ArticlePDF Available

Bridging Simulations and Experiments in Microstructure Evolution

Authors:

Abstract and Figures

We demonstrate the importance of anisotropic interface properties in microstructure evolution by comparing computed evolved microstructures to final experimental microstructures of 5170 grains in 19 thin aluminum foil samples. This is the first time that a direct experimental validation of simulation has been performed at the level of individual grains. We observe that simulated microstructures using curvature-driven grain boundary motion and anisotropic interface properties agree well with experimentally evolved microstructures, whereas agreement is poor when isotropic properties are used.
Content may be subject to copyright.
Bridging Simulations and Experiments in Microstructure Evolution
M. C. Demirel,
1,2,3
A. P. Kuprat,
2
D. C. George,
2
and A. D. Rollett
3
1
Materials Science and Technology, MST-8, Los Alamos National Laboratory, Los Alamos, New Mexico 87545
2
Theoretical Division, T-1, Los Alamos National Laboratory, Los Alamos, New Mexico 87545
3
Carnegie Mellon University, Department of Materials Science & Engineering, Pittsburgh, Pennsylvania 15213
(Received 28 May 2002; published 9 January 2003)
We demonstrate the importance of anisotropic interface properties in microstructure evolution by
comparing computed evolved microstructures to final experimental microstructures of 5170 grains in
19 thin aluminum foil samples. This is the first time that a direct experimental validation of simulation
has been performed at the level of individual grains. We observe that simulated microstructures using
curvature-driven grain boundary motion and anisotropic interface properties agree well with experi-
mentally evolved microstructures, whereas agreement is poor when isotropic properties are used.
DOI: 10.1103/PhysRevLett.90.016106 PACS numbers: 68.35.–p, 68.37.Hk, 68.55.Jk, 81.40.–z
This paper seeks to extend previous statistical compari-
sons of predicted and experimentally observed grain
boundary network evolution by demonstrating agreement
at the scale of individual grains, provided that the anisot-
ropy of interfacial energy and mobility is included. With
very few exceptions [1], only statistical comparisons have
been made such as determining the exponent in the
power-law relationship between average radius and time.
In addition to this general aim, we consider coarsening in
networks of low angle grain boundaries (subgrains),
which has been the subject of some controversy. Some
authors have postulated that the interfaces are essentially
sessile because they are themselves networks of lattice
dislocations and that coarsening occurs by rotation of
individual subgrains [2,3]. Rotation is driven by minimi-
zation of interfacial energy and tends to eliminate the
misorientation between certain pairs of adjacent grains,
thus causing coalescence and coarsening to occur. Gleiter,
on the other hand, showed that low angle grain boundaries
migrate under curvature driving forces just as observed
for high angle grain boundaries [4]. Clearly, both mecha-
nisms are feasible for coarsening of subgrain networks,
and so it is of some importance to determine which one is
applicable to real materials. Since the torque that drives
grain rotation is supralinear in (inverse) grain size, it is
reasonable to expect that curvature-driven migration
dominates at the large grain sizes observed in this ex-
periment whereas rotation should dominate in nanoscale
grain sizes.
We have previously shown a strong agreement between
small-scale grain growth experiments and anisotropic
three-dimensional simulations [5] obtained from elec-
tron backscatter diffraction (EBSD) measurements [6].
Using the same technique, we obtained data for 5170
grains from 19 thin aluminum foil samples with colum-
nar grain structure (thereby avoiding serial section-
ing) and compared our computational results with
experiments.
Our simulation model uses curvature-driven motion
implemented by
GRAIN3D [7], a three-dimensional,
gradient-weighted moving finite elements code. We as-
sume that the grain boundary motion is proportional to
the local mean curvature of the interface,
v
n
; (1)
where v
n
is the normal velocity of the interface, is the
grain boundary mobility, and is the interface energy
per unit area. is the sum of principal curvatures, i.e.,
twice the mean curvature; in these simulations, the cur-
vature is equal to the curvature observed in the plane.
The interfaces are represented as piecewise linear. For
details of the simulation method, the reader should refer
to Ref. [7].
The interfacial anisotropy is based on a previous deter-
mination of grain boundary energy, , and mobility, ,
from a statistical/multiscale analysis of triple junction
geometry and crystallography in aluminum [8–10], and
the result for energy is in very good agreement with
the Read-Shockley model [11] as expected. In the highly
textured, columnar grain structure aluminum foil we
investigated, most interfaces are low angle boundaries
(misorientation <15
). The grain boundary mobility is
low for small misorientations but undergoes a sharp
transition to high mobilities between 10
and 15
in
misorientation which is in agreement with the literature
[12–14]. For our simulations, we assume that all high
angle grain boundaries (> 15
) have the same values of
energy and mobility. The occurrence of high angle
boundaries is very low in this material, which means that
only a small error is introduced by this assumption. This
experiment permitted a verification of curvature-driven
interface motion [4], as compared with the competing
mechanism of grain rotation [3] leading to coalescence
[2]. We found a standard deviation of 0:98
in misorien-
tation angle between the initial and final experimental
PHYSICAL REVIEW LETTERS
week ending
10 JANUARY 2003
VOLUME 90, NUMBER 1
016106-1 0031-9007=03=90(1)=016106(4)$20.00 2003 The American Physical Society 016106-1
states, implying insignificant rotation of the boundary
[15] and that interface motion is indeed curvature driven.
To generate input for the simulation from EBSD ex-
perimental results, grain boundaries and triple junctions
are located from the recorded image, and the grain
boundary structure on the section plane is extruded to
create the three-dimensional mesh [16,17]. We use the
EBSD measurements to assign a single average orienta-
tion for each grain and thus a single interface misorien-
tation for each pair of grains. Using the results of Yang
et al., the interface misorientation gives a relative value
of and for each interface. Note that use of a constant
value of along the boundary from one triple junction to
the other means that our model does not include boundary
inclination dependence. The total volume of the structure
is kept constant and exterior boundaries are assumed to
be quasiperiodic [5]. Figure 1(A) shows a representative
experimental input. The simulation was initialized with
this state and was run until the area match function,
explained in the next paragraph, began to decrease. The
final experimental microstructure is shown in Fig. 1(B),
and is to be compared to the computed evolved micro-
structure shown in Figs. 1(C) and 1(D). A strong cube
texture, f100gh001i, was observed for both the initial and
the final experimental microstructures. This is as ex-
pected because of the small amount of coarsening that
has taken place during the experiment.
Simulations were performed with three different as-
sumptions for the interface properties. First, and
were assumed constant and equal to unity (isotropic).
Second, was allowed to depend on the misorientation
angle (anisotropic mobility). Third, and were al-
lowed to depend on the misorientation angle (anisotropic
mobility and energy). For quantitative comparison, we
introduce a regular grid of sampling points sx; y of
between 2500 and 10 000 points depending on the size
of the microstructure. A quantitative measure for the
comparison is then given by a normalized area match
function (NAMF). One may think of this comparison
method as overlaying the final experimental state onto
successive simulation snapshots and measuring at each
snapshot time the amount of area where the orientations
match. The higher the value, the better the match.
NAMF
i
1
Z
X
x;y
s
i
x; ys
exp
x; y; (2)
where for simulation snapshot i, s
i
is the crystal orien-
tation, and s
exp
is the crystal orientation for the final
experimental configuration, is the delta function, and Z
is the number of sampling points. If the crystal orienta-
tion for the spatial point x; y is the same as the final
experimental crystal orientation at the same location, the
delta function is one, otherwise it is zero. NAMF
i
is then
given by summing over the set of all x; y points and
dividing by the total number of points. Perfect agreement
is denoted by NAMF 1 and perfect disagreement by
NAMF 0. Since the final experimental state evolved
from the initial experimental state, which is identical to
the initial simulation state, the comparison begins with a
high degree of match. In a representative microstructure,
we find a 62% match between the initial simulation
state and the final experimental state (NAMF 0:62).
Figure 2(A) shows the results of isotropic and anisotropic
simulations for one sample. In the isotropic case, the
maximum NAMF gives a 68% match, whereas both
anisotropic cases reach a significantly higher maximum
NAMF. Figures 1(E) and 1(F) illustrate the pixelized
comparison of experimental to simulated final states
from which NAMF is calculated. All simulations even-
tually exhibit a decreased match as they were run longer
than the equivalent final experimental time. It is clear
from these results that anisotropic interface properties
play an important part in microstructure evolution.
With regard to simulations involving anisotropic energy,
we observed violations of the wetting condition (Young’s
equation) at some triple junctions. These force imbalances
occur at triple junctions where one of the three energies is
FIG. 1 (color online). (A) Initial experimental microstructure
(top view) with dimensions 2000 m 2000 m (339 grains).
(B) Final experimental microstructure. (C) Isotropic mobility
simulation final state. (D) Anisotropic mobility simulation final
state. (E) Pixelized subtractions of (C) from (B). (F) Pixelized
subtraction of (D) from (B).
PHYSICAL REVIEW LETTERS
week ending
10 JANUARY 2003
VOLUME 90, NUMBER 1
016106-2 016106-2
greater than the sum of the other two and can result from
using a constant along the entire interface. To prevent
this, we constrained the dihedral angles so that they are
always larger than 90
by adding an isotropic minimum
energy to all boundaries [15]. Because of the uncertainty
of the effects of adjusting in this manner, we consider
only the anisotropic mobility simulations in the remain-
der of this Letter.
Scaling the 19 simulation times so that the average of
all simulations matches the experimental time of 20 min
at 550
C, we find a low standard deviation, 1.80 min,
suggesting that simulated evolution and experimental
grain growth are consistent. In Fig. 2(B), for each of
the 19 simulations using anisotropic mobility proper-
ties, we display a vector representing the amount of
experimental microstructure evolution matched by the
simulation. Each vector origin gives NAMF
0
or the
amount of match between the final experimental state
and the initial microstructure state. Each vector tip
shows the NAMF
max
or the amount of match between
the experimental and simulated final states. To com-
pare the observed experimental microstructure evolution
with the anisotropic simulated evolution, we calculate
the mean (0.5) and standard deviation (0.172) of
NAMF
max
NAMF
0
=1 NAMF
0
, showing that our
simulation model accounts for 50% of the microstructure
evolution observed in experiment. The mean and standard
deviation for isotropic simulations are 0.17 and 0.147,
respectively.
The comparison between the experiment and compu-
tation at the individual grain level extends our knowledge
of grain evolution. First, microstructural evolution in
columnar aluminum foils can be correctly modeled using
anisotropic parameters. The results of three-dimensional
computer simulations using anisotropic mobility agree
well with experimental grain growth. Second, we have
shown that isotropic modeling has very little predictive
value. Using NAMF as a metric, we have shown that
isotropic modeling explains at most 17% of the evolution,
whereas anisotropic modeling accounts for 50% of the
evolution. The failure of our model to account for all the
evolution may be due to one or more of the following:
boundary inclination dependence, computational indeter-
minacy due to grain topology changes [7], dislocation
density, stored deformation energy, or by sample error
such as surface defects, sample preparation, local defor-
mations, and grains that are not columnar. Additionally,
in these mesoscale experiments, we observed a 0:98
standard deviation in misorientation angle between ini-
tial and final experimental states implying insignificant
grain rotation and that microstructure evolution is curva-
ture driven.
This research is support by Los Alamos National
Laboratory (DOE, W-7405-ENG-36), the MRSEC pro-
gram of NSF (DMR-0079996), and the Computational
Materials Science Network (US-DOE).
[1] T. Baudin, P. Paillard, and R. Penelle, Scr. Mater. 40,
1111 (19 99).
[2] H. Hu, in Recovery and Recrystallization of Metals,
edited by L. Himmel (Interscience, New York, 1963),
p. 362.
[3] G. Martin, Phys. Status Solidi B 172, 121 (1992).
[4] H. Gleiter, Philos. Mag. 20, 821 (1969).
[5] M. C. Demirel et al., Interface Sci. 10, 137 (2002).
[6] M. C. Demirel et al.,inElectron Backscatter Diffraction
in Materials Science, edited by M. K. Adam, J. Schwartz,
and Brent L. Adams (Kluwer Academic, New York,
2000), p. 65.
[7] A. P. Kuprat, SIAM J. Sci. Comput. 22, 535 (2000).
[8] B. L. Adams et al., Scr. Mater. 38, 531 (1998).
01020304050
0.6
0.7
0.8
0.9
isotropic energy(σ ) and mobility(µ)
anisotropic mobility (µ)
anisotropic mobility (µ) and energy(σ)
NAMF
Equivalent Experimental Time (minute)
0 2 4 6 8 10 12 14 16 18 20
0.5
0.6
0.7
0.8
0.9
1.0
NAMF
max
NAMF
0
NAMF
0
, NAMF
max
Microstructure number
(A) (B)
FIG. 2 (color online). (A) Results of a representative simulation with different interface properties: The isotropic case reaches a
maximum NAMF value of 0.68, accounting for 16% of the match; both anisotropic cases reach maximum NAMF values between
0.82 and 0.77, accounting for 53% and 40% of the match, respectively. (B) NAMF
max
and NAMF
0
values for 19 microstructures.
PHYSICAL REVIEW LETTERS
week ending
10 JANUARY 2003
VOLUME 90, NUMBER 1
016106-3 016106-3
[9] D. Kinderlehrer et al.,inProceedings of the Twelfth
International Conference on Textures of Materials
(NRC Research Press, Montre
´
al, Canada, 1999), p. 1643.
[10] C.-C. Yang, W.W. Mullins, and A. D. Rollett, Scr. Mater.
44, 2735 (2001).
[11] W. T. Read and W. Shockley, Phys. Rev. 78, 275 (1950).
[12] R. Viswanathan and C. L. Bauer, Acta Metall. 21,1099
(1973).
[13] R. Sandstrom, Acta Metall. 25, 897 (1977).
[14] D. A. Smith et al., Grain Boundary Structure and
Kinetics (American Society for Metals, Metals Park,
Ohio, 1980).
[15] M. C. Demirel, in Materials Science and Engineering
(Carnegie Mellon University, Pittsburgh, 2002), p. 127.
[16] B. R. Schlei, in SPIE Conference Proceedings, Seattle,
2002, Vol. XI (to be published).
[17] M. C. Demirel et al., J. Appl. Phys. (to be published).
PHYSICAL REVIEW LETTERS
week ending
10 JANUARY 2003
VOLUME 90, NUMBER 1
016106-4 016106-4
... There have been relatively few attempts in the past to compare the results of grain growth simulations with experiments on a grain-by-grain basis. For example, in 2003 Demirel et al. [44] validated an anisotropic grain growth simulation in 2D using GB curvature as a driving force and it explained 50% of the experimental grain growth. They also showed that an isotropic simulation with 17% explaining power had very poor matching with experimental evolution. ...
Preprint
Full-text available
The three-dimensional microstructure of Ni, observed after five annealing intervals, was compared to simulations of grain growth using the threshold dynamics method with the assumption of capillarity as the only driving force. A grain-by-grain comparison made it possible to identify the sources of differences between the simulation and experiment. The most significant difference was for grains of the smallest sizes, which the simulation predicted would lose volume and disappear at a greater rate than observed in the experiment. The loss of grains created errors in the numbers of neighbors of the remaining grains, and it was found that errors in the simulated grain volume were correlated to errors in the number of near neighbors. While anisotropic grain boundary properties likely play a role in the differences, the size dependence of the errors suggest that it might be necessary to include a size dependence in the model for grain boundary migration kinetics.
... There have been relatively few attempts in the past to compare the results of grain growth simulations with experiments on a grain-by-grain basis. For example, in 2003 Demirel et al. [44] validated an anisotropic grain growth simulation in 2D using GB curvature as a driving force and it explained 50% of the experimental grain growth. They also showed that an isotropic simulation with 17% explaining power had very poor matching with experimental evolution. ...
Article
Full-text available
The three-dimensional microstructure of Ni, observed after five annealing intervals, was compared to simulations of grain growth using the threshold dynamics method with the assumption of capillarity as the only driving force. A grain-by-grain comparison made it possible to identify the sources of differences between the simulation and experiment. The most significant difference was for grains of the smallest sizes, which the simulation predicted would lose volume and disappear at a greater rate than observed in the experiment. The loss of grains created errors in the numbers of neighbors of the remaining grains, and it was found that errors in the simulated grain volume were correlated to errors in the number of near neighbors. While anisotropic grain boundary properties likely play a role in the differences, the size dependence of the errors suggest that it might be necessary to include a size dependence in the model for grain boundary migration kinetics.
... Nevertheless, recent comparisons of experimental-and simulation-based results for the curvature-driven motion of SG coarsening in aluminum foils are in excellent agreement. 17) Herein, numerical simulation is carried out using the multi-phasefield (MPF) model developed by Steinbach et al.,2) which can reproduce the boundary motion controlled by local mean curvature. In the MPF model, a set of continuous order parameters ϕ i (i = 1, 2, ..., N) are defined to distinguish the orientation of the SGs, where ϕ i (r, t) represents the existence ratio of each orientation at position r and time t and N is the total number of order parameters. ...
Article
Full-text available
A unified theory for continuous and discontinuous annealing phenomena based on the subgrain growth mechanism was proposed by Humphreys around twenty years ago. With the developments in the unified subgrain growth theory, a number of Monte Carlo, vertex, and phase-field (PF) simulations have been carried out to investigate the nucleation and growth mechanisms of recrystallization by considering the local alignment of the subgrain structure. In this study, the effects of the microstructural inhomogeneities created in the deformed state on recrystallization kinetics and texture development were investigated. Numerical simulations of static recrystallization were performed in three-dimensional polycrystalline structures by coupling the unified subgrain growth theory with PF methodology. To prepare the initial microstructures, two-dimensional electron back scattering diffraction (EBSD) measurements were carried out on 90% and 99.8% cold-rolled pure iron. Our previous experimental study has shown that there are large differences in the texture formation processes during the recrystallization of cold-rolled iron samples. In cold-rolled iron with 90% reduction, the simulated texture exhibited nucleation and growth of γ-fiber (ND//<111>) grains at the cost of α-fiber (RD//<011>) components, where ND and RD denote normal direction and rolling direction, respectively. In contrast, the simulation results for cold-rolled iron with 99.8% reduction reproduced the high stability of the rolling texture during recrystallization. As a result, we conclude that the simulation results agreed with the experimentally observed textures in both the samples. Fullsize Image
... While this might be a reasonable approximation in some cases, it often fails to predict the motions of specific boundaries. 8,9 Of these physical quantities, perhaps the least is known about the curvature. In most work, H is assumed to be the inverse of the mean spherical equivalent grain radius. ...
Article
Full-text available
By mapping grain orientations on parallel serial sections of a SrTiO3 ceramic, it was possible to reconstruct three‐dimensional orientation maps containing more than 3000 grains. The grain boundaries were approximated by a continuous mesh of triangles and mean curvatures were determined for each triangle. The integral mean curvatures of grain faces were determined for all grains. Small grains with fewer than 16 neighbors mostly have positive mean curvatures while larger grains with more than 16 neighbors mostly have negative mean curvatures. It is also possible to correlate the mean curvature of individual triangles with the crystallographic characteristics of the grain boundary. The mean curvature is lowest for grain boundaries with (100) orientations and highest for grain boundaries with (111) orientations. This trend is inversely correlated to the relative areas of grain boundaries and directly correlated to the relative grain boundary energy. The direct correlation between the energy and curvature is consistent with the expected behavior of grain boundaries made up of singular orientations. Furthermore, because both the relative energy and curvature of grain boundaries with (100) orientations are minima in the distributions, these boundaries also have the lowest driving force for migration.
Article
Full-text available
Grain boundaries in polycrystalline materials migrate to reduce the total excess energy. It has recently been found that the factors governing migration rates of boundaries in bicrystals are insufficient to explain boundary migration in polycrystals. We first review our current understanding of the atomistic mechanisms of grain boundary migration based on simulations and high-resolution transmission electron microscopy observations. We then review our current understanding at the continuum scale based on simulations and observations using high-energy diffraction microscopy. We conclude that detailed comparisons of experimental observations with atomistic simulations of migration in polycrystals (rather than bicrystals) are required to better understand the mechanisms of grain boundary migration, that the driving force for grain boundary migration in polycrystals must include factors other than curvature, and that current simulations of grain growth are insufficient for reproducing experimental observations, possibly because of an inadequate representation of the driving force. Expected final online publication date for the Annual Review of Materials Research, Volume 53 is July 2023. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
Article
The grinding surface mechanical property is important for its working life span, which is determined by the microstructure distribution. To improve the machining mechanical performance, the greenly grinding generated heat is utilized to generate the coating layer without extra liquid. However, the coating layer microstructure composite is difficult to predict due to the non-steady grinding temperature. A digital complicated microstructure transformation (DCM) model is established and coupled with the transient grinding heat based on the validated improved finite difference method. The austenization, the ferrite reverted transformation, the cementite precipitation, the bainite transformation and the martensite transformation are all considered and automatically distinguished according to the grinding coating CCT. The analytic results indicate that the larger grinding depth mainly increases the bainite content, which can be restricted by improving the feeding rate, and the quicker feeding rate can increase the martensite content by three times. Combined with the measurement of corresponding representative mechanical performances (hardness and the residual stress), the achieved model is verified as an effective tool to realize the microstructure prediction. The research brings about a novelty macro mechanical property control of the grinding coating layer and is significant for achieving higher mechanical performance in surface strengthening manufacturing components.
Article
As part of rolled steel design and optimization, computational modeling and simulation is increasingly applied for predicting the properties and microstructure of final steel products. The time and cost of resorting to traditional trial‐and‐error based methods can be significantly reduced using the computational method. Among the numerical approaches, the finite element (FE) method has been customarily utilized in various cold and hot rolling processes over the years for its numerical efficiency and precision. The rolling process involves a thermo‐mechanical‐metallurgical phenomenon, and thus the interdisciplinary coupling of these is a challenge to rolling simulation and modeling the scientific community. Herein, recent studies on the FE modeling for predicting rolled shape, microstructural, and mechanical properties, among other, in steel rolling are extensively reviewed. An overview of the recent progress of the computational methods in multiscale computational architecture, which models the microstructural evolution and phase transformation by various sources, is particularly highlighted. This article is protected by copyright. All rights reserved.
Article
Revising grain growth Polycrystalline materials will often coarsen during annealing, a process that can modify the physical properties. A generally accepted theory ties the grain boundary curvature to the velocity at which grain boundaries migrate. Bhattacharya et al . measured this relationship for 52,000 grains in a nickel sample. They did not observe any relationship between curvature and velocity, suggesting that this accepted theory is not robust for polycrystalline samples. The observations require grain-coarsening models to be updated to more accurately predict the effects of annealing. —BG
Article
An improved phase field method by using statistical learning theory based optimization algorithm is developed for solving the phase field equations through building simple relationships between the key phase field variables and the phase evolution driving force, and using statistical analysis of mass computed data during phase field simulation. Phase field simulation results of growth of R phase and the B2–R phase transformation in a Ni-rich Ni50.5Ti49.5 alloy by using the proposed statistical strategy algorithm are compared with that using the conventional numerical algorithm, which demonstrates that with coupling the statistical learning theory, i.e., by means of the optimization algorithm, the credible simulated microstructure is obtained while maintaining high accuracy, and meanwhile the computational time has been significantly reduced.
Article
Full-text available
A unified theory for continuous and discontinuous annealing phenomena based on the subgrain growth mechanism was proposed by Humphreys about twenty years ago. With the developments in the unified subgrain growth theory, a number of Monte Carlo, vertex and phase-field (PF) simulations have been performed to investigate the nucleation and growth mechanisms of recrystallization by considering the local alignment of the subgrain structure. In this study, the effects of the microstructural inhomogeneities created in the deformed state on recrystallization kinetics and texture developments were investigated. Numerical simulations of static recrystallization were performed in 3D polycrystalline structures by coupling the unified subgrain growth theory with PF methodology. In order to prepare the initial microstructures, 2D electron back scattering diffraction (EBSD) measurements were performed on 90% and 99.8% cold-rolled pure iron. Our previous experimental study has shown that the texture formation processes in the recrystallization of those samples have large difference. In cold-rolled iron with 90% reduction, simulated texture exhibited nucleation and growth of γ-fiber (ND//<111>) grains by consuming α-fiber (RD//<011>) components, where ND and RD denote normal direction and rolling direction respectively. On the other hand, in cold-rolled iron with 99.8% reduction, simulation results reproduced the high stability of the rolling texture during recrystallization. As a consequence, the simulation results were in good agreements with experimentally observed textures in the both samples. Fullsize Image
Article
Full-text available
Experimental results on grain boundary properties and grain growth obtained using the Electron Backscattered Diffraction (EBSD) technique are compared with the Finite Element simulation results of an Al-foil with a columnar grain structure. The starting microstructure and grain boundary properties are implemented as an input for the three-dimensional grain growth simulation. In the computational model, minimization of the interface energy is the driving force for the grain boundary motion. The computed evolved microstructure is compared with the final experimental microstructure, after annealing at 550°C. Good agreement is observed between the experimentally obtained microstructure and the simulated microstructure. The constitutive description of the grain boundary properties was based on a 1-parameter characterization of the variation in mobility with misorientation angle.
Article
Full-text available
Microstructure evolution, where grain boundaries evolve by mean curvature motion, is modeled in three dimensions (3-D) using gradient-weighted moving finite elements (GWMFE). To do this, we modify and extend an existing 2-D GWMFE code to create a new code GRAIN3D which makes the 3-D microstructure modeling possible. The right-hand side term which drives the GWMFE motion can be viewed as surface tension forces, that is, as the negative gradient of the surface integral of a constant energy density µ on the triangular interfacial grid. Extensions to the method include equations for the motion of tetrahedra that are conformally attached to the moving piecewise linear triangular facets which represent the GWMFE discretization of the evolving grain boundaries. We present some new regularization terms which control element quality, as well as preventing element collapse in the simulation. New capabilities for changing the mesh topology are used to keep the grid edge lengths below a maximum allowable length hmax and to mimic actual changes in the physical topology, such as collapse and disappearance of individual grains. Validating runs are performed on some test cases that can be analytically solved, including collapse of a spherical grain and the case of columnar microstructure. In the spherical collapse case, the GWMFE method appears to have an error in the surface area collapse rate − dA dt which is O((∆θ) 2), where ∆θ is a measure of the angular resolution of the mesh. Finally, a run is presented where a true 3-D microstructure (possessing triple lines and quadruple points in the interior and triple points on the exterior boundaries) is evolved to a "2-D" columnar microstructure and finally evolved down to a single grain.
Chapter
Full-text available
Automated orientation measurement on a local basis is now widely accepted for characterization of materials. The technique relies upon indexing of electron backscatter diffraction patterns in a scanning electron microscope. In order to exploit the available information, it is important to understand its limitations with respect to accuracy. Experiments were carried out to measure orientation fields from a silicon single crystal. The orientation dispersion was 1°. Disorientation correlation maps revealed anisotropy in the spatial variation in measured orientation.
Article
A model is presented for the recovery of dislocations in the subgrain interiors and in the subboundaries. In addition to the pairwise annihilation the transfer of dislocations between the interior and the boundaries is considered. It is demonstrated that the recovery of the dislocations in the subgrain interiors in general occurs in two steps: first by pairwise annihilation and then by absorption at the boundaries, the latter process being essentially slower. The contribution from two mechanisms for subgrain coalescence is derived: extraction and emission of dislocations from the subboundaries. These contributions are compared to that from migration of boundary dislocations derived by Li. A comparison is also made to the subgrain growth rate due to subboundary migration. The two latter contributions which both involve the collective migration of dislocations, and are therefore conveniently referred to as such, are identical up to a numerical factor and occur consequently always together. Comparison to experiments will be made in a forthcoming paper.
Article
Kinetics of grain boundary migration in high-purity copper bicrystals with common [001] rotation axes have been investigated as a function of driving force, angle of misorientation and temperature. In all cases, grain boundary displacements obey a parabolic rate law and concomitant velocities are linearly related to driving force implying that intrinsic mobilities have been measured. Migration of low-angle boundaries is characterized by an activation energy of 49 kcal/mole. Close agreement between this value and 47 kcal/mole for lattice self diffusion in copper indicates that motion of low-angle boundaries is controlled by dislocation climb via a lattice self-diffusion mechanism. Migration of random high-angle boundaries, on the other hand, is characterized by activation energies ranging from 26 to 30 kcal/mole indicating that motion is controlled by direct atom transfer across the boundary. For a given misorientation, it has been possible to study boundary migration either in a pure tilt configuration or with varying degrees of twist character merely by varying orientation of the [001] rotation axes relative to the specimen surfaces. It is observed that values of the critical misorientation for transition from low to high-angle migration kinetics are about 9° and 13° for mixed and pure tilt boundaries respectively. Differences in mobility occur between the two configurations even at large misorientations, presumably due to the anisotropic nature of boundary diffusion.
Article
In classical simulation of abnormal grain growth, the microstructure is generally assumed to be homogeneous in size and the texture randomly distributed on such a theoretical microstructure which characterizes the primary recrystallized state. Experimentally, this last assumption could not be verified in Fe3%Si sheets (grade HiB). Consequently, a new simulation procedure has been implemented using a real microstructure measured by Orientation Imaging Microscopy (OIM) as an input data set. In this first work, the influence of the texture inhomogeneity has been analyzed from a theoretical point of view using Orientation Distribution Function (ODF) calculations and Monte-Carlo simulations without correlation with experimental measurements. The present paper deals with firstly the evolution of both microstructural and textural inhomogeneities and secondly the comparison, grain by grain, at several time steps of the experimental and simulated results obtained during secondary recrystallization.
Article
This paper describes a method for extracting, from measurements of a polycrystal, the relative excess free energy and the relative mobility of the grain boundaries as functions of the crystallographic type (five degrees of freedom) and relevant thermodynamic variables. The method requires the simultaneous measurement of both the geometry and the crystallography of a large number of grain boundary intersections; the crystallographic information may be obtained from orientation imaging microscopy (OIM). For simplicity, the intersections will be assumed to be triple junctions (three intersecting boundaries). The energies and mobilities are obtained as ratios to respective standards that must be determined independently. In section 2, the authors review Herring`s equation relating the grain boundary energies at a triple junction to the intersection angles. In section 3 they present the equation connecting the mobilities. In section 4, the mathematical method will be illustrated for the energy case, using the simplifying assumption that the energy is independent of the grain boundary inclination and that the misorientation can be specified by one angle as in a fiber texture. The illustration comprises 1.3 {times} 10{sup 5} randomly generated triple junctions with an assumed hypothetical free energy function.
Article
The relative grain boundary energies and mobilities in Al foil are investigated. These energies and mobilities are extracted from the geometry and crystallography of triple junctions. This approach is based on the assumption of local equilibrium of grain boundary energies at triple junctions.
Article
It is shown by transmission electron microscopy that the migration of small angle boundaries occurs by the movement of steps which have a high mobility along the boundaries. A dislocation model for the steps is proposed. According to this model the steps are singularities in the small angle boundary. The steps migrate by short-circuit diffusion of atoms along the core of the dislocation which forms the step. In the parts of the small angle boundary between the steps migrate by volume diffusion. It is shown that coarsening of sub-grains is due to the migration of small angle boundaries.
Article
The orientation dependence of the grain-boundary energy is a driving force for grain rotation across the boundary in a bicrystal. This has been observed experimentally. Since the energy gained during such a rotation must be entirely transformed into heat, and since this process occurs because of the friction of the grain-boundary dislocations on the lattice, a simple expression of the rate of energy dissipation as a function of the rate of rotation can be established. It is then demonstrated that the rate of rotation of the grains is the product of a "torque" (gain of energy per unit change of misorientation) times a mobility coefficient. The latter is a function of the misorientation angle, a feature which had been overlooked in the classical theory. Close to zero misorientation, where the grain-boundary energy exhibits a cusp, the mobility decays algebraically with the misorientation angle, while the torque diverges logarithmically only. As a consequence, the rate of rotation decreases with the misorientation, in agreement with experiment. Existing quantitative data are reasonably well fitted by this model.