ProQuestDocuments-2014-03-24 (1)

15
 _______________________________________________________________   _______________________________________________________________  Report Information from ProQuest 24 March 2014 08:04  _______________________________________________________________  24 March 2014 ProQuest

description

proqustr

Transcript of ProQuestDocuments-2014-03-24 (1)

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    1/15

    _______________________________________________________________

    _______________________________________________________________

    Report Information from ProQuest24 March 2014 08:04

    _______________________________________________________________

    24 March 2014 ProQuest

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    2/15

    able of contents1. Effect of Concurrent Precipitation on Recrystallization and Evolution of the P-Texture Component in a

    Commercial Al-Mn Alloy.................................................................................................................................. 1

    24 March 2014 ii ProQuest

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    3/15

    Document 1 of 1

    Effect of Concurrent Precipitation on Recrystallization and Evolution of the P Texture Component in aCommercial Al Mn AlloyAuthor: Tangen, S; Sjlstad, K; Furu, T; Nes, EProQuest document link

    Abstract: The recrystallization behavior of a cold-rolled Al-Mn alloy was investigated, focusing on the effect ofconcurrent precipitation on nucleation and growth of recrystallization and the formation of the P- ({011}[left angle

    bracket]111[right angle bracket]) and ND-rotated cube ({001}[left angle bracket]310[right angle bracket]) texture

    components. It was observed that if precipitation took place prior to or simultaneously with recovery and

    recrystallization processes, i.e., by concurrent precipitation, this resulted in a delayed recrystallization, a coarse

    and elongated grain structure, and an unusually sharp P-texture component. The P-texture component

    sharpened with increasing initial cold rolling reduction, increasing initial supersaturation of Mn, and decreasing

    annealing temperature. The P- and ND-rotated cube nucleation sites have an initial growth advantage

    compared to the particle-stimulated nucleation (PSN) sites due to their 40 deg [left angle bracket]111[right angle

    bracket]-rotation relationship to the Cu component of the deformation texture. The boundaries between such

    sites and the surrounding matrix will be of the 7 type, and it is assumed that such highly perfect boundaries

    will be less affected by solute segregation and precipitation, resulting in early growth advantage. It was further

    observed that dispersoids present prior to cold rolling and annealing had a weaker effect on the recrystallized

    grain size and texture compared to concurrent precipitation, even though the average dispersoid density was

    higher in the pre-precipitation cases. The finer grain size was explained by the wider dispersoid free zones

    surrounding the large constituent particles compared to the concurrent precipitation cases. Subsequent growth

    of the nucleated grains, however, was more hindered due to the Zener drag, consistent with the higher

    dispersoid densities. [PUBLICATION ABSTRACT]

    Full text: HeadnoteThe recrystallization behavior of a cold-rolled Al-Mn alloy was investigated, focusing on the effect of concurrent

    precipitation on nucleation and growth of recrystallization and the formation of the P- ({011}[left angle

    bracket]111[right angle bracket]) and ND-rotated cube ({001}[left angle bracket]310[right angle bracket]) texture

    components. It was observed that if precipitation took place prior to or simultaneously with recovery and

    recrystallization processes, i.e., by concurrent precipitation, this resulted in a delayed recrystallization, a coarse

    and elongated grain structure, and an unusually sharp P-texture component. The P-texture component

    sharpened with increasing initial cold rolling reduction, increasing initial supersaturation of Mn, and decreasing

    annealing temperature. The P- and ND-rotated cube nucleation sites have an initial growth advantage

    compared to the particle-stimulated nucleation (PSN) sites due to their 40 deg [left angle bracket]111[right angle

    bracket]-rotation relationship to the Cu component of the deformation texture. The boundaries between such

    sites and the surrounding matrix will be of the 7 type, and it is assumed that such highly perfect boundaries

    will be less affected by solute segregation and precipitation, resulting in early growth advantage. It was further

    observed that dispersoids present prior to cold rolling and annealing had a weaker effect on the recrystallized

    grain size and texture compared to concurrent precipitation, even though the average dispersoid density was

    higher in the pre-precipitation cases. The finer grain size was explained by the wider dispersoid free zones

    surrounding the large constituent particles compared to the concurrent precipitation cases. Subsequent growth

    of the nucleated grains, however, was more hindered due to the Zener drag, consistent with the higherdispersoid densities.

    DOI: 10.1007/s11661-010-0265-8

    The Minerals, Metals &Materials Society and ASM International 2010

    24 March 2014 Page 1 of 13 ProQuest

    http://search.proquest.com/docview/863468633?accountid=46437http://search.proquest.com/docview/863468633?accountid=46437
  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    4/15

    (ProQuest: ... denotes formulae omitted.)

    I. INTRODUCTION

    ALUMINUM-MANGANESE alloys (AA3xxx) dominate in the automobile heat exchanger industry. More than 95

    pct of all heat exchangers, such as condensers, evaporators, radiators, and oil coolers, are based on Al-Mn

    alloys due to their excellent combination of strength, ductility, corrosion resistance, brazeability, and

    affordability. Grain-size control is essential for heat exchanger applications. A coarse grain size is desirable formost heat exchanger applications, as this gives a high corrosion resistance, high sagging resistance, and good

    brazeability. In contrast, a fine grain size is preferred for products that require high formability as, for example,

    thin wall tubes. To get the optimal grain size, it is crucial to control the alloy composition, the thermomechanical

    processing, the size, the distribution of small and large particles, and the amount of elements in supersaturated

    solid solution throughout the entire process route.

    It is well established that second-phase particles have a strong effect on the recrystallization kinetics, final grain

    size, and texture.[1-6] Deformation zones may develop around large particles (>1 m) during deformation,

    which may activate particle-stimulated nucleation (PSN) of recrystallization,[1,2] whereas small, closely spaced

    dispersoids have the ability to retard (Zener pinning) both low- and high-angle grain boundary motion.[7,8]

    During thermomechanical processing or annealing of a deformed and supersaturated material, recovery as well

    as recrystallization may be influenced by the precipitation reaction, a phenomenon which commonly is referred

    to as concurrent precipitation. The different precipitation phenomena that may occur in the concurrent

    precipitation regime were first reported by Hornbogen[9] and Kster,[10] while Nes and Embury[3] were the first

    to report on the abnormal grain size that may result from this reaction. Later works by Nes and co-workers[5,6]

    focused on the texture aspects associated concurrent precipitation. This phenomenon will here be subjected to

    a new and focused study made possible by the recent developments in the electron backscattering diffraction

    (EBSD) technique in high-resolution scanning electron microscopy (SEM).

    Since the recrystallization texture determines the plastic anisotropy of the material, it has received increasing

    interest over the last decades. For hot-rolled and annealed sheets, the most typical recrystallization texture

    observed for materials with high stacking fault energy, such as aluminum, is the well-known cube texture. The

    cube texture develops during hot deformation, as the cube orientation is metastable at high temperatures (e.g.,

    References 11 and 12). After annealing of cold-rolled sheets, however, more random textures, which are

    nucleated at grain boundaries and by PSN, are commonly observed. Over recent years, the P- and ND-rotated

    cube textures have occasionally been observed in aluminum alloys after cold rolling and annealing.[6,12-18] In

    general, these latter orientations have been observed after annealing of cold-rolled Al-Mn alloys with a highly

    supersaturated solid solution, large cold rolling reductions, and low annealing temperatures.

    The main objective of the present work was to investigate the texture formation during recrystallization of a

    supersaturated commercial Al-Mn alloy. The focus was on establishing an understanding of the effect ofconcurrent precipitation on nucleation and growth of the P- and ND-rotated cube orientations, and the resulting

    grain size. Furthermore, in order to contrast the effect due to concurrent precipitation with that caused by a fine

    dispersion of particles present prior to annealing, this study also includes cases where a high number of Mn-

    bearing dispersoids were precipitated prior to cold rolling and final annealing.

    II. EXPERIMENTAL PROCEDURE

    The examined material was a commercial DC-cast AA3103 extrusion ingot (0.57 wt pct Fe, 1.0 wt pct Mn, 0.12

    wt pct Si, and other elements

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    5/15

    temperature-dependent term of Matthiessen's rule, the relationship between electrical conductivity and the solid

    solution content (wt pct) for the alloy conditions studied here is given by (Althenpohl[19]):

    ... [1]

    where is the electrical conductivity (m/ohm mm^sup 2^). It should be emphasized that the values of Mn in

    solid solution given in Table I are only estimates, as the true effects of other elements in solid solution, particle

    shape, number density, and type are not fully accounted for in the calculations.After the different heat treatments, the materials were water quenched to room temperature, and the four

    material variants were deformed by cold rolling to true strains of 0.5, 1.5, and 3.0. Finally, the rolled sheets were

    isothermally annealed (flash annealed) at different temperatures and times. Both hardness and electrical

    resistivity measurements were carried out on the sheet surfaces in order to follow the softening and precipitation

    reactions during annealing. Time-temperaturetransformation (TTT) diagrams were constructed on the basis of

    these measurements, where 25 pct softening was defined as the onset of recrystallization and a 2.5 pct

    increase in the electrical conductivity was defined as the start of precipitation. The recrystallized grain sizes

    have been measured in the longitudinal cross section defined by the rolling and normal directions (RD-ND

    plane) of the sheet midthickness, both by polarized light optical microscopy and the EBSD technique in a

    scanning electron microscope. Grain-size diagrams have also been constructed. These sheet materials will in

    the following be referred to as the "supersaturated sheets."

    In addition, a set of samples from the as-cast (0) and A-homogenized (A) materials were pretreated to give

    sheet variants with extremely high dispersoid number densities and nearly no Mn in supersaturated solid

    solution. The pretreatment consisted of cold rolling to a true strain of 0.5 (39 pct reduction), annealing at either

    573 K (300 C) or 623 K (350 C) for 10^sup 6^ s, and finally, cold rolling to an accumulated true strain of 3.0

    (95 pct accumulated strain). An overview of these material conditions and their denotations is given in Table II.

    These material variants will in the following be referred to as the "solute-free sheets" in order to distinguish them

    from the supersaturated sheets mentioned previously.

    The precipitation and softening behavior during annealing was investigated in greater detail by field emission

    gun-scanning electron microscopy (FESEM), and the recrystallization textures were monitored by EBSD and X-

    ray diffraction (XRD). Hence, both the local and bulk textures could be studied. In case of XRD, orientation

    distribution functions (ODFs) were calculated from four incomplete pole figures, namely, the {200}, {111}, {220},

    and {311} pole figures, according to the series expansion method.[20] All ODFs were further ghost corrected

    using the method suggested by Lcke et al.[21]

    III. EXPERIMENTAL RESULTS

    A. Supersaturated Sheets

    1. Characterization of the deformed state

    The deformed state will have a strong influence on the subsequent annealing behavior and it is therefore ofspecial importance to characterize the microstructure and texture of this state. The average subgrain size (d)

    and the misorientation across the subgrain boundaries ([varphi]) were calculated from EBSD orientation maps.

    The critical misorientation to define a subgrain boundary was set to 1.5 deg. In the case of the C material (0.31

    wt pct Mn in solid solution), increasing the true strain from 0.5 to 1.5, and 3.0, resulted in subgrain sizes of ...,

    respectively. The corresponding misorientation between the subgrains varied from 2.4 to 4.1 deg and 4.5 deg.

    Hence, the driving pressure for recrystallization, provided by the stored energy of the materials ..., generally

    increases with strain and most of the increase in the stored energy is associated with the decreasing subgrain

    size. The investigation showed that there was only a minor difference in the subgrain size and misorientation

    between the different material conditions (0, A, B, and C) for a constant strain.

    The particle size distributions of the deformed material conditions were measured by FESEM imaging and

    automated image processing. These results showed that the as-cast sheet material contained considerably

    fewer large particles and more small particles than the homogenized sheets. This difference will have an effect

    24 March 2014 Page 3 of 13 ProQuest

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    6/15

    on the subsequent recrystallization annealing, as the Zener pinning pressure due to the small particles will

    retard nucleation of recrystallization, while the largest particles will act as nucleation sites for recrystallization.

    Further, it could be seen that there were no significant differences in the particle size distributions of the

    homogenized and deformed conditions, indicating only minor particle breakup during rolling.

    The development of the sheet midthickness deformation texture with increasing strain for the A-material

    condition (0.72 wt pct Mn in solid solution), measured by XRD, is summarized as -fiber plots in Figure 1. Forthe investigated alloy, the deformation texture develops according to the theory of plane strain deformation of

    aluminum alloys, starting with a random texture followed by a gradual buildup of the and fibers as the strain

    increases. The deformation textures of the as-cast, B-homogenized, and C-homogenized material variants were

    similar to those of the A condition. The only difference was a minor texture sharpening with increasing amount

    of Mn in solid solution. These observations indicate that the Mn solid solution level has only a minor influence

    on the texture formation during cold deformation of non-heat-treatable Al-Mn alloys. The strength of the cube

    texture component after cold rolling was generally weak, around 1 to 3 times random, for all cold-rolled variants.

    2. Softening and precipitation during annealing

    The cold-rolled materials were isothermally annealed in salt baths, in the temperature range of 523 K to 773 K

    (250 C to 500 C) and for times ranging from 5 to 10^sup 6^ s. The softening reactions were followed by

    means of hardness testing, while the precipitation reaction was monitored by electrical conductivity

    measurements. The interaction between precipitation and softening, in terms of the overall transformation

    kinetics for the four Mn concentration levels, is summarized by the TTT diagrams in Figure 2. The background

    experimental curves for generating these TTT diagrams can be found in Reference 22. These diagrams show

    the effect of deformation on the softening and precipitation reactions for each of the four levels of Mn in solid

    solution. The thick solid lines indicate the onset of precipitation, while the thin solid and broken lines represent

    start and finished recrystallization, respectively.

    A critical temperature, T^sub C^, was defined as the temperature where the precipitation curve of the TTT

    diagram crosses the curve that indicates the start of recrystallization. Above this critical temperature,

    recrystallization is little affected by concurrent precipitation of Mn-bearing dispersoids, while at temperatures

    below T^sub C^, concurrent precipitation will retard recovery and recrystallization. Hence, the TTT diagrams

    predict when recrystallization occurs prior to precipitation and when the softening reactions are slowed and

    retarded by concurrent precipitation. The TTT diagrams confirm that an increased strain leads to a larger driving

    pressure for recrystallization but simultaneously also to an increased driving pressure for precipitation of Mn-

    bearing dispersoids. This is explained by a larger amount of stored deformation energy and a higher density of

    microstructural heterogeneities available for nucleation of recrystallization and dispersoids. It can be seen that

    the precipitation nose, i.e., the temperature where precipitation occurs most rapidly, is shifted toward shortertimes and higher temperatures with increasing initial solute Mn concentration. This results in a shift in the

    finished recrystallization time toward longer times and higher temperatures. Consequently, recrystallization of

    the as-cast strip (0.72 wt pct Mn in ss) becomes most retarded by concurrent precipitation, while the C strip

    (0.31 wt pct Mn in ss) is more or less unaffected by concurrent precipitation at temperatures >573 K (300 C).

    The information given by the TTT diagrams with respect to concurrent precipitation and the characteristic critical

    temperature, T^sub C^, are summarized in Figure 3. T^sub C^ is plotted vs the nominal concentration of Mn in

    solid solution prior to annealing for the three respective strain levels. The figure clearly demonstrates that the

    critical temperature increases with initial supersaturation of Mn in solid solution. As a consequence, the

    annealing temperature, which is necessary to avoid concurrent precipitation, increases with the initial solute Mn

    concentration. No distinct variation in the critical temperature with strain is observed.

    3. Effect of boundary misorientation on concurrent precipitation

    When studying partially annealed sheet samples undergoing concurrent precipitation, it was observed that the

    24 March 2014 Page 4 of 13 ProQuest

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    7/15

    dispersoids precipitated densely on some boundaries, while others were nearly dispersoid free (Figure 4 taken

    from the A variant (0.47 wt pct Mn in ss) after annealing at 623 K (350 C) for 10^sup 3^ s). The combination of

    backscattering electron (BSE) imaging and EBSD in an FESEM was used for this study. These investigations

    demonstrated that the dispersoids precipitated preferentially on high-angle grain boundaries, typically with

    misorientations >10 deg (cf. Figure 4). Microstructural observations of sheet samples annealed for longer times

    showed, qualitatively, that the dispersoids were more homogeneously distributed on all boundaries with varyingboundary misorientations. These observations are is in accordance with recent work by Somerday and

    Humphreys.[23]

    The micrograph in Figure 4 also shows some quite thick, bright boundaries (arrows in the figure), which

    correspond to high-angle grain boundaries. These boundaries are referred to as solute-rich zones. The reason

    why these boundaries appear bright in the BSE micrograph is most probably due to a local clustering/

    accumulation of elements with a higher atomic number than aluminum, i.e., clustering of Mn and Si. The grain

    boundaries act as sinks for supersaturated elements due to a reduction of the total energy of the system. The

    width of these boundaries/zones was typically ~1 to 10 nm, and they had a flakelike shape when viewed in

    three dimensions, covering large parts of the boundaries. It is reasonable to believe that, with time at

    temperature, the concentration of Mn eventually becomes so high at these zones or boundaries that Mn-bearing

    dispersoids precipitate. Further, the dispersoids will grow as a result of continuous diffusion from the matrix

    along the heterogeneities to the dispersoids. Hence, we can state that the solute zones are actually the

    precursors of the grain boundary dispersoids. These solute zones and dispersoid effects will successfully retard

    the softening processes by reducing the mobility of the grain boundaries. Since the elements diffuse along all

    heterogeneities/ boundaries, there will be a retarding drag even on the low-angle grain boundaries. However,

    the strongest drag will be on the most highly misoriented boundaries containing both a high concentration of

    accumulated elements and a high number of dispersoids. As a result, the most potent high-angle grain

    boundary nucleation sites experience the strongest Zener drag during annealing of a supersaturated material,

    and the softening processes become efficiently retarded.

    4. Recrystallization

    a. Grain size and texture. The profound effect of concurrent precipitation on the recrystallized grain shape and

    size is illustrated in Figure 5. When the sheets are annealed at temperatures above the critical temperature, TC,

    the structure recrystallizes into a fine grain size (... cf. Figure 5(a)). In contrast, annealing below TC results in an

    inhomogeneous and coarse grain structure, and the grains achieve a characteristic pancake shape (cf. Figure

    5(b)). These observations are in accordance with earlier works (cf. References 3 through 6, 24, and 25). The

    pancake-shaped grains resulting at low annealing temperatures are formed due to the lamellar alignment of the

    dispersoids during concurrent precipitation. Because of the pancake shape of the deformed grains, the

    dispersoids precipitate along the RD/TD plane, and hence, the growing grains experience the largest Zenerdrag in the direction normal to the rolling plane.[8] When annealing at temperatures below T^sub C^, the

    recrystallized grain size generally increases with strain and with the initial solute Mn concentration, i.e.,

    becoming coarsest for the as-cast variant (up to ...) and finest for the C material.

    The recrystallized grain-size diagram in Figure 6 expresses in a condensed form the various grain structures

    that may be expected for the current alloy system after cold deformation. The diagram indicates that in order to

    achieve a fine-grained structure after isothermal annealing, regardless of the recrystallization temperature, only

    ~0.3 to 0.2 wt pct Mn can remain in solid solution after the thermal homogenization treatment.

    The recrystallization textures were measured by means of XRD and in special cases by EBSD in an FESEM.

    During isothermal annealing of a deformed and supersaturated condition, the selected annealing temperature is

    crucial for the final recrystallization texture and grain structure in Al-Mn alloys. Figure 7 summarizes the

    recrystallization textures achieved for the as-cast condition and the three homogenization variants (A, B, and C)

    after cold rolling to strains of 3.0, as a function of the annealing temperature. The recrystallization textures are

    24 March 2014 Page 5 of 13 ProQuest

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    8/15

    seen to follow the same tendency as the recrystallized grain size with respect to the initial solute content and

    annealing temperature. At high annealing temperatures (T >T^sub C^), concurrent precipitation is avoided and

    the recrystallization textures become more or less random with weak P ({011}[left angle bracket]111[right angle

    bracket]) and cube ({001}[left angle bracket]100[right angle bracket]) components. The cube component is

    generally seen to strengthen with decreasing initial concentration of Mn in solid solution and with increasing

    annealing temperature, becoming as strong as 8 times random in the low solute B and C cases (0.38 wtpctMnand 0.31 wt pctMn in ss, respectively). It is interesting to note that the P component dominates with a

    strength of 4 times random in the as-cast case (0.72 wt pctMn in ss) recrystallized at 773 K (500 C). At lower

    annealing temperatures (T T^sub C^), the

    recrystallized grain structures consisted of relatively small equiaxed grains with weak, predominantly cube

    textures. Microstructural observations indicated that nucleation of recrystallization was controlled by PSN and

    grain boundary nucleation under such circumstances (Figure 8). In contrast, the material conditions that

    recrystallized under strong Zener drags from concurrent precipitation became coarse grained, with a sharp

    texture controlled by the P- and ND-rotated cube components.

    To study the nucleation efficiency of the particles under conditions of concurrent precipitation, the deformation

    structure surrounding coarse constituent particles has to be considered. Figure 9(a) displays the deformation

    structure around a constituent intermetallic particle after cold rolling to a strain of 3.0. The white broken lines in

    the figure illustrate the extension of the characteristic deformation zone, which surrounds the particle after

    deformation. Characteristic flow pattern can be seen to form around the hard particle. The deformation zone is,

    according to Humphreys,[1] characterized by rotated zones to the left and right of the particle, and distorted

    zones below and above the particle. A micrograph showing a closeup of a rotated zone close to a coarseparticle is presented in Figure 9(b). Such rotated zones correspond to a refined subgrain structure and high

    lattice rotations, giving a high local stored energy, which is crucial for nucleation of recrystallization.

    The micrograph in Figure 10, taken from the A-homogenized variant cold rolled to a strain of 3.0 and

    subsequently annealed to a recovered state at 623 K (350 C) for 10^sup 3^ s, shows the formation of a

    dispersoidfree zone surrounding the constituent eutectic particle after partial annealing. Note that the

    surrounding matrix contains a very high number density of dispersoids, which to a large extent are situated on

    the highangle grain boundaries. The dispersoid-free zone arises due to the depletion of Mn in supersaturated

    solid solution around the constituent particles both during casting, homogenization, and annealing after cold

    rolling (cf. References 26 through 28). Supersaturated Mn inside this zone will be drawn to the constituent

    particle and contribute to its phase transformation and growth. The depleted and dispersoid-free zones will

    further lead to a locally lower Zener drag during annealing of the deformed material. Consequently, nucleation

    of recrystallization should be less affected by precipitation in these zones, as compared to the potential

    24 March 2014 Page 6 of 13 ProQuest

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    9/15

    nucleation sites in the surrounding matrix that are hindered by a very high dispersoid number density.

    B. Solute-Free Sheets

    1. Characterization of the deformed state

    The solute-free sheets were in contrast to the "supersaturated sheets" processed in such a way as to precipitate

    a very high number density of small Mn-bearing dispersoids, which should in principle be randomly distributed,

    prior to the final recrystallization annealing of the coldrolled strips. In Table II, the four different variants of thismaterial are denoted by their accumulated true strain, their initial state, and their intermediate annealing

    temperature. Table II summarizes the calculated solute Mn contents in the initial state (after casting or

    homogenization) and after thermomechanical processing to the deformed state (accumulated strain of 95 pct). It

    can be seen that the two variants, which were intermediately annealed at 623 K (350 C), result in a very low

    solute content in the final cold-rolled state (0.15 wt pct) and can hence be called solute-free. The two variants,

    which were intermediately annealed at 573 K (300 C), do on the other hand contain a considerable soluteMn

    content after processing (0.30 to 0.35 wt pct), which is relatively similar to the C variant (0.31 wt pct Mn in ss) of

    the supersaturated sheets. This might result in some concurrent precipitation during final annealing and may

    therefore enhance the total Zener drag.

    The FESEM micrographs in Figure 11 represent the dispersoid structures resulting after completed processing

    of the as-cast and A-homogenized variants, intermediately annealed at 623 K (350 C) for 106 s and cold rolled

    to accumulated true strains of 3.0. It is clear that this process route results in a very high number density of

    small (d ~ 50 to 100 nm) dispersoids, especially in the case of the as-cast variant, which initially had the highest

    supersaturation of Mn prior to processing. The dispersoid number density per area was measured by counting

    dispersoids on a number of micrographs for each of the four solute-free sheets (Table II). The dispersoid

    number density of the two as-cast variants becomes around 21 to 23 m^sup -2^, which is higher than the

    situation for the supersaturated sheet. In comparison the as-cast condition, after cold rolling and concurrent

    precipitation annealing, typically gave a number density of around 10 to 16 m^sup -2^. The A-homogenized

    variants of the solute-free sheets have dispersoid number densities of around 6 to 9 m^sup -2^ after annealing,

    while annealing of the corresponding variant of the supersaturated sheet gave 1 to m^sup -2^.

    2. Recrystallized structures and textures

    The solute-free sheets were subsequently isothermally annealed in order to recrystallize the material for texture

    and grain structure analyses. The effect of the high dispersoid number density is clearly seen in Figure 12,

    where the softening kinetics of solute-free and supersaturated sheets at 673 K (400 C) are illustrated in a

    hardness vs time plot. An annealing temperature above 723 K to 773 K (450 C to 500 C) was necessary to

    fully soften the solute-free sheets within 106 s. The EBSD maps in Figure 13 represent the grain structure after

    isothermal annealing at 773 K (500 C). The mean grain length (DRD) was measured to ~50 m in the as-cast

    variants (Figure 13(a)) and ~12 m in the A-homogenized variant (Figure 13(b)).FESEM investigations of partially annealed sheet samples were performed in order to study the nucleation

    behavior in the solute-free sheets containing a high number density of dispersoids prior to deformation and

    annealing. Figure 14 shows a micrograph of a partially annealed specimen, and it is seen that nucleation of

    recrystallization has taken place in bands along the constituent particles, while no nucleation is observed in the

    areas without large particles. The image to the right shows an enlarged part of the area without any nucleation,

    which contains a high fraction of dispersoids that effectively pin the structure and prevent nucleation of

    recrystallization.

    A more detailed study of the early stages of nucleation of recrystallization is shown in the FESEM micrograph in

    Figure 15, illustrating the nucleation along bands of constituent particles in a partially annealed as-cast variant.

    The dispersoid number density is clearly lower in the zones surrounding the coarse particles. These dispersoid-

    free zones form during casting and further develop during intermediate annealing. Hence, similarly to the

    supersaturated sheets, nucleation of recrystallization becomes less affected by Zener drag at the constituents

    24 March 2014 Page 7 of 13 ProQuest

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    10/15

    than in the surrounding matrix.

    The global recrystallization textures were measured by means of EBSD. In general, the recrystallization texture

    becomes weaker for the solute-free sheets than for the supersaturated sheets. The two as-cast material

    variants developed weak ND-rotated cube and P components, with intensities around 5 and 3 times random,

    respectively. The two A-homogenized sheet variants, with a very fine recrystallized grain size, generally display

    the same texture components but with a lower intensity, both around 2 times random. The appearance of theND-rotated cube and P components after recrystallization of the solute-free sheets confirm that there is indeed

    an effect of the present dispersoids on the recrystallization texture evolution, although it is much weaker than in

    the concurrent precipitation case described previously.

    IV. DISCUSSION

    The salient experimental observations regarding the annealing characteristics of the cold-rolled AA 3103 alloy

    variants can be summarized as follows.

    Annealing at temperatures so high that no concurrent precipitation took place (T >T^sub C^) resulted in a

    finegrained recrystallized structure with a weak P texture and medium to strong cube texture. The strength of

    the cube component increased with decreasing initial Mn in supersaturated solid solution.

    Annealing at lower temperatures (T T^sub C^, become

    so effectively restrained by concurrent precipitation at temperatures T

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    11/15

    due to a higher mobility of the reaction front than growing grains of random orientations. If that had been the

    case, strong P and rotated cube textures observed in the present work should have resulted from

    recrystallization also at temperatures T >T^sub C^. The idea proposed by Daaland and Nes was that the 40 deg

    [left angle bracket]111[right angle bracket] grains had a nucleation and early growth advantage compared to

    grains of more random orientations. This idea was based on the observation shown in Figure 16 that the

    randomly oriented grains stopped growing after reaching an average size of about 5 to 7 m as a result ofconcurrent precipitation, while this precipitation reaction had less influence on the 40 deg [left angle

    bracket]111[right angle bracket] grains. This idea will now be discussed in the context of the present, more

    detailed metallographic observations.

    EBSD examination by Daaland and Nes[6] demonstrated that in the very beginning of recrystallization the

    density of randomly oriented grains, believed to be the result of PSN, was about an order of magnitude higher

    than the density of grains with the rotated cube or P orientation. In the fully transformed condition, however, the

    volume fraction of rotated cube grains was about 50 pct, while that of randomly oriented grains amounted to

    only 9 pct. The fact that these randomly oriented grains indeed could be associated with PSN has been nicely

    confirmed by the present investigation. During concurrent precipitation, the deformation zone recrystallizes

    readily, while growth of recrystallization beyond and into the matrix is effectively stopped. This situation was

    illustrated in Figure 10, taken from the A-homogenized condition. The solute depletion of the area just around

    the particles is enhanced due to the homogenization treatment. In order to support this trend, an additional

    experiment has been performed, selecting the as-cast variant cold rolled to a strain of 3.0 and subsequently

    annealed at 623 K (350 C) for 10^sup 3^ s. The combination of FESEM BSE imaging and EBSD mapping was

    used in order to display the recrystallized grains at an early stage of nucleation and growth and to measure their

    corresponding orientations.

    The EBSD maps in Figure 17 show a few examples of nucleated P- and ND-rotated cube grains. P- and ND-

    rotated cube grains are circled with white and black dashed lines, respectively. The white nonindexed areas in

    the EBSD maps correspond to constituent particles. Since the as-cast condition (0.72 wt pct Mn in ss) contains

    a significantly higher initial amount of Mn in solid solution compared to the A-homogenized case (0.47 wt pct Mn

    in ss) described previously, the effect of concurrent precipitation will be more severe. Consequently, an

    annealing treatment of 103 s at 623 K (350 C) caused the as-cast variant to soften by only a 20 pct reduction in

    hardness compared to 33 pct in the homogenized material, while the conductivity changes were 22 and 8 pct for

    the as-cast and homogenized versions, respectively. Still, however, in this as-cast condition, cases of relatively

    large recrystallized grains of the P and rotated cube orientations were found, as illustrated in Figure 17. This

    confirms the hypothesis proposed by Daaland and Nes[6] that the nucleation and early growth of grains with this

    40 deg h111i orientation relationship was much less affected by concurrent precipitation than grains of other

    orientations. The nature of the nucleation sites for these orientations, however, has not yet been fully identified.It has been speculated that the P- and ND-rotated cube-oriented grains nucleate in the vicinity of the constituent

    intermetallic particles by PSN (refer to the work of Engler and co-workers,[16,30] Ryu and Lee,[17] and

    Sjlstad).[29] These investigators performed TEM and SEM studies and observed P-oriented grains to be

    associated with coarse particles. However, one has to be aware that for any new grain of a few microns in

    radius, the probability that such a grain will be in contact with a coarse particle is about 1, given that the density

    of particles in commercial aluminum alloys with a size larger than 1 m in diameter is about 10^sup 16^ m^sup -

    3^.

    It is difficult to understand how 40 deg [left angle bracket]111[right angle bracket] grains can nucleate from the

    turbulent zones of deformation zones. It is well established that 40 deg [left angle bracket]111[right angle

    bracket] grains nucleate from bandlike features in commercial aluminum alloys.[5,6,31] However, still some

    evidence has been provided that the grains of orientations close to the P have been detected in the deformation

    zone[16,17] surrounding the large constituent particles. The explanation could very well be that the deformation

    24 March 2014 Page 9 of 13 ProQuest

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    12/15

    zones recrystallized into a random orientation distribution due to the low Zener drag (depleted zones). However,

    the growth of recrystallization beyond the deformation zones, in general, is stopped by a high solute and Zener

    drag effect. Only the recrystallized deformations zones of the P and rotated cube orientations, which in addition

    have a copper deformation component as their next neighbor, will be able to grow beyond the deformation

    zones due to their 40 deg [left angle bracket]111[right angle bracket] orientation relationship. The reason why

    these particular orientation relationships are less affected by concurrent precipitation is argued in the nextparagraph. One could also argue that P and rotated cube grains originate from matrix-type heterogeneities and

    subsequently use a neighboring particle deformation zone as a growth environment.

    The present observations clearly show that the P and rotated cube-oriented grains have an early growth

    advantage. A plausible explanation for this can be found in the concurrent precipitation pattern shown

    previously. First, solute elements segregate toward highangle boundaries, and then these boundaries become

    decorated by precipitates. Somewhat later, precipitation is observed to take place also at low-angle boundaries.

    On the assumption that the 40 deg [left angle bracket]111[right angle bracket] grains originate from matrix sites

    (i.e., sites located with the Cu-deformation texture component as next neighbor), these sites will be separated

    from the surrounding matrix by 7-type high-angle boundaries. It is reasonable to assume that such highly

    perfect boundaries will be less affected by solute segregation and precipitation. Or, in other words, 40 deg h111i

    grains will be less affected by concurrent precipitation, the result being a nucleation and early growth advantage

    for these grains compared with grains of other orientations.

    B. On the Effect of Dispersoids Present Prior to Annealing vs Concurrent Precipitation during Annealing

    It was observed that annealing of the solute-free sheets gave a considerably finer grain size and a much weaker

    recrystallization texture than the supersaturated sheets, even though the total number of dispersoids was

    higher. On the other hand, the time to complete recrystallization was considerably longer in these "sheets" as

    compared with the supersaturated sheets, which were affected by concurrent precipitation. Regarding the finer

    grain size, nucleation of recrystallization was observed to take place during the early stages of annealing of the

    solute-free sheets and thus mainly inside the dispersoid-free zones surrounding the constituent particles (cf.

    Figure 15). It is believed that the high nucleation frequency at the constituent particles is caused by the

    coarsening of the surrounding dispersoid-free zones during the intermediate annealing for 10^sup 6^ s, which

    the solute-free sheets were subjected to prior to cold rolling. These zones are considerably larger and contain

    fewer dispersoids than the corresponding zones in the supersaturated sheets. This might also explain why

    these materials develop a rather weak texture after recrystallization. A higher number of PSN sites, with a large

    variety of orientations, are able to grow to an overcritical size and become recrystallized grains. However, even

    though the grains are easily nucleated and reach an overcritical size for growth in the dispersoid-free zones

    close to the constituent particles, the continued expansion will be strongly retarded as the matrix contains a very

    high dispersoid number density (cf. Figure 12). It should, however, be mentioned that the total driving pressurefor recrystallization is somewhat lower in the case of the solute-free sheets due to the "intermediate"

    precipitation annealing at a true strain of 0.5. This can explain some of the difference in the recrystallization

    kinetics between the two sheets seen in Figure 12.

    Burger et al.[32] investigated the recrystallization texture after rolling of an Al-Mn alloy, pretreated differently to

    achieve various dispersoid number densities prior to recrystallization annealing. A high temperature was used to

    soften the material conditions. They found the P texture (termed RX texture in the actual article) to sharpen up

    to a maximum intensity of 2 to 3 times random as the dispersoid density increased. The P texture was observed

    together with the ND-rotated cube component, which was found to be less rotated as the dispersoid content

    decreased. These observations further confirm that dispersoids present prior to recrystallization annealing have

    a considerably less prominent effect on the development of the P-texture component compared to concurrent

    precipitation.

    V. CONCLUSIONS

    24 March 2014 Page 10 of 13 ProQuest

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    13/15

    The details around the formation of the precipitation and recrystallization behavior of Al-Mn alloys can be

    summarized as follows.

    Concurrent precipitation in a supersaturated Al-Mn alloy strongly affects the recrystallization behavior, leading to

    inhomogeneous and coarse-grained recrystallized structures. The degree of concurrent precipitation increases

    with rising initial supersaturation of Mn and with raising prior cold reduction due to an enhanced diffusion rate

    and higher number of heterogeneities available for nucleation of dispersoids.The recrystallization texture in the case of concurrent precipitation of Mn-rich dispersoids is dominated by the P-

    and ND-rotated cube components. Unusually sharp P textures were observed in the present work. The intensity

    of the P- and ND-rotated cube textures strengthens with increasing initial cold rolling strain, supersaturation of

    Mn, and dispersoid density, but decreases with increasing annealing temperature.

    The P- and ND-rotated cube nucleation sites have an initial growth advantage compared to the PSN sites due

    to their 40 deg h111i-rotation relationship to the Cu component. The boundaries between such sites and the

    surrounding matrix will be of the 7 type, and it is assumed that such highly perfect boundaries will be less

    affected by solute segregation and precipitation, resulting in the early growth advantage.

    Dispersoids present prior to annealing were seen to have a much weaker effect on the recrystallized grain size

    and texture than concurrent precipitation, although the total dispersoid number density was much higher. The

    growth of the recrystallized grains was thus more hindered than the nucleation process, giving a high nucleation

    frequency but a very slow recrystallization rate. Thus, the sheets recrystallized into a rather fine grain size with a

    weak texture. The high nucleation rate was attributed to the randomization of the dispersoids during rolling in

    combination with the very large precipitate-free zones surrounding the coarse constituent particles after finalized

    rolling, which made PSN relatively easy.

    ACKNOWLEDGMENTS

    This research was carried out as a part of the NFRproject Heat Treatment Fundamentals KMB project (Project

    No. 143877/213) in the subproject Nucleation of Recrystallization. Funding by the industrial partners, Hydro

    Aluminium, Raufoss ASA, and Elkem, is gratefully acknowledged.

    ReferencesREFERENCES

    1. F.J. Humphreys: Acta Metall., 1977, vol. 25, pp. 1323-44.

    2. F.J. Humphreys and M. Hatherly: Recrystallization and Related Annealing Phenomena, Elsevier Science

    Ltd., Oxford, United Kingdom, 1995.

    3. E. Nes and J.D. Embury: Z. Metallkd., 1975, vol. 66, pp. 589-93.

    4. O. Engler, J. Hirsch, and K. Lcke: Acta Metall. Mater., 1995, vol. 43, pp. 121-38.

    5. H.E. Vatne, O. Engler, and E. Nes: Mater Sci. Technol., 1997, vol. 13, pp. 93-102.

    6. O. Daaland and E. Nes: Acta Mater., 1996, vol. 44, pp. 1413-35.7. C.S. Smith: Trans. TMS-AIME, 1948, vol. 175, pp. 15-51.

    8. E. Nes, N. Ryum, and O. Hunderi: Acta Metall., 1985, vol. 33, pp. 11-22.

    9. E. Hornbogen: Praktischee Metallographie, 1970, vol. 9, p. 349.

    10. O. Kster: Recrystallization of Metallic Materials, F. Heassner, ed., Dr. Rieber Verlag, Stuttgart, 1971, pp.

    215-48.

    11. H. Weiland and J.R. Hirsch: Text. Microstruct., 1990, vols. 14-18, pp. 647-52.

    12. C. Maurice and J.H. Driver: Acta Metall. Mater., 1993, vol. 41, pp. 1653-64.

    13. S.E. N ss: Dr. Sci. Thesis, University of Oslo, Oslo, Norway, 1990.

    14. J. Liu and J.G. Morris: Metall. Mater. Trans. A, 2003, vol. 34A, pp. 2029-32.

    15. O. Engler, X.W. Kong, and K. Lcke: Acta Mater., 2001, vol. 49, pp. 1701-15.

    16. O. Engler: Mater. Sci. Technol., 1996, vol. 12, pp. 859-72.

    17. J.H. Ryu and D.N. Lee: Mater. Sci. Eng., 2002, vol. A336, pp. 225-32.

    24 March 2014 Page 11 of 13 ProQuest

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    14/15

    18. W.C. Liu and J.G. Morris: Metall. Mater. Trans. A, 2005, vol. 36A, pp. 2829-48.

    19. D. Altenpohl: Aluminium und Aluminiumlegierungen, Springer- Verlag, New York, NY, 1965.

    20. H.J. Bunge: Texture Analysis in Materials Science, Butterworth &Co., London, 1982.

    21. K. Lcke, J. Pospiech, K.H. Virnich, and J. Jura: Acta Metall., 1981, vol. 29, pp. 167-85.

    22. S. Tangen: Ph.D. Thesis, Norwegian University of Science and Technology, Trondheim, Norway, 2004.

    23. M. Somerday and F.J. Humphreys: Mater. Sci. Technol., 2003, vol. 19, pp. 20-29.24. E. Nes: Aluminium, 1976, vol. 52, pp. 560-63.

    25. P. Furrer and H. Warlimont: Aluminium, 1978, vol. 54, pp. 135-42.

    26. P.L. Morris and B.J. Duggan: Met. Sci., 1978, vol. 12, pp. 1-7.

    27. P. Furrer and G. Hausch: Met. Sci., 1979, pp. 155-62.

    28. C. Sigli: Proc. ICAA 4, T.H. Sanders and E.A. Starke, Jr., eds., Georgia Institute of Technology, Atlanta, GA,

    1994, pp. 513-20.

    29. K. Sjlstad: Ph.D. Thesis, Norwegian University of Science and Technology, Trondheim, Norway, 2003.

    30. O. Engler, P. Yang, and X.W. Kong: Acta Mater., 1996, vol. 44, pp. 3349-69.

    31. J. Hjelen, R. rsund, and E. Nes: Acta Metall., 1991, vol. 39, pp. 1377-1404.

    32. G. Burger, P. Wycliffe, C. Gabryel, and D.J. Lloyd: Mater. Sci. Forum, 1997, vol. 242, pp. 101-09.

    AuthorAffiliationS. TANGEN, Project Manager, and T. FURU, Senior Advisor, are with Hydro Aluminium, Research and

    Technology Development, N-6601 Sunndalsora, Norway. Contact e-mail: stian.tangen@hydro. com K.

    SJLSTAD, Project Manager, and E. NES, Professor, are with the Department of Materials Technology, NTNU,

    N-7491 Trondheim, Norway.

    Manuscript submitted March 12, 2009.

    Article published online July 15, 2010

    Subject:Alloys; Metallurgy; Cold; Temperature; Studies; Scanning electron microscopy; Annealing; Corrosionresistance; Heat exchangers; Aluminum;

    Publication title: Metallurgical and Materials Transactions

    Volume: 41A

    Issue: 11

    Pages: 2970-2983

    Number of pages: 14

    Publication year: 2010

    Publication date: Nov 2010

    Year: 2010

    Publisher: Springer Science & Business Media

    Place of publication: Warrendale

    Country of publication: Netherlands

    Publication subject: Metallurgy, Engineering

    ISSN: 10735623

    CODEN: MMTAEB

    Source type: Scholarly Journals

    24 March 2014 Page 12 of 13 ProQuest

  • 5/27/2018 ProQuestDocuments-2014-03-24 (1)

    15/15

    Language of publication: English

    Document type: Feature

    Document feature: Tables Graphs Photographs References

    ProQuest document ID: 863468633

    Document URL: http://search.proquest.com/docview/863468633?accountid=46437Copyright: Copyright Springer Science & Business Media Nov 2010

    Last updated: 2012-08-20

    Database: ProQuest Science Journals,ProQuest Research Library

    _______________________________________________________________

    Contact ProQuestCopyright 2014 ProQuest LLC. All rights reserved. - Terms and Conditions

    24 March 2014 Page 13 of 13 ProQuest

    http://search.proquest.com/docview/863468633?accountid=46437http://www.proquest.com/go/contactsupporthttp://www.proquest.com/go/contactsupporthttp://search.proquest.com/info/termsAndConditionshttp://search.proquest.com/info/termsAndConditionshttp://www.proquest.com/go/contactsupporthttp://search.proquest.com/docview/863468633?accountid=46437