 to the activation energy which equals to 117±5 kJ/mol for diffusion of Ge or Si in
8 aluminum [90-92]. Figures 11 (a) – (d) illustrate aging of this alloy as follows:
9 • All of the curves collapse to a master curve for all tested temperatures and times,
10 Figures 11 (a) – (d).
11 • The precipitate size increases monotonically with θ (time and temperature),
12 Figure 11 (b). The SHT data point (black filled circle) corresponds to aging for
13 two weeks (1.2×106 s) at room temperature after solution heat treatment and
14 quenching.
15 • The spacing of precipitates is at a minimum at the PA condition. It is
16 approximately equal in the UA and OA conditions, Figure 11 (c).
17 • The volume fraction of precipitates increases abruptly near the PA condition
18 where θ is ~ 10-13 hour (4.6×10-10 s), saturating at the thermodynamically

20
1 expected value, Figure 11 (d).
2

70
0
513 K(240 C) UA PA OA
0
473 K(200 C)
0
65 433 K(160 C)
Vicker's Hardness (HV)

0
373 K(120 C)
SHT
60

55

50

Al-Ge-Si
45
-14 -12 -10 -8 -6
10 10 10 10 10
θ (s)
3
4 (a)
5

21
30
513 K(240 C) UA PA OA
0
0
473 K(200 C)

Precipitate Diameter, d (nm)


0
433 K(160 C)
0
20 373 K(120 C)
SHT

10

Al-Ge-Si
0
-14 -12 -10 -8 -6
10 10 10 10 10
θ (s)
1
2 (b)

22
500
SHT UA PA OA

Precipitate Spacing, λ (nm)


400

300 0
513 K(240 C)
0
473 K(200 C)
0
433 K(160 C)
0
373 K(120 C) Al-Ge-Si
200
-14 -12 -10 -8 -6
10 10 10 10 10
θ (s)
1
2 (c)

23
1.5
513 K(240 C)
0 UA PA OA

Precipitate Volume Fraction, f (pct)


0
473 K(200 C)
0
433 K(160 C)
0
1 373 K(120 C)
SHT

0.5

Al-Ge-Si
0
-14 -12 -10 -8 -6
10 10 10 10 10
θ (s)
1
2 (d)
3 Figure 11. Evolution of precipitate characteristics: (a) the master aging curve, (b) size,
4 (c) spacing, and (d) volume fraction.
5
6 B. Tensile results
7
8 The engineering stress strain curves for the UA, PA, OA, and SHT conditions are shown
9 in Figure 12, along with that for annealed pure aluminum [93]. The high UA strain
10 hardening is presumably related to the absence of precipitates, which tend to increase the
11 yield stress and reduce high-strain strain hardening [94,95]. SHT and pure Al are
12 similarly devoid of precipitates but the SHT exhibits solid-solution strengthening.
13
14 The Orowan bypass mechanism (for hard pins) associates the yield stress increase from
15 aging to particle spacing [31,32,96] as follows:
16 σ = σ o + α ' μb / L (Equation 5)
17 where σo is the yield stress with no precipitates, α’ is a constant on the order of unity, µ is
18 the shear modulus, b is the Burger’s vector and L is the particle spacing. The shear
19 modulus for aluminum is 26 GPa, and its Burger’s vector is 2.86×10-10 m.
20

24
160 UA
PA
Engineering Stress (MPa)
OA
120

80
SHT

40
Pure Al

0
0 0.1 0.2 0.3
Engineering Strain
1
2 Figure 12. Tensile curves of Al-Ge-Si aged at 473 K (200 °C), as compared to pure Al.
3
4 Table III applies Orowan analysis using Equation 5 to the UA, PA, and OA conditions,
5 starting from the SHT condition with the assumption that particle spacing is infinite for
6 the SHT condition. α’ is a fit parameter (=3) such that the predicted yield stress for the
7 UA condition agrees with the measurement; using this same value the strengths of the PA
8 and OA conditions are predicted as shown. It is shown that the measured (σmeas) and
9 predicted (σpred) yield stresses agree reasonably well, implying that the hard pin model is
10 applicable for Ge/Si precipitates in all three aging conditions.
11
12 Table III also shows that the yield stress is maximized for the PA condition, consistent
13 with the hardness measurement, and the OA and UA conditions have approximately the
14 same, lower yield stress, also consistent with hardness.
15
16 The strain hardening rates of the aged materials are plotted against strain and σ-σy in
17 Figures 13 (a) and (b). Most remarkable is the identical strain hardening behavior of the
18 OA and PA conditions, which is significantly different from the UA condition. That is,
19 strain hardening behavior develops during aging, but does not reverse upon overaging in
20 lockstep with hardness and yield stress. The extended strain hardening range for the UA
21 condition is consistent with observations that solutes assist in the multiplication of
22 dislocations over large strain ranges [94,95]. Perhaps another way to state this is that
23 solutes do not inhibit the large-strain dislocation multiplication like precipitates do,

25
1 consistent with them being soft pins, not hard pins that serve as impenetrable obstacles.
2

3000

2000
dσ/dε (MPa)

OA

1000 UA SHT
PA

Al
0
0 0.05 0.1 0.15
True Strain
3
4 (a)

26
3000

2000
dσ/dε (MPa)

1000 UA
PA SHT
Al OA

0
0 50 100 150 200
σ−σ (MPa)
y
1
2 (b)
3 Figure 13. The evolution of work hardening rate of Al-Ge-Si aged at 473 K (200 °C),
4 as compared to pure aluminum: (a) dσ/dε vs. ε and (b) dσ/dε vs. σ-σy.
5
6 In dilute alloys the increase in yield stress from solid solution strengthening [97-99] is
7 proportional to Cα where C is the concentration of the solutes, and α is a constant. A
8 number of theories and experiments have found a 2/3 power of yield stress on alloy
9 concentration [100-103], while others support a 1/2 power [104-106]. Either is
10 consistent with the sparse data shown in Figure 14.
11
12
13

27
60

50 0.5
σ =398xC Al-1pct Ge-Si
ss

40
σ (MPa)

Al-0.2pct Ge-Si
30
ss

20
Al-0.4pct Ge-Si
10

0
0 0.4 0.8 1.2
Total Solute Concentration (atomic pct)
1
2 Figure 14. The yield stresses of the solution heat treated Al-Ge-Si with respect to the
3 total alloy concentrations.
4
5
6 C. CT results
7
8 The Bauschinger factors, i.e., β, γ, and ΔE for aged Al-Ge-Si and for annealed pure
9 aluminum 8 are plotted as a function of aging time in Figures 15 (a)-(c). The
10 over-over-aged condition (OOA) in these figures refers to the aging condition of two
11 weeks (1.2×106 s) at 433 K (160 °C) plus one week (6.0×105 s) at 513 K (240 °C). For the
12 SHT condition and for annealed pure Al, arbitrary aging times were assigned in order to
13 plot the data together with the other, normally-characterized conditions.
14
15

8
Some of the original stress-strain data was lost after the original experiments were performed, thus
making impossible the consistent calculation of some Bauschinger Factors for some cases.

28
Time (s)
0 2 4 6
10 10 10 10
0.4
Al-Ge-Si OA
PA

0.3
β
0.2
OOA
SHT UA
0.2
β

PA OA OOA

β UA
0.4

0.1

Al SHT

0 -14 -12 -10 -8 -6


10 10 10 10 10
θ (s)
1
2 (a)

29
Time (s)
0 2 4 6
10 10 10 10
0.04
Al-Ge-Si
PA

0.03 OA

PA OOA

OA
0.02 γ
γ

UA 2x
γ
s
PA
SHT
0.01 UA OOA

SHT OA
γ
UA 4x
OOA
SHT
0.00
-14 -12 -10 -8 -6
10 10 10 10 10
θ (s)
1
2 (b)

30
Time (s)
0 2 4 6
10 10 10 10
1
Al-Ge-Si PA
OA
0.8 ΔE
0

PA
OA

0.6 UA
ΔE (MPa)

ΔE PA
0.2 OA
0.4
UA

0.2 UA ΔE
0.4

0
-14 -12 -10 -8 -6
10 10 10 10 10
θ (s)
1
2 (c)
3 Figure 15. Measured “Bauschinger Factors” of pure aluminum and Al-Ge-Si materials
4 at SHT, UA, PA and OA conditions (a) Bauschinger Stress Factor, β (b) Bauschinger
5 Strain Factor, γ and (c) Bauschinger Energy Factor, ΔE.
6
7 In general, Figures 15 show that all characteristics of the Bauschinger effect are
8 maximized when the strength is maximized, that is, at the PA condition, consistent with
9 several other Al alloys [58,59]. However, it is clear that the Bauschinger effect declines
10 very little in the OA condition. The Bauschinger effect for OA is similar to PA, much
11 stronger than for the UA condition, contrary to tensile measures of hardness and yield
12 stress. The value of β determined using the 0.4 pct offset criterion and some of the ΔE
13 results maintain the PA value throughout the OA and even into the OAA conditions,
14 consistent with some reports in the literature [64,65]. Another striking factor is the
15 similarity of β values for pure Al and the SHT alloy condition, particularly in view of
16 the large disparity of flow stresses; these values are consistent with forward and reverse
17 torsion results from the literature [107]. The lack of effect of solid solution
18 strengthening on the magnitude of the Bauschinger effect is consistent with a dislocation
19 pile-up or bow-out mechanism: solutes are not hard pins and therefore do not induce
20 dislocation pinning or looping.
21
22 The Bauschinger Stress Factor β is shown in Figure 15 (a). The variation of β0.2
23 corresponds roughly that of precipitate spacing or, equivalently, yield stress. The

31
1 variation of β0.4 corresponds more closely with volume fraction of precipitate. The
2 Bauschinger Strain Factor γ is presented in Figure 15 (b). All three of the
3 strain-magnitude measures are qualitatively similar: they increase initially from SHT to
4 UA and PA, but drop after PA. The Bauschinger Energy Factor ΔE is shown in Figure 15
5 (c). The results are consistent with the results of β with the development of ΔE apparently
6 related to both precipitate spacing and precipitate volume fraction, Figure 11 (d).
7
8 Table IV summarizes the Bauschinger Factors β,γ, and ΔE, for selected choices for
9 criterion. The UA material has a similar particle spacing as the OA case (as expected
10 from the same hardness and yield stress), but all characterization parameters of the
11 Bauschinger effect are significantly smaller in UA as compared to the equal-strength OA.
12 The PA and OA conditions have similar Bauschinger factors even though the precipitate
13 spacing differs by 60 pct. These observations suggest that the Bauschinger factors in a
14 stable population of non-shearable precipitates are controlled both by the precipitate
15 spacing, and by the volume fraction of precipitates.
16
17 The reverse hardening behavior of aged Al-Ge-Si is compared with that of commercial Al
18 2524 alloy in Figures 16. The difference among the aging conditions for the
19 commercial alloy is much greater in view of the greater alloy content. Figure 16 (a)
20 shows that the OA sample for Al-Ge-Si experienced nearly the same Bauschinger
21 transition as the PA sample, although its precipitate size and spacing are 60 pct larger.
22 Figure 16 (b) [65] shows that the OA condition for Al 2524 has a larger Bauschinger
23 effect than the PA condition. This contradicts to the Orowan theory of the Bauschinger
24 effect [32,96], where the precipitate spacing would be the only controlling parameter for
25 dilute hard-pin precipitates.
26
27
28

32
150
Al-Ge-Si
0
473 K (200 C) Aging

100
Δσ (MPa)

OA
50

PA

UA
0
0 0.01 0.02 0.03
Δε
1
2 (a)

33
300
Al 2524
0
493 K (220 C) Aging

200
OA
Δσ (MPa)

PA
100

UA

0
0 0.01 0.02 0.03

Δε
1
2 (b)
3 Figure 16. Reduced CT hardening curves for various aging conditions: (a) Al-Ge-Si,
4 and (b) Al 2524 [83]
5
6 It is well-known that any factor that causes heterogeneous deformation inside of the
7 sample will contribute to the Bauschinger effect [108,109]. For the aged Al-Ge-Si
8 materials, the likely sources of inhomogeneity are precipitates, grain misorientations, and
9 grain boundaries. The results for these alloys show that precipitates are the main source
10 of the Bauschinger effect, even for these dilute alloys. Their development during aging
11 triples the magnitude of the Bauschinger effect as compared with SHT (see Table IV).
12
13 IV. SIMULATIONS AND DISCUSSION
14
15 The experiments and simple analyses presented above show that non-shearable
16 precipitates are the major source of the Bauschinger effect in these aged dilute alloys.
17 However, the results also show that the Bauschinger effect cannot be understood solely
18 through Eshelby and the Orowan theories. Particle size, spacing and volume fraction all
19 have impacts on the magnitude of the Bauschinger effect.
20

34
1 As introduced in Section I, most attempts to predict the Bauschinger effect fall into two
2 broad categories, continuum models [23-29] and semi-phenomenological dislocation
3 models [19,20,53,54] including dislocation bow-out models [35-38]. In order to test these
4 approaches for the current, well-characterized, microstructures, a representative volume
5 element (RVE) and corresponding FE model were developed as shown in Figures 17.
6

35
1
2 (a)

3
4 (b)
5 Figure 17. RVE FE model of Al-Ge-Si alloy for (a) UA/PA/OA conditions with Al
6 matrix and embedded hard particles(assumed particle size, volume fraction and spacings
7 are also listed in the table below the images; f=volume fraction, λ=spacing) (b) Pure
8 Al/SHT condition with 4 grains and corresponding grain boundaries.
9
10 A. Representative Volume Elements (RVE’s)

36
1
2 Distinct RVEs were constructed for three aging conditions: UA, PA, and OA. Figure
3 17 (a) shows a cross-sectional image for a tensile simulation. (For CT simulations,
4 tension was followed by the -4 pct of pre-strain.) The model has an Al matrix and
5 evenly-distributed elastic Ge-Si spherical particles chosen to represent approximately the
6 measured size distributions, spacings, and volume fractions as presented in Table IV and
7 Figure 9. The Hall-Petch effect is generally insignificant for Al alloys having normal
8 grain sizes. (The Hall-Petch slope or Al alloys is < 0.1 MN/m-3/2 [110,111].)
9
10 It is difficult to represent the polycrystalline and precipitate nature in RVE models for the
11 aged alloys considering the orders of magnitude of difference of scale between the
12 nano-scaled precipitates and micro-scaled grains [110,111]. Therefore, the Al matrix in
13 the RVE model for the aged alloys represents a single crystal. Ignoring the coarse grains
14 relative to the fine precipitates is reasonable in view of the size difference and the
15 minimal Hall-Petch slope. On the contrary, the RVE for the pure Al and SHT condition
16 considers only the grain boundary effect (grain size of 25 µm9), not the precipitate effect
17 (which is absent in any case since no precipitates are found), as shown in Figure 17 (b).
18
19 Symmetry boundary conditions were applied to the x-, y-, and z- planes. The bonding
20 between the elastic particle and the matrix is assumed to be perfect i.e., no interface
21 separation or sliding is allowed. A displacement boundary condition is applied to the
22 loading direction.
23
24 B. Alternate Constitutive Models
25
26 Three constitutive models were developed: 1) a standard continuum model, 2) a standard
27 Crystal Plasticity (CP) model based on slip system geometry [112-116] and 3) a novel
28 Super Dislocation (SD) model essentially similar to the CP model but accounting for the
29 discrete dislocation-dislocation, dislocation–obstacle elastic interactions10.
30
31 The SD model has an inherent length scale in the form of the magnitude of the real
32 Burgers vector, not an arbitrarily determined parameter [111]. Therefore, absolute particle
33 size and spacing have meaning, in addition to the volume fraction, which is the only
34 important microstructural geometric variable for continuum and CP models. The
35 precipitates are assumed to be elastically compatible with the Al matrix (same elastic
36 constants and orientation), but exhibit no plastic deformation, that is, they have an infinite
37 hardness.
38
39 Continuum Model - The Al matrix in the continuum model is elastically-isotropic
40 (Young’s modulus, E=70GPa, Poisson’s ratio, ν=0.3) with a plastic response obeying a

9
The reported grain sizes of Al or Al alloys are in broad range. The assumed size was arbitrary chosen
from literature [110,132].
10
In fact, while discrete-dislocation interactions are reflected in the formulation, they are implemented in
terms of lumped “Superdislocations” (hence the name of the technique) at the center of each finite
element.

37
1 von-Mises isotropic yield function and an isotropic hardening rule. The plastic strain
2 hardening of Al matrix was modeled using a Swift hardening law:
3
σ = K (ε + ε )
n
4 (Equation 6)
5
6 where K is material constant, ε 0 is strain at yield and n is strain hardening exponent.
7 The constants are selected to match the measured stress-strain behavior of the PA case,
8 with Ge/Si elastic properties being adopted as Young’s Modulus E=100GPa and Poisson’s
9 ratio ν=0.22 (average of pure Ge and Si).
10
11 CP and SD models - The SD model is a modification of standard CP models widely used
12 for finite element modeling of textured polycrystals. A summary of the SD model and
13 its implementation here is presented in the Appendix to this paper; for full information,
14 the original references are recommended [112,117-120].
15
16 The original SD model [118] was constructed for treating the elastic interaction of
17 discrete dislocations with each other and with obstacles on slip systems where they
18 intersect with grain boundaries. Later versions successfully incorporated other
19 dislocation slip obstacles, such as large martensite islands in dual-phase steels [120],
20 where the interaction is primarily at a macro/continuum/composite scale. In the current
21 work, the SD model is further extended to dilute Al alloys, where the precipitates have
22 such a small volume fraction as to be inconsequential in a composite/continuum sense;
23 only the Orowan interactions are significant.
24
25 An important aspect of the SD model (and the CP model as implemented here) is that the
26 constitutive parameters, with only one exception, are determined from single-crystal
27 properties. In this way, they are predictive such that the predictions can be usefully
28 compared with experiments to evaluate the fidelity of the constitutive approach – and,
29 therefore, the source of the physical phenomena. However, the strength of the alloy
30 must be used to determine one of the single-crystal properties. This is readily done using
31 one point of a tensile test result. Thus, predicting the Bauschinger effect uses no fit
32 parameters from the CT test itself. The two results are independent.
33
34 The SD (and CP) model discretizes a polycrystal into finite elements, many elements per
35 grain. The equilibrium boundary value problem is solved in the usual way in
36 Abaqus/Standard with the choice of constitutive models being implemented through user
37 subroutines. The solution of the equilibrium equation is implicit, but the coupling at
38 each time step with the user subroutines is explicit. That is, there is no iteration between
39 Abaqus and subroutines at each time step, but there are iterations at each time step within
40 Abaqus and within each of the subroutines.
41
42 The SD model consists of two such subroutines:
43 a) a nearly-standard CP model implemented in Abaqus/Standard, and,
44 b) a special meso-scale subroutine that computes the local backstress among
45 dislocation populations from element to element and updates dislocation densities

38
1 on each slip system in each element (along with corresponding local slip
2 resistances) based on Orowan’s equation.
3 The only difference in the SD and CP implementations here is the use of the “b”
4 subroutine above in order to compute the local back stresses caused by the dislocation
5 populations on the similar slip systems in each element within a grain and the
6 modification of the dislocation populations to conform with Orowan’s equation and the
7 plastic strain increments.
8
9 C. Simulation Results and Comparison with Experiment
10
11 Typical simulation results for the PA condition are shown in Figures 18. Figure 18 (a)
12 presents the stress-strain experimental results for a tension test and CT test as points, with
13 CT simulation results using the SD, CP, and Continuum models as lines. While the SD
14 model predicts the correct form and approximate quantitative details as well, the
15 Continuum and CP models show virtually no Bauschinger effect. This simple
16 comparison provides a strong basis for understanding the origin of the Bauschinger effect
17 in this material as arising from dislocation-dislocation and dislocation-precipitate
18 interactions, which are accounted for in the SD model but ignored in the CP and
19 Continuum models.
20

200
CP, Continuum
Absolute True Stress (MPa)

150
Data (CT)

SD (CT)
100 SD (T)

50

Data (T)
Al-Ge-Si (PA)
0
0 0.02 0.04 0.06 0.08 0.1
Absolute True Strain
21
22 (a)

39
150
Al-Ge-Si (PA)

100
Δσ (MPa)

Data (CT)
50 SD (CT)

CP, Continuum
0
0 0.01 0.02 0.03
Δε
1
2 (b)
3 Figure 18. Comparison of CT transient hardening, experimental and simulated (3
4 models): (a) stress-strain, and (b) Reduced Δσ vs. Δε form.
5
6 Figure 18 (b) is a more detailed comparison of measured and simulated CT results in
7 terms of Δσ−Δε, in the format already presented experimentally in Figure 2 (b) and
8 Figures 16 (a) and (b). The predicted hardening transient hardening is an excellent
9 match: it has all of the qualitative features of the experiment and most of the quantitative
10 features, including the Δσ value at yield and at a subsequent strain of 0.03. On the
11 contrary, the CP and Continuum simulations show essentially no transient disturbance
12 following the compression-tension transition: no change of stress or strain hardening after
13 yield.
14
15 Figures 19 and Table V present the full simulation and experimental results in terms of
16 the Bauschinger Factors introduced above. The results are consistent across the various
17 heat-treatment conditions in terms of form and general magnitudes. As in the
18 experiments, the predicted Bauschinger effect peaks at the PA condition and drops off
19 only slowly thereafter. The Bauschinger effect for precipitate-free SHT and pure Al
20 conditions is minimal by comparison – it corresponds to the small effect from grain
21 boundaries alone. The UA condition is closer to SHT in terms of the magnitude of the
22 Bauschinger effect, much smaller than for the similar-strength OA condition.
23

40
Time (s)
0 2 4 6
10 10 10 10
0.3
Al-Ge-Si
PA
OA OOA

0.2
Data
UA
0.4

OA
β

SD
0.1 UA
SHT

Al SHT CP/Continuum
Al
β =0 for all conditions
0.4
0
-14 -12 -10 -8 -6
10 10 10 10 10
θ (s)
1
2 (a)

41
Time (s)
0 2 4 6
10 10 10 10
0.03
Al-Ge-Si PA

Data OA

PA
0.02
UA
OA
2k
γ

SD
UA
0.01 SHT OOA

Al SHT CP/Continuum
γ =0.002-0.004
2k

0
-14 -12 -10 -8 -6
10 10 10 10 10
θ (s)
1

2 (b)

42
Time (s)
0 2 4 6
10 10 10 10
1
Al-Ge-Si
PA
0.8 Data
OA
PA
0.6
ΔE (MPa)

UA
0

OA
0.4 SD
UA

0.2
SHT
Al

0
-14 -12 -10 -8 -6
10 10 10 10 10
θ (s)
1

2 (c)

3 Figure 19. Comparison of Bauschinger factors from experiments and simulation as a


4 function of aging time: (a) β0.4 (b) γ2k and (c) ΔE0 (CP and Continuum results are not
5 shown: they like on the abscissa with ΔE=0.00-0.03. See Table V)
6
7 In view of Figures 19 and Table V, it is clear that the SD model captures the essence of
8 the transition of each case, with good quantitative accuracy while the standard CP and
9 Continuum simulations predict essentially no Bauschinger effect or transient behavior
10 following the reverse yield. The latter result was readily predicable, at least for
11 continuum simulations: such a small volume fraction of hard particles (even infinitely
12 hard, as treated here) cannot have a significant effect on the continuum-predicted
13 hardening. The same turns out to be true even including slip system effects, elastic
14 effects, and other such inhomogeneities. Clearly, the essence of the Bauschinger effect
15 lies with its origin in dislocation-dislocation interactions and dislocation-obstacle
16 interaction, in this case the latter being with precipitates.
17
18
19 V. CONCLUSIONS
20

43
1 The aging of Al-Ge-Si alloys and their subsequent mechanical characterization and
2 simulation were used to reveal the role of non-shearable precipitates on mechanical
3 behavior, thus separating the transition from shearable to non-shearble precipitates in
4 typical Al age-hardening alloys. Two kinds of conclusions were reached, the first with
5 respect to purely microstructural features associated with aging of Al-Ge-Si alloys, and
6 the second with respect to experimental and mechanistic aspects of the reverse yield and
7 hardening (generalized Bauschinger effect).
8
9 Aging of Al-Ge-Si
10
11 1. The aging curves 1 atomic percent Al-Ge-Si at three temperatures fall on a single
12 master curve with little scatter using a reduced temperature-time variable based
13 on published values of the activation energy of diffusion of germanium in
14 aluminum.
15
16 2. Consistent with the Orowan equation, the precipitate spacing passes through a
17 minimum at the peak aging condition, as defined by maximum hardness and
18 yield stress. Concurrently, the volume fraction of precipitates increases
19 dramatically to its full value, and it maintains a plateau thereafter, representing a
20 condition with nearly all the solute out of the matrix.
21
22 3. While the peak hardness and yield strength occur at the PA condition, and the OA
23 and UA conditions represent equal-but-lower hardness and strength, the strain
24 hardening is almost identical for OA and PA conditions, and the Bauschinger
25 effect is similar.
26
27 4. Under most aging conditions tested, the morphology of the Ge-Si precipitates
28 remains nearly equiaxed. The precipitate size increases monotonically with
29 aging, while the size distribution widens.
30
31 5. Some large particles (up to 10 times larger than intra-grain particles) form at
32 grain boundaries for the aging temperatures studied, accompanied by adjacent
33 precipitate-free regions.
34
35 6. Al-0.2pct Ge-Si and Al-0.4pct Ge-Si have inhomogenously distributed
36 precipitates and more irregularly-shaped laths and plates at all the studied aging
37 temperatures. These alloys were not characterized extensively.
38
39 Bauschinger Effect -Experimental
40
41 7. Bauschinger Factors have been introduced to characterize the nature of reverse
42 hardening in terms of stress, strain and combined stress/strain (energy). These
43 tend to rise and fall together. This result may not be universal [108].
44
45 8. Approximately 70 pct of the Bauschinger effect for the PA and OA conditions is
46 attributable to precipitates, with the remainder attributable to other obstacles,

44
1 primarily grain boundaries.
2
3 9. The Bauschinger effect is generally maximized when the strength is maximized,
4 i.e., the PA condition. However, it declines only slowly with further aging - the
5 OA and PA conditions exhibit similar Bauschinger Factors although the
6 precipitate spacing differs by 60 pct.
7
8 10. The Bauschinger effect for OA is much greater than UA, even though they
9 exhibit similar strength. These observations imply that the Bauschinger effect for
10 the non-shearable precipitates is not directly related to simple Orowan effects
11 arising from pin spacing alone.
12
13 Bauschinger Effect – Mechanistic
14
15 11. Simulations show that the dominant mechanism controlling the Bauschinger
16 effect is the generation of internal stresses by polarized dislocations that are
17 “piled-up” (in a general sense) against obstacles and which interact to create a
18 back stress that either resists strain (forward direction) or assists it (reverse
19 direction). Whether the actual dislocation configuration can be characterized as
20 pile-ups of parallel dislocations, bow-out of many dislocations between hard pins,
21 or another configuration, cannot be answered by the current work.
22
23 12. An RVE with properties determined by single-crystal, precipitate-free behavior
24 and an “SD” constitutive model accounting for elastic dislocation-dislocation,
25 dislocation-precipitate, and dislocation-grain boundary interactions predicted all
26 of the significant features of the observed Bauschinger effect. Such simulations
27 have no artificial length scales; instead they rely on measured microstructural
28 scales and the Burgers vector. Corresponding Continuum and CP simulations,
29 which have no real length scale, predict essentially no Bauschinger effect for
30 these alloys.
31
32 13. The reverse yielding behavior and flow of the very dilute Al alloys with
33 non-shearable precipitates cannot be predicted by composite models (Eshelby
34 theory), but is instead attributable to the interaction between precipitates and
35 dislocations.
36
37 14. The RVE based FE simulations for pure Al, SHT, PA, UA and OA conditions
38 reveal that the observed Bauschinger behavior is related not only to precipitate
39 spacing, but also to volume fraction. The observed behavior is reproduced
40 closely by the SD model predictions.
41
42 15. In Al alloys with non-shearable precipitates, even very small volume fractions,
43 the effect of grain boundaries on the Bauschinger effect is minor, contrary to
44 larger effect in ferrous alloys. Nonetheless, for the SHT condition or for pure Al
45 with normal grain sizes, the Bauschinger effect is still well-predicted by the SD
46 model.

45
1
2 ACKNLOWLEDGEMENTS
3
4 This research is supported by the U.S. Department of Energy, Office of Science, Office of
5 Basic Energy Sciences, Materials Science and Engineering Division, and particularly
6 Program Director John S. Vetrano, who are thanked for their support of the current work
7 under Award Number DE-SC0012483 and DE-SC0012587. Thanks also to the Alcoa
8 Technical Center, particularly Robert Hyland (currently at U.S. Steel), for partially
9 funding the original experiments many years ago and for providing the
10 aluminum-germanium-silicon alloys.
11

46
1 APPENDIX: SD and CP Models and Implementation
2
3 The interested reader is directed to the much longer, original, papers [112,117-120]
4 developing the SD Model and testing it with detailed microstructural characterization.
5 The development below is intended to provide sufficient background to under the method
6 and its particular implementation in the current work.
7
8 A. SD and CP Methods
9
10 In both CP and SD implementations, a classical crystal plasticity framework is employed
11 [112,121-123]. The viscoplastic shear rate of the power-law form was adopted to
12 represent γ (α ) as follow [122]:
13
1/ m
(α )  τ (α ) 
14 γ = γ0  (α )  sign(τ (α ) ) (Equation A1)
g 
15
16 where τ (α ) is the resolved shear stress on the slip system α , γ0 is a reference shear
17 rate, m is a strain rate sensitivity, and g (α ) is the slip resistance on the slip system α .
18
19 In both CP and SD implementations, a dislocation density-based constitutive equation for
20 single crystals is adopted [117]. The initial dislocation density is equally divided on all
21 slip systems in each element. The slip resistance of slip system α , g (α ) , has initial
22 value of g0 and evolves with plastic shear strain on all slip systems according to
23
NS
24 g (α ) = g0 + Aμb 
β
hαβ ρ β
=1
( )
(Equation A2)

25
26 where μ is the shear modulus, b is the Burgers vector, ρ (α ) is the dislocation
27 density in slip system α , hαβ = n(α )ξ ( β ) represent interaction cosine where n(α ) and
28 ξ ( β ) are the slip plane normal of slip system α and the dislocation line vector of slip
29 system β , respectively. The details of geometric formulation is given elsewhere [117].
30 The parameter A is the material constant that ranges from 0.3-0.6 [124-125]. A value of
31 0.4 is assumed in the SD and CP models.
32
33 The dislocation density evolves according to a well-known dislocation density evolution
34 equation [126]:
35
 
β
NS
1 ρ (β )  γ (α )
(α ) (α )
36 ρ =  − kb ρ 
(Equation A3)
b ka 
 
37

47
1 where k a and kb are material parameters related to dislocation generation and
2 annihilation, respectively.
3
4 In the SD treatment, the dislocation density for each slip system in each element is treated
5 as a single super-dislocation located at the centroid of the element. Mobile dislocation
6 content at the end of each time step is calculated to accommodate the strain gradient and
7 ia then redistributed according to the Orowan equation [127] in the view of the shear
8 strain increment throughout the body:
9
(α ) 1 (α )
10 ρ pass = γ (Equation A4)
bl (α )
11
(α )
12 where ρ pass is the rate of dislocation density passing through the element, l (α ) is the
13 length of the element parallel to the slip plane and γ (α ) is the shear rate.
14
15 The redistributed dislocation densities, in the form of the magnitude of the
16 Superdislocations on each slip system in each element are used to calculate the elastic
17 interaction forces among them using analytical solutions for the anisotropic elastic fields
18 of parallel dislocation segments [118,128]. These are introduced as backstresses on each
19 slip system, τ b(α ) , into Equation A1 is modified as follows:
20
1/ m
(α )
 τ eff
(α )
 (α )
21 γ = γ0  (α ) 
g 
sign(τ eff ) (Equation A5)
 
22
(α )
23 where τ eff = τ (α ) − τ b(α ) .
24
25 For interactions with grain boundaries, the SD subroutine enforces a critical local stress
26 for passage of slip through a grain boundary [118] or phase boundary as follows:
27
28 τ obs = (1 − N )τ * (Equation A6)
29
30 where τ obs is the obstacle strength when the slip across a grain boundary occurs, τ * is
31 a generally-unknown maximum obstacle strength – its value has been estimated, and
32 adopted here, as 5 times of macroscopic yield stress [129], and N is a geometrical
33 transmissivity factor determined based on the SWC (Shen, Wagoner, Clark [129]) 2nd
34 Criterion as follows:
35
36 N = ( L1 ⋅ Li ) × ( g1 ⋅ gi ) (Equation A7)
37

48
1 where L1 and Li are the intersection line vectors between grain boundary and slip
2 planes, and g1 and gi are the slip direction vectors of incoming and transmitted
3 dislocations, respectively.
4
5 The obstacle strength was incorporated for grain boundary elements by following
6 equations:
7
1/ m
(α ) (α ) (α )
 τ (α ) − τ (α )  (α ) (α )
8 τ eff >τ obs γ = γ0  eff (α ) obs  sign(τ eff
 g 
− τ obs ), (Equation A8)
 
9 τ eff(α ) ≤ τ obs
(α ) (α )
γ =0
 (Equation A9)
10
11 Equations A6-9 were also applied to phase boundaries in the current work, i.e. between
12 the Al matrix and Ge/Si precipitates.
13
14 B. Parameter Identification
15
16 The following constitutive parameters for the CP (and SD) models must be determined:
17 k a i , kb i , ρ0i (Equation A3) / g0i (Equation A2) (i=1,2 for Al, Ge-Si). When
18 Equations A6 and A7 are adopted and the value of τ * is taken to be 5 times the
19 macroscopic yield stress (as it is here), the SD Model introduces no undetermined
20 constants relative to a standard CP scheme. However, values of the parameters will vary
21 widely between SD and CP models because of the strong effect of elastic dislocation
22 interactions in SD that are ignored, and thus incorporated obscurely, in the standard CP
23 approach.
24
25
26 SD model (Micro Approach)
27
28 In the primary method for setting parameters, only single crystal properties for the phases
29 are used. Several parameters were adopted from the literature for Al as follows:
30 anisotropic elasticity constants of C11=107, C12=61 and C44=28 GPa [112] and an initial
31 dislocation density ρ0 Al (=1×1012/m2) for the Al matrix [130]. The strain hardening
32 parameters k a Al and kb Al (Equation 9) were obtained by fitting to stress-strain curves
33 of single crystal Al data [131]. τ* was set at 5 times of macroscopic yield stress of PA
34 sample by following earlier works11 [118,119], i.e., 575 MPa.
35

11
The assumed value (5 times of macroscopic yield stress) was actually assumed for grain-to-grain
interfaces in the earlier works, but same assumption was employed for grain-to-particle interface in this
study.

49
1 For the Ge-Si precipitates, the anisotropic elasticity constants, k a Ge − Si and k b Ge − Si were
2 taken the same as those for the Al while a value of ρ 0Ge− Si of 1×1016/m2 (104 times
3 higher than in Al 12 ) was adopted to simulate an elastic-only precipitate. The
4 transmissivity factor, N, to calculate τ obs was set to be 0. That virtually means that
5 dislocations coming from precipitates are not transmitted toward the Al matrix.
6
7 With all other constitutive parameters set by microstructural measurements, only one
8 parameter, g0i , was set using the measured macro yield stress of the alloy. This value
9 (same for Al and Ge/Si) was determined by matching the simulated yield with measured
10 yield for the PA condition, 115 MPa.
11
12 CT simulations using the resulting constitutive model are shown in Figure A1. Note that
13 the tensile strain hardening is not predicted perfectly by the micro approach, i.e. the
14 constitutive model with only one constant matched to tensile behavior (i.e. yield stress).
15 This discrepancy, as shown before, turns out to be unimportant for the main goal of
16 predicting the Bauschinger effect, but it makes it difficult to separate the monotonic strain
17 hardening from the reverse hardening effects. In order to simplify the appearance of the
18 predictions, an alternate macro approach to parameter identification was also used, as
19 outlined below.
20
21

12
The assumed initial dislocation density in Ge/Si particles imposes ~300 MPa yield strength increase,
higher than the ultimate tensile strength of the alloy. Other simulations were conducted with 105 times
higher, corresponding to ~1 GPa yield strength increase. No difference in overall stress-strain or
Bauschinger behavior was predicted.

50
200
CP (CT)
Continuum (CT)
SD-singleX fit (T)
Absolute True Stress (MPa) 150
Data (CT)
SD-singleX fit (CT)
100 SD-macro fit (T)
SD-macro fit (CT)

50

Data (T)
Al-Ge-Si (PA)
0
0 0.02 0.04 0.06 0.08 0.1
Absolute True Strain
1
2 Figure A1. Comparison of simulated and experimental CT tests. Three constitutive
3 models compared
4
5
6 SD model (Macro Approach)
7
8 In the macro approach to parameter identification, the constitutive parameters
9 representing dislocation multiplication and strain hardening, k a Al and k b Al , were fit
10 using the measured stress-strain curve for the alloy rather than relying on single-crystal
11 measurements from the literature. As shown in Figure A1, following this approach
12 reproduces the observed tensile strain hardening much more closely, thus making it easier
13 to see the transient following the path reversal. Most importantly, the line labelled “SD
14 Model” in Figure 18 (b), in the main body of this paper, represents the identical results
15 from both methods. This shows that there is no difference to the predicted Bauschinger
16 effect using the two sets of parameters, one obtained from microstructural properties only,
17 the other including parameters set by measured alloy strength and strain hardening from a
18 tensile test. In either case, the CT test data is independent of the prediction – that is,
19 nothing from the CT test is used in predicting the transient response.
20
21 As noted above, and as shown in Figure A1, the macro approach to parameter
22 identification captures the strain hardening much more closely than the micro approach.

51
1 But, there is no significant difference in terms of Bauschinger effect predictions. This is
2 shown by the common line marked “SD Model” on Figure 18 (b) that corresponds to both
3 sets of parameters.
4
5 CP Model (Macro Approach)
6
7 Using standard methods for fitting CP model parameters, k a i , kb i , ρ0i were optimized
8 to reproduce the tensile test stress-strain curve; nothing was used from CT test results. All
9 the other parameters were set identically to the SD models. All of the final parameters
10 are shown in Table AI.
11
12

52
1 REFERENCES
2 1. J. Bauschinger: Technischen Hochschule in Munchen, 1886, vol. 13, pp. 31.

3 2. L. Geng, Y. Shen and R.H. Wagoner: Int. J. Plast., 2002, vol. 18, pp. 743-67.

4 3. L. Geng and R.H. Wagoner: Int J Mech Sci, 2002, vol. 44, pp. 123-48.

5 4. J. Lee, J.-Y. Lee, F. Barlat, R. H. Wagoner, K. Chung, and M.-G. Lee, Int. J. Plast.,

6 2013 vol. 45, pp. 140-59.

7 5. J.-Y. Lee, J.-W. Lee, M.-G. Lee, and F. Barlat, Int. J. Solids Struct., 2012, vol. 49, pp.

8 3562-72.

9 6. R. H. Wagoner, H. Lim, M. G. Lee, Int. J. Plast., 2013, Vol. 45, pp. 3-20.

10 7. R. H. Wagoner, J. F. Wang and M. Li: ASM Handbook, ASM, Materials Park, OH,

11 2006, pp. 733-55.

12 8. D. Raabe, F. Roters, F. Barlat and L.-Q. Chen, Continuum Scale Simulation of

13 Engineering Materials, Fundamentals - Microstructures - Process Applications, eds.,

14 Wiley-VCH, Weinheim, 2004, pp. 777-794.

15 9. W. D. Carden, L. M. Geng, D. K. Matlock and R. H. Wagoner, Int. J. Mech. Sci.,

16 2002, vol. 44, pp. 79- 101.

17 10. K.P. Li , W.P. Carden, and R.H. Wagoner, Int. J. Mech. Sci., 2002, vol. 44, pp. 103-22.

18 11. A. B. Doucet and R. H. Wagoner, Metall. Trans. A, 1989, vol. 20A, pp. 1483-93.

19 12. J. V. Laukonis and R. H. Wagoner, Metall. Trans. A, 1984, vol. 15A, pp. 421-25.

20 13. R. H. Wagoner and J. V. Laukonis, Metall. Trans. A, 1983, vol. 14A, pp. 1487-95.

21 14. R. H. Wagoner, Metall. Trans. A, 1982, vol. 13A, pp. 1491-1500.

22 15. C. Zhou, Z. Chen, J.W. Lee, M.G. Lee, R. H. Wagoner, Int. J. Plast., 2015, vol. 75, pp.

23 121-40.

53
1 16. Z. Chen, U. Gandhi, J. Lee, R. H. Wagoner, J. Mater. Proc. Technol., 2016, vol. 227,

2 pp. 227-243.

3 17. Z. Chen, H. J. Bong, D. Li and R. H. Wagoner, Int. J. Plast., 2016, vol. 83, pp.

4 178-201.

5 18. L. Sun, R. H. Wagoner, Int. J. Plast., 2011, vol. 27, pp. 1126-44.

6 19. J. da Costa Teixeira, L. Bourgeois, C. W. Sinclair, and C. R. Hutchinson, Acta Mater.,

7 2009, vol. 57, pp. 6075-89.

8 20. Y. Chen, M. Weyland, and C. R. Hutchinson, Acta Mater., 2013, vol. 61, pp. 5877-94.

9 21. W. Prager: J. Appl. Mech. Trans. ASME, 1956, vol. 23, pp. 493-6.

10 22. H. Ziegler: Q. Appl. Math., 1959, vol. 17, pp.

11 23. E.H. Edwards and J. Washburn: Journal of Metals, 1955, vol. 6, pp. 1239-42.

12 24. M.Z. Wang, S. Lin, C.H. Li and e. al.: Scripta Mater., 1996, vol. 35, pp. 1183-88.

13 25. B. Johannesson, S.L. Ogin, M.K. Surappa, P. Tsakiropoulos, S. Brynjolfsson and I.O.

14 Thorbjornsson: Scripta Mater., 2001, vol. 45, pp. 993-1000.

15 26. A.P. Reynolds and J.S. Lyons: Metall Mater Trans A, 1997, vol. 28, pp. 1205-11.

16 27. L.M. Brown and W.M. Stobbs: Phil. Mag., 1971, vol. 23, pp. 1185-99.

17 28. L.M. Brown and W.M. Stobbs: Phil. Mag., 1971, vol. 23, pp. 1201-33.

18 29. M.T. Ma, B.Z. Sun and Y. Tomota: ISIJ INT, 1989, vol. 29, pp. 74-77.

19 30. E.A.J. Starke: in: A. K. Vasudevan, R. D. Doherty (Eds.), Aluminum Alloys -

20 Contemporary Research and Aplications, Academic Press, San Diego, CA, 1989, pp.

21 35-63.

22 31. E. Orowan: Discussion on Internal Stresses, The Institute of Metals, London, 1948.

54
1 32. E. Orowan: in: G. M. Rassweiler, W. L. Grube (Eds.), Symposium on Internal Stresses

2 and Fatigue in Metals, Detroit, MI, 1958.

3 33. J.D. Eshelby: Proc. Royal. Soc. London A, 1957, vol. 241, pp. 376-96.

4 34. W.F. Hosford and R.H. Zeisloft: Metall. Trans. A, 1972, vol. 3, pp. 113-21.

5 35. P. van Liempt and J. Sietsma, Mater. Sci. Eng. A, 2016, vol. 662, pp. 80-87.

6 36. D. L. McDowell, Mater. Sci. Eng. R Reports, 2008, vol. 62, pp. 67-123.

7 37. Tiwari, G. J. Tucker, and D. L. McDowell, Philos. Mag., 2013, vol. 93, pp. 478-98.

8 38. M. Kato, T. Fujii, and S. Onaka, Mater. Trans., 2008, vol. 49, pp. 1278-83.

9 39. E.F. Rauch and J.H. Schmitt: Mater. Sci. Eng. A, 1989, vol. 113, pp. 441-48.

10 40. T. Hasegawa, T. Yako and U.F. Kocks: Mater. Sci. Eng., 1986, vol. 81, pp. 189-99.

11 41. Z. Hu, E.F. Rauch and C. Teodosiu: Int. J. Plasticity, 1992, vol. 8, pp. 839-56.

12 42. E.V. Nesterova, B. Bacroix and C. Teodosiu: Mater. Sci. Eng. A, 2001, vol. 309-310,

13 pp. 495-9.

14 43. J.H. Schmitt, E. Aernoudt and B. Baudelet: Mater Sci Eng, 1985, vol. 75, pp. 13-20.

15 44. E.F. Rauch: Solid State Phenom, 1992, vol. 23-24, pp. 317-448.

16 45. U.F. Kocks, T. Hasegawa and R.O. Scattergood: Scripta Metall., 1980, vol. 14, pp.

17 449-54.

18 46. H. Mughrabi: Acta Metall, 1983, vol. 31, pp. 1367-79.

19 47. D.V. Wilson and P.S. Bate: Acta Metall. Mater., 1994, vol. 42, pp. 1099-111.

20 48. Y. Strauven and E. Aernoudt: Acta Metall, 1987, vol. 35, pp. 1029-36.

21 49. J. A. Benito, J. Jorba, J. M. Manero, and A. Roca, Metall. Mater. Trans. A, 2005, vol.

22 36, pp. 3317-24.

55
1 50. B. Peeters, M. Seefeldt, C. Teodosiu and a. others: Acta Mater., 2001, vol. 49, pp.

2 1607-19.

3 51. B. Peeters, B. Bacroix, C. Teodosiu, P. Van Houtte and E. Aernould: Acta Mater.,

4 2001, vol. 49, pp. 1621-32.

5 52. S. Bouvier, J.L. Alves, M.C. Oliveira and L.F. Menezes: Comp Mater Sci, 2005, vol.

6 32, pp. 301-15.

7 53. A. Simar, Y. Bréchet, B. de Meester, A. Denquin, and T. Pardoen, Acta Mater., 2007,

8 vol. 55, pp. 6133-43.

9 54. G. Fribourg, Y. Bréchet, A. Deschamps, and A. Simar, Acta Mater., 2011, vol. 59, pp.

10 3621-35.

11 55. W. Z. Han, A. Vinogradov, and C. R. Hutchinson, Acta Mater., 2011, vol. 59, pp.

12 3720-36.

13 56. D. Bardel, M. Perez, D. Nelias, S. Dancette, P. Chaudet, and V. Massardier, Acta

14 Mater., 2015, vol. 83, pp. 256-68.

15 57. G. Vincze, E.F. Rauch, J.J. Gracio and e. al.: Acta Mater., 2005, vol. 53, pp. 1005-13.

16 58. F. Barlat and J. Liu: Mater. Sci. Eng. A, 1998, vol. pp. 47-61.

17 59. P.S. Bate, W.T. Roberts and D.V. Wilson: Acta Metall., 1981, vol. 29, pp. 1797-814.

18 60. J. da Costa Teixeira, L. Bourgeois, C. W. Sinclair, and C. R. Hutchinson, Acta Mater.

19 57, 6075 (2009).

20 61. T. Dorin, F. De Geuser, W. Lefebvre, C. Sigli, and A. Deschamps, Mater. Sci. Eng. A.,

21 2014, vol. 605, 119-26.

22 62. A. Deschamps, B. Decreus, F. De Geuser, T. Dorin, M. Weyland, Acta Mater., 2013,

23 vol. 61, pp. 4010-21.

56
1 63. H. Proudhon, W. J. Poole, X. Wang, and Y. Bréchet, Philos. Mag., 2008, vol. 88, pp.

2 621-40.

3 64. R.K. Boger, R.H. Wagoner, F. Barlat and W. Gan: Metall Mater Trans A, 2016, ( in

4 Preparation).

5 65. R.K. Boger: Ph. D. thesis, Mater. Sci. Eng., The Ohio State University, Columbus,

6 Ohio 43210, 2005.

7 66. Y. Chen, M. Weyland, and C. R. Hutchinson, Acta Mater., 2013, vol. 61, pp. 5877-94.

8 67. A. Deschamps, D. Dumont, Y. Brechet and e. al.: in: M. Tiryakioglu (Ed.), Proceeding

9 of the James T. Stanley honorary symposium on aluminum alloys, Indianapolis, USA,

10 2001.

11 68. E. Hornbogen, A.K. Mukhopadhyay and E.A. Starke: Z Metallkd, 1992, vol. 83, pp.

12 577-84.

13 69. E. Hornbogen, A.K. Mukhopadhyay and E.A. Starke: Scripta Metall Mater, 1992, vol.

14 27, pp. 733-38.

15 70. V. Radmilovic, D. Mitlin, A.J. Tolley and e. al.: Metall Mater Trans A, 2003, vol. 34,

16 pp. 543-51.

17 71. F. Barlat, Materials Science Division, Alcoa Technical Center, 100 Technical Drive,

18 Alcoa Center, PA 15069, USA, 2005.

19 72. E.E. Underwood and E.A. Starke: in, Fatigue Mechanisms, Proceedings of an

20 ASTM-NBS-NSF symposium, ASTM STP 675, Kansas City, MO, 1978, pp. 633-82.

21 73. S.M. Allen: Philos Mag, 1981, vol. 43, pp. 325-35.

22 74. P.M. Kelly, A. Jostsons, R.G. Blake and J.G. Napier: Phys Status Solidi (A) Appl Res.,

23 1975, vol. 31, pp. 771-80.

57
1 75. J. W. Martin ‘Precipitation hardening’, 2nd edn. 58-59,18: 1998, Oxford,

2 Butterworth-Heinemann.

3 76. ASTM Standard E8 / E8M - 08: ASTM International, West Conshohocken, PA, DOI:

4 10.1520/E0008_E0008M-08.

5 77. R.K. Boger, R.H. Wagoner and e. al.: Int. J. Plast., 2005, vol. 21, pp. 2319-43.

6 78. W. Gan, K. Okamoto, S. Hirano, K. Chung, C.M. Kim and R.H. Wagoner: J Eng

7 Mater-T ASME, 2008, vol. 130, pp. 031007-1-15.

8 79. V. Balakrishnan: MS Thesis, Department of Materials Science and Engineering, The

9 Ohio State University, Columbus, Ohio, 1999.

10 80. Kun Piao, J. K. Lee, J. H. Kim, H. Y. Kim, K. Chung, F. Barlat, R. H. Wagoner, Int. J.

11 Plast., 2012, vol. 38, pp. 27-46.

12 81, K. Piao, K. Chung, M. G. Lee, R. H. Wagoner, Metall. Mater. Trans. A, 2012, vol.

13 43A, pp. 3300-13

14 82. Laser extensometer, model LE-01, Electronic Instrument Research, PO Box 678 o

15 Irwin, PA 15642, 2005.

16 83. A. Abel: Mater. Forum, 1987, vol. 10, pp. 11-26.

17 84. A. Abel and H. Muir: Phil. Mag., 1972, vol. 26, pp. 489-504.

18 85. X.Y. Lou: MS thesis, The Materials Science and Engineering, The Ohio State

19 University, Columbus, Ohio 43210, 2005.

20 86. J.L. Murray and A.J. McAlister: J Phase Equilib, 1984, vol. 5, pp. 74-84.

21 87. S.I. Fujikawa and Y.I. Izeki: Metall Trans A, 1993, vol. 24, pp. 277-82.

22 88. H.M. Kagaya, K. Imazawa, M. Sato and T. Soma: Physica B, 1998, vol. 245, pp.

23 252-55.

58
1 89. J.W. Martin: Precipitation Hardening, 2nd, Butterworth-Heinemann, 1998.

2 90. T. Zumkley and H. Mehrer: Z Metallkd, 1998, vol. 89, pp. 454-63.

3 91. N.L. Peterson and S.J. Rothman: Phys Rev B, 1970, vol. 1, pp. 3264-&.

4 92. D. Simonovic and M.H.F. Sluiter: Phys Rev B, 2009, vol. 79, pp. -.

5 93. N. Hansen: Acta Metall, 1977, vol. 25, pp. 863-69.

6 94. A. Deschamps, D. Dumont, Y. Brechet and e. al.: in: M. Tiryakioglu (Ed.), Proceeding

7 of the James T. Stanley honorary symposium on aluminum alloys, Indianapolis, USA,

8 2001.

9 95. L.M. Cheng, W.J. Poole, J.D. Embury and D.J. Lloyd: Metall Mater Trans A, 2003,

10 vol. 34, pp. 2473-81.

11 96. U.F. Kocks and H. Mecking: Acta Metall., 2003, vol. 29, pp. 1865-75.

12 97. P. Haasen: in: R. W. Cahn, P. Haasen (Eds.), Physical Metallurgy, North-Holland,

13 1996.

14 98. M.Z. Butt and P. Feltham: J Mater Sci, 1993, vol. 28, pp. 2557-76.

15 99. U.F. Kocks: Metall Mater Trans A, 1985, vol. 16, pp. 2109-29.

16 100. M.Z. Butt, I.M. Ghauri, R. Qamar, K.M. Hashmi and P. Feltham: Acta Metall., 1981,

17 vol. 29, pp. 829-34.

18 101. R. Labusch: Acta Metall., 1972, vol. 20, pp. 917.

19 102. F.R.N. Nabarro: Phil. Mag., 1977, vol. 35, pp. 613.

20 103. P. Kratochvil, P. Lukac and B. Sprusil: Czech. J. Phys. B, 1673, vol. 23, pp. 621.

21 104. R.L. Fleischer: Acta Metall., 1963, vol. 11, pp. 203-09.

22 105. T.H. Courtney: Mechanical behavior of materials, McGraw-Hill, Boston, 2000.

23 106. T.H. Wille, W. Gieseke and C. Schwink: Acta Metall., 1987, vol. 35, pp. 2679-93.

59
1 107. M.G. Stout and A.D. Rollett: Metall. Trans. A, 1990, vol. 21, pp. 3201-13.

2 108. W. Gan, P. Zhang, R.H. Wagoner and G.S. Daehn: Metall Mater Trans A, 2006, vol.

3 37A, pp. 2097-106.

4 109. X.Y. Hu, W. Chao, H. Margolin and e. al.: Scripta Metall. Mater., 1992, vol. 27, pp.

5 865-70.

6 110. N. Hansen, Acta Metall., 1977, vol. 25, pp. 863-69.

7 111. T. H. Courtney, Mechanical Behavior of Materials: Second Edition (Waveland Press,

8 2005).

9 112. S. R. Kalidindi, Ph. D. thesis, Massachusetts Institute of Technology, 1992.

10 113.P. R. Dawson, Int. J. Solids Struct., 2000, vol. 37, 115-30.

11 114.D. Raabe, P. Klose, B. Engl, K.-P. Imlau, F. Friedel, and F. Roters, Adv. Eng. Mater.,

12 2002, vol. 4, pp.169-80.

13 115.A. Ma, F. Roters, and D. Raabe, Acta Mater., 2006, vol. 54, pp. 2169-79.

14 116.D. Raabe, D. Ma, and F. Roters, Acta Mater., 2007, vol. 55, pp. 4567-83.

15 117. M. G. Lee, H. Lim, B. L. Adams, J. P. Hirth, and R. H. Wagoner, Int. J. Plast., 2010,

16 vol. 26, pp. 925-38.

17 118. H. Lim, M. G. Lee, J. H. Kim, B. L. Adams, and R. H. Wagoner, Int. J. Plast., 2011,

18 vol. 27, pp. 1328-54.

19 119. H. Lim, S. Subedi, D. T. Fullwood, B. L. Adams, and R. H. Wagoner, Mater. Trans.,

20 2014, vol. 55, pp. 35-38.

21 120. H. J. Bong, H. Lim, M.-G. Lee, D. Fullwood, E. Homer, and R. H. Wagoner, Comput.

22 Mater. Sci., vol. pp. in Preparation

23 121. AJ. J. R. Rice, J. Mech. Phys. Solids, 1971, vol. 19, pp. 433-55.

60
1 122. D. Peirce, R. J. Asaro, and A. Needleman, Acta Metall., 1982, vol. 30, pp.

2 1087-1119.

3 123. AK. R. J. Asaro, Advances in Applied Mechanics, 1983, vol. 23, pp. 1-115.

4 124. J. Gubicza, N. Q. Chinh, J. L. Lábár, S. Dobatkin, Z. Hegedűs, and T. G. Langdon, J.

5 Alloys Compd., 2009, vol. 483, pp. 271-74.

6 125. G. Schoeck and R. Frydman, Phys. Status Solidi (b), 1972, vol. 53, pp. 661-73.

7 126. U. F. Kocks, J. Eng. Mater. Technol. Trans. ASME 98, 1976, vol. 98, p. 76.

8 127. E. Orowan, Proc. Phys. Soc., 1940, vol. 52, pp. 8-22.

9 128. Hirth, J.P., Lothe, J., 1969. Theory of Dislocations. McGraw-Hill, 1969, New York.

10 129. Z. Shen, R. H. Wagoner, and W. A. T. Clark, Scr. Metall.. 1986, vol. 20, pp. 921-26.

11 130. R. Bagheriasl, D. R. Tari, S. Kurukuri, and M. J. Worswick, in Compr. Mater.

12 Process., edited by S. Hashmi, 1st Editio (Elsevier, 2014), p. 235.

13 131. W. Hosford, R. Fleischer, and W. Backofen, Acta Metall., 1960, vol. 8, pp. 187-99.

14 132. A. Turnbull and E. R. de los Rios, Fatigue Fract. Eng. Mater. Struct., 1995, vol. 18,

15 pp. 1343-54.

16

61
1 Figure Caption

3 Figure 1. Tensile properties and Vickers hardness (HV) correlations for various

4 aluminum alloys: (a) σ y vs. HV, (b) σ UTS vs. HV, (c) 0.5( σ y + σ UTS ) vs. HV, and (d)

5 σ UTS / σ y vs. HV.

6 Figure 2. Schematics illustrating the definition of Bauschinger factors, using (a) CT

7 stress-strain plots and (b) Δσ-Δε plots.

8 Figure 3. TEM micrographs of Al-0.2pct Ge-Si, aged at 473 K (200 °C) for 8 hours

9 (2.9×104 s): (a) lower magnification, and (b) higher magnification.

10 Figure 4. Aging curves of Al-Ge-Si alloy at three temperatures.

11 Figure 5. TEM micrograph of Al-Ge-Si, aged at 513 K (240 °C) for 2 hours (7.2×103

12 s)

13 Figure 6. Microstructure of Al-Ge-Si, as-received: (a) lower magnification, and (b)

14 higher magnification.

15 Figure 7. TEM micrographs of Al-Ge-Si, aged for one week (6.0×105 s) at 393 K

16 (120 °C): (a) grain junction, (b) grain boundary area, (c) grain interior, and (d) fine

17 precipitates inside a grain.

18 Figure 8. TEM micrographs of Al-Ge-Si, aged at 473 K (200 0C) for: (a) 0.5 hours

19 (3.6×103 s) (UA), (b) 12 hours (4.3×104 s) (PA), and (c) one week (6.0×105 s) (OA)

20 Figure 9. Precipitate size distribution of Al-Ge-Si aged at 473 K (200 °C). The

21 average diameter (d) and standard deviation are included.

22 Figure 10. Aging curves for Al-Ge-Si with quantitative precipitate measurement results.

62
1 Figure 11. Evolution of precipitate characteristics: (a) the master aging curve, (b) size,

2 (c) spacing, and (d) volume fraction.

3 Figure 12. Tensile curves of Al-Ge-Si aged at 473 K (200 °C), as compared to pure Al.

4 Figure 13. The evolution of work hardening rate of Al-Ge-Si aged at 473 K (200 °C),

5 as compared to pure aluminum: (a) dσ/dε vs. ε and (b) dσ/dε vs. σ-σy.

6 Figure 14. The yield stresses of the solution heat treated Al-Ge-Si with respect to the

7 total alloy concentrations.

8 Figure 15. Measured “Bauschinger Factors” of pure aluminum and Al-Ge-Si materials

9 at SHT, UA, PA and OA conditions (a) Bauschinger Stress Factor, β (b) Bauschinger

10 Strain Factor, γ and (c) Bauschinger Energy Factor, ΔE.

11 Figure 16. Reduced CT hardening curves for various aging conditions: (a) Al-Ge-Si,

12 and (b) Al 2524 [83]

13 Figure 17. RVE FE model of Al-Ge-Si alloy for (a) UA/PA/OA conditions with Al

14 matrix and embedded hard particles(assumed particle size, volume fraction and spacings

15 are also listed in the table below the images; f=volume fraction, λ=spacing) (b) Pure

16 Al/SHT condition with 4 grains and corresponding grain boundaries.

17 Figure 18. Comparison of CT transient hardening, experimental and simulated (3

18 models): (a) stress-strain, and (b) Reduced Δσ vs. Δε form.

19 Figure 19. Comparison of Bauschinger factors from experiments and simulation as a

20 function of aging time: (a) β0.4 (b) γ2k and (c) ΔE0 (CP and Continuum results are not

21 shown: they like on the abscissa with ΔE=0.00-0.03. See Table V)

22 Figure A1. Comparison of simulated and experimental CT tests. Three constitutive

23 models compared

24
63
1 Tables
2

3 Table I. Atomic compositions of three Al-Ge-Si alloys (unit in pct)


Al Ge Si Fe Cu Zn Ti
Al-1pct Ge-Si Balance 0.49 0.50 0.03 0.01 0.03 0.02
Al-0.4pct Ge-Si Balance 0.19 0.16 0.03 0.00 0.01 0.01
Al-0.2pct Ge-Si Balance 0.07 0.16 0.02 0.00 0.01 0.01
4
5 Table II. Atomic, weight and volume percentages of the Ge and Si elements in three
6 Al-Ge-Si alloys
Atomic Percent Weight Percent Volume Percent
Ge Si Ge+Si Ge Si Ge+Si Ge Si Ge+Si
Al-1pct Ge-Si 0.49 0.50 0.99 1.30 0.51 1.81 0.66 0.59 1.25
Al-0.4pct
Ge-Si 0.19 0.16 0.36 0.51 0.17 0.68 0.26 0.20 0.46
Al-0.2pct
Ge-Si 0.07 0.16 0.24 0.19 0.17 0.36 0.10 0.20 0.29
7
8 Table III. Measured and predicted yield stresses of aged Al-Ge-Si
σmeas (MPa) σpred (MPa) σmeas − σpred
(MPa)
SHT 46.9 / /
UA 93.5 93.5 (fit) 0
PA 114.6 127.0 -12.4
OA 96.4 95.6 0.8
9
10 Table IV Bauschinger variables for aged Al-Ge-Si materials
d (nm) λ (nm) f (pct) β0.2 β0.4 γ2x ΔE0
SHT - - - 0.21 0.06 0.008 -
UA 9.4 390.5 0.2 0.25 0.15 0.017 0.54
PA 15.7 227.4 1.2 0.34 0.21 0.026 0.90
OA 25.9 371.7 1.2 0.34 0.21 0.023 0.82
11
13
12 Table V. Predicted Bauschinger variables for aged Al-Ge-Si materials
Condition
Al SHT UA PA OA
Experiment 0.05 0.06 0.15 0.21 0.21
β0.4 SD 0 0.03 0.09 0.23 0.17
Standard CP 0 0 0 0 0

13
Some simulations produce small negative values of β indicating no Bauschinger effect. These values
are shown simply as “0” (zero) in Table V.

64
Continuum 0 0 0 0 0
Experiment - 0.008 0.017 0.026 0.023
SD 0.004 0.007 0.012 0.019 0.018
γ2x Standard CP 0.003 0.003 0.004 0.005 0.004
Continuum 0.002 0.003 0.004 0.004 0.004
Experiment - - 0.54 0.90 0.82
SD 0.05 0.07 0.29 0.64 0.39
ΔE0 (MPa) Standard CP 0.02 0.01 0.03 0.01 0.03
Continuum 0.00 0.03 0.01 0.00 0.00

1
2 Table AI. Constitutive parameters for the three constitutive models
Constitutive kaAl kbAl ρ0Al g0Al τ*
Ge-Si Ge-Si Ge-Si 5
Model (=ka ) (=kb ) (=ρ0 /10 ) (=g0Ge-Si) (MPa)
(m) (/m2) (MPa)
12
Al Standard CP 26 110b 4.7×10 0 -
SD 50 100b 1.0×1012 0 35
13
SHT Standard CP 15 141b 1.5×10 16 -
SD 50 100b 1.0×1012 16 235
12
UA Standard CP 15 153b 4.0×10 41 -
12
SD 50 100b 1.0×10 41 168
PA Standard CP 12 120b 2.5×1013 46 -
12
SD (Approach 1) 50 100b 1.0×10 46 573
SD (Approach 2) 60 500b 1.0×1012 46 573
13
OA Standard CP 10 150b 1.0×10 41 -
SD 50 100b 1.0×1012 41 482
3

65
Tables

Table I. Atomic compositions of three Al-Ge-Si alloys (unit in pct)


Al Ge Si Fe Cu Zn Ti

Al-1pct Ge-Si Balance 0.49 0.50 0.03 0.01 0.03 0.02

Al-0.4pct Ge-Si Balance 0.19 0.16 0.03 0.00 0.01 0.01

Al-0.2pct Ge-Si Balance 0.07 0.16 0.02 0.00 0.01 0.01

Table II Atomic, weight and volume percentages of the Ge and Si elements in three Al-Ge-Si
alloys
Atomic Percent Weight Percent Volume Percent

Ge Si Ge+Si Ge Si Ge+Si Ge Si Ge+Si

Al-1pct Ge-Si 0.49 0.50 0.99 1.30 0.51 1.81 0.66 0.59 1.25

Al-0.4pct Ge-Si 0.19 0.16 0.36 0.51 0.17 0.68 0.26 0.20 0.46

Al-0.2pct Ge-Si 0.07 0.16 0.24 0.19 0.17 0.36 0.10 0.20 0.29

Table III. Measured and predicted yield stresses of aged Al-Ge-Si


σmeas (MPa) σpred (MPa) σmeas-σpred (MPa)

SHT 46.9 / /

UA 93.5 93.5 (fit) 0

PA 114.6 127.0 -12.4

OA 96.4 95.6 0.8

Table IV Bauschinger variables for aged Al-Ge-Si materials


d (nm) λ (nm) f (pct) β0.2 β0.4 γ2x ΔE0

SHT - - - 0.21 0.06 0.008 -

UA 9.4 390.5 0.2 0.25 0.15 0.017 0.54

PA 15.7 227.4 1.2 0.34 0.21 0.026 0.90

OA 25.9 371.7 1.2 0.34 0.21 0.023 0.82


Table V Predicted Bauschinger variables for aged Al-Ge-Si materials
Condition

Al SHT UA PA OA

Experiment 0.05 0.06 0.15 0.21 0.21

SD 0 0.03 0.09 0.23 0.17

β0.4 Standard CP 0 0 0 0 0

Continuum 0 0 0 0 0

Experiment - 0.008 0.017 0.026 0.023

SD 0.004 0.007 0.012 0.019 0.018

γ2x Standard CP 0.003 0.003 0.004 0.005 0.004

Continuum 0.002 0.003 0.004 0.004 0.004

Experiment - - 0.54 0.90 0.82

SD 0.05 0.07 0.29 0.64 0.39

ΔE0 (MPa) Standard CP 0.02 0.01 0.03 0.01 0.03

Continuum 0.00 0.03 0.01 0.00 0.00

Table AI. Constitutive parameters for the three constitutive models


Constitutive kaAl kbAl ρ0Al g0Al τ*
Model (=kaGe-Si) (=kbGe-Si) (=ρ0Ge-Si/105) (=g0Ge-Si) (MPa)
(m)
(/m2) (MPa)

Al Standard CP 26 110b 4.7×1012 0 -

SD 50 100b 1.0×1012 0 35

SHT Standard CP 15 141b 1.5×1013 16 -

SD 50 100b 1.0×1012 16 235

UA Standard CP 15 153b 4.0×1012 41 -

SD 50 100b 1.0×1012 41 168

PA Standard CP 12 120b 2.5×1013 46 -


SD (Approach 1) 50 100b 1.0×1012 46 573

SD (Approach 2) 60 500b 1.0×1012 46 573

OA Standard CP 10 150b 1.0×1013 41 -

SD 50 100b 1.0×1012 41 482

S-ar putea să vă placă și