Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2012 Nov 20;109(50):20314–20319. doi: 10.1073/pnas.1210456109

Directional grain growth from anisotropic kinetic roughening of grain boundaries in sheared colloidal crystals

Shreyas Gokhale a,1, K Hima Nagamanasa b,1, V Santhosh c, A K Sood a,c, Rajesh Ganapathy c,2
PMCID: PMC3528504  PMID: 23169661

Abstract

The fabrication of functional materials via grain growth engineering implicitly relies on altering the mobilities of grain boundaries (GBs) by applying external fields. Although computer simulations have alluded to kinetic roughening as a potential mechanism for modifying GB mobilities, its implications for grain growth have remained largely unexplored owing to difficulties in bridging the widely separated length and time scales. Here, by imaging GB particle dynamics as well as grain network evolution under shear, we present direct evidence for kinetic roughening of GBs and unravel its connection to grain growth in driven colloidal polycrystals. The capillary fluctuation method allows us to quantitatively extract shear-dependent effective mobilities. Remarkably, our experiments reveal that for sufficiently large strains, GBs with normals parallel to shear undergo preferential kinetic roughening, resulting in anisotropic enhancement of effective mobilities and hence directional grain growth. Single-particle level analysis shows that the mobility anisotropy emerges from strain-induced directional enhancement of activated particle hops normal to the GB plane. We expect our results to influence materials fabrication strategies for atomic and block copolymeric polycrystals as well.

Keywords: colloids, grain boundary migration, anisotropic grain growth


The motion and rearrangement of grain boundaries (GBs) is central to our understanding of recrystallization (1), grain growth and its stagnation (2), and superplasticity (3) in a broad class of polycrystalline materials including metals (4), ceramics (5), colloidal crystals (6), and block copolymers (7, 8). Polycrystals are pervasive as engineering materials and elucidating mechanisms that determine the structure and dynamics of their GBs continue to be a central goal of materials research (9). Advances in computer simulation methods (2, 10, 11) and experiments (12, 13) have provided substantial insights into the microscopic origins of these mechanisms. In conventional materials, nevertheless, establishing a direct link between the dynamics at the single/few atom length scale and the collective behavior of the many thousands of atoms that constitute GBs and grains poses a serious challenge (14).

A bridging of length scales is vital for grain growth studies. The key parameter in grain growth and its stagnation is the mobility of GBs, which is determined by their roughness (2, 15). Although transmission electron microscopy (TEM) is well-suited for grain growth measurements, no technique exists that can nonintrusively quantify the atomic scale roughness of buried GBs (14). These experimental shortcomings are compounded in driven polycrystals, which assume practical significance in the fabrication of functional materials via grain growth engineering (2, 7). Driven GBs (16) are thought to kinetically roughen (17), which may result in significantly enhanced mobilities (15) and influence grain growth. Further, in nanocrystalline materials, which often possess superior mechanical properties compared with their coarse-grained counterparts (18), kinetic roughening is expected to have a profound impact on mechanical behavior (16). Despite its technological relevance, the implications of kinetic roughening for GB mobility and grain growth remain poorly understood. Although molecular dynamics simulations can model atomistic phenomena realistically, they often are limited to small system sizes (19) and short time scales (11), which may influence measurements of GB roughness (15) and predictions of grain growth laws (2, 20). Kinetic Monte Carlo simulations (21) and phase field models (22) are insensitive to microscopic details. Thus, there is a need for a complementary multiscale approach that can probe kinetic roughening and its link to grain growth in driven polycrystals.

In this work, we combined techniques in colloid science to investigate the dynamics of GBs in sheared 3D colloidal crystals over length scales spanning from the single-particle level to that of the GB network. Our single-particle resolution experiments reveal that for large oscillatory strains, boundaries with normals oriented parallel to shear are preferentially kinetically roughened. We find that the equilibrium capillary fluctuation method (CFM) allows us to extract effective mobilities, even under shear. Analysis of dynamics at the single-particle level reveals that strain-induced anisotropy in particle displacements is manifested as a preferential enhancement of the effective mobilities. Most strikingly, by imaging the GB network under shear, we show that this shear-induced anisotropy in the mobility leads to directional grain growth. Because colloid models are valuable analogs of atomic systems (2327), insights gained from the present studies should be relevant for atomic and block copolymeric polycrystals as well.

Our system consisted of thermo-responsive size-tunable poly(N-isopropylacrylamide) (PNIPAm) colloids tagged with the fluorophore rhodamine 6G for confocal imaging (Materials and Methods). We adjusted the suspension volume fractions ϕ such that at 311 K, where the particle diameter Inline graphic μm, Inline graphic, and the equilibrium phase is a liquid. Annealing the samples to 296 K led to particle swelling, with Inline graphic μm, resulting in a Inline graphic, and the suspension crystallized into a polycrystalline state. A particular advantage of PNIPAm colloids is that they allow us to control the grain size systematically by altering the annealing rate (SI Text and Fig. S1). To mimic the nanocrystalline regime in which GB-mediated response is predominant (12, 18) and grain sizes typically are 50–200 atom diameters across, we aimed for average grain sizes of the order of 100 colloid diameters. We restricted our attention to high-angle grain boundaries (HAGBs) formed when adjacent grains that straddle the boundary have a misorientation angle Inline graphic. Geometrically, GBs can be described by a planar array of misfit dislocations (28). In HAGBs, the density of geometrically necessary dislocations is large and the dislocation cores overlap, leading to a continuous disordered interface.

Results and Discussion

Using our confocal rheometer, we simultaneously imaged the real-space particle-level microstructure of colloidal polycrystals subjected to a controlled shear deformation (Materials and Methods, SI Text, and Fig. S2). Samples were loaded in a custom-designed temperature-controlled parallel-plate shear cell with a typical plate separation of Inline graphic μm, corresponding to roughly 60 crystalline layers. Thermal annealing of the colloidal suspension at a rate of 0.125 K/min yielded the desired grain size with the Inline graphic crystal planes parallel to the walls of our shear cell; therefore, our studies are confined to pure tilt boundaries. The plane Inline graphic in Fig. 1A shows a snapshot of an HAGB. We quantify the dynamics of the GB interface (SI Text) by the height function Inline graphic, which is the 1D string formed by the intersection of the GB plane n with the imaging plane Inline graphic (thick white line in Fig. 1A). The image plane Inline graphic typically is 8–10 layers above the bottom plate to avoid wall effects. To investigate shear-induced anisotropic effects, experiments consisted of studying the dynamics of Inline graphic at different oscillatory shear strains, γ, for two GB configurations: Inline graphic (denoted by GB) and Inline graphic (denoted by GB), where v is the velocity (Movie S1). Inline graphic was changed in our experiments by keeping the oscillation frequency fixed at Inline graphic rad/s and changing the strain amplitude Inline graphic. The values of Inline graphic quoted here have been corrected to incorporate the effect of wall slip (SI Text and Fig. S3). GB properties sensitively depend on Inline graphic (4, 27, 28), and hence we focused on HAGBs within a narrow range of Inline graphics, Inline graphic, to facilitate quantitative comparison of results from experiments on GB and GB.

Fig. 1.

Fig. 1.

Kinetic roughening of HAGBs. (A) Schematic of the shear geometry. The planes Inline graphic and Inline graphic form the top and bottom plates of our shear cell. The imaging plane Inline graphic shows two crystallites with Inline graphic. The GB plane is the shaded region denoted by normal n. This figure corresponds to the configuration Inline graphic. (BE) Probability distribution of height fluctuations Inline graphic for increasing Inline graphic: (B) Inline graphic, (C) Inline graphic, (D) Inline graphic, and (E) Inline graphic. In CE, solid curves and dashed curves are gaussian fits to histograms that correspond to GB|| and GB, respectively. The circles correspond to w, and the width of the gray-shaded region corresponds to the w for Inline graphic.

Kinetic Roughening of GBs.

To characterize the roughness of the interface, we quantify a related parameter: the interface width, w, defined as the root-mean-squared (rms) fluctuations of Inline graphic (29). In Fig. 1 BE, we plot the normalized histograms of height fluctuations Inline graphic for increasing Inline graphic. Here, Inline graphic. For all Inline graphic, Inline graphic is gaussian for GB (solid curves) and GB (dashed curves), allowing us to estimate w (shown by circles) directly from the fits (Fig. 1 CE). For Inline graphic, Inline graphic (Fig. 1 B and C). Remarkably, we find that for Inline graphic, GB undergoes preferential kinetic roughening with Inline graphic (Fig. 1 D and E). Earlier studies addressed equilibrium roughening of HAGBs, in the context of faceting–defaceting transition (30), and kinetic roughening of Inline graphic GBs (16). An important finding from the present study is that general nonfaceted HAGBs also undergo kinetic roughening (15). Further, the migration mechanism of GBs is expected to change from a step-type motion to a continuous motion across the roughening transition. In Fig. S4, we plot the GB center-of-mass, defined as Inline graphic, with t. Although it is difficult to estimate the roughening transition temperature for our HAGBs, we indeed find that the motion of GBs at zero-shear and GB at Inline graphic is stepped and that of the kinetically roughened GB (Inline graphic) is continuous (Fig. S4).

Shear-Induced Anisotropy in GB Mobility.

Having quantified the roughness of GBs, we next focused on determining their mobility, M, which relates the migration velocity of the boundary Inline graphic to the driving force F (4). The net driving force acting on a boundary is composed of two parts, Inline graphic, where the first term is due to intrinsic boundary curvature, κ, and the second term is from externally imposed stresses. Here, Inline graphic is the GB stiffness (31). To calculate M, we first determined Inline graphic using the CFM (3234). In the present studies, a 1D section of the 2D GB interface is considered, which is not unreasonable as recent experiments have shown that Inline graphic for an equilibrium crystal–melt interface obtained by ignoring the third dimension completely is in good agreement with that obtained from 3D computer simulations (35). The equilibrium roughness of an interface is a tradeoff between the interfacial free energy, which is minimum for a smooth interface, and thermal energy. In CFM, thermally excited capillary fluctuations of the interface are decomposed into normal modes and the amplitude of the modes decays as Inline graphic. Here, Inline graphic is the thermal energy, k is the wavevector, and L is the interface length. Albeit CFM is an equilibrium statistical mechanics approach, studies on a driven colloidal liquid–gas interface have shown that shear leaves the k dependence of Inline graphic unchanged, which allows the definition of an effective shear-dependent stiffness, even when the system is out of equilibrium (36).

We perform a Fourier decomposition of Inline graphic (SI Text and Fig. S5) and plot Inline graphic vs. k for GB (solid circles) and GB (hollow circles) for increasing Inline graphic (Fig. 2 AC). In close parallel with the sheared colloidal liquid–gas interface (36), the equilibrium CFM analysis is valid for driven solid–solid interfaces as well (Fig. S6). We find that Inline graphic is nearly a constant Inline graphic over a decade in k, for both boundary configurations and for all Inline graphic. Notably, we find that Inline graphic, obtained by averaging Inline graphic over the shaded region (Fig. 2 AC), is preferentially lowered for GB (solid circles) compared with GB (hollow circles) for Inline graphic (Fig. 2D and Fig. S5A). An anisotropic lowering of Inline graphic for GB vs. GB is a result of anisotropic enhancement in the interface roughness and is consistent with Fig. 1 CE. It is important to note that in ref. 36 Inline graphic depends on the shear rate Inline graphic, whereas for HAGBs we find that Inline graphic depends strongly on Inline graphic (SI Text and Fig. S7).

Fig. 2.

Fig. 2.

Shear-induced anisotropy in GB mobility. Inline graphic vs. k for (A) Inline graphic, (B) Inline graphic, and (C) Inline graphic for GB|| (●) and GB (○). (D) Stiffness Inline graphic as a function of Inline graphic obtained from AC by averaging Inline graphic over the shaded region. (E) Mobility M and (F) reduced mobility Inline graphic as a function of Inline graphic for GB|| (●) and GB (○). In DF, (Inline graphic) corresponds to Inline graphic.

For capillary fluctuations that decay as Inline graphic, the dynamic correlation function Inline graphic (SI Text, Fig. S8, and Fig. S9 AC) yields the product Inline graphic (37). Here, ξ is the lateral correlation length (38) obtained from exponential fits to the height–height correlation function Inline graphic (Fig. S9 D and E) and Inline graphic is the complementary error function. Fig. S9 AC shows Inline graphic for GB (solid circles) and GB (hollow circles) for increasing Inline graphic. Plugging in the value of Inline graphic (Fig. 2D), obtained earlier, into the complementary error function fits of the data (black lines in Fig. S9 AC) yields an effective shear-dependent mobility M. We find that the increase in M with Inline graphic for the kinetically roughened boundary, GB (Fig. 2E, solid circles), is more rapid compared with GB (Fig. 2E, hollow circles) and is almost an order of magnitude larger at Inline graphic (Fig. S5B). It generally is believed that kinetic roughening alters GB mobilities and dictates microstructure evolution in driven polycrystals. However, even in the zero-driving force limit, a connection between roughness and mobility has been shown only in computer simulations (15) and its extension to nonzero driving forces has not been explored. By integrating fast confocal microscopy with rheology, we have shown that the shear-induced anisotropy in M is a direct consequence of preferential kinetic roughening.

Will shear-induced anisotropy in mobility, Inline graphic, lead to directional grain growth? To answer this question, we take a closer look at the mobility relation Inline graphic (31). Focusing on the second term, one would naively expect that an anisotropy in M implies an anisotropy in v. Although it is difficult to characterize Inline graphic acting on GBs in a driven polycrystal (2), it chiefly consists of contributions from the applied bulk stress and its coupling to the elastic anisotropy in adjacent grains (39). In our experiments, we expect the driving force due to bulk stress to be zero because of the oscillatory nature of the applied strain. Further, the driving force due to elastic anisotropy becomes significant only for grain sizes larger than about 1 μm (∼3,000σ) in atomic systems (39), which is much larger than the grain sizes used in our experiments (∼100σ). It therefore is likely that in our experiments, Inline graphic and hence Inline graphic. We note that with increasing Inline graphic, M increases (Fig. 2E) whereas Inline graphic is found to decrease (Fig. 2D) for both GB and GB. Thus, in the approximation Inline graphic, we expect an anisotropy in GB motion and grain growth only if the reduced mobility, Inline graphic, is itself anisotropic. Remarkably enough, Fig. 2F shows that anisotropic kinetic roughening indeed results in anisotropic-reduced mobilities with Inline graphic at large Inline graphic. Although we considered a narrow range of Inline graphics to allow comparison across the two GB configurations, given that HAGBs exhibit qualitatively similar dynamics over a broad range of Inline graphics (27, 40), we expect the anisotropy in M and Inline graphic to be a generic feature of these boundaries.

Directional Grain Growth Under Shear.

Even as our single-particle resolution measurements revealed a shear-induced anisotropy in both M and Inline graphic, to link this to grain growth it is necessary to probe GB dynamics under shear at the grain network length scale. To this end, we performed Bragg diffraction microscopy (BDM), which is the analog of TEM for colloidal crystals (41), simultaneously with rheometry (Materials and Methods, SI Text, and Fig. S2B). Here, the colloidal crystal sample, confined between the shear plates, is illuminated by a white-light source and a detector is placed at the first Bragg diffracted spot. A perfect single crystal would result in uniform intensity at the detector. For a polycrystal, however, the Bragg condition is met only by crystallites of a particular orientation, resulting in a diffraction contrast image. We note here that BDM does not yield an orientation map of the polycrystal but allows us to distinguish between low-angle GBs, which appear as arrays of discrete spots corresponding to dislocation cores, and HAGBs, which appear as continuous curves (42). A typical BDM snapshot of the colloidal polycrystal is shown in Fig. 3A, with GBs shown by solid lines and GBs shown by dashed lines. Fig. 3 B and C shows Inline graphic vs. t for Inline graphic. In complete concord with results thus far, anisotropy in M and Inline graphic indeed results in anisotropic grain growth with Inline graphic on an average (Movie S2; see also SI Text, Fig. S10, and Movie S3). As a consistency check, we estimated the expected GB velocity by plugging in the value of Inline graphic and assuming Inline graphic, where d is the average grain radius, which is ∼100 μm. The calculated GB velocity is in close agreement with those obtained from BDM measurements.

Fig. 3.

Fig. 3.

Directional grain growth under shear. (A) Snapshot of a colloidal polycrystal obtained using BDM. The white arrow labeled v shows the shear velocity direction. The curves highlight GBs for which migration velocities were measured. Solid curves indicate GB||s and the dashed curves indicate GBs. The arrows point along the direction of GB motion, and their length is proportional to the GB migration velocity Inline graphic. (B) Inline graphic vs. t for GB||s. (C) Inline graphic vs. t for GBs. The lines are least-squares fits to the data. The numbers shown adjacent to the GB center-of-mass profiles vs. time in B and C correspond to those in A. Boundary 4 was used as a reference for the drift correction for GB||s and GBs.

Microscopic Origins of Directional Grain Growth.

Our observations cannot be rationalized within the theoretical framework of shear-coupled GB migration because our HAGBs possess no identifiable structural units, the resolved shear stress has no component in the glide plane of boundary dislocations (4345), and the capillary fluctuation spectrum decays as Inline graphic (46). Instead, we chose to exploit the analogy between HAGBs and glass-forming liquids (27, 40). For HAGBs, the rate-controlling events for migration are single-atom hops across the boundary plane (47, 48) and the characteristic time associated with these hops is the cage-breaking time Inline graphic. We extracted Inline graphic from the inflection point in the mean first passage time Inline graphic (48), defined as the average time taken by a particle to traverse a distance r for the first time. We find that t* for γo = 2% is greater than t* for γo = 8.3% for both GB and GB (Fig. 4 A), implying that the activation barrier for cage breaking is lowered at a higher Inline graphic (SI Text and Fig. S11). This lowering of activation barriers is reminiscent of the strain-induced deformation of the potential energy landscape of sheared glasses (49). For a migrating boundary, the self-part of the van Hove correlation function Inline graphic for Inline graphic develops a characteristic peak at an intermediate distance r with Inline graphic. Here, Inline graphic corresponds to the cage size. However, Inline graphic (Fig. 4 B and C) do not exhibit peaks between Inline graphic and σ for both GB and GB even for Inline graphic, in striking resemblance to observations on stationary GBs (48). This is not entirely surprising given that over the duration of our confocal rheology experiments (420 s), the net displacement of Inline graphic is only about Inline graphic, even for GB||. Further, although displacements in all directions contribute to the peak at Inline graphic in Inline graphic (Fig. 4 B and C), hops normal to the GB plane are the ones that primarily influence boundary mobility. To determine whether shear biases activated hops, we plot the angular distribution of displacements contributing to the peak in Inline graphic for GB and GB. For Inline graphic, the displacements are more isotropically distributed (Fig. S12) compared with Inline graphic for which we find a significant enhancement along the shear direction for both GB and GB (Fig. 4 D and E). For GB, these shear-enhanced hops are normal to the boundary plane and therefore lead to a preferential increase in Inline graphic (Fig. S6) and M (Fig. 2E). Collectively, our observations show that strain enhances HAGB interface fluctuations without contributing to a bulk driving force for migration, consistent with our earlier assumption of Inline graphic, and is therefore analogous to temperature. As expected, similar to the variation of M and Inline graphic with temperature (50), the dependence of M on Inline graphic is much stronger than that of Inline graphic (Fig. 2). Not only does this rationalize the success of the CFM in extracting Inline graphic and M for sheared HAGBs, but it also accounts for the anisotropic enhancement in Inline graphic and therefore directional grain growth.

Fig. 4.

Fig. 4.

Single-particle dynamics at GBs. (A) Mean first passage time Inline graphic for GB|| [Inline graphic (■), Inline graphic (●)] and GB [Inline graphic (Inline graphic), Inline graphic (○)]. The dashed curves are Hill function fits used to extract the inflection point Inline graphic, shown as solid lines for GB|| and dashed lines for GB. (B and C) Self part of the van Hove correlation function Inline graphic for Inline graphic for GB|| (B) and GB (C) at Inline graphic, Inline graphic, and Inline graphic. (D and E) Angular distribution of displacements contributing to the nearest-neighbor peak of Inline graphic for Inline graphic over Inline graphic for GB|| (D) and GB (E). In D and E, the solid arrow denotes the shear direction and the dashed arrow denotes the GB normal. About 50Inline graphic of particle hops happen within a 40° window centered on the shear direction for both GB|| and GB.

Conclusions

By bringing together experimental techniques in colloid science, we have bridged vastly disparate length and time scales to unravel the microscopic underpinnings of grain growth in sheared polycrystals. We have shown conclusively that preferential kinetic roughening (Fig. 1 CE) and the resulting anisotropy in effective mobilities (Fig. 2E) ultimately lead to directional grain growth (Fig. 3). We find that strain enhances only the amplitudes of capillary fluctuation modes (Fig. 2 AC) but leaves the fluctuation spectrum itself qualitatively unchanged (Fig. S6). This strongly suggests that the paradigm of shear as an effective temperature, routinely adopted to describe driven glasses (51), is germane even to the HAGB interfaces investigated here (27). We observe that the strain-induced lowering of activation barriers and the concomitant anisotropic enhancement of particle displacements satisfactorily explain the observed anisotropy in GB mobility (Fig. 4). Owing to the close similarity in the physics of GBs across diverse systems (4, 8, 27, 37), we expect our results can be extended to atomic and block copolymeric polycrystals as well. Further, by combining the temperature-tunability of PNIPAm colloids with template-directed growth (25, 41), in principle it should be possible to design bicrystals and polycrystals of controlled grain crystallography and size and investigate their response to shear deformation. More importantly, our experiments exemplify a multiscale approach that can be applied readily to elucidate fundamental as well as technologically relevant phenomena, such as grain rotation and coalescence, GB sliding, and texture evolution in driven polycrystals.

Materials and Methods

PNIPAm colloids of diameter 950 nm (polydispersity <5%) were synthesized by the standard emulsion polymerization route. All the chemicals used were purchased from Sigma-Aldrich and had a purity in excess of 98%. Particles were purified using the method described in ref. 27. The purified samples were concentrated to yield Inline graphic at 296 K.

Confocal Rheology.

To facilitate confocal imaging under shear, we integrated a fast confocal microscope (Visitech VT-Eye) with a commercial rheometer (MCR-301, Anton Paar) mounted on a homemade mechanical stage (SI Text and Fig. S1A). Samples were imaged using a Leica objective (Plan Apochromat 100× N.A. 1.4, oil immersion) and a laser excitation centered at 514 nm. The field of view was a 54 × 54-μm slice containing ∼3,200 particles. Images were captured at 10 frames per second (fps) for Inline graphic = 2% and 8.3% and 26 fps for Inline graphic = 18.3%. The temperature was maintained at 296 K for all experiments. Standard codes were used for particle tracking (52), and subsequent analysis was performed using algorithms developed independently. The particle-tracking resolution as calculated from the micrometer to pixel ratio is 0.07 μm.

BDM Under Shear.

For grain growth measurements under shear, we integrated BDM with rheometry (SI Text). The sample was illuminated by a white light LED source incident at an angle θ = 56˚ (see Fig. S1B for a schematic). A low-magnification Leica objective (Plan Apochromat 10× N.A. 0.4, dry) was used to image the GB network. The temperature was maintained at 296 K, and an oscillatory strain of frequency ω = 1 rad/s and amplitude Inline graphic = 3.9% was applied to the sample. The 0.915 × 0.680-mm field of view contained ∼25 grains. Images were captured at 1 fps for 4 hours, and GB profiles were tracked manually at periodic intervals to generate their center-of-mass time series. To subtract drift contributions from the GB motion, a flat immobile boundary was used as a reference.

Supplementary Material

Supporting Information

Acknowledgments

The authors thank Jack Douglas and Vikram Deshpande for useful discussions. The authors also thank the anonymous reviewers for their valuable suggestions. S.G. thanks the Council for Scientific and Industrial Research (CSIR) India for a Shyama Prasad Mukherjee Fellowship, K.H.N. thanks CSIR India for a Senior Research Fellowship, A.K.S. thanks CSIR India for a Bhatnagar Fellowship, and R.G. thanks the International Centre for Materials Science and the Jawaharlal Nehru Centre for Advanced Scientific Research for financial support.

Footnotes

The authors declare no conflict of interest.

This article is a PNAS Direct Submission. J.A.W. is a guest editor invited by the Editorial Board.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1210456109/-/DCSupplemental.

References

  • 1.Humphreys FJ, Hatherly M. Recrystallization and Related Annealing Phenomena. 2nd Ed. Oxford, UK: Elsevier; 2004. [Google Scholar]
  • 2.Holm EA, Foiles SM. How grain growth stops: A mechanism for grain-growth stagnation in pure materials. Science. 2010;328(5982):1138–1141. doi: 10.1126/science.1187833. [DOI] [PubMed] [Google Scholar]
  • 3.Lu L, Sui ML, Lu K. Superplastic extensibility of nanocrystalline copper at room temperature. Science. 2000;287(5457):1463–1466. doi: 10.1126/science.287.5457.1463. [DOI] [PubMed] [Google Scholar]
  • 4.Gottstein G, Shvindlerman LS. Grain Boundary Migration in Metals: Thermodynamics, Kinetics, Applications. Boca Raton, FL: Taylor & Francis; 2010. 2nd Ed. [Google Scholar]
  • 5.Powers JD, Glaeser AM. Grain boundary migration in ceramics. Interface Sci. 1998;6(1):23–39. [Google Scholar]
  • 6.Palberg T, Monch W, Schwarz J, Leiderer P. Grain size control in polycrystalline colloidal solids. J Chem Phys. 1995;102(12):5082–5087. [Google Scholar]
  • 7.Torija MA, Choi SH, Lodge TP, Bates FS. Large amplitude oscillatory shear of block copolymer spheres on a body-centered cubic lattice: Are micelles like metals? J Phys Chem B. 2011;115(19):5840–5848. doi: 10.1021/jp202468y. [DOI] [PubMed] [Google Scholar]
  • 8.Ryu HJ, Fortner DB, Rohrer GS, Bockstaller MR. Measuring relative grain-boundary energies in block-copolymer microstructures. Phys Rev Lett. 2012;108(10):107801-1–5. doi: 10.1103/PhysRevLett.108.107801. [DOI] [PubMed] [Google Scholar]
  • 9.Thompson CV. Structure evolution during processing of polycrystalline films. Annu Rev Mater Sci. 2000;30(1):159–190. [Google Scholar]
  • 10.Janssens KGF, et al. Computing the mobility of grain boundaries. Nat Mater. 2006;5(2):124–127. doi: 10.1038/nmat1559. [DOI] [PubMed] [Google Scholar]
  • 11.Trautt ZT, Upmanyu M, Karma A. Interface mobility from interface random walk. Science. 2006;314(5799):632–635. doi: 10.1126/science.1131988. [DOI] [PubMed] [Google Scholar]
  • 12.Rupert TJ, Gianola DS, Gan Y, Hemker KJ. Experimental observations of stress-driven grain boundary migration. Science. 2009;326(5960):1686–1690. doi: 10.1126/science.1178226. [DOI] [PubMed] [Google Scholar]
  • 13.Wang Z, et al. Atom-resolved imaging of ordered defect superstructures at individual grain boundaries. Nature. 2011;479(7373):380–383. doi: 10.1038/nature10593. [DOI] [PubMed] [Google Scholar]
  • 14.Van Swygenhoven H. Polycrystalline materials. Grain boundaries and dislocations. Science. 2002;296(5565):66–67. doi: 10.1126/science.1071040. [DOI] [PubMed] [Google Scholar]
  • 15.Olmsted DL, Foiles SM, Holm EA. Grain boundary interface roughening transition and its effect on grain boundary mobility for non-faceting boundaries. Scr Mater. 2007;57(12):1161–1164. [Google Scholar]
  • 16.Lee SB, et al. Kinetic roughening of a ZnO grain boundary. Appl Phys Lett. 2010;96(19):1919061-3. [Google Scholar]
  • 17.Halpin-Healy T, Zhang Y-C. Kinetic roughening phenomena, stochastic growth, directed polymers and all that. Aspects of multidisciplinary statistical mechanics. Phys Rep. 1995;254(4):215–414. [Google Scholar]
  • 18.Meyers MA, Mishra A, Benson DJ. Mechanical properties of nanocrystalline materials. Prog Mater Sci. 2006;51(4):427–556. [Google Scholar]
  • 19.Cherkaoui M, Capolungo L. Atomistic and continuum modeling of nanocrystalline materials. In: Hull R, Parisi J, Osgood RM Jr., Warlimont H, editors. Springer Series in Material Science. Vol 112. New York: Springer Science+Business Media; 2009. p. 81. [Google Scholar]
  • 20.Farkas D, Mohanty S, Monk J. Linear grain growth kinetics and rotation in nanocrystalline Ni. Phys Rev Lett. 2007;98(16):165502-1–4. doi: 10.1103/PhysRevLett.98.165502. [DOI] [PubMed] [Google Scholar]
  • 21.Anderson MP, Srolovitz DJ, Grest GS, Sahni PS. Computer simulation of grain growth—I Kinetics. Acta Metall. 1984;32(5):783–791. [Google Scholar]
  • 22.Chen L-Q. Phase-field models for microstructure evolution. Annu Rev Mater Res. 2002;32(1):113–140. [Google Scholar]
  • 23.Suresh S. Crystal deformation: Colloid model for atoms. Nat Mater. 2006;5(4):253–254. doi: 10.1038/nmat1621. [DOI] [PubMed] [Google Scholar]
  • 24.Schall P, Weitz DA, Spaepen F. Structural rearrangements that govern flow in colloidal glasses. Science. 2007;318(5858):1895–1899. doi: 10.1126/science.1149308. [DOI] [PubMed] [Google Scholar]
  • 25.Ganapathy R, Buckley MR, Gerbode SJ, Cohen I. Direct measurements of island growth and step-edge barriers in colloidal epitaxy. Science. 2010;327(5964):445–448. doi: 10.1126/science.1179947. [DOI] [PubMed] [Google Scholar]
  • 26.Han Y, et al. Geometric frustration in buckled colloidal monolayers. Nature. 2008;456(7224):898–903. doi: 10.1038/nature07595. [DOI] [PubMed] [Google Scholar]
  • 27.Nagamanasa KH, Gokhale S, Ganapathy R, Sood AK. Confined glassy dynamics at grain boundaries in colloidal crystals. Proc Natl Acad Sci USA. 2011;108(28):11323–11326. doi: 10.1073/pnas.1101858108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Read WT, Shockley W. Dislocation models of crystal grain boundaries. Phys Rev. 1950;78(3):275–289. [Google Scholar]
  • 29.Barabasi A-L, Stanley HE. Fractal Concepts in Surface Growth. New York: Cambridge Univ Press; 1995. [Google Scholar]
  • 30.Yoon DY, Cho YK. Roughening transitions of grain boundaries in metals and oxides. J Mater Sci. 2005;40(4):861–870. [Google Scholar]
  • 31.Lobkovsky AE, Karma A, Mendelev MI, Haataja M, Srolovitz DJ. Grain shape, grain boundary mobility and the Herring relation. Acta Mater. 2004;52(2):285–292. [Google Scholar]
  • 32.Hoyt JJ, Asta M, Karma A. Atomistic and continuum modeling of dendritic solidification. Mater Sci Eng Rep. 2003;41(6):121–163. [Google Scholar]
  • 33.Trautt ZT, Upmanyu M. Direct two-dimensional calculations of grain boundary stiffness. Scr Mater. 2005;52(11):1175–1179. [Google Scholar]
  • 34.Foiles SM, Hoyt JJ. Computation of grain boundary stiffness and mobility from boundary fluctuations. Acta Mater. 2006;54(12):3351–3357. [Google Scholar]
  • 35.Nguyen VD, Hu Z, Schall P. Single crystal growth and anisotropic crystal-fluid interfacial free energy in soft colloidal systems. Phys Rev E Stat Nonlin Soft Matter Phys. 2011;84(1):011607-1–6. doi: 10.1103/PhysRevE.84.011607. [DOI] [PubMed] [Google Scholar]
  • 36.Derks D, Aarts DGAL, Bonn D, Lekkerkerker HN, Imhof A. Suppression of thermally excited capillary waves by shear flow. Phys Rev Lett. 2006;97(3):038301-1–4. doi: 10.1103/PhysRevLett.97.038301. [DOI] [PubMed] [Google Scholar]
  • 37.Skinner TOE, Aarts DGAL, Dullens RPA. Grain-boundary fluctuations in two-dimensional colloidal crystals. Phys Rev Lett. 2010;105(16):168301-1–4. doi: 10.1103/PhysRevLett.105.168301. [DOI] [PubMed] [Google Scholar]
  • 38.Werner A, Schmid F, Muller M, Binder K. Anomalous size-dependence of interfacial profiles between coexisting phases of polymer mixtures in thin-film geometry: A Monte Carlo simulation. J Chem Phys. 1997;107(19):8175–8188. [Google Scholar]
  • 39.Haslam AJ, et al. Stress-enhanced grain growth in a nanocrystalline material by molecular-dynamics simulation. Acta Mater. 2003;51(7):2097–2112. [Google Scholar]
  • 40.Zhang H, Srolovitz DJ, Douglas JF, Warren JA. Grain boundaries exhibit the dynamics of glass-forming liquids. Proc Natl Acad Sci USA. 2009;106(19):7735–7740. doi: 10.1073/pnas.0900227106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Schall P, Cohen I, Weitz DA, Spaepen F. Visualizing dislocation nucleation by indenting colloidal crystals. Nature. 2006;440(7082):319–323. doi: 10.1038/nature04557. [DOI] [PubMed] [Google Scholar]
  • 42.Pieranski P. In: Physics of Defects. Balian R, Kleman M, Poirier JP, editors. Amsterdam: North-Holland; 1981. pp. 183–200. [Google Scholar]
  • 43.Cahn JW, Mishin Y, Suzuki A. Coupling grain boundary motion to shear deformation. Acta Mater. 2006;54(19):4953–4975. [Google Scholar]
  • 44.Cahn JW, Taylor JE. A unified approach to motion of grain boundaries, relative tangential translation along grain boundaries, and grain rotation. Acta Mater. 2004;52(16):4887–4898. [Google Scholar]
  • 45.Trautt ZT, Adland A, Karma A, Mishin Y. Coupled motion of asymmetrical tilt grain boundaries: Molecular dynamics and phase field computer simulations. Acta Mater. 2012;60(19):6528–6546. [Google Scholar]
  • 46.Karma A, Trautt ZT, Mishin Y. Relation between equilibrium fluctuations and shear-coupled motion of grain boundaries. Phys Rev Lett. 2012;109(9):095501-1–5. doi: 10.1103/PhysRevLett.109.095501. [DOI] [PubMed] [Google Scholar]
  • 47.Zhang H, Srolovitz DJ, Douglas JF, Warren JA. Atomic motion during the migration of general [001] tilt grain boundaries in Ni. Acta Mater. 2007;55(13):4527–4533. [Google Scholar]
  • 48.Zhang H, Srolovitz DJ, Douglas JF, Warren JA. Characterization of atomic motion governing grain boundary migration. Phys Rev B Condens Matter. 2006;74(11):115404-1–10. [Google Scholar]
  • 49.Sollich P, Lequeux F, Hebraud P, Cates ME. Rheology of soft glassy materials. Phys Rev Lett. 1997;78(10):2020–2023. [Google Scholar]
  • 50.Zhang H, Du D, Srolovitz DJ, Mendelev MI. Determination of grain boundary stiffness from molecular dynamics simulation. Appl Phys Lett. 2006;88(12):121927-1–3. [Google Scholar]
  • 51.Haxton TK, Liu AJ. Activated dynamics and effective temperature in a steady state sheared glass. Phys Rev Lett. 2007;99(19):195701-1–4. doi: 10.1103/PhysRevLett.99.195701. [DOI] [PubMed] [Google Scholar]
  • 52.Crocker JC, Grier DG. Methods of digital video microscopy for colloidal studies. J Colloid Interface Sci. 1996;179(1):298–310. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information
Download video file (4.7MB, mov)
Download video file (939.9KB, mov)
Download video file (615.7KB, mov)

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES