art%3a10.1007%2fs10853-014-8099-1.pdf

13
Localized strain and heat generation during plastic deformation in nanocrystalline Ni and Ni–Fe T. Chan Y. Zhou I. Brooks G. Palumbo U. Erb Received: 29 November 2013 / Accepted: 31 January 2014 / Published online: 15 February 2014 Ó Springer Science+Business Media New York 2014 Abstract Room temperature tensile testing was performed on a coarse-grained polycrystalline Ni (32 lm), a nanocrys- talline Ni (23 nm) and two nanocrystalline Ni–Fe (16 nm) electrodeposits at two strain rates of 10 -1 and 10 -2 /s. Strain localizations and local temperature increases were simulta- neously recorded during tensile testing. For all materials, higher loads or higher strain rate generally resulted in higher peak temperature with the highest temperatures recorded in the fracture regions. The maximum temperature for the nanocrystalline materials was just over 80 °C, which is sig- nificantly below the reported temperatures for the onset of thermally activated grain growth. Therefore, the previously reported grain growth observed on similar materials after tensile deformation is likely not thermally activated but a stress-induced phenomenon. Despite the wide grain range from 16 nm to 32 lm, all samples exhibited similar strain localization behavior. Local strain variations initiated in the early stage of macroscopic uniform deformation, subsequent necking and fracture took place in the region of initial strain localization. While the coarse-grained polycrystalline Ni exhibited little strain rate sensitivity, gradually increased strain rate sensitivity was observed for the 23 nm Ni and the two 16 nm Ni–Fe samples, suggesting that both dislocation- mediated and grain-boundary-controlled mechanisms were operative in the deformation of the nanocrystalline Ni and Ni–Fe samples. Introduction Many past papers have reviewed the effect of grain size on the mechanical properties of nanocrystalline metals (e.g., [13]). As per the Hall–Petch grain size hardening mech- anism, nanometals exhibited significant increases in yield strength and hardness (e.g., [46]). However, limited ten- sile ductility was usually observed for nanocrystalline materials in comparison with their conventional coarse- grained counterparts due to their low capacity for strain hardening [2, 7]. Low tensile ductility of nanomaterials is almost invariably associated with early onset of localized deformation, e.g., necking and/or shear bands [811]. While a great deal of effort has focused on the underlying deformation mechanisms (e.g., [1217]), there is limited experimental investigation on the phenomena of localized strain and heat generation during plastic deformation of these materials, which could have a significant effect on the microstructure and mechanical properties [18]. There is considerable debate over the possibility of thermally activated grain growth due to heat generation associated with localized strain of specimens during defor- mation. Previous studies have reported grain growth during tensile testing in nanocrystalline Ni [10], Ni–Fe [18] and Co–P [19], all produced by Integran’s electrodeposition process [2022]. For example, nanocrystalline Ni (grain size: 20 nm) was tested in tension at strain rates between 10 -5 and 10 -1 /s [10]. Transmission electron microscopy in one of the shear band regions at the higher strain rates revealed considerable grain growth. It was speculated that grain growth in nanocrystalline Ni was due to the kinetic energy released during high-speed deformation, which could result in local hot spots within the shear band reaching temperatures on the order of 300 °C[10]. At this tempera- ture, significant grain growth is expected to take place in the T. Chan Y. Zhou (&) U. Erb Department of Materials Science and Engineering, University of Toronto, 184 College Street, Suite 177, Toronto, ON M5S 3E4, Canada e-mail: [email protected] I. Brooks G. Palumbo Integran Technologies Inc, 6300 Northam Drive, Mississauga, ON L4V 1H7, Canada 123 J Mater Sci (2014) 49:3847–3859 DOI 10.1007/s10853-014-8099-1

Upload: prasad-raikar

Post on 30-Sep-2015

5 views

Category:

Documents


1 download

TRANSCRIPT

  • Localized strain and heat generation during plastic deformationin nanocrystalline Ni and NiFe

    T. Chan Y. Zhou I. Brooks G. Palumbo

    U. Erb

    Received: 29 November 2013 / Accepted: 31 January 2014 / Published online: 15 February 2014

    Springer Science+Business Media New York 2014

    Abstract Room temperature tensile testing was performed

    on a coarse-grained polycrystalline Ni (32 lm), a nanocrys-talline Ni (23 nm) and two nanocrystalline NiFe (16 nm)

    electrodeposits at two strain rates of 10-1 and 10-2/s. Strain

    localizations and local temperature increases were simulta-

    neously recorded during tensile testing. For all materials,

    higher loads or higher strain rate generally resulted in higher

    peak temperature with the highest temperatures recorded in

    the fracture regions. The maximum temperature for the

    nanocrystalline materials was just over 80 C, which is sig-nificantly below the reported temperatures for the onset of

    thermally activated grain growth. Therefore, the previously

    reported grain growth observed on similar materials after

    tensile deformation is likely not thermally activated but a

    stress-induced phenomenon. Despite the wide grain range

    from 16 nm to 32 lm, all samples exhibited similar strainlocalization behavior. Local strain variations initiated in the

    early stage of macroscopic uniform deformation, subsequent

    necking and fracture took place in the region of initial strain

    localization. While the coarse-grained polycrystalline Ni

    exhibited little strain rate sensitivity, gradually increased

    strain rate sensitivity was observed for the 23 nm Ni and the

    two 16 nm NiFe samples, suggesting that both dislocation-

    mediated and grain-boundary-controlled mechanisms were

    operative in the deformation of the nanocrystalline Ni and

    NiFe samples.

    Introduction

    Many past papers have reviewed the effect of grain size on

    the mechanical properties of nanocrystalline metals (e.g.,

    [13]). As per the HallPetch grain size hardening mech-

    anism, nanometals exhibited significant increases in yield

    strength and hardness (e.g., [46]). However, limited ten-

    sile ductility was usually observed for nanocrystalline

    materials in comparison with their conventional coarse-

    grained counterparts due to their low capacity for strain

    hardening [2, 7]. Low tensile ductility of nanomaterials is

    almost invariably associated with early onset of localized

    deformation, e.g., necking and/or shear bands [811].

    While a great deal of effort has focused on the underlying

    deformation mechanisms (e.g., [1217]), there is limited

    experimental investigation on the phenomena of localized

    strain and heat generation during plastic deformation of

    these materials, which could have a significant effect on the

    microstructure and mechanical properties [18].

    There is considerable debate over the possibility of

    thermally activated grain growth due to heat generation

    associated with localized strain of specimens during defor-

    mation. Previous studies have reported grain growth during

    tensile testing in nanocrystalline Ni [10], NiFe [18] and

    CoP [19], all produced by Integrans electrodeposition

    process [2022]. For example, nanocrystalline Ni (grain

    size: 20 nm) was tested in tension at strain rates between

    10-5 and 10-1/s [10]. Transmission electron microscopy in

    one of the shear band regions at the higher strain rates

    revealed considerable grain growth. It was speculated that

    grain growth in nanocrystalline Ni was due to the kinetic

    energy released during high-speed deformation, which

    could result in local hot spots within the shear band reaching

    temperatures on the order of 300 C [10]. At this tempera-ture, significant grain growth is expected to take place in the

    T. Chan Y. Zhou (&) U. ErbDepartment of Materials Science and Engineering,

    University of Toronto, 184 College Street, Suite 177,

    Toronto, ON M5S 3E4, Canada

    e-mail: [email protected]

    I. Brooks G. PalumboIntegran Technologies Inc, 6300 Northam Drive,

    Mississauga, ON L4V 1H7, Canada

    123

    J Mater Sci (2014) 49:38473859

    DOI 10.1007/s10853-014-8099-1

  • 20 nm Ni electrodeposits. Calorimetric studies have shown

    that peak growth temperatures for major grain growth are in

    the range from 280 to 290 C for the Ni electrodeposits withaverage as-deposited grain size between 15 and 30 nm [23].

    Moreover, grain growth activity was observed in a sequence

    of relaxation and subgrain coalescence processes when

    20 nm Ni electrodeposits were subjected to annealing at

    lower temperatures, as low as 120 C [24]. Higher thermalstability was observed in nanocrystalline NiFe. For

    example, the Ni-10 wt% Fe (grain size 15 nm) and

    Ni-20 wt% Fe (grain size 13 nm) displayed peak tempera-

    tures of grain growth around 334 and 379 C, respectively,indicating a stabilizing effect by alloying nanocrystalline Ni

    electrodeposits with Fe [23]. These experimental observa-

    tions provide a basis to assess the possibility of having

    potential grain growth resulting from localized heating in

    nanocrystalline nickel-based electrodeposits during room

    temperature deformation.

    The main focus of the current study was (1) to establish

    a correlation between localized strain and heating, and (2)

    to assess the possibility of having thermally activated grain

    growth in this type of nanomaterials. Heat dissipation and

    strain localization during in situ tensile testing at different

    strain rates were investigated on nanocrystalline Ni and

    NiFe electrodeposits produced through the same produc-

    tion facility and using the same synthesis process as those

    for the samples used in Refs. [10, 18, 19].

    Experimental

    The materials used in this study consist of four batches of

    samples, coarse-grained polycrystalline nickel (commercially

    available Ni 200), and three nanocrystalline electrodeposits,

    Ni and two NiFe alloys with different Fe contents. All

    nanocrystalline electrodeposits were prepared by Integran

    Technologies Inc, which used the same process to produce

    the samples for the investigations in Refs. [10, 18, 19].

    Samples from the same nanocrystalline batches were already

    characterized in previous studies [25, 26], using transmission

    electron microscopy and X-ray line broadening measure-

    ments (Scherrer approach) [27], as well as energy-dispersive

    X-ray spectroscopy (EDS). For the polycrystalline Ni, the

    grain size was determined using standard metallography.

    Tensile coupons of 1.0 mm thickness were machined by

    the electric discharge machining (EDM) method to produce

    dog-bone-shape tensile testing samples with dimension

    details as shown in Fig. 1. There was a minor modification in

    the geometry of tensile coupons in the current study with

    respect to ASTM E8 standard. The coupon tab width was

    widened from 10 to 20 mm (Fig. 1) to provide increased

    gripping area after slipping problems owing to the extremely

    high material strength were experienced, while the rest,

    especially the critical gauge section, was kept identical to that

    of ASTM E8.

    All materials were tested at tensile strain rates of 10-2

    and 10-1/s on a MTS servo-hydraulic testing machine

    combined with the in situ measurements of local strain and

    temperature across the gauge length, using the digital

    image correlation (DIC) technique and an infrared camera,

    respectively. Prior to tensile testing, each sample was

    prepared by spraying a layer of black paint on one side to

    equalize emissivity for better infrared detection, thus more

    accurate temperature measurement. On the opposite side,

    samples were lightly abraded using 320 silicon carbide

    paper for the measurement of localized strains. Localized

    strain patterns during deformation were captured by a high

    resolution DIC Camera (Allied Vision Technologies),

    while the global strain was obtained based on the crosshead

    displacement of the MTS machine frame. Samples with

    abraded surfaces facing toward the CCD camera were

    illuminated by a source of white light during the entire

    tensile testing. The images were post-processed through a

    DIC system to calculate the displacement and the localized

    strain distributions. Localized strains contours were then

    mapped along the sample gauge over a length of 40 mm.

    Changes of temperature in the samples were measured

    using a pre-calibrated high resolution infrared camera

    (Deltatherm 1410) during tensile deformation. Frame rates

    of the infrared detector were set to 15 frames per second to

    best capture the temperature profiles along sample gauges.

    Images captured during the experiment were post-pro-

    cessed through a Thermoelastic Stress Analysis (TSA)

    system to locate the temperature increase along the gauge

    section during tensile testing. The maximum temperature

    along the sample gauges was recorded for all samples.

    Results and discussion

    Grain size and chemical composition

    Table 1 summarizes the chemical compositions and grain

    sizes for the samples used in the current study. For the

    coarse-grained Ni, i.e., commercial Ni 200, the average

    grain size of 32 lm was obtained after analysis of a seriesof optical micrographs and measurement of over 400

    grains. For the nanocrystalline materials, the nano-Ni had a

    Fig. 1 Tensile coupon dimensions in mm

    3848 J Mater Sci (2014) 49:38473859

    123

  • grain size of 23 nm, while the two NiFe samples both had

    a grain size of 16 nm as per XRD line broadening mea-

    surements [25, 26].

    Deformation at different tensile strain rates

    All materials were tested at strain rates of 10-1 and 10-2/s.

    Figure 2 shows the respective engineering stressstrain

    curves for all samples.

    At the rate of 10-2/s, the coarse-grained Ni (grain size:

    32 lm) showed a yield strength (YS) of 208 MPa, ultimatetensile strength (UTS) of 442 MPa and considerable duc-

    tility with total elongation (or total global strain) of 50.4 %.

    With the refinement of grain sizes, there is a significant

    increase in YS and UTS due to the HallPetch grain size

    strengthening mechanism as indicated in Fig. 2a. The

    nanocrystalline Ni (grain size: 23 nm) showed values of

    YS (950 MPa) and UTS (1504 MPa), about 4 times higher

    than the coarse-grained Ni. The two remaining 16 nm

    NiFe samples exhibited different strength values, YS of

    1010 and 1130 MPa, as well as UTS of 1579 and 1741 MPa,

    for the Ni-2.6 wt% Fe and Ni-8.5 wt% Fe, respectively. The

    fact that higher strength was observed at higher Fe content

    suggests that there is some solution-strengthening effect due

    to Fe solute for the NiFe samples. In addition, all

    nanocrystalline samples showed limited ductility with

    measured elongation to fracture not exceeding about 8 %.

    The same trends in the tensile properties were found at

    the strain rate of 10-1/s. Significant strength enhancements

    were attained, whereas the total elongation or ductility was

    limited for the nanocrystalline samples in comparison with

    the coarse-grained Ni (Fig. 2b). These observations of high

    strength and reduced ductility are in agreement with

    numerous previous experimental studies summarized in

    previous review papers (e.g., [2, 28]). While the HallPetch

    relationship accounts for the strengthening, low tensile

    ductility is generally attributed to the lack of strain hard-

    ening required to resist macroscopic strain localization and

    extend uniform plastic flow for nanocrystalline materials

    (e.g., [9, 11]).

    Table 2 summarizes tensile property values of all sam-

    ples at the two strain rates, 10-2 and 10-1/s. For the coarse-

    grained Ni (32 lm), strain rate sensitivity is insignificant;the key tensile properties, yield strength, ultimate tensile

    strength and elongation remain virtually unchanged. When

    varying the strain rate from 10-2 to 10-1/s, the nanocrys-

    talline Ni (23 nm) showed YS increase from 950 to

    970 MPa but essentially no change in UTS, indicating

    some minor strain rate sensitivity. In comparison, the two

    NiFe samples with smaller grain size of 16 nm displayed

    evident sensitivity to the strain rate, showing an increase in

    both YS and UTS at the higher strain rate.

    Although a detailed analysis of strain rate sensitivity

    was not the major focus of the current study, it should be

    pointed out that the observed grain size dependent strain

    rate sensitivity is in general agreement with previous

    studies on nickel-based alloys (e.g., [28]). In a recent study

    [29], the strain rate sensitivity, m, was examined over a

    wide grain size range from 10 nm to 100 lm for Ni andNi-based alloys based upon extensive results from various

    Table 1 Sample grain size and composition

    Sample Fe content

    (wt%)

    Average

    grain size

    1. Coarse-grained Ni (Ni 200) 0 32 lm

    2. Nanocrystalline Ni 0 23 nm

    3. Nanocrystalline NiFe 1 2.6 16 nm

    4. Nanocrystalline NiFe 2 8.5 16 nm

    Fig. 2 Stress-strain curves of all samples deformed at a strain rate of a 10-2/s and b 10-1/s

    J Mater Sci (2014) 49:38473859 3849

    123

  • studies with different testing methods, e.g., tensile tests at

    different stain rates, stress relaxation [30] and strain rate

    jump tests [31]. It was observed that the m values increased

    significantly and ranged from 0.011 to 0.033 for materials

    with average grain sizes between 10 and 50 nm, while in

    the broad 50 nm100 lm range, the m value showed lim-ited variation and remained below 0.005 [29] and within

    the reported 0.0010.004 range for coarse-grained Ni [10].

    This explains the insensitivity toward the strain rate for the

    coarse-grained Ni (32 lm), and gradually increaseddependence on the strain rate for the 23 nm Ni and the two

    16 nm NiFe alloys (Table 2).

    The significant m value increase at very small grain

    sizes for Ni-based materials can be attributed to deforma-

    tion mechanisms other than the well-established disloca-

    tion-mediated (DM) mechanisms for coarse-grained

    materials, owing to the close association between strain

    rate sensitivity and underlying deformation mechanisms

    (e.g., [3236]). Numerous experimental and simulation

    studies have provided convincing evidence that dislocation

    activity is negligible and deformation is dominated by

    grain-boundary-controlled (GBC) mechanisms, such as

    grain boundary sliding or grain rotation for nanocrystalline

    materials with average grain sizes less than 10 nm (e.g.,

    [28, 37, 38]). Therefore, it is plausible that in the current

    study, DM mechanisms dominated in the deformation of

    the coarse-grained Ni, whereas the 23-nm Ni and the two

    16-nm NiFe samples deformed through both DM and

    GBC mechanisms.

    Strain localization

    For conventional metallic materials, plastic deformation

    during uni-axial tensile testing is normally expected to be

    distributed uniformly at applied stresses lower than the

    UTS; macroscopic localized deformation, necking, occurs

    at the stress beyond the UTS [39]. In reality, the amount of

    plastic deformation can vary from one location to another

    well before the onset of necking. Digital image correlation

    (DIC) strain measurement [40] has been shown to be an

    effective approach to examine in situ strain localization of

    materials, i.e., it is capable of capturing the evolution of

    localized strain during tensile testing. For example in an

    earlier study, rapid evolution of shear banding was cap-

    tured using the DIC approach for fully dense electrode-

    posited nanocrystalline Ni (grain size: 20 nm) during an

    in situ tensile testing at quasi-static strain rate of 10-4/s

    [41].

    Figure 3 shows the results of DIC measurements at

    different stages of deformation represented by five global

    strains for the coarse-grained polycrystalline Ni of 32 lmduring tensile testing at a strain rate of 10-1/s. In the

    contour map (Fig. 3a), the levels of local strain from low to

    high are indicated by colors in a sequence of purple, blue,

    green, yellow and red and the local strain field contours of

    the coarse-grained Ni are mapped along the gauge section

    at selected global strains indicated at the bottom of Fig. 3a,

    up to the fracture strain of just over 50 %. Although uni-

    form deformation is expected for the sample within the

    global strain up to UTS at 45 % (Table 2), noticeable

    variations of strain are displayed within the gauge length

    even at low global strain of 15 % as indicated by two

    colors, purple and blue.

    As the global strain increases, the colors in the gauge

    section keep changing due to intensified localized defor-

    mation. At the global strain of 35 %, still well within the

    range of uniform deformation, the gauge section already

    shows a range of different contour colors, indicating sig-

    nificant levels of localized strain. The highest level of

    localization occurred during necking when imposed

    deformation was concentrated essentially only in the

    necking region. This is demonstrated in the contour map at

    the fracture strain of 50 %, whereby the red area corre-

    sponds to the region of necking with very high local strain,

    leading to fracture at the end of the tensile test (Fig. 3a).

    Corresponding strain distributions along the gauge

    length at the same global strains, i.e., 15, 25, 35, 45 and

    50 %, respectively, are shown in Fig. 3b to present more

    detailed information on the evolution of strain localization.

    Note that the global strains do not always match with local

    Table 2 Tensile properties at strain rate of 10-2 and 10-1/s

    Samples Grain size Strain

    rate (/s)

    Yield strength

    (MPa)

    UTS

    (MPa)

    Elongation

    at UTS (%)

    Total

    elongation (%)

    Poly-Ni 32 lm 10-2 208 442 45 50.4

    10-1 208 445 45 51

    Nano-Ni 23 nm 10-2 950 1504 5.5 8.2

    10-1 970 1508 5.5 8

    Nano-Ni-2.6 % Fe 16 nm 10-2 1010 1579 5.5 7.6

    10-1 1100 1763 5.5 7.5

    Nano-Ni-8.5 % Fe 16 nm 10-2 1130 1741 5.5 7.7

    10-1 1150 1867 5.5 7.5

    3850 J Mater Sci (2014) 49:38473859

    123

  • strain values, especially at low stress/strain levels. This

    difference is somehow expected because of different

    methods used in the strain measurements; global strain was

    obtained indirectly from the crosshead displacement of the

    tensile machine frame, while local strain values were

    measured directly in the sample gauge section through the

    DIC technique. It has been shown that strain values based

    on crosshead displacement include contributions from

    other sources, e.g., machine compliance and materials

    outside the gauge length (e.g., those close to grip sections),

    in addition to the strain of the sample gauge length [42]. As

    a result, the measured strain based on the crosshead dis-

    placement almost always exceeds the actual strain in the

    sample gauge section [42]. However, a discussion of

    crosshead strain vs DIC strain is beyond the scope of this

    study. The focus here is instead on the strain localization

    within the stage of macroscopic uniform elongation, i.e.,

    up to the global strain at UTS, a characteristic deformation

    feature that can be reliably identified using crosshead

    displacement. Hence, the disparity between global and

    local strain has little impact in addressing strain localiza-

    tion for the samples in the current study.

    Examination of the strain distribution curves reveals that

    there are two deformation regions within the gauge section:

    the region at the two ends, Region 1 and the region in the

    center, Region 2. Due to its proximity to the grip section,

    Region 1 is strongly influenced by its geometric confine-

    ment and has to accommodate the transition from close to 0

    strain at the grip section to the imposed strain on the

    sample. In other words, two factors contribute to the local

    strain in Region 1, geometric conformation and material

    properties. It appears that the geometric contribution

    dominates the local strain in Region 1 as demonstrated by

    the largely linear relationship between gauge location and

    local strain with a slope increasing with the respective

    imposed strain or global strain (Fig. 3b). This linear rela-

    tionship of Region 1 ends where the strain transition to the

    imposed value is accomplished; the end locations depend

    Poly Ni (0.1/s)

    0

    20

    40

    60

    80

    0 10 20 30 40Gauge length, mm

    Loca

    l str

    ain,

    %

    50%45%35%25%15%

    (a)

    (b)

    Region 2Region 1 Region 1

    Fig. 3 Strain distribution at different global strains for the coarse-grained Ni (grain size: 32 lm), a strain field contour maps, andb strain distributions within the gauge length at five levels of global

    strain. Region 2 is approximately located between the two red-dotted

    lines (Color figure online)

    J Mater Sci (2014) 49:38473859 3851

    123

  • on the global strain and manifest themselves in the area

    with evident slope change. For example at the UTS global

    strain of 45 %, the locations are approximately at the gauge

    locations of 10 and 33 mm, respectively, as shown in

    Fig. 3b. Region 2 is located between these two points,

    where the geometric contributions diminish and local

    strains result entirely from the material properties. Hence,

    the analysis of strain localization evolution should focus on

    the strain distributions in Region 2, identified approxi-

    mately between the two dotted lines in Fig. 3b.

    For the coarse-grained Ni, strain distribution appears

    fairly uniform in most locations of Region 2 at the global

    strain of 15 %. Slightly different local strain was displayed

    between the gauge locations of 26 and 32 mm (Fig. 3b).

    This strain disparity became more visible in the same

    location at the global strain of 25 %, and gradually evolved

    into a broader valley, with a local strain significantly dif-

    ferent from the rest in Region 2 at 35 % global strain, even

    though the sample was still well within the strain range of

    macroscopic uniform deformation (\45 %). When theapplied stress was above UTS, necking took place and

    subsequent deformation concentrated in this area as dem-

    onstrated in the local strain distribution at the global strain

    of 50 % (Fig. 3b). Immediately thereafter, fracture occur-

    red at the location with peak strain of about 80 %, sub-

    stantially higher than the global fracture strain just over

    50 % (Fig. 3a, b).

    A similar development of localized strain is observed in

    the nanocrystalline Ni sample with 23 nm grain size. As

    shown in the strain field contour maps of Fig. 4a, deforma-

    tion appears to be fairly uniform across the gauge length in

    the early stage up to a global strain of 2.5 %. The sample

    exhibited notable differences in local strains at a global strain

    of 3.5 % as indicated by two colors, dark and light blue. The

    local strain became more pronounced at the UTS strain of

    5.5 % (Fig. 4a). Upon necking, imposed deformation was

    concentrated in the center, and at about 8.0 % global strain,

    fracture occurred in the necking region (red area) at an angle

    Nano Ni (0.1/s)

    0

    3

    6

    9

    12

    15

    0 5 10 15 20 25 30 35 40

    Gauge length, mm

    Loca

    l str

    ain,

    %

    7.5%5.5%4.5%3.5%2.5%

    (b)

    (a)

    Region 2Region 1

    Region 1

    Fig. 4 Strain distribution at different global strains for the nanocrystalline Ni (grain size: 23 nm), a strain field contour maps, and b straindistributions within the gauge length. Region 2 is approximately located between the two red-dotted lines (Color figure online)

    3852 J Mater Sci (2014) 49:38473859

    123

  • of 62 (dashed line) as shown in Fig. 4a. In comparison withthe coarse-grained Ni, necking on the nanocrystalline Ni

    sample is less pronounced. Details on strain localization

    evolution can be more clearly demonstrated from strain

    distributions along the gauge length at given global strains.

    Figure 4b shows the distributions at five global engineering

    strain levels of 2.5, 3.5, 4.5, 5.5 and 7.5 % for the nano-

    crystalline Ni sample tested at a strain rate of 10-1/s.

    Similar to the deformation of coarse-grained Ni, the

    strain distribution in Fig. 4b can be divided into two regions

    for the nanocrystalline Ni: Region 1 at the ends of the gauge

    length with strong influence from sample geometric con-

    finement and Region 2 (identified approximately between

    the two red-dotted lines) representing intrinsic material

    properties. It can be seen that the nanocrystalline Ni already

    exhibited slight local strain variation somewhere between

    the gauge length of 17 and 25 mm in Region 2 even at the

    global strain as low as 2.5 %. As the global strain increased

    from 3.5 to 5.5 %, i.e., still within the strain range of

    macroscopic uniform deformation, the strain disparity

    became more and more pronounced between this area and

    the rest in Region 2. In the final deformation stage of

    instability, the imposed deformation concentrated in this

    area in the form of necking with the peak local strain

    reaching about 14 % at the global strain of 7.5 %. At the

    slightly higher global strain of 8.0 %, the sample fractured

    at the location with the peak local strain (Table 2; Fig. 4b).

    For the two 16 nm NiFe electrodeposits and all mate-

    rials tested at a strain rate of 10-2/s, strain localization was

    observed in a similar manner. Generally, all samples

    showed not much difference in the development of strain

    localization during tensile testing at both strain rates.

    Despite the wide grain size range (16 nm to 32 lm) andvery different tensile strengths and measured elongation to

    fracture values, strain localization invariably started during

    the stage of macroscopic uniform deformation, much

    earlier than the occurrence of the macroscopic strain

    localization in all of them. Subsequent necking took place

    close to the region where the strain localization initiated.

    There is a good agreement on the strain localization

    behavior observed here for the nanocrystalline samples and

    a recent DIC investigation, also on electrodeposited

    nanocrystalline Ni [41]. During tensile testing of a 20 nm

    Ni, initiation of shear localization was observed during the

    uniform deformation stage. Rapid shear-banding evolu-

    tion occurred immediately after necking, the macroscopic

    strain localization and was concentrated in the middle of

    the gauge area, leading ultimately to sample fracture [41].

    Stress-induced heat generation

    When metals are plastically deformed, part of the plastic

    work is dissipated as heat during the deformation process

    (e.g., [43, 44]). In the current study, temperature change

    due to heat release was measured and recorded by the IR

    detector at different global strain levels. Figure 5a shows

    selected infrared images of coarse-grained Ni at various

    global strain levels, indicated at the bottom in Fig. 5a of

    the individual snapshots at a strain rate of 10-1/s.

    With the increase of global engineering strain from 15 to

    45 %, the localized temperature along the sample gauge

    increased continuously as indicated by the color changes

    from light blue to yellow and orange. Temperatures are

    distributed in a relatively symmetrical fashion along the

    gauge with the maximum located in the central region; the

    dashed line at a global strain of 50 % indicates the position

    of fracture, which coincides with the area of the highest

    temperature (i.e., red color) and/or most heat dissipation.

    Figure 5b shows the relationship between global strain

    and the instantaneous increase in the maximum tempera-

    ture, DTmax, on the gauge section during tensile testing atthe strain rate of 10-1/s. During the stage of macroscopic

    uniform deformation, DTmax reached approximately42 C at the UTS strain of 45 % (eu in Fig. 5b, i.e., theonset strain of necking). For the initial range of deforma-

    tion (045 %), the average rate of temperature increase was

    low, \1 C for each percent of global strain. Subsequentdeformation resulted in a rapid local temperature increase

    in the necking region up to 70 C for the DTmax at fracturestrain of 51 %. With the addition of the room temperature

    of 25 C, the recorded maximum temperature in the gaugelength, Tmax, was approximately 95 C for coarse-grainedNi at the end of room temperature tensile testing.

    The measurements for the 23 nm grain size nanocrys-

    talline Ni during tensile testing at the strain rate of 10-1/s

    are summarized in Fig. 6. At global strains below 3.5 %,

    there was a very limited amount of heat generation as

    indicated by little color change in the corresponding

    infrared images (Fig. 6a). The nanocrystalline Ni displayed

    a discernible temperature increase starting at a global

    engineering strain of 4.5 %. In the final stage of defor-

    mation, a narrow region of high temperature formed around

    the fracture location (the dashed line at an angle of 62along the loading axis), indicating a highly concentrated

    hot spot for the nanocrystalline Ni. This is somewhat dif-

    ferent from the coarse-grained Ni, for which a much

    broader and more or less uniform temperature distribution

    was observed in the deformation stage close to fracture

    (Fig. 5a).

    Regarding the evolution of the DTmax, the nanocrystal-line Ni exhibited a qualitatively similar profile as the

    coarse-grained Ni. As shown in Fig. 6b, the temperature

    increase was low during the uniform deformation stage.

    A steep climb in temperature initiated at the onset of

    necking, 5.5 % global strain (eu), and all the way to thepoint of fracture (8 %). Specifically, the net increase in

    J Mater Sci (2014) 49:38473859 3853

    123

  • temperature was approximately 12 C during the first stageof deformation from 0 to 5.5 % global strain, while a

    maximum DTmax of 58 C was recorded at the strain veryclose to the fracture strain of 8.0 % (Fig. 6b). With the

    addition of 25 C for room temperature, the recordedmaximum temperature, Tmax, was therefore 83 C for thenanocrystalline Ni during tensile testing.

    The two NiFe samples generally exhibited heat gen-

    eration behavior similar to the nanocrystalline Ni; the ini-

    tial temperature increase before necking was low, and a

    rapid increase was observed after necking up to fracture. In

    addition, the generated heat was concentrated in the narrow

    area of the gauge center, surrounding the ultimate fracture

    line for the two nanocrystalline NiFe samples.

    Table 3 summarizes the maximum recorded tempera-

    tures, Tmax, along with the global elongation recorded at

    both strain rates for all materials. It is evident that all

    samples experienced local temperature increases during

    tensile testing, showing Tmax values noticeably higher than

    the ambient testing temperature. Furthermore for a given

    sample, the attained maximum temperature, Tmax, is

    reduced at lower tensile strain rate. This apparent effect of

    strain rate on Tmax reveals the critical influence of heat

    conduction on the process of strain-induced heat

    generation.

    As shown in Table 2, the yield strength, tensile strength

    and elongation to fracture of the coarse-grained polycrys-

    talline Ni remained virtually the same at the two strain

    rates employed, namely 10-2 and 10-1/s. For the material

    under scrutiny, same amount of strain-induced heat should

    be generated during tensile testing at the both strain rates

    due to identical sample size. Because of strain localization,

    heat generation was generally not uniform and varied from

    one location to another, resulting in heat flow from higher

    to lower temperature regions. During tensile testing, the

    sample grip sections experienced little deformation and

    thus functioned mainly as heat sinks for any strain-induced

    heat. In the final stage of necking, most of the heat was

    generated in the necking region, i.e., the gauge center, and

    was then conducted toward the grip sections at both ends of

    the samples. Due to the finite heat conduction of Ni, tem-

    perature distribution became a strong function of testing

    0

    100

    200

    300

    400

    500

    0 10 20 30 40 50 60

    Global strain/Elongation, %

    Eng.

    stre

    ss, M

    Pa

    0

    20

    40

    60

    80

    Tm

    ax ,

    0 C

    u

    (a)

    (b)

    600

    550

    500

    450

    400

    350

    300

    250

    200

    Fig. 5 Temperature change incoarse-grained Ni, a infraredimages with the dashed line at

    50 % strain indicating the final

    fracture position, and

    b engineering stressstraincurve and the maximum

    temperature increase, DTmax, ata strain rate of 10-1/s. Note that

    the numbers on the color scale

    bar are temperature in Celsius

    (Color figure online)

    3854 J Mater Sci (2014) 49:38473859

    123

  • duration, and thus, the recorded maximum temperature

    depended on strain rate. Higher strain rate is equivalent to

    shorter test duration and therefore more concentrated

    in situ temperature distribution, leading to higher maxi-

    mum temperature for a given material with the same input

    of mechanical energy. Hence, the coarse-grained Ni

    exhibited higher Tmax at the higher strain rate (Table 3).

    Likewise, the different Tmax values observed at the two

    strain rates for the 23 nm Ni and the two 16 nm NiFe

    samples can be mainly attributed to the heat conduction. In

    addition, temperature distribution/gradients in the nano-

    crystalline materials (Fig. 6a) are more concentrated/stee-

    per than those in the coarse-grained counterpart (Fig. 5a).

    This likely resulted from the facts that for the nanocrys-

    talline electrodeposits, the duration of tensile testing was

    shorter due to their reduced ductility, and the thermal

    conductivity is lower, e.g., up to 25 % reduction as

    observed in our recent study for a series of nanocrystalline

    Ni with average grain sizes of 2550 nm [45].

    Implication of the measured heat generation

    Returning to the main focus of this study, the measured

    Tmax values observed as a result of strain-induced heat

    generation do not support thermal activation as an under-

    lying mechanism for the grain growth during room tem-

    perature tensile testing reported in previous studies (e.g.,

    [10, 18]). For instance, the highest temperature recorded in

    the 23 nm grain size nanocrystalline Ni was approximately

    83 C during tensile testing, well below 120 C, the

    0

    400

    800

    1200

    1600

    0 2 4 6 8Global strain/Elongation, %

    Eng.

    stre

    ss, M

    Pa

    0

    10

    20

    30

    40

    50

    60

    70

    Tm

    ax ,

    0 C

    u

    (a)

    (b)

    600

    550

    500

    450

    400

    350

    300

    250

    200

    Fig. 6 Temperature change inthe 23-nm nanocrystalline Ni,

    a infrared images with thedashed line at 8 % strain

    indicating fracture location, and

    b engineering stressstraincurve versus the maximum

    temperature increase at a strain

    rate of 10-1/s. Note that the

    numbers on the color scale bar

    are temperature in Celsius

    (Color figure online)

    Table 3 Localized strain and maximum temperature recorded duringtensile testing

    Samples Grain

    size

    Strain

    rate (/s)

    Global strain (%) Tmax(C)

    Total At UTS

    Poly-Ni 32 lm 10-2 50.4 45 50

    10-1 51 45 95

    Nano-Ni 23 nm 10-2 8.2 5.5 59

    10-1 8 5.5 83

    Nano-Ni-2.6 % Fe 16 nm 10-2 7.6 5.5 57

    10-1 7.5 5.5 79

    Nano-Ni-8.5 % Fe 16 nm 10-2 7.7 5.5 59

    10-1 7.5 5.5 81

    J Mater Sci (2014) 49:38473859 3855

    123

  • reported onset temperatures for the initiation of micro-

    structure evolution, e.g., by subgrain coalescence, for

    20 nm grain size Ni electrodeposits [24]. Moreover, the

    highest measured Tmax value was approximately 80 C forthe two nanocrystalline NiFe samples, while their grain

    growth temperature was reported to be 50100 C higherthan that of the nanocrystalline Ni due to the stabilizing

    effect of the alloying element Fe (e.g., [23]). Evidently

    based on the measured Tmax alone, the strain-induced heat

    release is likely not sufficient to activate migration of grain

    boundaries.

    One uncertainty regarding the measurement of Tmax is the

    spatial resolution of the infrared camera. As shown in Ref.

    [10], grain growth during tensile testing may be confined in

    a narrow band area less than half a micrometer across. It is

    unlikely that the infrared camera with a spatial resolution of

    about 50 lm used in the current study can detect tempera-ture differences within a submicron area; the measured Tmaxwas actually an average value over a larger area. It is pos-

    sible that a small area in the deformation core attained an

    instant local temperature higher than the measured Tmaxduring the tests. The temperature difference between the

    maximum and minimum value in the resolution unit area,

    DTRes, may be estimated using a 1D heat conduction model,defined by Fouriers law, along the cross-section normal,

    DQDt

    jA DTDx

    1

    where j and A are, respectively, thermal conductivity andcross-sectional area; DQ/Dt is the heat flux, leading to thetemperature gradient, DT/Dx, along the heat conductionpath. The heat flux is associated with the rate of

    temperature change, b = DT/Dt, through heat capacity Cpat constant pressure,

    DQDt

    Cp DTDt Cpb 2

    Then, DTRes in the area can be assessed as follow bycombining Eqs (1) and (2), replacing Cp with specific heat

    cp through Cp = mcp = qALcp, and assuming asymmetrical areal temperature distribution (Dx = L/2)and negligible ambient heat exchange,

    DTRes qcpL2

    2jb 3

    where L and q are the feature length of the resolution area(50 lm) and material density, respectively. Some of theparameters, e.g., cp and b, can have a range of values. In such

    cases, the values are chosen to maximize DTRes for thenanocrystalline Ni. Here, the upper value of cp = 0.56 J/g K

    is used, based on the reported cp value range of

    0.460.56 J/g K within 25200 C for nanocrystalline Ni[46]. The handbook value of q = 8.89 g/cm3 for coarse-

    grained Ni [47] can be used here because the density of fully

    dense nanocrystalline Ni is grain size independent [48]. The

    reported value of j is 90 W/m K for coarse-grained Ni [49].Grain size reduction to about 25 nm resulted in 25 %

    reduction of j or j = 67.5 W/m K for the nanocrystallineNi-based on recent observations [45]. The value of b can be

    obtained from the slope in the plot of temperature vs strain.

    For example in Fig. 6b, the largest slope section is located

    at the top or the final stage of deformation, corresponding to

    a rate of 635 K/s. Consequently, these values can be used to

    calculate an upper bound value of DTRes of approximately0.1 C. This is insignificant with respect to the measuredTmax of 83 C for the nanocrystalline Ni. In other words, theuncertainty due to the spatial resolution of the infrared

    camera is expected to be insignificant in the Tmax mea-

    surements in the current study.

    Another critical parameter is the time or duration avail-

    able for thermally activated grain growth. In Ref. [10], the

    observed growth from 20 to 200 nm grain size during room

    temperature tensile testing at a strain rate of 103/s on the

    nanocrystalline Ni occurred over a duration shorter than

    10-4 s, corresponding to a rapid growth rate of more than

    106 nm/s. This is in contrast to the actual growth rate

    observed in an in situ TEM study on electrodeposited Ni

    (as-deposited grain size: 20 nm) isothermally annealed at a

    much higher temperature, 420 C [50]. In this study, it wasshown that grain boundary (GB) migration occurred as a

    series of discontinuous steps followed by periods of

    boundary stagnation; the GB velocity profile was repre-

    sented by a series of delta functions. With a video sampling

    rate of 30 frames per second, the highest recorded instan-

    taneous GB migration rate was 700 nm/s, while most grain

    boundaries migrated at a rate below 300 nm/s for the

    nanocrystalline Ni at 420 C [50]. Much lower averagegrowth rates were observed in an ex-situ TEM study on a

    series of nanocrystalline Ni-based electrodeposits, whereby

    the same Ni sample as in Ref. [45] showed the fastest

    average growth rate of only 33 nm/s after 5 s of isothermal

    annealing at 420 C [51]. Note that the nanocrystalline Nisamples used in Refs. [10, 50, 51] were all produced by the

    same Integran electrodeposition process and exhibited the

    same as-deposited average grain size (20 nm), similar grain

    size distributions and likely similar impurity (e.g., S and C)

    contents. Therefore, not much difference is expected in the

    thermal behavior and stability for the Ni samples of the

    three studies. Yet huge differences in growth rates, at least 3

    orders of magnitude (106 vs. 700 nm/s), were observed

    among the three studies. This indicates a different activation

    mechanism for the grain growth shown in Ref. [10] than the

    thermally activated growth observed in Refs. [50, 51].

    Overall, the measurements of Tmax in the current study

    and the growth kinetics comparison between previous

    studies on very similar materials suggest that thermally

    3856 J Mater Sci (2014) 49:38473859

    123

  • activated grain growth during room temperature tensile

    testing for nanocrystalline nickel is highly unlikely. Hence,

    the reported grain growth during tensile testing in previous

    studies may be attributed to other contributing processes,

    e.g., stress-induced grain growth.

    In early studies on superplastic deformation, stress-

    induced grain growth was reported in conventional coarse-

    grained materials (e.g., [52, 53]). Given the significance of

    grain size in nanostructures, this type of grain growth has

    been an on-going focus of numerous experimental and

    molecular dynamics simulation (MD) studies on nano-

    crystalline materials in recent years, [5466]. Evidence has

    been presented for stress-induced grain growth through

    direct experimental observations. For instance, stress-

    induced grain rotation and growth were observed during

    in situ TEM tensile testing on nanocrystalline Ni electro-

    deposits [54, 63]. Again, the materials used in Ref. [63]

    were produced by the same Integran electrodeposition

    process as the one in the current study. In another in situ

    TEM study, grain growth occurred upon nanoindentation in

    nanocrystalline Al films [57]. Additionally, grains were

    found to grow faster at cryogenic temperature than at room

    temperature in a study on strained nanocrystalline Cu [58].

    Further, it was recently concluded, based on the results of

    tensile testing coupled with transmission electron micros-

    copy on nanocrystalline Al films that grain growth was a

    shear stress-driven process [64]. Generally, the experi-

    mental observations were in agreement with the results

    from MD studies, e.g., on nanocrystalline Pd [54], Cu [56]

    and Ni [62, 66].

    Summary

    Temperature increases were measured during tensile test-

    ing of both coarse-grained Ni and nanocrystalline Ni and

    NiFe electrodeposits due to strain-induced heat genera-

    tion. Generally, higher strain values and/or higher strain

    rates resulted in higher observed peak temperature for each

    material, with the highest temperatures recorded in the

    regions of subsequent fracture. For the coarse-grained

    polycrystalline Ni, the maximum attained mechanically

    induced temperature was approximately 95 C, whereasthe highest temperature for the nanocrystalline electrode-

    posits was approximately 83 C. This is significantly belowthe reported threshold temperatures for the onset of ther-

    mally activated grain growth in nanocrystalline Ni and

    NiFe alloys. Therefore, thermally activated grain growth

    is highly unlikely for nanocrystalline Ni and NiFe during

    room temperature tensile testing. Previously reported grain

    growth events observed in similar materials are likely due

    to stress-driven grain boundary migration.

    As per DIC measurements, the samples with grain sizes

    ranging from 16 nm to 32 lm exhibited similar strainlocalization behavior. Local strain variations initiated in

    the early stages of macroscopic uniform deformation

    and became more pronounced close to the point of peak

    load or UTS. Subsequent necking took place in the region

    where significant strain localization initiated. Material

    grain size showed little influence on strain localization

    behavior. Some grain size dependent strain rate sensitivity

    was observed during the tensile testing at two strain rates of

    10-1 and 10-2/s. While the coarse-grained Ni exhibited

    insensitivity to strain rate, gradually increased strain rate

    sensitivity was demonstrated in the 23 nm grain size Ni

    and the two 16 nm grain size NiFe samples. This sensi-

    tivity suggests that both dislocation-mediated and grain-

    boundary-controlled mechanisms were operative in the

    deformation of the nanocrystalline Ni and NiFe samples.

    Acknowledgements The authors would like to thank Dr. DavidBackman, Mr. Richard Bos and Mr. Thomas Sears from the National

    Research Council, Institute of Aerospace Research for their help and

    many valuable suggestions during the DIC and infrared experiments.

    YZ would like to thank Mr. Cho for the fruitful discussion on thermal

    and electric transport in nanocrystalline metals. Highly appreciated is

    the financial support by the Natural Sciences and Engineering

    Research Council of Canada (NSERC) and the Ontario Research

    Fund (ORF).

    References

    1. Weertman JR (2007) Mechanical behavior of nanocrystalline

    metals. In: Koch CC (ed) Nanostructured materials, 2nd edn.

    William Andrew Publications, Norwich, NY, pp 537564

    2. Koch CC (2007) Structural nanocrystalline materials: an over-

    view. J Mater Sci 42:14031414

    3. Whang SH (ed) (2011) Nanostructured metals and alloys.

    Woodhead Publ Ltd, Oxford

    4. Chokshi AH, Rosen A, Karch J, Gleiter H (1989) On the validity

    of the Hall-Petch relationship in nanocrystalline materials.

    Scripta Metall Mater 23:16791684

    5. Siegel RW, Fougere GE (1995) Mechanical properties of nano-

    phase metals. Nanostr Mater 6:205216

    6. Erb U, Aust KT, Palumbo G (2007) Electrodeposited nanocrys-

    talline metals, alloys, and composites. In: Koch CC (ed) Nano-

    structured materials, 2nd edn. William Andrew Publications,

    Norwich, NY, pp 235292

    7. Kumar KS, Van Swygenhoven H, Suresh S (2003) Mechanical

    behavior of nanocrystalline metals and alloys. Acta Mater

    51:57435774

    8. Ebrahimi F, Zhai Q, Kong D (1998) Deformation and fracture of

    electrodeposited copper. Scripta Mater 39:315321

    9. Ebrahimi F, Bourne GR, Kelly MS, Matthews TE (1999)

    Mechanical properties of nanocrystalline nickel produced by

    electrodeposition. Nanostruct Mater 11:343350

    10. Dalla Torre F, Van Swygenhoven H, Victoria M (2002) Nano-

    crystalline electrodeposited Ni: microstructure and tensile prop-

    erties. Acta Mater 50:39573970

    11. Ma E (2003) Instabilities and ductility of nanocrystalline and

    ultrafine-grained metals. Scripta Mater 49:663668

    J Mater Sci (2014) 49:38473859 3857

    123

  • 12. Panin AV, Panina AA, Ivanov YF (2008) Deformation macro-

    localisation and fracture in ultrafine-grained armco iron. Mater

    Sci Eng A 486:267272

    13. Cheng S, Ma E, Wang YM, Kecskes LJ, Youssef KM, Koch CC,

    Trociewitz UP, Han K (2005) Tensile properties of in situ con-

    solidated nanocrystalline Cu. Acta Mater 53:15211533

    14. Wei Q, Jia D, Ramesh KT (2002) Evolution and microstructure of

    shear bands in nanostructured Fe. Appl Phys Lett 81:12401242

    15. Fu H, Benson DJ, Meyers MA (2004) Computational description

    of nanocrystalline deformation based on crystal plasticity. Acta

    Mater 52:44134425

    16. Sansoz F, Dupont V (2007) Atomic mechanism of shear locali-

    zation during indentation of a nanostructured metal. Mater Sci

    Eng C 58:15091513

    17. Joshi SP, Ramesh KT (2008) Grain size dependent shear insta-

    bilities in body-centered and face-centered cubic materials. Mater

    Sci Eng A 493:6570

    18. Fan GJ, Fu LF, Choo H, Liaw PK, Browning ND (2006) Uniaxial

    tensile plastic deformation and grain growth of bulk nanocrys-

    talline alloys. Acta Mater 54:47814792

    19. Fan GJ, Fu LF, Qiao DC, Choo H, Liaw PK, Browning ND

    (2006) Grain growth in a bulk nanocrystalline Co alloy during

    tensile plastic deformation. Scripta Mater 54:21372141

    20. Erb U, El-Sherik AM (1994) US Patent # 5,353,266, October

    1994

    21. Erb U, El-Sherik AM, Cheung CKS, Aus MJ (1995) US Patent #

    5,433,797, July 1995

    22. Erb U, Palumbo G, McCrea JL (2011) The processing of bulk

    nanocrystalline metals and alloys by electrodeposition. In:

    Whang SH (ed) Nanostructured metals and alloys: processing,

    microstructure, mechanical properties and applications. Cam-

    bridge Woodhead Publishing Ltd., Oxford, pp 118151

    23. Hibbard G, Erb U, Aust KT, Klement U, Palumbo G (2002)

    Thermal stability of nanostructured electrodeposits. Mater Sci

    Forum 386388:387396

    24. Klement U, Erb U, Aust KT (1995) Investigations of the grain

    growth behavior of nanocrystalline nickel. Nanostruct Mater

    6:581584

    25. Brooks I, Lin P, Palumbo G, Hibbard GD, Erb U (2008) Hard-

    ness-tensile strength relationships for nanocrystalline materials.

    Mater Sci Eng A 491:412

    26. Brooks I, Palumbo G, Hibbard GD, Wang Z, Erb U (2011) On the

    intrinsic ductility of electrodeposited nanocrystalline metals.

    J Mater Sci 46:7713

    27. Cullity BD (1978) Elements of X-ray diffraction, 2nd edn.

    Addison-Wesley, Reading, MA, p 102

    28. Meyers MA, Mishra A, Benson DJ (2006) Mechanical properties

    of nanocrystalline materials. Prog Mater Sci 51:427556

    29. Gu CD, Lian JS, Jiang Q, Zheng WT (2007) Experimental and

    modeling investigations on strain rate sensitivity of an electrode-

    posited 20 nm grain sized Ni. J Phys D Appl Phys 40:74407446

    30. Wang YM, Ma E (2004) On the origin of ultrahigh cryogenic

    strength of nanocrystalline metals. Appl Phys Lett 85:27502752

    31. Dalla Torre F, Spatig P, Schaublin R, Victoria M (2005) Defor-

    mation behaviour and microstructure of nanocrystalline electro-

    deposited pressure torsioned nickel. Acta Mater 53:2337234932. Wei Q, Cheng S, Ramesh KT, Ma E (2004) Effect of nano-

    crystalline and ultrafine grain sizes on the strain rate sensitivity

    and activation volume: fcc versus bcc metals. Mater Sci Eng A

    381:7179

    33. Asaro RJ, Suresh S (2005) Mechanistic models for the activation

    volume and rate sensitivity in metals with nanocrystalline grains

    and nano-scale twins. Acta Mater 53:33693392

    34. Gu CD, Lian JS, Jiang Q, Jiang ZH (2007) Ductile-brittle-ductile

    transition in an electrodeposited 13 nanometer grain sized

    Ni-8.6wt.%Co alloy. Mater Sci Eng A459:7581

    35. Gu P, Dao M, Asaro RJ, Suresh S (2011) A unified mechanistic

    model for size-dependent deformation in nanocrystalline and

    nanotwinned metals. Acta Mater 59:68616868

    36. Dao M, Lu L, Asaro RJ, De Hosson JTM, Ma E (2007) Toward a

    quantitative understanding of mechanical behavior of nanocrys-

    talline metals. Acta Mater 55:40414065

    37. Van Swygenhoven H, Spaczer M, Caro A (1999) Microscopic

    description of plasticity in computer generated metallic nano-

    phase samples: a comparison between Cu and Ni. Acta Mater

    47:31173126

    38. Conrad H (2003) Grain size dependence of the plastic deforma-

    tion kinetics in Cu. Mater Sci Eng A341:216228

    39. Dieter GE (1988) Mechanical metallurgy, SI Metric. McGraw-

    Hill Publications, New York, NY

    40. Chu TC, Ranson WF, Sutton MA, Peters WH (1985) Application

    of digital-image-correlation techniques to experimental mechan-

    ics. Exp Mech 25:232244

    41. Zhu R, Zhou J, Jiang H, Lui Y, Ling X (2010) Multi-scale

    modeling of shear banding in fully dense nanocrystalline Ni

    sheet. Mater Sci Eng A 527:17511760

    42. Zhao YH, Guo YZ, Wei Q, Topping TD, Dangelewicz AM, Zhu

    YT, Langdon TG, Lavernia EJ (2009) Influence of specimen

    dimensions and strain measurement methods on tensile stress

    strain curves. Mater Sci Eng A 525:6877

    43. Taylor GI, Quinney H (1934) The latent energy remaining in

    metal after cold working. Proc R Soc Lond Ser A 143:307326

    44. Bever MB, Holt DL, Titchener AL (1973) The stored energy of

    cold work. Prog Mater Sci 17:5177

    45. Wang S, Zhou Y, Erb U et al Thermoelectrical transport in

    nanocrystalline Ni (in preparation)

    46. Turi T, Erb U (1995) Thermal expansion and heat capacity of

    porosity-free nanocrystalline materials. Mater Sci Eng A203:3438

    47. Gale WF, Totemeier TL (eds) (2004) Smithhells metals reference

    book, 8th edn. Elsevier Butterworth-Heinemann, Burlington,

    MA, p 1425

    48. Erb U (2012) Size effects and structure-sensitivity of properties in

    electrodeposited nanomaterials. Mater Sci For 706709:16011606

    49. Callister WD (2001) Fundamentals of materials science and

    engineering, 5th edn. Wiley, New York, p S-251

    50. Hibbard GD, Radmilovic V, Aust KT, Erb U (2008) Grain

    boundary migration during abnormal grain growth in nanocrys-

    talline Ni. Mater Sci Eng A494:232238

    51. Hibbard GD, Aust KT, Erb U (2008) On interfacial velocities

    during abnormal grain growth at ultra-high driving forces.

    J Mater Sci A43:64416452

    52. Ghosh AK, Hamilton CH (1979) Mechanical behavior and

    hardening characteristics of a superplastic Ti-6Al-4 V alloy.

    Metall Trans 10A:699706

    53. Senkov ON, Myshlyaev MM (1986) Grain growth in a super-

    plastic Zn-22 % Al alloy. Acta Metall 34:97106

    54. Shan Z, Stach EA, Wiezorek JMK, Knapp JA, Follstaedt DM,

    Mao SX (2004) Grain boundary-mediated plasticity in nano-

    crystalline nickel. Science 305:654657

    55. Haslam AJ, Moldovan D, Yamakov V, Wolf D, Phillpot SR, Gleiter

    H (2003) Stress-enhanced grain growth in a nanocrystalline material

    by molecular-dynamics simulation. Acta Mater 51:20972112

    56. Schitz Jakob (2004) Strain-induced coarsening in nanocrystal-

    line metals under cyclic deformation. Mater Sci Eng A

    375377:975979

    57. Jin M, Minor AM, Stach EA, Morris JW (2004) Direct obser-

    vation of deformation-induced grain growth during the nanoin-

    dentation of ultrafinegrained Al at room temperature. Acta Mater

    52:53815387

    58. Zhang K, Weertman JR, Eastman JA (2005) Rapid stress-driven

    grain coarsening in nanocrystalline Cu at ambient and cryogenic

    temperatures. Appl Phys Lett 87:061921

    3858 J Mater Sci (2014) 49:38473859

    123

  • 59. Gianola DS, Van Petegem S, Legros M, Brandstetter S, Van

    Swygenhoven H, Hemker KJ (2006) Stress assisted discontinuous

    grain growth and its effect on the deformation behavior of

    nanocrystalline aluminum thin films. Acta Mater 54:22532263

    60. Cahn JW, Mishin Y, Suzuki A (2006) Coupling grain boundary

    motion to shear deformation. Acta Mater 54:49534975

    61. Brandstetter S, Zhang K, Escuadro A, Weertman JR, Van

    Swygenhoven H (2008) Grain coarsening during compression of

    bulk nanocrystalline nickel and copper. Scripta Mater 58:6164

    62. Farkas D, Mohanty S, Monk J (2008) Strain-driven grain

    boundary motion in nanocrystalline materials. Mater Sci Eng

    A493:3340

    63. Wang YB, Li BQ, Sui ML, Mao SX (2008) Deformation-induced

    grain rotation and growth in nanocrystalline Ni. Appl Phys Lett

    92:011903

    64. Rupert TJ, Gianola DS, Gan Y, Hemker KJ (2009) Experimental

    observations of stress-driven grain boundary migration. Science

    18:16861690

    65. Xu W, Dai P, Wu X (2010) Effect of stress-induced grain growth

    during room temperature tensile deformation on ductility in

    nanocrystalline metals. Bull Mater Sci 33:561568

    66. Tucker GJ, Foiles SM (2013) Molecular dynamics simulations of

    rate-dependent grain growth during the surface indentation of

    nanocrystalline nickel. Mater Sci Eng A571:207214

    J Mater Sci (2014) 49:38473859 3859

    123

    Localized strain and heat generation during plastic deformation in nanocrystalline Ni and Ni--FeAbstractIntroductionExperimentalResults and discussionGrain size and chemical compositionDeformation at different tensile strain ratesStrain localizationStress-induced heat generationImplication of the measured heat generation

    SummaryAcknowledgementsReferences