Marco
Pinna
*a,
Javier
Diaz
*bc,
Christopher
Denison
a,
Andrei
Zvelindovsky
a and
Ignacio
Pagonabarraga
bc
aSchool of Mathematics and Physics, College of Health and Science, Centre for Computational Physics, University of Lincoln, Brayford Pool, Lincoln, LN6 7TS, UK. E-mail: mpinna@lincoln.ac.uk
bDepartament de Física de la Matèria Condensada, Universitat de Barcelona, Martí i Franqués 1, 08028 Barcelona, Spain. E-mail: jdiazbranas@ub.edu
cUniversitat de Barcelona Institute of Complex Systems (UBICS), Universitat de Barcelona, 08028 Barcelona, Spain
First published on 16th December 2024
The potential applications of block copolymer thin films, utilising their self-assembly capabilities, are enhanced when achieving long-range ordering. In this study we explain the experimental alignment of lamellae under shear flow findings [S. Pujari et al. Soft Matter, 2012, 8, 5258] and classify the alignment mechanisms based on shear rate and segregation, uncovering similarities to the systems subjected to electric fields, suggesting a common pathway of lamellae orientations. However, the presence of thin films surfaces introduces distinct features in the lamellae orientation under shear compared to electric fields. Notably, we observe the emergence of a three-dimensional rotation alongside the conventional two-dimensional rotation. Furthermore, a transient regime has been identified within the melting mechanism, which confirms the existence of the checkboard pattern proposed by Schneider et al. [Macromolecules, 2018, 51, 4642]. These findings significantly enhance our understanding of block copolymer alignments and shed light on the intricate interplay between external fields and the lamellar structure.
Experimentally, long range orientation applying oscillatory shear on poly(ethylene-propylene)-poly(ethylethylene) (PEPPEE) diblock copolymer melts (in bulk systems) has been obtained by Koppi et al.35,36 They found two distinct orientations: lamellae oriented perpendicular to the shear plane and coincident with the shear direction, and lamellae parallel to the plane of shear. These orientations were dependent on the polymer chemistry, frequency and temperature of the studied system.
Later, register's group studied experimentally the lamellae morphology in neutral ultra-thin films under a shear flow. They found alignment of perpendicular lamellae to the shear plane along the direction of the applied shear flow.37 Computationally such orientation has been obtained in bulk by dynamic density functional theory.38 The stability of this configuration has been studied, in bulk, using dissipative particle dynamics (DPD).39 In this study, they found that for higher shear rates the long-range orientational order of the lamellae inside the grains disintegrates while at lower shear rates there is a shrinking of the unstable grains.39 Experimentally, it has also been demonstrated that a stress level exceeding 10 KPa is necessary to align the lamellae in the direction of shear flow, even when the distance between domains is less than one. In the case of thicker layers, researchers have noted the presence of a polystyrene layer at the top of the film (parallel lamellae), which can be eliminated by etching the film.37
In this work, we will perform a systematic study of the aligning mechanism of lamellar-forming BCP in ultra-thin neutral films under shear flow. The thickness of the film will be explored in order to understand the appearance of parallel lamellar orientation in experiments of thicker films,37 as well as the degree of segregation of the lamellar domains. A mesoscopic cell dynamic simulation (CDS) method will be used to capture the dynamic behaviour of BCP under shear stress. We use a simple model where the hydrodynamic interactions are neglected in this work. In a weakly segregated system, the interfaces are diffuse and the viscosity variation is minimal.40 Thus, the linear velocity profile is a first approximation, disregarding viscosity mismatches and interfacial tension-driven flows. The general solution for the velocity field can be obtained from hydrodynamics. It has been shown that for a layered (lamellae) system with different polymer viscosities, the solution adopts the linear velocity profile.40 This refinement is essential for describing the temperature-dependent orientation of the mesostructure lattice in the gradient-vorticity plane at high shear rates (see ref. 41). Nevertheless, in the weak segregation regime, the main aspects of flow alignment can be accurately described by a diffusion-convection equation with the imposed velocity profile from eqn (3). Numerous studies have demonstrated that this equation effectively captures many relevant phenomena in sheared inhomogeneous fluids.,40–55 Furthermore, if the polymers have equal bulk viscosity, the approximation is even more accurate. Using this approximation, we have a model that is considerably fast allowing us to explore large systems and phase diagrams where the migration of defects plays an important role to the different mechanisms. This particular theoretical model has an established record on reproducing experimental results as seen in ref. 15 and 55–57.
(1) |
(2) |
A shear flow with velocity field v = (vx,vy,vz) is considered, leading to the modified Cahn–Hilliard dynamics43,59
(3) |
Eqn (3) can be made explicit as
(4) |
Wi = tc | (5) |
Periodic boundary conditions are imposed across the x and the y horizontal directions, while hard walls are placed at the top and bottom of the system (z = 0 and z = Lz). Dirichlet boundary conditions have been imposed on ψ, with ψ(0) = ψbottoms and ψ(Lz) = ψtops and a Neumann boundary conditions for the chemical potential μ(r) = δF/δψ(r) at the surface δμ/δz (0) = δμ/δz(Lz) = 0. For simplicity we will limit to the case of ψbottoms = ψtops = ψs that has identical walls (see Fig. 2-left side). Fig. 2 on the left side shows the 3D studied system while the right side shows the obtained stable lamellae configuration once the steady shear flow has been applied.
The order parameter time evolution, eqn (4), is numerically solved using a CDS61,62 for which the Laplacian is approximated as where the term a0 represents the grid size unit and 〈〈ψ〉〉 is given by (ref. 61)
(6) |
Fig. 3 A stencil for Laplacian, where NN denotes nearest neighbours, NNN-next-nearest neighbours and NNNN-next next nearest neighbours to the point r. |
In order to compare simulations with experiments, it is necessary to use large simulation boxes, which can be reached thanks to the computational efficiency of the CDS model and a parallel implementation of the algorithm.63,64
Structure factor S(q,t) = |ψq(t)|2 is used to analyse the morphology of the BCP melt, with ψq(t) being the Fourier transform of the order parameter ψ. In particular, the 2D structure factor can be calculated from the 2D slice at a distance z = Lz/2 in the middle of the thin film. The corresponding 2D structure factor S(qxy,ϕ) can be averaged radially to obtain the angular scattering function S(ϕ) where ϕ is the polar angle in the x−y plane.
Mean absolute value of the order parameter 〈|ψ|〉 provides information on the degree of segregation of the BCP melt.65 A disordered, homogeneous BCP state is characterised by ψ = 0, while a phase-separated BCP, in the bulk, takes values |ψ| ∼ ψeq. This global parameter 〈|ψ|〉 provides information on the degree of phase separation of the BCP over time.
Nematic order parameters Sα = 〈(3((r)α)2 − 1)/2〉 describe the nematic order of the BCP with each axis α = x,y,z to provide information on the nematic alignment of the BCP with respect to the external shear flow, which is applied in the x direction. The local orientation of the lamella domains is calculated as (r) = ∇ψ/(|∇ψ|) and the average is weighted with (∇ψ)2.
Fig. 5 shows the different mechanism of alignment of the lamellar systems under shear flow in terms of the shear rate and the temperature-like parameter τ. The initial configuration of the lamellar morphology is similar to the method performed in ref. 33 and 56 where a pre-aligned lamellar system with two main grains oriented roughly perpendicular to the applied shear direction and to the shear flow plane has been applied. The presence of the ordered grains, grain boundaries and defects helps to initiate the different alignment mechanism. All systems have been equilibrated for 2 × 104 steps, followed by the application of a shear flow in the x direction. The shear flow can align the lamellae system via three main mechanisms: rotations, nucleation and growth (NG) and selective disorder or partial melting (PM),33,34,56 denoted respectively by squares, circles and rhombi. Three different lamellar thicknesses have been simulated: Lz/H0 = 0.5, 1, and 2, to investigate the role of vertical confinement (see Fig. 5). Dotted lines represent curves of constant Weissenberg number Wi and can be seen to roughly correspond to the boundaries between different main mechanisms, which are the primary factors in the alignment of the lamellae. While the main mechanisms dominate, other mechanisms are always present but with less prominence.
This suggests that the ratio of relaxation time scale of the BCP and the external shear flow determines the dominant mechanism of alignment.
The differences across the three diagrams in Fig. 5 reveals the role of the thin film thickness: a lamellar phase in a thin film with small thickness at Lz/H0 = 0.5 is highly frustrated, which results in a quasi-2D behaviour for the alignment mechanism. A similar mechanism diagram has been found for BCP under electric fields.28–30,33,34,55,56,66 When the shear flow is weak only a rotation mechanism can be observed (black squares) in Fig. 5. For intermediate shear rates, nucleation and growth (red circles) is identified while for larger shear rates the system undergoes partial melting (blue diamonds).
When the thickness of the film is increased an extra mechanism (3D rotation) is found. The 3D rotation mechanism is indicated by grey squares in Fig. 5b and c. This new mechanism had not been observed previously, either in the 2D computational simulations or by experiments when an electric field was applied.28–30,34,55,56 This mechanism is only observed when the system is far above the order–disorder transition. Fig. 5 also shows that the region of the nucleation and grow mechanism is larger for thicker films (see Fig. 5c, red circles). In the rest of this section, we quantify each of the mechanisms and their differences.
To better understand the different mechanisms a time evolution of the mean absolute value of the order parameter ψ is shown in Fig. 6, with 〈|ψ|〉 ≪ ψeq. The vertical line in Fig. 6 shows the starting of the shear flow application. For rotation mechanisms, the order parameter is roughly constant, indicating little changes in the degree of separation during the mechanism. A considerable drop is shown in Fig. 6 for the blue curve (PM) indicating an increased disorder present in the partial melting while the other mechanisms display a better order during the time evolution of the system. The PM mechanism can also be identified using Minkowski functionals as shown in Fig. S1 in the ESI.†
Fig. 6 Time evolution of the mean absolute value of the order parameter 〈|ψ|〉/ψeq scaled with the equilibrium bulk value for four representative simulations corresponding to the phase diagram in Fig. 5(b) for Lz/H0 = 1 with: τ = 0.25 and = 0.0316 (partial melting – PM); τ = 0.4 and = 0.0316 (nucleation and growth); τ = 0.25 and = 0.001 (rotation in 2D); and τ = 0.4 and = 0.001 (rotation in 3D). The vertical dotted line indicates the start point of shearing at t = 2000. |
The alignment mechanism has been examined through the azimuthal angle of the scattering peaks. A brief representation of how the azimuthal angle was calculated is shown in Fig. 7. The analysis begins with real-time simulations (see Fig. 7a), followed by an investigation of the scattering peaks (an example is shown in Fig. 7b). The angle and intensity of these peaks for each azimuthal angle over time are then reported in detail in Fig. 7c.
The alignment pathway for a representative system undergoing the PM mechanism with τ = 0.19 and = 0.18 for a Lz/H0 = 1 is shown in Fig. 8. The phase diagram of the PM mechanism has been shown in Fig. 5(b). Before the application of the shear flow (t = 2000), four scattering intensity peaks are visible S(ϕ), corresponding to the two grains of lamellar orientation. In all the scattering intensity plots, each peak exhibits a reflection at an angle of π, a feature that arises from the nematic nature of the lamella alignment (see Fig. 7b). This π symmetry causes each grain orientation in the lamellar mesophase to appear twice in the scattering intensity plots (see Fig. 7c). Upon the application of the shear flow in the direction of the red arrow in Fig. 8 (t = 2051), these four initial peaks fade, followed by a short period without any clear peak. This corresponds to the partial melting of the system, as also seen by the drop in 〈|ψ|〉 shown in Fig. 6. Rapidly, two new peaks re-emerge, shown in Fig. 8 and corresponding to the shear-aligned orientation of the lamellae.
2D snapshots of the top-view on the x–y plane at z = Lz/2 in Fig. 8 indicate the formation of a homogeneous region (in gray), corresponding to melting. Only lamellae that were initially aligned with shear flow (grain boundaries and defects) remain not melted, indicated by the red circle in Fig. 8 (t = 2051). Fig. 9 shows a closer look at the order–disorder–order revealing the emerging of a transient morphology before alignment of lamellae perpendicular to the shear plane. In order to reveal the pattern of the BCP, the limits of the colorbar are decreased to 0.01 which allows to identify (and explicitly for the first time - to our knowledge) a chessboard-like configuration as an intermediate state between the initial state and the alignment along the shear direction. Fig. 9 shows also the time evolution of the chessboard pattern. At t = 2000 there are no chessboard patterns as the shear flow has not yet been applied. The formation of the chessboard starts to be visible after t = 2051, only after 51 time steps from the application of the shear flow. A zoomed-in view, on the right side of Fig. 9 has been provided to better visualize the pattern. Such pattern for the transition, from unstable to stable grains, was proposed by Schneider et al.39 using particle-based DPD simulations in Fig. 26.
Fig. 9 Chessboard-like transient configuration of the bottom part of the snapshots in Fig. 8. Snapshots show the bottom part of the full system, displaying the 2D slice at z = Lz/2. Note that the limits of the colormap are smaller (ψ = 0.01) than for the main text corresponding image 8. The right side of the figure shows a zoom of the red square on the left. |
An example of NG mechanism is shown in Fig. 10 under shear flow for = 0.013, τ = 0.325 and Lz/H0 = 1. The shear started after a pre-aligned configuration obtained at t = 2000 (same as in Fig. 8). This can be seen from the scattering intensity peaks shown in Fig. 10. After the shear was applied, two main peaks are formed: one at the position of the old peak due to the pre-aligned system and a new one due the shear flow direction. While the simulation continues, the old peak loses intensity while the new peak grows in intensity (see scattering intensity between t = 2000 and 2300 in Fig. 10). After t = 2400 the system consists of lamellae domains oriented parallel to the shear flow direction and some domains with random orientation. It is still possible to identify some rotation of grains due to the change between the two peaks even though the main mechanism is given by the growth of domains of lamellae oriented parallel to the shear direction. As the nucleation and growth proceeds, the peaks which initially coexist together change till there are only domains left, parallel to the shear flow direction. Snapshots of Fig. 10 (see t = 3374 and 4124) show the decreasing of domains for lamellae non-parallel to the shear flow direction and after t = 11000 only parallel domains to the shear flow directions remain. In this case, partial melting is not observed.
Fig. 10 Alignment of a lamellar mesophase via NG with τ = 0.325, and Lz/H0 = 1 for a shear rate = 0.013 (Wi = 6.2 × 10−2). The colourmap of the scattering intensity S(ϕ) is shown along with snapshots of the two-dimensional plane x–y at z = Lz/2 (with time shown above each picture). The colorbar on the right indicates the strength of the scattering peaks, in logarithmic scale, with red areas indicating alignment peaks. The initial configuration is omitted, being the same as in Fig. 8. |
An example of the 2D rotation mechanism under an applied shear flow for = 0.0024 for τ = 0.29 and Lz/H0 = 1 is shown in Fig. 11. The system is initialised from the same configuration as Fig. 8. This can be seen from the scattering intensity peaks shown in Fig. 11. Following the application of the shear flow, the two main peaks remain until t ∼ 104 after which, they shift to ϕ = 0.5π (1.5π). The shifting of the intensity peaks corresponds to the rotation of grains. Fig. 11 inset shows an event of grain rotation starting from t = 2108 originated at defects in the lamellar morphology. Afterwards, defects move and merge to align parallel along the shear flow.
Previous 2D studies28,34,55,56 have reported the same mechanism as the ones described so far: PM, NG and 2D rotation. However, the shear flow in a 3D system allows for non-2D mechanisms that involve vertical rotation. The new 3D rotation mechanism has been found for moderately thick films as shown in Fig. 5 (grey squares). Fig. 12 shows the scattering intensity for = 0.00183, τ = 0.36 for a Lz/H0 = 2.0, which is similar to the mechanism shown in Fig. 11 for 2D rotation. The snapshots suggest the appearance of extended parallel lamellar domains (A or B polymer blocks described by black or white areas in Fig. 12). The time evolution of S(ϕ) of Fig. 12 shows that the main peaks do not disappear as in Fig. 8 and 10 or as peak shift as in Fig. 11. Before the shift of the peaks, in fact it is possible to notice the merging of the two main peaks while these are fading (around t = 200000). However, the 2D scattering intensity is not sufficient to understand completely the out-of-plane alignment mechanism.
Fig. 13 shows the vertical out-of-plane nematic ordering Sz for = 0.001 and two representative values of the temperature: τ = 0.25 and τ = 0.4, corresponding to 2D and 3D rotation. For the higher temperature, τ = 0.25 the system displays no significant change in the nematic order parameter, Sz ∼ −0.5, which indicates a prevailing perpendicular alignment of the lamellar domains throughout the film. For the lower temperature τ = 0.4, a clear change in sign in Sz ∼ 0.6 indicates that the lamellar domains are temporarily orienting parallel to the surfaces (along the shear plane). A 3D top-view of the out-of-plane rotation mechanism (shear plane) is shown in Fig. 12 for each snapshot, which was found experimentally and can be compared with AFM images in Fig. 4 of ref. 37 for films with comparable thicknesses.
Fig. 13 Nematic alignment Sz of the BCP perpendicular to the thin film for two representative runs with Lz/H0= 2 and = 0.001 (see Fig. 5(c)): τ = 0.25 for the 2D rotation and τ = 0.4 for the 3D rotation mechanism. The early and long-time values of Sz ∼ −0.5 indicate perpendicular lamellar alignment (see miniatures) while Sz ∼ 0.6 indicate the partial transient parallel alignment of the lamellar domains. Vertical dotted line indicates the starting point of the shear flow, while the time is scaled with the characteristic shear time ts = 1/. |
However for thin films with small thickness Lz/H0 = 0.5, (Fig. 5(a)), the partial melting region (for low degree of segregation, small τ, bottom of the diagram) overlaps into the region of moderate Wi values. This can be due to surface effects, because of the proximity of the two neutral surfaces, which promote a lower segregation regime and therefore facilitates the order–disorder–order transition.
The 3D rotation shown in Fig. 5 is, to the best of our knowledge, a novel observation not previously documented in the literature. While lamellar alignment under electric fields has been extensively studied, it is important to note that alignment mechanisms under electric fields and shear flow are fundamentally different. Specifically, the electric field-induces alignment is reminiscent of extensional deformation under extensional flow condition. In contrast, the mechanisms observed under shear flow, such as the 3D rotation reported here, represent a distinct dynamic effect that cannot be directly related to alignment under electric fields. Experiments on films have shown lamellae aligned parallel to the surface under shear flow, as demonstrated in Fig. 4 of ref. 37. In our study, parallel lamellar alignment becomes prominent when the 3D rotation mechanism is present, particularly for thicker films (similar to experiments), as illustrated in Fig. 5(b) and (c). This alignment is more pronounced for thicker films with Lz/H0 = 2 and occurs at higher values of τ. The parallel alignment of lamellar domains along the shear flow (and, in this context, parallel to the shear flow plane) has been identified as metastable when compared to the perpendicular alignment.39 Analytical studies have shown that for bulk systems, perpendicular and parallel lamellar orientations exhibit distinct stability regions depending on temperature and shear rate.40 However, for thin films under shear flow, analytical results are limited to systems with selective walls.38 The considerable similarity between the mechanism diagram for BCP under shear and under electric fields suggests that the alignment mechanism of lamellar-forming BCP under external fields are universal for thin films with small thickness, and intrinsic to the material. Our findings suggest that the observed 3D rotation (mechanism that can be explained by the specific 3D nature of the shear flow) introduces a novel mechanism that drives lamellar alignment, distinct from those induced by electric fields or previously-known effects in shear flow.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4sm01241k |
This journal is © The Royal Society of Chemistry 2025 |