The smectic order of wrinkles (2024)

Energy of wrinkled patterns

Before introducing the physical model, we must start with choosing a convenient coordinate system to represent the configuration of a thin wrinkled sheet. Such representation must be able to account for long-wavelength deformations to capture the overall geometry and, as we shall see, short wavelength wrinkles that take up the excess area. For this purpose we choose surface coordinates in which the configuration, assumed to be lying near the plane z=0, takes the form

The smectic order of wrinkles (1)

where a and φ are scalar fields that in general depend on u and v. While this choice of Cartesian-like coordinates {u,v} may seem unintuitive, it can be rationalized by considering a uniform wrinkle pattern. Let the amplitude a(u,v) and wavenumber φ(u,v) be constant, that is, a(u,v)=a and φ(u,v)=qu. Then the arclength of a curve along u is just

The smectic order of wrinkles (2)

When the amplitude is small The smectic order of wrinkles (3), ds/du is approximately constant while still capturing the overall excess length taken up by the wrinkle. With the representation in equation (1), slow variations in a and φ do not alter the local wrinkle features. Calculations in terms of u and v are greatly simplified because the baseline, wrinkled metric does not have a short wavelength dependence on u, v. This is the major advantage of using these coordinates, and the reason for introducing the O(a2) correction to equation (1). In addition, it is clear from equation (1) that for small amplitudes the coordinates {u, v} are close to the Cartesian {x, y} coordinates of the projection of f onto the plane.

We will subsequently assume that the amplitude a(u,v) and phase φ(u,v) represent a slowly varying envelope of the wrinkled pattern, namely The smectic order of wrinkles (4) and The smectic order of wrinkles (5). We further assume small amplitude wrinkles, namely the excess length to be The smectic order of wrinkles (6). These assumptions directly relate to properties of the solution rather than to controlled parameters in the problem, but we will later show that they are consistent with results obtained from the model in a wide parameter regime and with solutions observed in experiments and simulations. In the following text we use the notation to indicate equality up to higher order terms in one or more of these small parameters.

With the above assumptions, the 2D metric and curvature tensors of configuration (1) take the form

The smectic order of wrinkles (7)

The smectic order of wrinkles (8)

with i, j{u,v}, and n the unit vector normal to the surface. Keeping our assumed scale separation in mind, we will be interested in looking at coarse-grained versions of different fields, namely ones which are averaged over oscillations at the wrinkle length scale |φ|−1, and have only features which persist across at least several wrinkles. For that purpose we define a coarse-graining operator The smectic order of wrinkles (9), where Ω is a region of radius |φ|−1 around each point and X can be any tensor field. Scale separation renders this definition unambiguous; the sheet’s reference metric (and of course the planar metric) can be approximated as Euclidean at the scale of Ω (since rad(Ω)<<D) and therefore integration, even of tensor fields, is well-defined. Since on the scale of Ω the envelope of a, φ can be considered constant, while φ itself changes significantly, this coarse-graining operation will be manifested in the vanishing of all high-frequency modes einφ(u,v), where n≥1, in X. For tensor fields this can be done element-wise in coordinates. Applying this to the metric (3) and curvature (4) tensors yields

The smectic order of wrinkles (10)

Let us now assume a scenario in which a thin elastic non-Euclidean shell is put atop a bulk of liquid (Fig. 2). The system energy H takes contributions from both the elastic energy of the thin sheet—stretching and bending17,18—and from the gravitational potential energy associated with displacing liquid elements (perfect wetting of the sheet is assumed). It can therefore be written in the form

(a) A spherical shell laid atop a flat body of water may develop wrinkles as means to overcome the geometric incompatibility. We denote by D the typical size of the shell and by R its radius of curvature. (b) We choose coordinates {u, v}, which at large scales follow the Cartesian coordinates {x, y} of the projection, however, at sub-wrinkle scales are proportional to arclength coordinates on the sheet. The amplitude a(u, v) and wave vector The smectic order of wrinkles (12) (written as in terms of a phase field φ(u, v)) may significantly vary across the sheet, however, the wavelength The smectic order of wrinkles (13) has a universal preferred value λ0. (c) Experiment: a hexagonal D=6 cm patch was cut out of a 30 μm-thick PDMS spherical shell, R=6 cm, before laid atop water. Shown is a top view schlieren image of the equilibrium configuration. (d) Simulation: a finite-element code with the physical parameters of the experiment (c). Shown is the vertical displacement from the plane. A single wrinkle (solid line) may bend, that is, change its in-plane direction, with wrinkle curvature encoded in the second derivatives of the phase field φ. In the shown example, the obtained pattern breaks into six domains, such as the lightened region on the right. Wrinkle bending focuses into domain walls (thicker solid segment of the wrinkle line). We denote by L the typical domain size. (e) The height squared profile of (d), at two different slices (horizontal and vertical), normalized by The smectic order of wrinkles (14). The envelope of these curves is by definition Δ(u, v). The black curve, The smectic order of wrinkles (15), plotted for R=6 cm (no fitted parameters), is anticipated for an axisymmetric pattern. (f) The full Δ(u, v) measured from (d). (g) Sketch of the phase field φ(u, v) for the pattern in (c,d). Level sets of φ indicate wrinkles.

The smectic order of wrinkles (16)

where

The smectic order of wrinkles (17)

The smectic order of wrinkles (18)

The smectic order of wrinkles (19)

For an isotropic material the stretching modulus is The smectic order of wrinkles (20) (with E Young’s modulus, t the thickness and ν Poisson’s ratio), bending modulus is The smectic order of wrinkles (21), and substrate modulus is K=ρg (fluid density and gravitational acceleration, respectively). The parameters The smectic order of wrinkles (22) and The smectic order of wrinkles (23) are the shell’s reference metric and curvature tensors, respectively, and The smectic order of wrinkles (24) where The smectic order of wrinkles (25) and tr denotes contraction with the contravariant reference metric tensor The smectic order of wrinkles (26). Note that the gravitational potential is not proportional to The smectic order of wrinkles (27), and that it is not fully determined by the surface’s fundamental forms, because of the non-covariance of the liquid (in the 2D geometry). The difference between its proper area form dx dy and our coordinate area form du dv is high-order in our small parameters and will be neglected in the following. Substituting equation (5) into equations (7)–(9) yields

The smectic order of wrinkles (28)

The smectic order of wrinkles (29)

The smectic order of wrinkles (30)

where (φ)ij=ijφ and (φφ)ij=iφ∂jφ (it is worth noting that The smectic order of wrinkles (31)). Under our assumed scale separation, the Hamiltonian (6) (or any large-scale integral) is unchanged by coarse-graining its integrand, thus we can rewrite equations (6) and (10)–(12):

The smectic order of wrinkles (32)

where

With our thinness and slowly varying envelope assumptions, the modulus Y that appears in the first term of equation (14) is much larger than the other moduli in equation (14). It follows that The smectic order of wrinkles (34), hence that the value of the local excess length Δ=a2|φ|2 is nearly constrained locally (see Supplementary Note 1 for a detailed derivation). Under this constraint, the bending and gravity terms in equation (14) are minimized at the well-studied finite wavenumber The smectic order of wrinkles (35), around which we can expand these terms.

The elastic/smectic energy functional

We now rearrange the energy density (14) and write it as a sum of two new components:

The smectic order of wrinkles (36)

The first component in equation (15), which we denote the elastic term, takes the form

The smectic order of wrinkles (37)

where we find an effective reference metric, The smectic order of wrinkles (38), that is slightly shifted from the actual reference metric of the shell towards a flat Euclidean metric according to the stretchability parameter η. The elastic energy component (16) penalizes deviations of the coarse-grained metric from this effective reference metric.

To illustrate the nature of this energy component, it is helpful to consider a wrinkled configuration of constant amplitude a. The coarse-grained metric (5) of such a configuration is identical to that of a surface whose topographic map is given by the wrinkled pattern, where wrinkles represent level-sets spaced The smectic order of wrinkles (39) in the z direction. Therefore, equation (16) penalizes deviations of that smooth surface’s geometry from the target geometry The smectic order of wrinkles (40). However, this geometry is the coarse-grained version of many different wrinkle patterns, for example, taking a topographic map of the same surface from different projection angles or with a different choice of level-set spacing (wrinkle amplitude).

This degeneracy is broken by the second component in equation (15), the smectic term, which takes the form

The smectic order of wrinkles (41)

The first term, smectic compression, penalizes deviations of wrinkle spacing from the preferred wavenumber k0. Its modulus comes from the bending-gravity interplay, but is also proportional to the local excess length Δ which may therefore be thought of as a (perhaps spatially varying) local smectic order parameter. The second term, smectic curvature, penalizes wrinkle bending (Fig. 2d). Interestingly it has nothing to do with elastic bending, and actually comes only from residual tangent-to-sheet elastic stretching. A geometric intuition can be given to this term—unlike straight wrinkles, bent wrinkles have nontrivial Gaussian curvature which oscillates at half the wrinkle wavelength. If the target geometry is smooth at the wrinkle scale, that implies strain which results in elastic stretching, which is the source for the second term in (equation 17).

A very similar analysis can be performed when the substrate is soft elastic solid rather than liquid; the wrinkle wavenumber k0 is then to be replaced with The smectic order of wrinkles (42), where Es is the substrate’s Young’s modulus and the substrate’s Poisson’s ratio is assumed 0.5 (ref. 8). The relevant dimensionless stretchability parameter η is The smectic order of wrinkles (43), which does not depend on the thickness (as is expected from dimensional analysis). The following results hold in either physical scenario.

We next explore some of the outcomes of this analogy between wrinkles and smectics. to test our predictions, we employ two-dimensional (2D) finite-element simulations (Abaqus) of an elastic spherical shell with bulk gravitational force. The elastic and gravitational moduli, as well as the sheet geometry, can be tweaked to give us full control over the relevant dimensionless parameters (η, γ). Fluid surface tension exists in the experiment and is neglected in the simulation, however, in the relevant parameter regime its effect is negligible (Supplementary Fig. 2). Figure 2c,d demonstrates the good agreement in equilibrium configurations between the simulation and an experiment of a PDMS shell on water, performed with the same physical parameters. An extensive experimental and numerical research of such systems is in progress.

Domain walls

From the two smectic moduli in equation (17), one can extract a smectic penetration length

The smectic order of wrinkles (44)

where λ0 is the wrinkle wavelength. The penetration length λpen is a typical length over which smectic layers—or wrinkles—may bend smoothly. Hence, the wall between two smectic domains, with an angle difference The smectic order of wrinkles (45) between their wrinkle directions, will be of typical width 2λpen/tan ω (ref. 19). From equation (18) one learns that domain wall width is proportional to The smectic order of wrinkles (46), hence proportional to the amplitude. In systems with a non-Euclidean underlying geometry, the amplitude may vary across the system (since The smectic order of wrinkles (47)). Consequently domain walls may vary in width, possibly even along one continuous wall, as can be well seen in Fig. 3a. Remarkably, equation (18), based entirely on the theory of smectic liquid crystals, provides a precise prediction for the relation between the width of domain walls and local amplitude in wrinkled patterns. This prediction is in excellent agreement with measurements taken on our wrinkled elastic sheet simulations (Fig. 3c).

(a,b) Equilibrium wrinkle patterns in a square section of a spherical shell on liquid substrate (simulation). Vertical displacement z (top-left) and mean curvature H (bottom-right) clearly indicate splitting of the patterns into four domains divided by domain walls of varying widths. (a) η≈2.4 × 10−4, all domain walls are of the ‘Chevron’ type, that is, wrinkle angle changes monotonically when moving along the wrinkle across a domain wall. (b) η≈5.7 × 10−5, domain walls in the high-amplitude wrinkles become Ω-shaped through the formation of localized d-cones (cusp is well seen in the mean curvature map). (c) The width of Chevron-shaped domain walls is proportional to the penetration length (equation (18)), hence to the amplitude. Ω-shaped walls have widths proportional to the wrinkle wavelength, regardless of the amplitude. Widths are extracted from simulations by fitting wrinkle angle to an arctangent, in compliance with19. Penetration lengths are calculated from system parameters and the measured amplitude of each wrinkle.

Full size image

As Fig. 3 suggests, the above prediction appears to be valid only above certain values of the parameter η. Nonetheless, its breakdown is also explained by smectic theory; indeed, many systems that exhibit layered patterns, ranging from diblock copolymer melts15,20,21 to Rayleigh–Bénard convection22,23, show a transition from ‘Chevron’-shaped domain walls at small wall angles to Ω-shaped walls at large angles. This bifurcation is achieved via the formation of a pair of disclinations at each layer, to better maintain layer spacing in the system. A similar transition was observed, though to the best of our knowledge not explicitly discussed, in wrinkled herringbone patterns24,25, seemingly via the formation of d-cones (ref. 26). The critical domain boundary angle for this transition is made wider when the thickness of the sheet (or equivalently η) increases, thus not allowing focusing of the curvature into small d-cones cores. Since the width of the Ω-shaped domain wall scales as the wrinkle wavelength λ0, and not as the penetration length, one expects domain walls to abruptly change their scaling with decreasing η, as seen in Fig. 3. For wider domain wall angles, the Chevron-type walls and consequently the width λpen relation are expected to persist onto narrower sheets.

Undulation instability

In smectic systems, certain boundary conditions may induce an undulation instability in the smectic structure27,28. At its onset, the wavelength of this undulation is

The smectic order of wrinkles (49)

where the d is a system scale at which boundary conditions are introduced. To test if similar behaviour exists in wrinkled systems, we designed a numerical experiment (Fig. 4) that follows the footsteps of ref. 27 in their smectic-A experiment. In our simulations, we hold a thin strip at its boundaries at a wavelength λB which is slightly mismatched with the preferred wrinkle spacing λ0=0.95λB. One resolution for such mismatch may be an undulation, which allows layer spacing in the bulk to be smaller than on the boundary, while paying some cost for layer bending. The predicted optimum for this exchange is given by equation (19). We examine a sequence of systems with fixed geometry and wrinkle wavelength, however, with penetration/undulation lengths varying between simulations (achieved via changing both fluid density and sheet thickness so that λ0 remains constant, however, η changes).

(a) Equilibrium wrinkle patterns in a flat thin ribbon residing on a flat body of water (simulation), held at its far ends with a fixed wavy boundary. The preferred wrinkle wavelength λ0 is 5% shorter than the boundary wavelength λB. Shown are the results for various values of the predicted undulation wavelength The smectic order of wrinkles (51) (equation (19)), where the ribbon’s width is used as a surrogate for the system scale d. (b) The measured undulation wavelength is plotted against the predicted one. While corroborating the predicted scaling, the measured wavelength is consistently larger than The smectic order of wrinkles (52) by a factor of ≈1.7, suggesting that the relevant system scale d may be different than the ribbon width by a constant factor. (c) Simulation results for a spherical cap with free boundary conditions. An undulation appears spontaneously, and its observed wavelength (black triangles) agrees with the predicted value, where the shell’s diameter appears to be a good surrogate for the system scale d.

Full size image

Indeed, an undulation in the wrinkling pattern appears (Fig. 4a). The observed undulation wavelength scales as the square root of the penetration length (Fig. 4b), as accurately predicted by equation (19). When using the ribbon width as a surrogate for the system scale d, we get a factor of ≈1.7 between the observed wavelength and the predicted one, asserting that ribbon width may not be the relevant system scale.

We further observe that wrinkle undulations spontaneously appear in certain regions of larger systems with free boundary conditions as a result of the global geometric incompatibility. We show examples for this observation in Fig. 4c and in Supplementary Fig. 1. We conjecture that this instability sets the scale for the herringbone patterns seen in biaxially compressed sheets on elastic substrates25,29,30, as shown in Fig. 1d. The scaling implied by this conjecture agrees with the stability analysis scaling presented in ref. 31.

Domain size scaling

Last, our smectic analogy approach can further give a coarse prediction for the typical size and number of domains in systems that wrinkle as a result of geometric incompatibility. The coarse-grained metric g(cg) of a domain-type patterns is approximately that of a piecewise-linear configuration that approximates the curved target metric The smectic order of wrinkles (53). The strain that results from the difference between the two induces the elastic energy (16), therefore preferring smaller domains (that is, a better piecewise-linear approximation). However, the smectic energy term (17) is concentrated into domain walls, thus penalizing many-domain solutions. The interplay between these two contributions sets the typical domain size.

We consider a spherical shell with radius of curvature R. For domains of typical size L, the maximal (average) elastic strain scales as L2R−2 (R−2 is the Gaussian curvature mismatch). Integrated over the entire sheet of typical size D, this will result in total elastic energy The smectic order of wrinkles (54). The energy per-unit-length of a domain wall is proportional to the smectic curvature modulus divided by the smectic penetration length19, that is, scales as Δ3/2η1/2λ0Δ3/2t. Upon introducing the typical excess length implied by the global geometric incompatibility, The smectic order of wrinkles (55), and the total length of the walls D2/L, we get the total smectic contribution The smectic order of wrinkles (56). Minimization of the total energy (15) suggests the scaling

The smectic order of wrinkles (57)

where γ is the Föppl–von Kármán number. This general scaling relation should be taken with caution, since the number of domains is discrete and is very much affected by boundary conditions. A quantitative experimental or numerical verification of equation (20) must therefore involve very large systems (so that boundary effects are irrelevant), with typical domain size extracted via statistical analyses of the pattern. Another experimental complication involves the very small exponent by which the domain size in equation (20) depends on physical parameters. Preliminary numerical observations may support relation (20) (Fig. 5), however, a robust numerical verification using a finite-element thin-sheet simulation is computationally inconceivable as a result of the extreme multi-scale nature of the problem. Nevertheless, we hope that a more efficient simulation based on our phase field φ approach may in the future allow verification of this scaling.

Typical equilibrium wrinkle patterns in a hexagonal section of a thin spherical shell on a liquid substrate. All simulation parameters are kept fixed, except for the sheet thickness. The Föppl–von Kármán number γ of the system is indicated above each pattern. While the top-left pattern is clearly divided into six domains, increasing γ seems to decrease the typical domain size. The scale bar to the right of each pattern shows the scaling of typical domain sizes predicted by equation (20), up to a constant factor (only the ratios between bars are meaningful since equation (20) suggests only the scaling of the typical size and leaves the prefactor unknown).

Full size image
The smectic order of wrinkles (2024)
Top Articles
Latest Posts
Article information

Author: Cheryll Lueilwitz

Last Updated:

Views: 6487

Rating: 4.3 / 5 (54 voted)

Reviews: 85% of readers found this page helpful

Author information

Name: Cheryll Lueilwitz

Birthday: 1997-12-23

Address: 4653 O'Kon Hill, Lake Juanstad, AR 65469

Phone: +494124489301

Job: Marketing Representative

Hobby: Reading, Ice skating, Foraging, BASE jumping, Hiking, Skateboarding, Kayaking

Introduction: My name is Cheryll Lueilwitz, I am a sparkling, clean, super, lucky, joyous, outstanding, lucky person who loves writing and wants to share my knowledge and understanding with you.