twinning-slip transitions in mg az31bli.mit.edu/stuff/rhw/upload/31.pdf · 2011. 9. 15. · 1...

56
1 Twinning-Slip Transitions in Mg AZ31B K. Piao 1 , K. Chung 2 , M. G. Lee 3 , R. H. Wagoner 1 * 1 Department of Materials Science and Engineering, 2041 College Road Ohio State University, Columbus, OH 43210, USA 2 Department of Materials Science and Engineering, Research Institute of Advanced Materials, Seoul National University, 599 Gwanak-ro, Gwanak-gu Seoul 151-742, Republic of Korea 3 Graduate Institute of Ferrous Technology, Pohang University of Science and Technology, Pohang 790-784, Republic of Korea Many metals, particularly ones with HCP crystal structures, undergo deformation by combinations of twinning and slip, the proportion of which depends on variables such as temperature and strain rate. Typical techniques to reveal such mechanisms rely on metallography, x-ray diffraction, or electron optics. Simpler, faster, less expensive mechanical tests were developed in the current work and applied to Mg AZ31B. The curvature of compressive stress-strain plots over a fixed strain range was found to be a consistent indicator of twinning magnitude, independent of temperature and strain rate. The relationship between curvature and areal fraction of twins was determined. Transition temperatures determined based on stress-strain curvature were consistent with ones determined by metallographic analysis and flow stresses, and depended on strain rate by the Zener-Hollomon parameter, a critical value for which was measured. The transition temperature was found to depend significantly on grain size, a relationship for which was established. Finally, it was shown that the transition temperature can be determined consistently, and much faster, using a single novel “Step-Temperature” test.

Upload: others

Post on 02-Feb-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

  • 1

    Twinning-Slip Transitions in Mg AZ31B

    K. Piao1, K. Chung2, M. G. Lee3, R. H. Wagoner1*

    1Department of Materials Science and Engineering, 2041 College Road Ohio State University, Columbus, OH 43210, USA

    2Department of Materials Science and Engineering, Research Institute of Advanced Materials, Seoul National University, 599 Gwanak-ro, Gwanak-gu Seoul 151-742,

    Republic of Korea

    3Graduate Institute of Ferrous Technology, Pohang University of Science and Technology, Pohang 790-784, Republic of Korea

    Many metals, particularly ones with HCP crystal structures, undergo deformation by

    combinations of twinning and slip, the proportion of which depends on variables such as

    temperature and strain rate. Typical techniques to reveal such mechanisms rely on

    metallography, x-ray diffraction, or electron optics. Simpler, faster, less expensive

    mechanical tests were developed in the current work and applied to Mg AZ31B. The

    curvature of compressive stress-strain plots over a fixed strain range was found to be a

    consistent indicator of twinning magnitude, independent of temperature and strain rate.

    The relationship between curvature and areal fraction of twins was determined.

    Transition temperatures determined based on stress-strain curvature were consistent with

    ones determined by metallographic analysis and flow stresses, and depended on strain

    rate by the Zener-Hollomon parameter, a critical value for which was measured. The

    transition temperature was found to depend significantly on grain size, a relationship for

    which was established. Finally, it was shown that the transition temperature can be

    determined consistently, and much faster, using a single novel “Step-Temperature” test.

  • 2

    Keywords: Deformation mechanisms; Magnesium AZ31 alloy sheet; Deformation twin;

    Dislocation slip; Transition temperature, Mechanical testing, Compression test, Cyclic

    test, Grain size, Strain rate.

    *Corresponding Author: R. H. Wagoner; Tel.: +1-614-292-2079, Fax: +1-614-292-6530,

    E-mail address: [email protected]

    Manuscript date: August 30th, 2011

    Submitted to Metallurgical and Materials Transactions A

  • 3

    1. INTRODUCTION

    As the lightest structural metal[1], magnesium with low density (1.74 g/cm3) and

    moderate ductility [2] is attractive to the automobile industry, among others, because of

    its light weight, high tensile strength (vs. Al) [1, 3] and high fatigue resistance (vs. Al) [2,

    4, 5]. Most magnesium products are made by die casting [6] because of its poor

    formability at room temperature, which is attributed to the low symmetry of the

    hexagonal close packed (HCP) crystalline structure and thus the limited slip systems [3].

    Magnesium wrought alloys have strong plastic anisotropy, especially in the strongly

    textured sheet alloys [2]. At room temperature, the yield stress in compression in the

    sheet plane is approximately half of that in tension [7, 8]; and the stress-strain curve in

    compression shows an unusual inflected shape, distinct from the normal concave-down

    aspect in tension. These phenomena are attributed to 1) the strong texture with the basal

    plane parallel to the plane of sheet resulting from the manufacture process (c-axis

    perpendicular to the sheet plane); and 2) the activation of { }2110 twinning in in-plane compression (with c-axis elongated) [9]. Un-twinning (or “detwinning”) also occurs in

    the subsequent reverse loading after compression and shows a similarly inflected stress-

    strain curve [8, 10].

    According to the Von Mises criterion [11, 12], an arbitrary homogeneous deformation of

    polycrystalline metals requires five independent slip systems [13]. In magnesium alloys,

    basal slip, prismatic slip and pyramidal slip provide four independent slip

    systems, facilitating deformation only in the basal plane. At room temperature,

  • 4

    { }2110 twinning, as the most common twinning mode in magnesium, can provide an independent deformation mechanism (in the c-direction) to satisfy the Von Mises

    criterion [14].

    Pyramidal slip, with high critical resolved shear stress (CRSS, ~90MPa), is

    difficult to activate at room temperature in magnesium [15, 16]. Twinning has a polar

    nature, which means that the { }2110 twin in magnesium can only be activated when the c-axis is elongated [17, 18]. The maximum strain along the c-axis is 0.064 [14, 19], with

    the basal planes in the twinned region rotated 86.3o about the intersection line with the

    twinning plane [20].

    At elevated temperatures, as the only two systems that provide for deformation in the c-

    axis direction, twinning and pyramidal slip are in competition with each other. The

    critical resolved shear stress (CRSS) for ac + slip decreases to ~35MPa at 200oC and

    to ~25MPa at 300oC [21-23], whereas that for twinning is expected to be relatively

    temperature insensitive [24]. Thus, as the testing temperature is increased, the difference

    between the activation stresses of these two mechanisms is reduced until a critical

    temperature, Tt, is reached where the two stresses are equal [25]. At temperatures higher

    than this, slip dominates the flow. Similar behavior has been identified for other twinning

    metals [24, 26]

    The activation of deformation twinning is correlated to the shape of stress-strain curves.

    For example, the slope of the normalized strain hardening rate ((dσ/dε)/G with respect to

  • 5

    deformation strain) changed from negative to positive, in various metals: brasses [27-29],

    stainless steels [27], Co-Ni alloys [27-30], Ti [31-34], Zr [35], and Mg [36]. The

    curvature of the stress-strain curve has been proposed to represent differences between

    slip-dominant flow (d2σ/dε20) [37], and was used

    to discern the grain-size effect in Mg alloys [38].

    Deformation twinning has been more directly revealed by various experimental methods:

    optical metallurgraphy [8, 39, 40], in-situ acoustic emission to detect twin activation [8,

    41], X-ray diffraction to measure texture changes [8, 42, 43], EBSD for similar results

    [44-48], and in-situ neutron diffraction to measure residual stress and local relaxed strain

    [41, 49-52]. Each of these methods has specific disadvantages for testing transition

    temperatures because of special equipment, specimen preparation, numerous samples,

    and high times and costs.

    The polar nature of { }2110 twinning in rolled, basal-textured Mg AZ31B sheet requires large in-plane compressive strains (corresponding to large through-thickness extensile

    strains) in order to reveal the shape of the stress-strain curve under conditions where

    twinning is possible. Large compressive strains are not readily obtained in sheet alloys

    because of buckling, but several schemes for avoiding buckling at room temperature have

    been introduced [53-56]. At least one of these methods [53] has been shown to permit

    continuous compressive and tensile testing, including cycling between the two, thus

    permitting the development of detailed constitutive equations [57-60], and the calibration

    of acoustic emission signatures [8], and anelasticity results after strain reversal [53]. In a

  • 6

    recent development by the co-authors, the Boger technique has been extended [61, 62] to

    allow elevated-temperature tension/compression testing, potentially allowing the

    possibility of determining transition temperatures mechanically.

    A schematic of the new device and specimen are presented in Figure 1, and will be

    described in more detail in the Experimental Procedures of this paper. Tension-

    Compression-Tension (T-C-T) tests were carried out from room temperature to 250oC at

    a strain rate of 10-3/s [62], with results summarized as shown in Figures 2. Figure 2(a)

    shows the stress-strain curve; and Figures 2(b)-(c) are the enlarged portions for

    compressive and tensile legs at 125oC and 150oC, revealing the distinctive features of

    slip-dominant (always concave-down) and twin-dominant (concave-down transitioning to

    concave-up) curves. A transition temperature Tt between these two deformation

    mechanisms (or at least between curve shapes) is clearly seen within the temperature

    range of 125oC to 150oC. This range of values of Tt was preliminarily verified by

    quantitative metallography showing that the areal fraction of twins after compression

    were 4.5% at 150oC and 43% at 125oC [62].

    It should also be noted that other evidence for twinning during the first compressive leg

    at temperatures below 150oC can be seen in the jerky flow during twinning [8] and the

    almost-athermal aspect of compressive flow stress with temperatures associated with

    twinning domination, particularly as compared with compressive curves at higher, slip-

    dominated temperatures. Finally, it is worth noting that these distinguishing aspects

  • 7

    largely disappear in the final tension cyclic, presumably because of the much lower

    nucleation stress of untwining [8].

    The new elevated-temperature, tension-compression test introduces the possibility of

    revealing the transitions among deformation mechanisms (twinning vs. slip)

    mechanically. The hypothesis underlying the approach is that the curvature of the stress-

    strain curve is related in an identifiable, consistent way to the predominant slip

    mechanism. In order to test this hypothesis and the mechanically-based approach that is

    suggested, a series of compression tests were conducted over a range of temperatures and

    strain rates, then analyzed both mechanically and metallographically. An even

    faster/simpler/cheaper “Step-Temperature” (“ST”) test was developed subsequently and

    results from it compared with the monotonic compression tests to evaluate its potential

    utility.

    2. EXPERIMENTAL PROCEDURES

    Mg AZ31B sheet alloy was mechanically tested at a temperature range of 110oC-250oC

    and strain rates of 0.0005/s-0.1/s using monotonic compression [62] and a novel “ST”

    test. The deformation mechanisms were revealed by standard optical metallography.

    2.1 Materials

    Six batches of Mg AZ31B alloy sheet from various suppliers were used. The chemical

    compositions, thicknesses, and grain sizes are shown in Table 1. The additions of Al and

  • 8

    Zn to Mg act as solid-solution strengthening agents [1]. Alloy AU was the primary

    material for the work in this paper; the other alloys were used primarily to estimate the

    role of grain size on deformation mechanism.

    2.2 Mechanical tests

    A test device, presented elsewhere in detail [61, 62] and shown schematically in Figure 1,

    uses side loading through heated plates to stabilize in-plane compressive deformation of

    an exaggerated dog-bone shape specimen. For all the tests presented here, a constant side

    force of 2.5kN was used and was found sufficient to prevent buckling. An automatic

    feedback system controls the temperature of the specimen in range of 100oC-250oC, with

    a temperature fluctuation with time within 1oC and a temperature difference throughout

    the gage length (38mm) less than 5oC at 250oC [62]. Heating times from room

    temperature range from 2 to 5 minutes depending on the target temperature. Tests were

    carried out using an MTS 810 testing machine with a 200kN load cell.

    For monotonic compression testing, a maximum strain of -0.08 was obtained at strain

    rates of 0.001/s, 0.01/s, and 0.1/s. Following deformation, samples were quenched into

    water. A non-contact EIRTM LE-05 laser extensometer was used to measure the specimen

    deformation over a gage length of 25.4mm. The friction and biaxial effects introduced by

    the side force were corrected for after testing [8, 53, 63]. To obtain a uniaxial-equivalent

    stress without friction, friction coefficients of 0.03 and 0.08 were determined by a series

    of tensile and compressive tests, respectively, with various side forces at room

    temperature using procedures described elsewhere [8, 53, 63].

  • 9

    A novel Step-Temperature test (“ST test”) was conceived of as a way to efficiently

    determine the transition temperature, Tt, between deformation twin and dislocation slip

    using a single sample. The concept started with a tension-compression cycle of

    deformation at a temperature T where T>Tt. The temperature would then be reduced by

    some decrement, say -10oC, and the deformation cycle repeated at progressively lower

    temperatures until the stress-strain curves indicated a twinning-dominated mechanism.

    The underlying assumption was that cyclic deformation by slip (i.e. at T>Tt) would not

    change the microstructure sufficiently to alter Tt.

    For the implementation attempted here, the cyclic deformation range was tension (to a

    strain of 0.04)-compression (to a strain of -0.04) for 5 cycles, each at a constant strain rate

    with the temperature for each compression-tension cycle decreased by 10oC from the

    previous cycle. Each cycle was preceded by a hold time to reach the new target

    temperature, typically 40s. The total testing strain range of 0.08 (absolute accumulated)

    was accomplished in strain control mode, with holding between deformation cycles in

    stroke control. A total strain range of 0.08 was chosen to be large enough to reveal stress-

    strain patterns characteristic of deformation twinning and to allow 5-cycle tests without

    fracture. Instead of standard true strain, the absolute value of each strain increment is

    summed to produce an accumulated absolute strain εΔ=ε 2 , where Δ ε is the true

    strain increment over a monotonic interval. Strain rates of 0.0005, 0.001, 0.005, 0.01,

    0.05, and 0.1/s were used. A few examples of an alternate form of ST test with

  • 10

    temperature increased for each cycle were done for comparison. The analysis of

    mechanical tests is presented in the “Results and Analysis” section of this paper.

    2.3 Metallography measurement

    Samples for optical metallography were cut from the gauge regions where the material

    was deformed and heated uniformly. Specimens were mounted using epoxy, then

    successively ground and polished, finishing with 1 μm diamond paste. Acetic picral

    solution (4.2 g picric acid, 10 ml acetic acid, 70 ml ethanol and 10 ml water) was used to

    etch for 5-10 seconds. Specimens were examined immediately after etching to avoid

    oxidation.

    In order to assess the scatter and reproducibility of twin fractions and grain sizes, analysis

    was conducted for 5 observation areas of 130μm × 108μm from a single specimen

    compressively deformed to a strain of -0.08 at 150oC and at 0.01/s. The areal fraction of

    deformation twins was determined by a point counting method according to ASTM

    standard E562-08 [64]. 320 points were uniformly distributed over each microstructural

    image. The grain size was calculated using a linear intercept method following [65]. At

    least 200 grains were imaged in each field of view. The standard deviations for the five

    images were 0.04 for areal fraction of twin, i.e., 0.266 ± 0.04, and 0.3μm for grain size,

    i.e., 4.8μm ± 0.3μm.

    3. RESULTS AND ANALYSIS

  • 11

    Monotonic compression tests at various temperatures and strain rates for the AU material

    (Figure 3) were first used to ascertain the estimated transition temperatures based on

    stress-strain curvature. These results were then compared with measured twin areal

    fractions to check for consistency. With that correlation established, the faster ST test

    was carried out to compare its ability to reproduce distinctions among dominant

    deformation mechanisms. Finally, the fast ST test was used to correlate the mechanical

    behaviors of 6 Mg31AZ alloys in terms of transition temperature and grain size.

    3.1 Monotonic compression testing and analysis

    Isothermal monotonic compression tests were carried out at strain rates of 10-1/s, 10-2/s,

    and 10-3/s, Figures 3 (a)-(c), respectively. Results at lower temperatures exhibit an

    inflected aspect (concave up) associated with twinning while higher temperatures exhibit

    the “normal” concave down curvature typical of slipping-only materials. To enable

    quantification of the shape of the stress-strain curve, a quadratic equation

    ( CBA 2 +ε+ε=σ ) was fit to the data for the effective strain range 0.03-0.08. The sign of

    the first coefficient, A, is equal to the second derivative, one measure of curvature:

    A=(d2σ/dε2). A positive value of A over this strain range corresponds to a concave-up

    (twinning-dominant) appearance while a negative value of A over this strain range

    corresponds to a concave down (slip dominant) appearance.

    A simple examination of Figures 3 reinforces the impressions gained from the tension-

    compression test results in Figures 2. The shape (curvature sign and magnitude) of

    compressive stress-strain curves over the strain range 0.03-0.08 are affected markedly by

    the temperature and strain rate of the deformation. Based on the hypothesis that the value

  • 12

    of A is directly related to the ratio of twinning strain to dislocation slipping strain, Figures

    3 show that higher temperatures favor slip while higher strain rates favor twinning. The

    hypothesized transition temperatures (from visual inspection of the curves) are presented

    on each plot. They are significantly affected by strain rate, with an increase of 65oC

    corresponding to a strain rate change of 2 orders of magnitude.

    Figure 4 quantifies the visual interpretation of Figure 3 using the least-square values of A

    (curvature). The lines shown in the figure are themselves best-fit quadratic

    representations of A vs. test temperature, and the intersection with the A=0 axis

    represents a hypothetical dividing line between twin dominated deformation (A greater

    than 0) and slip dominated deformation (A less than 0). The condition A=0 represents an

    estimate, based on mechanical behavior, of the transition temperature, Tt, between the

    two regimes. As shown in the figure, the values of Tt obtained in this way are 138oC,

    165oC, and 204oC, respectively, consistent with the visual estimates from Figure 3 of

    145oC, 170oC, and 210oC.

    In order to assess whether the mechanically-indicated transition temperatures correlate

    with other indications, optical metallography was carried out to measure the areal fraction

    of twins after deformation. Examples of such micrographs are shown in Figures 5 for

    temperatures below, at, and above the estimated transition temperature of 138oC as

    determined by mechanical analysis for a strain rate of 10-3/s. At 160oC, the areal fraction

    of twins, Af, is 0.02. This value is consistent with minimum values of approximately

    0.06 observed by Lou et. al. [8] after cyclic testing causing twinning and untwining. This

  • 13

    range of values presumably represents the minimum fractions of twins attainable by

    removal by mechanical means. At 140oC, near the transition temperature determined by

    mechanical means, Af is 0.09. At 120oC, Af is 0.27. These results are consistent with the

    transition temperatures from Figure 4, and furthermore are consistent with the hypothesis

    that curvature A is an indirect measure of twin fraction Af after a fixed amount of

    compressive deformation.

    Additional metallography was done to test and quantify the postulated relationship

    between A and Af. Figure 6 presents measured twin areal fractions, Af , and compressive

    stress-strain curvatures, A, for a range of temperatures and strain rates. Note that the data

    do not reveal a systematic variation with respect to temperature or strain rate. The best-

    fit line is represented by the following relationship:

    03.010.0)MPa(A)MPa(00005.0A 1Monof ±+=− (1)

    where the standard error of fit of 0.03.

    The A-intercept of Eq. (1) (Af=0.1 at A=0) represents the area fraction of twins observed

    after a monotonic compressive strain of -0.08 where the stress-strain curve is essentially

    linear (A=0). This value serves to calibrate phenomenologically the definition of

    transition between twin-dominated and slip-dominated deformation, as interpreted by the

    shape of the mechanical stress-strain curve.

    The fact that for a given test and strain range the areal fraction of twins is related to the

    curvature, but independent otherwise of both temperature and strain rate suggests that

  • 14

    temperature and strain rate might be representable by a single variable. The obvious

    choice to try is the Zener-Hollomon parameter [66], Z, which is represented by the

    following:

    =

    RTQZ expε (2)

    where Q is the activation energy for self diffusion, R is the ideal gas constant

    (8.31J/molK) and T is the absolute temperature in K. A value of Q equal to

    135,000J/mol has been found to correlate temperature and flow stress of magnesium

    alloys [47].

    In order to first test the usefulness of Equation 2, the compressive test data from Figures

    3 were rearranged and plotted in terms of the yield stress and Z, Figure 7(a). Note that

    the data represents a range of temperatures from 120oC to 250oC and strain rates from 10-

    3/s to 10-1/s correlating to a range of Z of 4 orders of magnitude. The data fall onto two

    straight lines which intersect at a transition value of 2.28.31Z

    Zln

    0

    trans ±=

    . The best-

    fit lines can be represented as follows, with the standard errors of fits shown:

    )MPa(5)MPa(55ZZln)MPa(5.5)MPa(

    0

    twiny ±−

    =σ (3)

    )MPa(4)MPa(187ZZln)MPa(6.9)MPa(

    0

    slipy ±−

    =σ (4)

    where Z0=1/s, and the two lines and corresponding yield stresses have tentatively been

    divided into twin-dominated deformation ( twinyσ ) and slip-dominated deformation (slipyσ )

    based on the intersection point of the two lines.

  • 15

    The existence of two lines suggest a mechanistic interpretation that is consistent with the

    literature, i.e., slip-dominated deformation at higher temperatures/lower strain rates, and

    twin-dominated deformation at lower temperatures/higher strain rates. In a simplistic

    interpretation, the lines represent the lower activation stress of the two deformation

    mechanisms, but of course in reality a combination of mechanisms can be, and usually is,

    required in order to satisfy the von Mises rule [11]. The slopes of the two lines in Figure

    7(a) are consistent with the observation that the flow stress for slip is more rate-

    dependent (or, equivalently, by Equation 2, temperature-dependent) than that for

    twinning [24]. The slope of the slip line is approximately 1.7 times that of twin line.

    This is qualitatively consistent with many reports of twinning occurring athermally when

    taking into account the combined nature of slip and twinning required at lower

    temperatures, where non-basal slip requires a much higher stress to activate [67]. If, in

    fact, twinning is completely athermal, the ratio of slopes suggests that the initial ratio of

    strain from twinning to slip would likely be in the range of 2/3 near the transition

    temperature. This can be compared with Lou’s results [8] at room temperature where the

    initial ratio of twinning strain to total strain was approximately 0.85. The sign of the

    difference is consistent with slip being favored at higher temperatures, thus

    reducing the initial fraction of deformation accomplished by twinning.

    While Figure 7(a) is conceptually appealing because it emphasizes the cross-over of

    activation stresses for the two combinations of deformation mechanisms (i.e. slip only

    and slip+twinning), a more direct and hopefully more sensitive measure is related to the

  • 16

    emerging concept in the current work of characterizing the combination of deformation

    mechanisms using the stress-strain curvature, A. Figure 7(b) is similar to Figure 7(a), but

    is based on the curvature rather than the yield stress. Again, two lines fit the data, with a

    transition determined from their intersection near the same value of ln(Ztrans/Z0) as for

    yield stress ( ( ) 2.28.31.vs,9.07.31Z/Zln 0trans ±±= ). The reduced scatter reflects the

    better consistency of curvature data as compared with yield stress data. This ln(Ztrans/Z0)

    value is also indistinguishable from the one for A=0, established earlier in this paper

    ( ) 4.02.32Z/Zln 0trans ±= . The best-fit straight lines are as follows, with the standard

    errors of fits shown:

    )MPa(586)MPa(51715ZZln)MPa(1604)MPa(A

    0

    twin ±−

    = (5)

    )MPa(480)MPa(14388ZZln)MPa(427)MPa(A

    0

    slip ±−

    = (6)

    Figure 7(c) is a plot similar to Figures 7(a) and (b) except using measured areal fractions

    of twins. The transition values, respectively from the intersection of the two lines

    ( ( ) 8.03.31Z/Zln 0trans ±= ) and from the Af=0.10 calibrated criterion

    ( ( ) 4.02.32Z/Zln 0trans ±= ) are identical within the scatter, and are consistent with the

    ones determined from Figures 7a and b. The best-fit equations are as follows, with the

    standard errors of fits shown:

    03.04.2ZZln079.0A

    0

    twinf ±−

    = (7)

    01.07.0ZZln023.0A

    0

    slipf ±−

    = (8)

  • 17

    Note that standard deviation from twin-fraction measurement is 0.04, which is consistent

    with the above fit errors. Using each of the various criteria obtains an average critical

    Zener-Hollomon parameter in a logarithm format as follows: ( ) 3.18.31Z/Zln 0trans ±= .

    If the yield stress criterion is eliminated because of its higher inherent scatter, the scatter

    is reduced but the average is unchanged: ( ) 8.08.31Z/Zln 0trans ±= . The scatter of these

    numbers is approximately the same as combined uncertainty of the experiments and

    fitting to obtain the lines and their intersections. Thus, it is concluded an identical value

    of Ztrans is obtained by all of the techniques used to reveal it, both mechanical and

    metallographic.

    3.2 ST testing and analysis

    Easily conducted and interpreted, the monotonic compressive test usually requires at least

    5 samples to determine a transition temperature, Tt, between slip and twin for one strain

    rate, excluding the trial samples to provide on estimate of Tt within approximately 10oC.

    Optical metallography is even more cumbersome, and requires a similar number of

    destructively-tested specimens. Based on the validity of the curvature of the stress-strain

    curve on determining Tt, a step-temperature (ST) cyclic T/C test was conceived and

    designed to potentially shorten the experimental time, and to reduce the required

    specimen number. Detailed experimental procedures have been described in the

    Experimental Procedures section of this paper.

    Figure 8 presents a typical ST test result for Mg (AU) alloy at a strain rate of 10-3/s. Both

    T and C legs show the expected transition from concave-down (slip) at higher

  • 18

    temperatures and inflected shapes (twinning or untwinning) at lower temperatures. A

    quadratic equation ( CBA 2 +ε+ε=σ ) was again fit to the true stress-strain curves for an

    incremental strain range of 0.03-0.08 following each stress reversal, Figures 9. The first

    tension leg is excluded in the plots because of its different strain range, 0.04 vs. 0.08.

    Figures 9a and b compare the various legs and the best-fit curvature values (and the

    visual goodness of fit).

    Figure 9c illustrates the variation of curvature with temperature obtained from an ST test

    in compressive legs ( CtT ) and tension legs (TtT ), respectively, and compares with a series

    of isothermal monotonic compression tests presented earlier. The transition temperatures

    from the ST test are 140oC from both tension and compression legs, as compared with

    145oC from the series of monotonic tests. Analysis of the expected combined uncertainty

    of Ztrans from all monotonic tests excluding yield stress characterizations (i.e.

    ( ) 8.08.31Z/Zln 0trans ±= ) gives the equivalent untertainty for Ttrans of C9C147 oo ±

    for this case. Therefore, the ST test offers a faster, more economical, material-conserving

    and equally-accurate alternative to performing a series of compression tests, with or

    without metallographic analysis.

    ST tests were performed over a range of strain rates and using temperature decrements of

    5oC and 10oC, with results shown in Table 2. The transition temperatures, Tt (defined as

    the average values of CtT and TtT from compression and tension legs of the ST test,

    respectively) are identical for the two decrements within the overall scatter of the

    technique (standard deviation of 3.6oC). For all subsequent results, the transition

  • 19

    temperatures are determined by the average results from tension and compression legs for

    ST testing.

    In Figure 10, the ST test results (points) are compared with the corresponding monotonic

    compression results presented previously (solid line, with dashed lines representing the

    scatter):

    ( )( ) ( )( ) ( ) ( ) ( )CCKCLnmolKJ

    molJCT oooot 9273//8.08.31/31.8/000,135)(

    0

    ±−−±×

    =εε

    (9)

    where s/10 =ε and the standard error of Tt is 9oC. Except at the highest strain rate of

    0.1/s (corresponding to a transition temperature of 231oC for ST tests), the ST tests and

    monotonic tests give identical results within the experimental scatter for each technique.

    The difference of transition temperature from the two techniques at 0.1/s will be

    discussed below.

    The time-efficient and specimen-efficient ST test, being now been confirmed as giving

    transition temperatures consistent with monotonic compression testing and

    metallographic methods (except at the highest strain rate, 0.1/s), allows the rapid testing

    of a range of material to look at other relationships. As a test of this capability, and to

    assist with understanding the discrepancy of testing at 0.1/s-230oC, five other batches of

    Mg AZ31B with varying grain sizes in the as-received (annealed) condition were

    subjected to ST testing, with results shown in Figure 11. The equation for the best-fit line

    shown in Figure 11 is as follows:

    ( ) ( ) ( ) ( )C8.6mdm/C5.6C6.118)C(T oooot ±μ×μ+= (10)

  • 20

    where d is the average grain size and the standard error of fit is 6.8oC. This dependence

    is consistent with similar reports in the literature [38]. Equation (10) can be converted to

    a Zener-Hollomon basis using the known strain rate of 0.001/s used in the tests

    represented in Figure 11:

    ( ) ( ) 6.0/546.01.34)/( 0 ±×−= mdmZZLn trans μμ (11)

    where Z0=1/s, and the standard error of Eq. (11) is 0.6, corresponding to the standard

    deviation of Tt, 6.8oC, for Eq. (10).

    With known relationships for transition temperatures or critical Zener-Hollmon

    parameters in terms of grain size (Eqs. 10 and 11), the question of the discrepancy of

    results between ST and monotonic tests at the highest tested strain rate can be addressed.

    Given the high transition temperature for ST testing at 0.1/s (231oC) an obvious

    possibility is grain growth occurring during the test, thus increasing the measured Tt. To

    test this hypothesis, the grain sizes for Mg AZ31B (AU) were measured after the

    monotonic tests on either side of the transition temperature, 200oC and 210oC, and for the

    single ST test which started at 250oC and ended at 210oC. The monotonic tests showed no

    grain growth, with measured values of 4.0μm and 4.1μm, respectively, for the two

    temperatures. (Recall that the initial grain size for Mg AZ31B (AU) was 4.0μm +/- 0.3

    μm.) Conversely, the grain size of the ST specimen after testing had increased to 7.6μm 1.

    Using Equation 11 to predict the role of a grain size change from 4.05μm to 7.6μm shows 1 In order to see if the grain growth rates during testing were consistent with regular grain growth kinetics, undeformed specimens of Mg AZ31B (AU) were heat-treated for 5 minutes at 150oC, 200oC, and 250oC. The specimens heat treated at 150oC and 200oC showed no grain growth (3.8μm and 4.0μm finally, compared with 4.0 +/- 0.3 initially) whereas the specimen heat treated at 250oC showed a grain size increased to 6.8μm. The grain growth behavior under deformation conditions at similar temperatures is similar.

  • 21

    an expected change of transition temperature from 202oC to 231oC, a difference of 29oC.

    That corresponds closely with the measured differences of Tt between the two tests of

    27oC (204oC vs. 231oC). Thus, the principal source of discrepancy between ST and

    monotonic tests at the highest strain rate (and thus highest tested temperatures) is caused

    by grain growth.

    In order to determine whether a similar one-to-one relationship between A and Af can be

    established for ST tests, metallographic analysis was conducted for ST tests stopped after

    1, 3 and 5 cycles. The results, Figure 12, are similar to those for monotonic compression

    tests conducted isothermally, Figure 6, in terms of exhibiting a one-to-one

    correspondence between A and Af, unaffected by temperature or strain rate. The actual

    relationship is not identical to the one observed for monotonic, isothermal tensile tests.

    Except at very small values of A and Af, the relationships are the same except for a fixed

    offset (of either A or Af). Differences such as are shown in Figure 12 are to be expected

    when considering the complex effects of cyclic deformation, time-at-temperature changes,

    and the different accumulated strain magnitudes between the two tests, and the possible

    effects on slip resistance and twin nucleation.

    In order to clarify the origins of the different relationship between A and Af for the two

    kinds of tests, modified ST tests starting a low temperature were conducted. The

    transition temperatures obtained from these “ST-up” tests were significantly higher than

    measured by the standard ST test, monotonic tests, or metallographic examination (all of

    which are consistent). The ST-up tests give Tt of 156oC, versus 141oC obtained from the

  • 22

    standard ST test (Table 2). This difference is likely attributable to long-lasting changes of

    microstructure based on the twinning at the initial, lower-temperature cycles. In a

    simplified view, twinning and untwining during cycling at temperatures below Tt make

    subsequent twinning easier by reducing the twin nucleation stress. This is consistent with

    acoustic emission and mechanical testing results presented by Lou (2007).

    Metallographic examination after an ST-up test at 0.001/s starting at 120oC and finishing

    at 160oC exhibited an areal twin fraction of 0.1. That is, twinning continued at

    temperatures above the value of Tt for the virgin material.

    As a final probe to discern the role of mechanical cycling separate from the thermal

    complexities of the ST test, isothermal “ST-no” tests were done. The observed curvatures

    of the ST-no tests are compared with the ST and ST-up tests in Figure13. The isothermal

    tests show that cycling alone increases the curvature of the stress-strain curve, with

    results intermediate between St-up and ST tests. This is consistent with the expectation

    that cycling reduces the activation stress for twinning (and untwining) [8], thus making

    twin-dominant deformation possible at higher temperatures than for monotonic tests. This

    tends to confirm the origin of the differences seen in the ST-up and standard ST tests.

    The conclusion is that the ST-up test is unlikely to give Tt values consistent with other

    measures because of the reduction of twinning activation stress by low-temperature

    mechanical cycling.

    4. CONCLUSIONS

  • 23

    The following conclusions were reached by mechanical testing in monotonic

    compression, cyclic tension/compression, and metallographic analysis of Mg AZ 31B:

    1. It is possible to measure transition temperature for dominant deformation

    mechanisms (slip vs. slip-plus-twinning) purely mechanically based on the

    curvature of the stress-strain relationship over fixed strain intervals. Procedures

    for carrying out such tests have been presented. Mechanical tests based on

    compressive yield stress are simpler to interpret and yield similar results, but are

    less accurate.

    2. Mechanical measurement of transition temperatures are consistent with, and have

    similar scatter to, metallographic determinations.

    3. Transition temperature depending on strain rate according to a critical Zener-

    Hollomon parameter, the value of which has been measured consistently using

    several kinds of tests. The activation energy is consistent with presentations in

    the literature [38].

    4. Transition temperatures are linearly related to grain size, in agreement with the

    literature [38].

    5. There is a one-to-one relationship between stress-strain curvature over a fixed

    strain interval and the area fraction of twins measured after such deformation.

    This relationship is independent of temperature and strain rate over the ranges

    used here. An explicit form has been presented.

    6. The deformation mechanism transition can defined mechanically by a zero-

    curvature stress strain response over a strain range 0.03-0.8 corresponding to a

  • 24

    final areal fraction of twins equal to 0.1 for monotonic compression tests and 0.04

    for “ST” (step temperature, cyclic tension/compression) tests.

    7. Cyclic deformation at temperatures above the transition temperature (i.e.slip-

    dominated) has little effect on the subsequent transition temperature. Conversely,

    cyclic deformation below the transition temperature (twin-dominated) increases

    the subsequent transition temperature, likely by a reduction of twin nucleation

    stress at proposed by Lou [8].

    8. Care must be taken with ST tests to avoid times-at-temperatures sufficient to

    induce grain growth. Grain growth raises the transition temperature relative to that

    for the virgin material.

    9. Mechanical tests are faster, simpler, less expensive, and consume less material

    than traditional techniques used for determining transition temperatures. They

    introduce the possibility of rapidly testing in a range of material using small

    quantities of each.

    ACKNOWLEDGMENTS

    This work was supported by the National Research Foundation of Korea (Grant NRF-

    2010-220-D00037). Many thanks to AUSTEM Co. for providing materials, to Dr. Lou

    Hector, Jr. (GM R&D Center) for providing materials, to Mr. Steve Bright for assistance

    with sample preparation for optical metallography, and to Professors Sean Agnew and

    Frederic Barlat for always-helpful discussions.

  • 25

    REFERENCES 1. Roberts, C.S., Magnesium and its alloys. 1960, New York and London: John

    Wiley.

    2. Agnew, S.R., Wrought Magnesium: A 21st Century Outlook. JOM, 2004: p. 20-21.

    3. Bettles, C.J. and Gibson, M.A., Current wrought magnesium alloys: strengths and

    weaknesses. JOM, 2005. 57(5): p. 46.

    4. Duygulu, O. and Agnew, S.R., The effect of temperature and strain rate on the

    tensile properties of textured magnesium alloy AZ31B sheet. Magnesium

    Technology 2003, 2003: p. 237-242.

    5. Wu, L., et al., The effects of texture and extension twinning on the low-cycle

    fatigue behavior of a rolled magnesium alloy, AZ31B. Materials Science and

    Engineering A, 2010. 527: p. 7057-7067.

    6. Luo, A.A., Recent Magnesium alloy development for elevated temperature

    applications. International Materials Reviews, 2004. 49(1): p. 13-30.

    7. Agnew, S.R. and Duygulu, O., Plastic anisotropy and the role of non-basal slip in

    magnesium alloy AZ31B. International Journal of Plasticity, 2005. 21: p. 1161-

    1193.

    8. Lou, X.Y., et al., Hardening evolution of AZ31B Mg sheet. International Journal

    of Plasticity, 2007. 23: p. 44-86.

    9. Beck, A., Technology of Magnesium and its Alloys. 1943, London: Kynock Press.

    23.

  • 26

    10. Nobre, J.P., et al., Deformation asymmetry of AZ31 wrought magnesium alloy.

    Key Engineering Materials, 2002. 230-232: p. 267-270.

    11. Von Mises, R., Mechanik der plastischen formanderung von kristallen. Z. Angew.

    Math. Mech., 1928. 8: p. 161-185.

    12. Taylor, G.I., Plastic strain in metals. Journal of The Institute of Metals, 1938.

    13. Agnew, S.R. and Duygulu, O., TEM investigation of dislocation mechanisms in

    Mg alloy AZ31B sheet. Magnesium Technology 2004, 2004: p. 61-65.

    14. Kocks, U.F. and Westlake, D.G., The important of twinning for the ductility of

    CPH polycrystals. Transactions of the Metallurgical Society of AIME, 1967. 239:

    p. 1107-1109.

    15. Agnew, S.R., Yoo, M.H., and Tome, C.N., Application of texture simulation to

    understanding mechanical behavior of Mg and solid solution alloys containing Li

    or Y. Acta Materialia, 2001. 49: p. 4277-4289.

    16. Yoo, M.H., et al., Nonbasal deformation modes of HCP metals and alloys: role of

    dislocation source and mobility. Metallurgical and Materials Transactions A,

    2002. 33: p. 813-?

    17. Reed-Hill, R.E. and Abbaschian, R., Physical Metallurgy Principles. 3rd ed. 1994,

    Boston: PWS Publ. Company.

    18. Hertzberg, R.W., Deformation and fracture mechanics of engineering materials.

    4th ed. 1996, New York: J. Wiley & Sons, Inc.

    19. Li, M., Constitutive modeling of slip, twinning, and untwinning in AZ31B

    magnesium, in Department of Material Science and Engineering. 2006, The Ohio

    State University: Columbus, OH. p. 128.

  • 27

    20. Wonsiewicz, B.C. and Backofen, W.A., Plasticity of magnesium crystals. Trans.

    TMS-AIME, 1967. 239: p. 1422-1431.

    21. Obara, T., Yoshiinga, H., and Morozumi, S., [112-2](11-23) Slip system in

    Magnesium. Acta Materialia, 1973. 21: p. 845-853.

    22. Yoshinaga, H. and Horiuchi, R., On the Flow Stress of α Solid Solution Mg-Li

    Alloy Single Crystals. Trans. JIM, 1963. 4: p. 134-141.

    23. Stohr, J.F. and Poirier, J.P., Etude en Microscopie Electronique de Glissement

    Pyramidal {11-22} dans le Magnesium. Philosophy Magzine, 1972. 25:

    p. 1313-1329.

    24. Meyers, M.A., Vohringer, O., and Lubarda, V.A., The onset of twinning in metals:

    a constitutive description. Acta Materialia, 2001. 49: p. 4025-4039.

    25. Avedesion, M.M. and Baker, H., ASM specialty handbook: Magnesium and

    Magnesium alloys. 1999: ASM International.

    26. Christian, J.W. and Mahajan, S., Deformation Twinning. Progress in Materials

    Science, 1995. 39: p. 1-157.

    27. El-Danaf, E., Kalidindi, S.R., and Doherty, R.D., Influence of Grain Size and

    Stacking-Fault Energy on Deformation Twinning in FCC Metals. Metallurgical

    and Materials Transaction A, 1999. 30A: p. 1223-1233.

    28. Asgari, S., et al., Strain Hardening Regimes and Microstructural Evolution

    during Large Strain Compression of Low Stacking Fault Energy FCC Alloys that

    Form Deformation Twins. Metallurgical and Materials Transaction A, 1997. 28A:

    p. 1781-1795.

  • 28

    29. El-Danaf, E., Kalidindi, S.R., and Doherty, R.D., Influence of Deformation Path

    on the Strain Hardening Behavior and Microstructure Evolution in Low SFE

    FCC Metals. International Journal of Plasticity, 2001. 17: p. 1245-1265.

    30. Mahajan, S. and Williams, D.F., Deformation Twinning in Metals and Alloys.

    International Metallurgical Reviews, 1973. 18: p. 43-61.

    31. Salem, A.A., Kalidindi, S.R., and Doherty, R.D., Strian Hardening Regimes and

    Microstructure Evolution during Large Strain Compression of High Purity

    Titanium. Scripta Materialia, 2002. 46: p. 419-423.

    32. Salem, A.A., Kalidindi, S.R., and Doherty, R.D., Strain Hardening of Titanium:

    Role of Deformation Twinning. Acta Materialia, 2003. 51: p. 4225-4237.

    33. Salem, A.A., et al., Strain Hardening Due to Deformation Twinning in α-

    Titanium: Mechanisms. Metallurgical and Materials Transaction A, 2006. 37A: p.

    259-268.

    34. Chichili, D.R., Ramesh, K.T., and Hemker, K.J., The High-Strain-Rate Response

    of Alpha-Titanium: Experiments, Deformation Mechanisms and Modeling. Acta

    Materialia, 1998. 46: p. 1025-1043.

    35. Song, S.G. and III, G.T.G., Influence of Temperature and Strain Rate on Slip and

    Twinning Behavior of Zr. Metallurgical and Materials Transaction A, 1995. 26A:

    p. 2665-2675.

    36. Khan, A.S., et al., Mechanical response and texture evolution of AZ31 alloy at

    large strains for different strain rates and temperatures. International Journal of

    Plasticity, 2011. 27: p. 688-706.

  • 29

    37. Remy, L., Kinetics of FCC Deformation Twinning and its Relationship to Stress-

    Strain Behavior. Acta Materialia, 1978. 26: p. 443-451.

    38. Barnett, M.R., et al., Influence of grain size on the compressive deformation of

    wrought Mg-3Al-1Zn. Acta Materialia, 2004. 52: p. 5093-5103.

    39. Barnett, M.R., Twinning and the ductility of magneisum alloys Part I: "Tension"

    twins. Materials Science and Engineering A, 2007. 464: p. 1-7.

    40. Jiang, L., et al., Influence of {10-12} extension twinning on the flow behavior of

    AZ31 Mg alloy. Materials Science and Engineering A, 2007. 445-446: p. 302-309.

    41. Muransky, O., et al., Investigation of deformation twinning in a fine-grained and

    coarse-grained ZM20 Mg alloy: Combined in situ neutron diffraction and

    acoustic emission. Acta Materialia, 2010. 58: p. 1503-1517.

    42. Barrett, C.S. and C T Haller, J., Twinning in Polycrystalline Magnesium. Metals

    Technology, 1946. 13(8).

    43. Wang, Y.N. and Huang, J.C., The role of twinning and untwinning in yielding

    behavior in hot-extruded Mg-Al-Zn alloy. Acta Materialia, 2007. 55: p. 897-905.

    44. Jiang, L., et al., Twinning and texture development in two Mg alloys subjected to

    loading along three different strain paths. Acta Materialia, 2007. 55: p. 3899-

    3910.

    45. Keshavarz, Z. and Barnett, M.R., EBSD analysis of deformation modes in Mg-

    3Al-1Zn. Scripta Materialia, 2006. 55: p. 915-918.

    46. Nave, M.D. and Barnett, M.R., Microstructures and textures of pure magnesium

    deformed in plane-strain compression. Scripta Materialia, 2004. 51(881-885).

  • 30

    47. Barnett, M.R., Nave, M.D., and Bettles, C.J., Deformation microstructures and

    textures of some cold rolled Mg alloys. Materials Science and Engineering A,

    2004. 386: p. 205-211.

    48. Jain, A., et al., Grain Size Effects on the Tensile Properties and Deformation

    Mechanisms of a Magnesium Alloy, AZ31B, Sheet. Materials Science and

    Engineering A, 2008. 486: p. 545-555.

    49. Agnew, S.R., et al., Study of Slip Mechanisms in a Magnesium Alloy by Neutron

    Diffraction and Modeling. Scripta Materialia, 2003. 48: p. 1003-1008.

    50. Clausen, B., et al., Reorientation and Stress Relaxation due to Twinning:

    Modeling and Experimental Characterization for Mg. Acta Materialia, 2008. 56:

    p. 2456-2468.

    51. Wu, L., et al., Internal Stress Relaxation and load redistribution during the

    twinning-detwinning-dominated cyclic deformation of a wrought magnesium alloy,

    ZK60A. Acta Materialia, 2008. 56: p. 3699-3707.

    52. Muransky, O., et al., In situ neutron diffraction investigation of deformation

    twinning and pseudoelastic-like behavior of extruded AZ31 magnesium alloy.

    International Journal of Plasticity, 2009. 25: p. 1107-1127.

    53. Boger, R.K., et al., Continuous, large strain, tension/compression testing of sheet

    material. International Journal of Plasticity, 2005. 21(12): p. 2319-2343.

    54. Tan, Z., Magnusson, C., and Persson, B., The Bauschinger Effect in Compression-

    tension of sheet materials. Materials Science and Engineering A, 1994. 183: p.

    31-38.

  • 31

    55. Kuwabara, T., Nagata, K., and Nakako, T., Measurement and analysis of the

    Bauschinger effect of sheet metals subjected to in-plane stress reversals.

    AMPT'01. 2001, University Carlos III de Madrid, Madrid.

    56. Cao, J., et al., Experimental and numerical investigation of combined isotropic-

    kinematic hardening behavior of sheet metals. International Journal of Plasticity,

    2009. 25: p. 942-972.

    57. Li, M., et al., An efficient constitutive model for room-temperature, low-rate

    plasticity of annealed Mg AZ31B sheet. International Journal of Plasticity, 2010.

    26: p. 820-858.

    58. Kim, J.H., Sung, J.H., and Wagoner, R.H., The Shear Fracture of Dual-Phase

    Steel. International Journal of Plasticity, 2011. 27: p. 1658-1676.

    59. Sung, J.H., Kim, J.H., and Wagoner, R.H., A plastic constitutive equation

    incorporating strain, strain-rate, and temperature. International Journal of

    Plasticity, 2010. 26: p. 1746-1771.

    60. Sun, L. and Wagoner, R.H., Complex unloading behavior: nature of the

    deformation and its consistent constitutive representation. International Journal of

    Plasticity, 2011. 27: p. 1126-1144.

    61. Piao, K., et al. Tension / Compression Test of Mg AZ31B at Elevated

    Temperatures. in North American Deep Drawing Research Group. 2008. Windsor,

    CA.

    62. Piao, K., et al., A Sheet Tension/Compression Test for Elevated Temperature.

    Submitted to International Journal of Plasticity, 2011.

  • 32

    63. Balakrishnan, V., Measurement of in-plane Bauschinger Effect in metal sheets.

    1999, The Ohio State University.

    64. ASTM-E562-08e1, Standard test method for determining volume fraction by

    systematic manual point count. 1999.

    65. ASTM-E112-10, Standard test methods for determining average grain size.

    ASTM International. 2010, 100 Barr Harbor Drive, PO Box C700, West

    Conshohocken, PA 19428-2959, United States.

    66. Zener, C. and Hollomon, J.H., Effect of Strain Rate Upon Plastic Flow of Steel.

    Journal of Applied Physics, 1944. 15: p. 22-32.

    67. Obara, T., Yoshinga, H., and Morozumi, S., {11-22} slip system in

    magnesium. Acta Metallurgica, 1973. 21: p. 845-853.

    68. F J Polesak, I., et al., Blind Study of the effect of Processing History on the

    Constitutive Behavior of Alloy AZ31B. TMS-AIME 2008, 2008.

    69. AUSTEM company, L., Korea. 1993.

  • List of Tables Table 1: Summary properties of Mg AZ31B alloys; compositions in weight percent.

    Table 2: Transition temperatures Tt for alloy AU obtained from ST tests.

  • Table 1: Summary properties of Mg AZ31B alloys; compositions in weight percent.

    Sheet* Al (%) Zn (%)

    Mn (%)

    Mg (%)

    Thickness (mm) gs (μm)

    AU 2.8 0.90 0.46 Balance** 2.02 4.0

    GM-A N/A N/A N/A Balance 1.05 6.0-10.0

    GM-M 2.6 0.71 0.32 Balance 1.01 11.0

    GM-N 3.0 0.74 0.35 Balance 0.99 8.2

    GM-O 3.0 0.74 0.32 Balance 1.00 5.8

    GM-X 2.9 0.95 0.53 Balance 0.98 8.3

    * Materials except for AU were provided by Dr. Lou Hector at General Motors, US. More information on the GM-designated alloys appear in the literature [68]. The AU material was provided by the AUSTUM company, Korea [69]. ** The AU alloy also contained 0.02% Ca, 0.005% Fe, and 0.0007% Cu. The remaining alloys were not tested for such trace quantities.

  • Table 2: Transition temperatures Tt for alloy AU obtained from ST tests.

    Strain Rate (/s) ΔT (

    oC) ttT (oC) ctT (

    oC) 2

    TTT

    ct

    tt

    t+

    = (oC)

    0.0005 10 129 137 133 0.0005 5 129 134 132 0.001 10 141 140 141 0.001 5 143 142 143 0.005 10 161 169 165 0.005 5 167 170 169 0.01 10 171 180 176 0.01 5 169 174 172 0.05 10 195 205 200 0.05 5 196 201 199 0.1 10 230 231 231 0.1 5 231 229 230

  • List of Figures

    Figure 1. Schematic of elevated-temperature T/C test and sample dimensions [53, 62] Figure 2. Isothermal cyclic T-C-T test results for Mg AZ31B(AU) at 0.001/s: (a) full results for testing at 25oC to 250oC [62]; (b) expanded compressive results after tension to a strain of 0.04 (not shown), and (c) expanded tension results after T( 04.0=ε ) and C( 08.0=ε ) intervals (not shown). Figure 3. Isothermal, monotonic compressive test results for Mg AZ31B (AU) at strain rates of (a) 0.001/s, (b) 0.01/s, and (c) 0.1/s. Figure 4. Determination of transition temperatures, Tt, of Mg AZ31B (AU) for isothermal monotonic compression tests at strain rates of 10-3, 10-2, 10-1/s. Figure 5. Optical micrographs of Mg AZ31B (AU) samples with deformation of ε=-0.08 at (a) 120oC, (b) 140oC, and (c) 160oC at a strain rate of 0.001/s. Figure 6. Experimental and best-fit linear correlation between the areal fraction of deformation twins and the stress-strain curvature, A=(d2σ/dε2), obtained from monotonic compressive hardening tests for the strain range -0.03 to -0.08 at various temperatures and strain rates.

    Figure 7. Correlation of quantitative indicators of deformation mechanism with Zener-Hollomon parameter Z at various temperatures and strain rates as shown: a) compressive yield stress (-0.002 offset), b) stress-strain curvature, A, strain range -0.03- -0.08, c) measured areal twin fraction, Af, after a compressive strain of -0.08. Figure 8. Step-temperature cyclic T/C test at temperature from 150oC to 110oC, at a strain rate of 0.001/s using Mg AZ31B (AU) material. Figure 9. Determination of transition temperature, Tt, of Mg AZ31B (AU) using an ST cyclic test at a strain rate of 0.001/s: a) fitting of curvatures in compression portions; b) fitting of curvatures in tension portions; and c) transition temperatures determined from compression and tension portions compared with the one from isothermal monotonic compression tests. Figure 10. Evolution of transition temperature with strain rate. The solid line is the best-fit representation of the combined monotonic compression data corresponding to

    ( ) 8.08.31Z/ZLn 0trans ±= , Eq. 9. The dashed lines represent the scatter of Eq. 9, i.e., C9o .

    Figure 11. Dependence of transition temperature with grain size determined from ST tests for six Mg AZ31B alloys.

  • Figure 12. Correlation between the areal fraction of deformation twin and the curvature, A (d2σ/dε2), obtained from ST cyclic T/C tests. Figure 13. Development of stress-strain curvature with mechanical tension-compression cycling in ST, ST-up, and ST-no tests.

  • Figure 1. Schematic of elevated-temperature T/C test and sample dimensions [53, 62].

  • 0

    100

    200

    300

    400

    0 0.05 0.1 0.15 0.2 0.25

    Abs

    olut

    e Tr

    ue S

    tres

    s (M

    Pa)

    Accumulated Absolute True Strain

    Mg AZ31B (AU), RD, TCT testThickness = 2.02 mmStrain rate = 0.001/s, μ = 0.03Side force = 2.5kN

    25oC50oC

    75oC

    100oC

    125oC

    150oC

    175oC

    200oC225oC

    250oCTension TensionCompression

    25oC 50oC

    75oC

    100oC

    125oC

    150oC175oC

    200oC225oC250oC

    (a)

    120

    130

    140

    150

    160

    170

    180

    0.05 0.06 0.07 0.08 0.09 0.1 0.11

    Abs

    olut

    e Tr

    ue S

    tres

    s (M

    Pa)

    Accumulated Absolute True Strain

    Mg AZ31B (AU), RD, TCT testThickness = 2.02 mmStrain rate = 0.001/s, μ = 0.03Side force = 2.5kN

    125oC

    150oC

    C after T

    (b)

  • 100

    125

    150

    175

    200

    0.12 0.14 0.16 0.18 0.2

    Abs

    olut

    e Tr

    ue S

    tres

    s (M

    Pa)

    Accumulated Absolute True Strain

    Mg AZ31B (AU), RD, TCT testThickness = 2.02 mmStrain rate = 0.001/s, μ = 0.03Side force = 2.5kN 125oC

    150oC

    T after T-C

    (c)

    Figure 2. Isothermal cyclic T-C-T test results for Mg AZ31B(AU) at 0.001/s: (a) full results for testing at 25oC to 250oC [62]; (b) expanded compressive results after tension to a strain of 0.04 (not shown), and (c) expanded tension results after T( 04.0=ε ) and C( 08.0=ε ) intervals (not shown).

  • 100

    120

    140

    160

    180

    200

    0 0.02 0.04 0.06 0.08 0.1

    Abs

    olut

    e Tr

    ue S

    tres

    s (M

    Pa)

    Absolute True Strain

    120oC

    130oC

    140oC

    150oC

    160oC

    Mg AZ31B (AU), Ct=2.02mm, gs=4.0μm, strain rate = 0.001/s

    (a)

    100

    120

    140

    160

    180

    0 0.02 0.04 0.06 0.08 0.1

    Abs

    olut

    e Tr

    ue S

    tres

    s (M

    Pa)

    Absolute True Strain

    Mg AZ31B (AU), Ct=2.02mm, gs=4.0μm, strain rate = 0.01/s

    150oC

    160oC

    170oC

    180oC

    190oC

    (b)

    C145T ot ≅

    C170T ot ≅

  • 100

    110

    120

    130

    140

    150

    160

    0 0.02 0.04 0.06 0.08 0.1

    Abs

    olut

    e Tr

    ue S

    tres

    s (M

    Pa)

    Absolute True Strain

    Mg AZ31B (AU), Ct=2.02mm, gs=4.0μm, strain rate = 0.1/s 190oC

    200oC

    210oC

    220oC

    230oC

    (c)

    Figure 3. Isothermal, monotonic compressive test results for Mg AZ31B (AU) at strain rates of (a) 0.001/s, (b) 0.01/s, and (c) 0.1/s.

    C210T ot ≅

  • -2x103

    -1x103

    0

    1x103

    2x103

    3x103

    4x103

    5x103

    100 120 140 160 180 200 220 240

    Cur

    vatu

    re A

    (MPa

    )

    Test Temperature (oC)

    Mg AZ31B(AU), Monotonic Ct=2mm, gs=4.0μm

    0.001/s0.01/s0.1/s

    0.001/s

    0.01/s

    0.1/s

    Tt=138oC 165oC 204oC

    Concave-up

    Concavedown

    Figure 4. Determination of transition temperatures, Tt, of Mg AZ31B (AU) for isothermal monotonic compression tests at strain rates of 10-3, 10-2, 10-1/s.

  • (a) 10-3/s, 120oC. Af=0.27

  • (b) 10-3/s, 140oC. Af=0.09

  • (c) 10-3/s, 160oC. Af=0.02

    Figure 5. Optical micrographs of Mg AZ31B (AU) samples with deformation of ε=-0.08 at (a) 120oC, (b) 140oC, and (c) 160oC at a strain rate of 0.001/s.

  • 0

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    -4000 -2000 0 2000 4000 6000

    Are

    al fr

    actio

    n of

    twin

    Curvature A (MPa)

    Mg AZ31B, RD, Monotonic Ct=2mm, gs=4.0mm

    0.001/s0.01/s0.1/s

    Fig. 5(a)

    120oC

    Fig. 5(b), 140oC

    Fig. 5(c)

    160oCTwin dominated

    Slip dominated

    150oC

    160oC

    170oC

    130oC

    150oC

    190oC

    200oC

    210oC220oC

    230oC

    180oC190oC

    Figure 6. Experimental and best-fit linear correlation between the areal fraction of deformation twins and the stress-strain curvature, A=(d2σ/dε2), obtained from monotonic compressive hardening tests for the strain range -0.03 to -0.08 at various temperatures and strain rates.

  • 80

    90

    100

    110

    120

    130

    140

    28 30 32 34 36

    Abs

    olut

    e Yi

    eld

    Stre

    ss, |

    σ y| (

    MPa

    )

    Ln(Z/Z0)

    0.001/s0.01/s0.1/s

    Twin dominant (A>0)Slip dominant (A0)

    0.1/s

    Twin dominated

    Slip dominated

    120oC

    130oC

    140oC

    150oC160oC

    150oC

    160oC

    170oC180oC

    190oC

    190oC

    200oC

    210oC

    240oC230oC

    220oC

    Slip dominant (A

  • 0

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    28 30 32 34 36

    Are

    al F

    ract

    ion

    of T

    win

    Ln(Z/Z0)

    Mg AZ31B (AU), RD, Monotonic Ct=2.02mm, gs=4.0μm

    0.001/s0.01/s

    Twin dominant (A>0)

    0.1/s

    Twin dominated

    Slip dominated

    120oC

    130oC

    140oC

    150oC

    160oC

    150oC

    160oC

    170oC180oC

    190oC

    190oC200oC

    210oC220oC

    240oC230oC

    Slip dominant (A

  • -200

    -100

    0

    100

    200

    300

    0 0.2 0.4 0.6 0.8

    True

    Str

    ess

    (MPa

    )

    Accumulated Absolute True Strain

    Mg AZ31B (AU) ST cyclic T/C test t=2.02mm, gs=4.0μm, strain rate=0.001/s

    150oC 140oC 130oC 120oC 110oC

    Figure 8. Step-temperature cyclic T/C test at temperature from 150oC to 110oC, at a strain rate of 0.001/s using Mg AZ31B (AU) material.

  • 120

    140

    160

    180

    200

    0 0.02 0.04 0.06 0.08 0.1

    Abs

    olut

    e Tr

    ue S

    tres

    s (M

    Pa)

    Absolute True Strain after Revesal

    110oC (A=4219MPa)

    120oC (A=2223MPa)

    130oC (A=1078MPa)

    140oC (A=-164MPa)

    150oC (A=-1956MPa)

    Mg AZ31B (AU) compressive cycles,t=2.02mm, gs =4.0μm, strain rate = 0.001/s

    Fit curve: Aε2+Bε+C=σ

    (a)

    120

    140

    160

    180

    200

    0 0.02 0.04 0.06 0.08 0.1

    Abs

    olut

    e Tr

    ue S

    tres

    s (M

    Pa)

    Absolute True Strain after Reversal

    110oC (A=16465MPa)

    120oC (A=10143MPa)

    130oC (A=3946MPa)

    140oC (A=27MPa)

    150oC (A=-2591MPa)

    Mg AZ31B (AU), Tensile cycles,t=2.02mm, gs=4.0μm, strain rate = 0.001/s

    Fit curve: Aε2+Bε+C=σ

    (b)

  • -5x103

    0

    5x103

    1x104

    1.5x104

    2x104

    100110120130140150160

    Cur

    vatu

    re A

    (MPa

    )

    Temperature (oC)

    Mg AZ31B (AU), strain rate=0.001/sT

    down, t=2.02mm, gs=4.0μm

    ST, Tension

    ST,Compression

    Monotonic,Compression

    Tt, Mono

    Tt, ST

    (c)

    Figure 9. Determination of transition temperature, Tt, of Mg AZ31B (AU) using an ST cyclic test at a strain rate of 0.001/s: a) fitting of curvatures in compression portions; b) fitting of curvatures in tension portions; and c) transition temperatures determined from compression and tension portions compared with the one from isothermal monotonic compression tests.

  • 120

    140

    160

    180

    200

    220

    240

    0.0001 0.001 0.01 0.1

    Tran

    sitio

    n te

    mpe

    ratu

    re (o

    C)

    Strain rate (/s)

    Mg AZ31B (AU), RD,t=2.02mm, gs=4.0μm

    ST Cyclic Data, ΔT=10oCST Cyclic Data, ΔT=5oCEq. 9 (Monotonic Tests)

    Figure 10. Evolution of transition temperature with strain rate. The solid line is the best-fit representation of the combined monotonic compression data corresponding to

    ( ) 8.08.31Z/ZLn 0trans ±= , Eq. 9. The dashed lines represent the scatter of Eq. 9, i.e., C9o .

  • 120

    140

    160

    180

    200

    2 4 6 8 10 12

    Tran

    sitio

    n te

    mpe

    ratu

    re (o

    C)

    Grain size (μm)

    Mg AZ31B, RD,Strain rate = 0.001/s

    Cyclic, ΔT=10oC

    Cyclic, ΔT=5oC

    Best-fit line

    AU

    GM-O

    GM-A

    GM-N

    GM-XGM-M

    Figure 11. Dependence of transition temperature with grain size determined from ST tests for six Mg AZ31B alloys.

  • 0

    0.04

    0.08

    0.12

    0.16

    -2000 -1000 0 1000 2000 3000

    Are

    al fr

    actio

    n of

    twin

    Curvature A (MPa)

    Mg AZ31B, RD, Cyclic Ct=2.02mm, gs=4.0mm

    0.001/s0.01/s0.1/s

    Twin dominated

    Slip dominated

    160oC-120oC

    160oC-140oC

    160oC

    190oC

    190oC-170oC

    190oC-150oC

    250oC

    250oC-230oC

    250oC-210oC

    ST testMonotonic test

    Figure 12. Correlation between the areal fraction of deformation twin and the curvature, A (d2σ/dε2), obtained from ST cyclic T/C tests.

  • -3x103

    -2x103

    -1x103

    0

    1x103

    2x103

    3x103

    0 1 2 3 4 5 6

    A (M

    Pa)

    Cycle

    Mg AZ31B (AU), strain rate=0.01/st=2.02mm, gs=4.0μm

    ST-up

    ST

    ST-no

    190oC

    180oC

    170oC

    170oC

    170oC

    170oC170oC

    170oC

    170oC160oC

    150oC

    150oC 190oC

    160oC180oC

    Figure 13. Development of stress-strain curvature with mechanical tension-compression cycling in ST, ST-up, and ST-no tests.

    Article FileTable 1-2Figure 1-13