Localized Structures in the Multi-dimensional Swift–Hohenberg Equation by Scott Gregory McCalla B.A., Cornell University; Ithaca, NY, Sc.M., Brown University; Providence, RI, A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in The Division of Applied Mathematics at Brown University PROVIDENCE, RHODE ISLAND May 2011 c Copyright 2011 by Scott Gregory McCalla This dissertation by Scott Gregory McCalla is accepted in its present form by The Division of Applied Mathematics as satisfying the dissertation requirement for the degree of Doctor of Philosophy. Date Björn Sandstede, Ph.D., Advisor Recommended to the Graduate Council Date John Mallet–Paret, Ph.D., Reader Date Arnd Scheel, Ph.D., Reader Approved by the Graduate Council Date Peter Weber, Ph.D., Dean of the Graduate School iii Vitae Professional Preparation Cornell University Mathematics and Physics B.A. cum laude, 2005 Brown University Applied Mathematics Sc.M., 2007 Brown University Applied Mathematics Ph.D., expected 2011 Appointments Research Assistant, Brown University Summer 2009 - Present Teaching Assistant, Brown University Spring 2009 Coline M. Makepeace Fellow, Brown University Fall 2007 - Spring 2008 Teaching Assistant, Brown University Spring 2007 Research Assistant, Los Alamos National Laboratory Spring 2000- Summer 2008 Publications McCalla, S.G., B. Sandstede. Snaking of radial solutions of the multi-dimensional SwiftHohenberg equation: a numerical study. Physica D, 239, 1581-1592 (2010). Lestone, J.P., S.G. McCalla. Statistical Model of Heavy-Ion Fusion-Fission Reac- tions. Phys. Rev. C, 79, 044611 (2009). McCalla, S.G., J.P. Lestone. Fission Decay Widths for Heavy-Ion Fusion-Fission Reactions. Phys. Rev. Lett., 101, 032702 (2008). iv Minisymposium Organization Organizing minisymposium on“Multi-dimensional Localized Patterns” for the SIAM Conference on Applications of Dynamical Systems (DS11) with David Lloyd of the University of Surrey. Presentations Localized structures in the multi-dimensional Swift-Hohenberg equation (invited minisymposium talk) SIAM Conference on Nonlinear Waves and Coherent Structures (NW10), Philadelphia, 2010 Radial solutions and the cessation of snaking for the multi-dimensional Swift–Hohenberg equation (contributed talk) 8th AIMS Conference on Dynamical Systems, Differential Equations and Applications, Dresden University of Technology, 2010. Radial solutions and the cessation of snaking for the multi-dimensional Swift-Hohenberg equation (invited talk) BU/Brown PDE Seminar, Boston University, 2010. Radial solutions and the cessation of snaking for the multi-dimensional Swift–Hohenberg equation (contributed talk) Applied Math Days, Rensselaer Polytechnic Institute, 2010. Statistical model calculations of heavy-ion induced fusion-fission reactions (poster) Student Symposium, Los Alamos National Laboratory, 2006. Statistical model calculations of heavy-ion induced fusion-fission reactions (invited talk) T-16 (Nuclear Physics Group) Seminar, Los Alamos National Laboratory, 2006. Statistical model calculations of heavy-ion induced fusion-fission reactions (talk) American Physical Society April Meeting, Dallas, 2006. v Awards SIAM Student Travel Award, Conference on Nonlinear Waves and Coherent Structures (NW10), 2010. Defense Programs Award of Excellence (NNSA), Los Alamos National Laboratory, 2006. Los Alamos Awards Program (LAAP) Award, Cash Award, Los Alamos National Laboratory, 2006. Dean’s List, Cornell University. Teaching Experience and Training Sheridan Teaching Certificate I, Brown University, 2010. The Sheridan Center is devoted to improving the quality of instruction at Brown; this was a year long teacher training course. AM65: Essential Statistics, Teaching Assistant, Brown University, Spring 2009 AM36: Methods of Applied Mathematics II, Teaching Assistant, Brown University, Spring 2007 Conferences Attended Joint SIAM/RSME-SCM-SEMA Meeting on Emerging Topics in Dynamical Systems and Partial Differential Equations, Universitat Politecnica de Catalunya, 2010. vi Acknowledgements Firstly, I want to thank my adviser Björn Sandstede. His guidance and support have been crucial to my research. He additionally allowed me the freedom to explore on my own and study a variety of topics. He treated me with respect and understanding. In addition to Professor Sandstede, I want to thank Professor Mallet–Paret and Professor Scheel for generously donating their time to serve on my committee. My research advisers at Los Alamos National Laboratory, Karen Hill and John Lestone, introduced me to scientific research. They were fantastic mentors and are still good friends. At Brown, I have made many friends among the graduate students, postdocs, staff and faculty. They have helped me to learn many things from math to barbecuing. In particular, I would like to thank Peter van Heijster for discussions on both mathematics and soccer. Hopefully the Dutch will win soon. Thanks also to Sunil Chhita for many shared meals and coffees. My family always provided a great environment to grow up in. My brothers’ interests are far removed from math and their willingness to share these interests with me has kept me well-rounded. My mom has always been supportive in every way imaginable. Finally, I would like to thank my wife Stephanie for her unflagging support and encouragement throughout graduate school. She is a wonderful partner and will be a fantastic mother. I look forward to the future. Thank you all for your help. vii Abstract of “ Localized Structures in the Multi-dimensional Swift–Hohenberg Equation ” by Scott Gregory McCalla, Ph.D., Brown University, May 2011 This goal of this thesis is to understand patterns in the Swift–Hohenberg equation. The patterns studied are localized, stationary and radially symmetric in dimensions one through three. The emphasis is placed on the existence of these structures through numerical evidence and analytic proofs. The bifurcation structure of localized stationary radial patterns of the Swift–Hohenberg equation is explored when a continuous parameter n is varied that corresponds to the underlying space dimension whenever n is an integer. In particular, this numerical investigation reveals how 1D pulses and 2-pulses are connected to planar spots and rings when n is increased from 1 to 2. It also elucidates changes in the snaking diagrams of spots when the dimension is switched from 2 to 3. A previously unknown spot solution is additionally uncovered. The second half of the thesis is devoted to rigorously proving this spot’s existence. The amplitude of the spot exhibits an unexpected scaling as the bifurcation parameter is reduced to zero. The spot is constructed by gluing two known solutions together, each scaling as the square root of the bifurcation parameter, but it has a much larger scaling. This behaviour is explained as a result of the proof. Contents Vitae iv Acknowledgments vii 1 Introduction 1 2 Preliminaries 2.1 Variation of constants . . . . . . . 2.2 Resonant terms and normal forms 2.3 Stable manifold theorem . . . . . . 2.4 Roughness theorem for exponential 2.5 Turing bifurcations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . dichotomies . . . . . . . 3 Numerical Exploration 3.1 Numerical algorithms . . . . . . . . . . . . . 3.2 Localized 2D states . . . . . . . . . . . . . . . 3.3 The connection between 2D and 3D branches 3.4 The connection between 1D and 2D branches 3.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 10 11 12 13 16 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 25 28 33 35 39 4 Existence Near Onset for Spot B in 2D 4.1 Geometry of the rings and spots . . . . . . . . . . . . 4.2 Life at the core . . . . . . . . . . . . . . . . . . . . . . 4.3 Normal forms . . . . . . . . . . . . . . . . . . . . . . . 4.4 The rescaling chart . . . . . . . . . . . . . . . . . . . . 4.5 The transition chart . . . . . . . . . . . . . . . . . . . 4.6 The fixed points . . . . . . . . . . . . . . . . . . . . . 4.7 The formal argument in two dimensions . . . . . . . . 4.8 The flow around the equilibria in the transition chart . 4.8.1 Transversality and the ring: . . . . . . . . . . . 4.8.2 P+ : . . . . . . . . . . . . . . . . . . . . . . . . 4.8.3 Connecting P+ to P− . . . . . . . . . . . . . . 4.8.4 P− : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 45 47 49 52 54 57 58 62 63 66 77 80 viii . . . . . . . . . . . . . . . . . . . . 4.9 Proof of the main theorem: matching the core . . . . . . . . . . . . . . . . . 4.10 The breakdown of monotonicity . . . . . . . . . . . . . . . . . . . . . . . . . 4.11 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 93 95 5 Conclusion 98 5.1 Summary of main results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 5.2 Open questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 A Proofs for Asymptotics A.1 The general approach . . . . . . . . . . . . A.2 Decay estimates for the ring in the far field A.3 Corollaries for different f (s) . . . . . . . . . A.3.1 Corollary 1: η(ε01 ) . . . . . . . . . . A.3.2 Corollary 2: c1 (δ− , δ+ ) . . . . . . . . ix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 103 104 106 106 106 List of Figures 1.1 1.2 1.3 1.4 Desert grass spots and rings from [35]. Gas discharges in [29]. Ferrosoliton from [17]. Hexagons from [31]. . . . . . . . . . . Patterns from [22]. . . . . . . . . . . . . . . . . . . . . . . . . The profiles for rings and spots. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1 The center panel contains the bifurcation diagram of 1D localized pulses. The symmetric profiles that correspond to parameters on the light-colored curve have a maximum at r = 0 as shown in panels (1), (2), and (5), while the symmetric profiles corresponding to the darkcolored branch have a minimum at r = 0 as illustrated in panel (3). As we move up on each branch, a pair of new rolls is added to the solution profile at every other fold bifurcation. The two different branches discussed above are connected by ladder branches that correspond to asymmetric profiles as indicated in panels (3)-(5). These asymmetric structures bifurcate at pitchfork bifurcations near each fold from the symmetric pulses. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Shown are profiles, representative color plots, and bifurcation branches of localized planar spot A solutions in the top row and of the two localized planar ring solutions in the bottom row. Profiles and color plots correspond to solutions at (µ, ν) = (0.005, 1.6). [Reproduced from [22]]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 The profiles of spots A and B with (µ, ν) = (0.005, 1.6) are compared in the left panel, while an enlarged plot of spot A is shown separately in the right panel. Note that spot B resembles an inverted spot A but with a much larger amplitude. The zeros of both profiles appear to align well for r 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4 Shown are the connected bifurcation curve of spot A and one of the ring solutions in panel (i) and the bifurcation branch of spot B and the second ring solution in panel (ii). In the upper right corner of panel (i), the branch oscillates between three folds aligned approximately at µ ≈ 0.18, 0.19, and 0.21, and we refer to the part of the branch that oscillates between the two rightmost folds as the secondary snaking structure. Note that the vertical L2r -axes in panels (i)-(ii) are scaled differently: in particular, the spot A branch reaches a larger value of the L2r -norm. The solution profiles at the points labelled (a)(d) are shown in Figure 3.5. . . . . . . . . . . . . . . . . . . . . . . . x 3 4 5 6 22 23 24 29 3.5 Panels (a)-(d) contain the solution profiles of spots and rings at the parameter values labelled (a)-(d) on the branches shown in Figure 3.4. As the spot and ring branches are traversed towards increasing L2r -norm, additional rolls are added at the right tail of the localized profiles. The maximal (minimal) amplitude of spot A (spot B) always occurs at r = 0 along the branch. For rings, u(0; µ) oscillates between positive and negative values as we move from one leftmost fold to the next on the branch; new rolls are created only at the tail but not near r = 0. We refer to the movies at http://www.dam.brown.edu/people/mccalla/SpotAmovie.mpg and http://www.dam.brown.edu/people/mccalla/SpotBmovie.mpg for further details on the behavior of spots and rings. . . . . . . . . . 3.6 Panel (i) shows in blue the connected snaking branch of the spot B and ring B solutions from Figure 3.4(ii) together with a stack of isolas, plotted in red and alternately in dashed and solid, along which profiles resemble those of spot B and ring B. Panel (ii) contains the spot A curve (in dark cyan) and the spot B branch (in blue) from Figure 3.4 together with the stacked isolas (in red) from panel (i). Note that the isolas align well with the secondary snaking structure visible near the top of the spot A branch, indicating that that they pinch off from the spot A branch as n is changed. . . . . . . . . . . . . . . . . . . . . . . 3.7 Shown is the first isola (in green) of a second family of stacked isolas that appears above the spot A branch (plotted in dark cyan). . . . . . 3.8 The lower parts of both panels contain the connected snaking branch of spot A and ring A (in dark cyan) from Figure 3.4. Above this branch, we found a family of stacked isolas (plotted in green) that include the isola shown in Figure 3.7. The stack of isolas extends only up to a value of the L2r -norm at which the profiles consist of approximately 38 rolls. Above this value, we found a single connected solution curve (drawn in brown) that consists of two intertwined branches that both snake, seemingly indefinitely. For clarity, we show only one of the two intertwined branches in the upper part of panel (ii). Solution profiles along the upper snaking curve can be found in the accompanying http://www.dam.brown.edu/people/mccalla/SpotABmovie.mpg. . . . 3.9 The two panels show log-log plots of the two leftmost and two rightmost folds of the high snaking branch shown in Figure 3.8, indicating that the snaking branch converges algebraically to the Maxwell point µ = 0.2 of 1D rolls. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.10 Panel (i), reproduced from [22], indicates that the amplitude of spot A 1 scales as µ 2 as µ approaches zero. As shown in panel (ii), the amplitude of spot B appears to scale approximately like µ0.374 . . . . . . . . . . . 3.11 To delineate the existence region of spot B, we continued spot B in the parameter ν for several fixed values of µ and visualize the resulting solution branches in two different ways: in the left panel, we plot ν versus the squared L2r -norm (the values of µ decrease from right to left), while the right panel shows log µ versusp ν. Note that the solution branches stay above the critical value ν = 27/38 and that the L2r norm of the associated profiles goes to infinity as ν approaches the lower end of each branch. . . . . . . . . . . . . . . . . . . . . . . . . . xi 29 30 31 31 33 33 34 3.12 The bifurcation curves of spot A and spot B solutions are presented in panels (i) and (ii), respectively, for different values of the dimension parameter n. The insets show the branches for n = 3 in more detail. . 3.13 Panel (i) contains the upper snaking branches of spots for n = 2 (in brown), n = 2.3 (in cyan), and n = 3 (in black). Panel (ii) contains the two arms of the snaking branch for n = 3 to illustrate that they do not overlap. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.14 The two curves plotted in cyan diamonds correspond to the limits at n = 1 of the lower planar spot A and spot B branches when continued in n. The profiles along these branches for n = 1 coincide with the 1D pulses shown in Figure 3.1. The solid figure-eight isolas plotted in red arise when we continue the two ring branches from n = 2 down to n = 1 using the methods outlined in §3.1. The profiles along each isola are symmetric 1D 2-pulses. . . . . . . . . . . . . . . . . . . . . . 3.15 The left panel contains the ring A branch for different values of n plotted in the planar L2r -norm. The curve for n = 1.2 is connected but clearly shows structures that will pinch off to become individual isolas for smaller values of n. These isolas continue to form and pinch off as the dimension is decreased further, thus leading to isolas of 2-pulses with a given L2r -norm and an arbitrary separation between the pulses. The right panel shows the ring A branch for n = 1.2 and n = 1.3 but now plotted in the one-dimensional L2x -norm. Note that the curve for n = 1.2 appears to cover an entire family of what will later become separate 2-pulse isolas at n = 1. . . . . . . . . . . . . . . . . . . . . . 3.16 The left panel contains four isolas at n = 1 that are found from the two planar ring branches through continuation in n. The right panel contains the solution profiles at the topmost intersection of these isolas with the line µ = 0.195: the profiles in panels (a)-(b) and B come from ring A, while the profiles in panels (c)-(d) arise from ring B. Since these profiles were computed with Neumann boundary conditions at r = 0, they can be reflected across r = 0 and therefore correspond to 2-pulses. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.17 The profiles shown here at n = 1 were found through continuation from rings. Due to the Neumann conditions imposed at r = 0, these solutions correspond to symmetric 2-pulses with different separation distances represented by the number s of small oscillations near r = 0. 3.18 When we continue an asymmetric 1D pulse that is centered some distance away from x = 0 in n, we obtain the isolas in panel (i) which shrink and eventually disappear. Panel (ii) contains continuation results in (µ, n) of the two upper and lower folds along the isolas. As n increases, the lower folds disappear in a cusp, thus making the isola more circular, while the collision of the remaining upper folds corresponds to the point at which the isola disappears. . . . . . . . . . . . 4.1 Spot A versus spot B. . . . . . . . . . . . . . . . . . . . . . . . . . . . xii 35 35 36 36 37 38 38 45 4.2 The pictured schematic represents the core manifold, the transition region, and the rescaling chart. The core manifold, at top left, is a two-dimensional manifold that captures the smooth bounded solutions within the interval [0, r0 ] for a fixed but finite r0 > 0. The transition chart, at bottom left, captures the algebraic growth and decay of solutions. The rescaling chart, at right, captures the exponential decay of solutions as the radius goes to ∞. The blue solid curves represent spot A and the ring on the left and right respectively. The dashed red line represents spot B. . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 The solid blue curve for the ring lies in the transverse intersection of s cu W∞ and W− . We can find starting data near (A+ , z+ ) = (−η, 0) in s W∞ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4 A cartoon of the orbit. We are looking for a ring solution that lies in the invariant plane where α1 = 0 and ε2 = 0. The remainder terms in both the transition and rescaling chart then drop out and we are left with (8.1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . s 4.5 The equilibrium at (z1 , A1 ) = ( 12 , 0) is pictured. A segment of W∞ is drawn in red. By choosing the appropriate point on this manifold, the matching equations are solved. We need to understand how this segment is affected by the flow between and around the two equilibria. s from the posi4.6 In this cartoon, we display the expected motion of W∞ tive equilibrium backwards in time towards the negative equilibrium. . s 4.7 Here the full motion of W∞ between P− (on the left) and P+ (on the right) is shown with the expected linear growth rates around and between the equilibria. Sections 4.8.2, 4.8.3 and 4.8.4 are equivalent to stepping across this picture from right to left. . . . . . . . . . . . . 4.8 The radial rescaling is done with γ = 1/2 corresponding to r̃ = εr. The absolute values of several solutions are plotted for various values of µ ∈ [0.04, 0.14]. Note the change in amplitude of the extrema as µ is varied: the amplitude equation can be visualized through these values. For the largest value of µ, the maximum amplitude is not at the origin. This is an example of the expected breakdown of monotonicity, though we are far from the valid range of our calculations. As µ is decreased, the second largest extremum shrinks while third largest extremum grows. When µ is made small enough, we expect a shift in which is largest. This can be seen in figure 4.9. . . . . . . . . . . . . . 4.9 This plot is similar to figure 4.8, only for a value of µ = 0.0059. As is highlighted by the dashed line, there is now a non-monotonicity in the amplitude with respect to r. . . . . . . . . . . . . . . . . . . . . . . . 4.10 The amplitude of spot B in three dimensions is shown as a function of µ. The data is insufficient to fully acquire the scaling: both a linear scaling and a linear scaling multiplied by a logarithmic term adequately fit the data. From the formal analysis, we expect the scaling to look 1 like µ 4 (log µ)α for some unknown constant α. . . . . . . . . . . . . . . xiii 46 59 64 67 78 80 94 94 96 Chapter One Introduction 2 Patterns arise constantly in nature. They are structured phenomena that come from some homogeneous or completely random background. The initial background does not dictate the final ordered state in any expected fashion. Recently, a great deal of effort has gone into understanding the formation of these patterns, culling ideas from mathematics, physics, biology, chemistry and material science. Mode-locked laser systems produce high energy pulses of light. Spots bundle magnetic field lines in ferromagnetic fluids. Grass spots and rings develop in deserts where resources are scarce. Hexagon patches form in cooling lava fields, such as the Giant’s Causeway in Northern Ireland. Oscillons, localized pulses that fluctuate periodically in time, appear in vertically vibrated trays of sand and clay. I will concentrate on studying these patterns as solutions to partial differential equations (PDEs), generally reaction-diffusion systems. Pattern can refer to many things from a structured solution that approaches a periodic curve at infinity, to an exponentially localized solution. In reaction-diffusion systems, these nonhomogenous states are often formed through a Turing, or pattern forming, bifurcation. These are generally instabilities that arise because of the diffusive terms. Normally diffusion is thought to stabilize a system, but this is not always the case. The general heuristic behind this is often described in terms of predator-prey relationships. The predator and prey densities are expected to diffuse from some initial distribution, but they diffuse at different rates (assume the prey moves more slowly). Also assume there is some equilibrium state between the predators and prey where their densities do not change in time. If initially there is some point where there is an abundance of prey then the predator population will increase at that region. However, the predators will diffuse quickly away, and the prey more slowly. This could, if the diffusion of the predator was fast enough, leave a region where the prey population remained above the equilibrium state. In the surrounding area, the predator population may have been high enough to further drive the prey below the equilibrium state as they diffused away. This would lead to a pattern from a system expected to be stable without diffusion. These patterns can be stationary, traveling, or oscillatory. The simplest example would 3 193 ecological complexity 4 (2007) 192–200 Vegetation patches [Sheffer et al.] Figure 1.1: Desert grass spots and rings from [35]. Gas discharges in [29]. 2 be a stationary Gaussian profile for a one dimensional system (i.e. u(x, t) = ae−bx ). This 2 2 could also travel, u(x, t) = ae−b(x−ct) , or oscillate, u(x, t) = a sin(ct)e−bx . The profiles could also be periodic. The next class of patterns consists of a periodic function modulated Fig. 1 – Ring patterns in nature. (a) Mixture of rings and spots of Poa bulbosa observed in the Northern Negev (250 mm yrS1). 2 (L.) Baker maritima (b) A ring of Asphodelus ramosus L. observed in the Negev desert (170 mm yrS1). (c) A ring of Urginea −c(x−dt) ramets observed in Wadi Rum, Jordan (50 mm yrS1). Photographs by E. Meron (a) and H. Yizhaq (b and c). by a localized profile, for example u(x, t) = a sin bxe . For two dimensional patterns, things can get more interesting. First, there are spots as seen in figures 1.1, 1.2, and 1.3: introduced by Gilad et al. (2004, 2007). The model extends earlier models (Rietkerk et al., 2004) in capturing the non-local nature of water uptake by plants’ roots, and the augmentation of the root system in response to biomass growth. The model has been used to study mechanisms of vegetation pattern formation and ecosystem engineering along environmental gradients, addressing in particular the question of resilience to disturbances (Yizhaq et al., 2005; Gilad et al., 2007). An extension of the model to plant communities, containing several vegetation functional groups, has recently been used to study transitions between competition and facilitation in woody-herbaceous systems along stress gradients (Gilad et al., in press). competition than those at the circumference of the patch. As these are radially symmetric patterns, often periodic modulated by a decaying a consequence a ‘‘latent ring’’, wherefunctions the biomass density at the patch core is smaller than the density at the periphery, a visible ring involving central die-back, is expected to envelope, with a large positive or or negative bump at their center that decays into the form. Methods far field. Then there are rings2.1.(figure 1.1) which look like target patterns: assuming At the beginning of the growing season (early winter) 2 stationarity and radial symmetryindividual they P.would look(10–15 likecmu(r) = acompletely cos bre−c(r−bπ) . Spots can bulbosa genets diameter), covered with green leaves, were transplanted into 4 L pots cm diameter) and to a greenhouse. Genets were collected be concentrated on a hexagonal (18 lattice to produce hexagon patterns as is seen in all three from a dry Mediterranean field site (Adulam, Israel 31816 N 2. 0 348250 E) with an annual average rainfall of 400 mm yr!1, after ca. 270 mm of precipitation. The pots were filled with vermiculite, a homogeneous artificial horticultural substrate, and distributed at random in a greenhouse for maximal uniformity. To investigate whether pattern formation is water dependent we uniformly irrigated the pots (after 2 weeks acclimation period) once a week for 13–14 weeks (until the end of the winter growth season), with water amounts equivalent to 0, 100, 300 and 500 mm rainfall yr!1 (0, 126, 380 and 630 mL per pot per week accordingly), 20 replications per water treatment. figures. Hexagons can fill the entire plane or form in infinite strips as well as in growing Experimental studies and shrinking patches. Three-dimensional spots, such as light bullets, also exist and could To investigate whether ring formation is water dependent we tested the influence of different water regimes on the growth of P. bulbosa L. genets in laboratory conditions. We hypothesized that non-uniform biomass distributions should result from competition of individual ramets over the limited water resource. Moreover, individuals in the central part of a genet patch should experience stronger be organized into coherent stacks like in a crystal lattice. In this thesis, I will study stationary radially symmetric solutions in the Swift–Hohenberg equation: ut = −(1 + ∆)2 u − µu + νu2 − κu3 + O(u4 ), x ∈ Rn . (0.1) I will concentrate on spots and rings. They are interesting structures in their own right, but additionally in two dimensions hexagons are known to appear from spots through a symmetry breaking bifurcation [23, 3]. I am interested in the existence and bifurcation structure of different solutions ranging from one to three dimensions. The planar 4 Figure 1.2: Ferrosoliton from [17]. Hexagons from [31]. Ferro-fluids [Richter] Swift–Hohenberg (SH) equation serves as a normal form for Turing bifurcations in reaction diffusion equations with small data. It is one of the simplest models that exhibits interesting pattern forming behavior and is studied as an archetypical system for these reasons. Under the assumptions of stationarity and radial symmetry, this equation can be simplified dramatically to (∂r2 + n−1 ∂r + 1)2 u = −µu + νu2 − κu3 , r r∈R (0.2) where n is the dimension. In the above equation, the dimension appears explicitly. I will treat n as a continuous parameter and study the equation’s solutions as the dimension varies. This will allow me to examine how certain interesting features of the solutions cease and adjust between the one dimensional, planar and spherical cases. The behavior for non-integer values of n is interesting for analytical reasons. For example, the one dimensional SH equation exhibits snaking: this is pictured in figure 3.1. This is characterized by a parameter region where an infinite family of solutions exist. These look like localized roll patterns, only each solution has a different number of rolls. The bifurcation curve for these solutions is connected and appears like a snake’s tracks in grass. The behavior disappears in two and three dimensions, and by treating the dimension as a continuous parameter I can examine how the Hamiltonian of the system changes for an infinitesimally small change from n = 1. The 1050 LLOYD, SANDSTED 5 Figure 1.3: Patterns from [22]. ability to adjust n continuously is helpful from a numerical standpoint as well. It allows Figure 1. (a) Localized stationary spots and hexagon patches of (1.1 me stationary to find solutions in twoand and stripes three dimensions from for the well dimensional spots of (1.5) (µ,known ν) =one (2.5, 4). Both images solutions versa. u(x, and y), vice with x plotted horizontally and y vertically, where the values o indicated in the color bars shown to the right of the color plots: The colo are incontinuation the samesoftware fashion. I usedproduced the numerical auto07p [14] to study the bifurcation structure of the stationary radially symmetric SH equation. Previous work had uncovered two snaking families of symmetric solutions in one dimension. Asymmetric branches of solu- focus on the region ν ≥ 0 since the case ν < 0 is then recovere trivial state u = 0 is stable for µ > 0 and destabilizes at µ = two-pulses lying on isolas between the snaking branches and ladders. Additionally, [22] that have nonzero finite spatial wavelength. At µ = 0, hexa uncovered two families of ring solutions and one family of spot solutions in the planar case; bifurcation from u = 0 for each ν >the0,twowhile rolls bifurcate in a these were believed to snake. The present work reveals that! dimensional solutions from u = 0 provided ν > νr :=and become 27/38 [44]. While the bif do not snake. The ring solutions eventually turn around spot solutions. This initially unstable for µ > 0, they stabilize in a subsequent s means a second spot, referred to as spot B (with the original spot being spot A), exists. to bistability between the nontrivial patterns and the trivial By following the solutions down to one dimension the spots can be seen to become the of trivial and patterned states opens symmetric up thebifurcation possibility of fi snaking branches while the rings become two pulses. The spherically patches of hexagons rolls those in Figure 1. structure is qualitatively similar to theor planar case. such For n > as 1 snaking seemsshown to terminate. we profiles shallforfocus paper. Sample the spotson and in ringsthis are presented in figure 1.4. We first review briefly the situation in one dimension and existence of spot A, spot B, and the rings is established numerically through aThemore extensive discussion. In one space dimension, the Swi the computations mentioned above. I use analytic methods coming from the study of localized structures, as shown in Figure 2. The patterns show dynamical system to establish the existence of some of these solutions analytically. Spot A eter space, and their width increases as we move up on the is referred to as snaking [92]. There are several interesting q patterns shown in Figure 2: can we predict for which values can we determine a priori which periodic pattern is selected t tions exist that connect the snaking curves. The work of [21] established the existence of 6 u Spot A Spot B Ring Spot B u r r Figure 1.4: The profiles for rings and spots. and the rings were shown to exist analytically in [22]. Using similar methods, I have proven the existence of spot A in three dimensions. However, this same approach does not work to show spot B exists. The existence of spot B in two dimensions is established in this thesis, and the proof relies on the existence results of [22]. In some sense, spot B forms by gluing the spot A and ring solutions together in a piecewise fashion, but it behaves differently from either the rings or spot A in some unexpected ways. The thesis is organized as follows. In chapter 2, I discuss the general analytic approach used to establish the existence of the spot solutions. This will include examples and explanations for many of the key ideas used in the proofs. Chapter 3 details the numerical experimentation using auto07p. The numerical methods will be briefly explained and most of the chapter is dedicated to the bifurcation structure and its variation with changing dimensions. Chapter 4 rigorously establishes the existence of spot B in two dimensions. The proof for spot B implies certain behavior should be seen in the solution profiles, and some additional calculations are done to verify this behavior. In the conclusions, I discuss some open problems exposed by the previous results and briefly give some ideas that might be useful to their resolution. Remark 0.1. Spot A, spot B and the rings exist for different values of the parameters ν and κ: see the following chart for details. Spot B, being a spot A and ring glued together, only exists for parameter values (ν, κ) where the rings exist. This restriction on ν is solely used to establish the rings for the proof. 7 Existence Regions Spot A Spot B Rings ν>0 q ν > 27κ 38 q ν > 27κ 38 Chapter Two Preliminaries 9 This thesis concentrates on the existence of stationary solutions to the Swift–Hohenberg equation: all the numerical and analytical techniques used arise from the study of dynamical systems. In order to apply dynamical systems techniques to partial differential equations (PDEs), I will use spatial dynamics. Spatial dynamics refers to singling out a single spatial coordinate from the ambient space for treatment as the dynamical variable. This technique is especially well-suited to studying stationary one-dimensional solutions to PDEs. Traveling wave solutions, after switching to a co-moving reference frame, submit to the same techniques. In chapter 4 and in [25], these techniques are used to prove the existence of radially symmetric solutions in two and three dimensions where the radius is the spatial variable of interest. First, we would like to discuss some basic properties of ordinary differential equations (ODEs). The fixed point (0, 0) of the trivial linear ODE d u 2 0 u = dt v 0 −1 v with solution 2t u(t) Cu e = v(t) Cv e−t has stable manifold given by the v direction and unstable manifold given by u. More generally, the stable manifold consists of all solutions that converge to the equilibrium as t → ∞, and the unstable manifold of all solutions that converge to the equilibrium as t → −∞. Many techniques exist to track these manifolds near the equilibrium, and, in certain perturbative regimes, far from the equilibrium. For our purposes, normal form coordinates and the variation of constants formula will suffice to do this. Invariant manifold theory forms a cornerstone in the study of dynamical systems and it will be our primary tool to prove the existence of the spot solutions. These methods will be further discussed below, and the roughness of exponential dichotomies will proved as an illustrative example. The section will finish with a discussion of Turing bifurcations. 10 2.1 Variation of constants The variation of constants formula, here quoted from [10, Proposition 2.37 (Variation of Constants Formula)], allows us to write down the full solution to an ODE when we know the evolution operator for the linear leading order system. For a smooth function g, we have the following formula. Proposition 1.1. Consider the initial value problem ẋ = A(t)x + g(x, t), x(t0 ) = x0 and let t 7→ Φ(t) be a fundamental matrix solution for the homogeneous system ẋ = A(t)x that is defined on some interval J0 containing t0 . If t 7→ φ(t) is the solution of the initial value problem defined on some subinterval of J0 , then φ(t) = Φ(t)Φ −1 (t0 )x0 + Φ(t) Z t Φ−1 (s)g(φ(s), s)ds. t0 We mostly deal with systems in a perturbative regime, where the leading order system is linear and hyperbolic. The nonlinearities are generally small perturbations away from this. This variation of constants formula allows us to write down a fixed point equation for our solutions, and then we can solve them using the contraction mapping principle. Generally, we can make the formula a contraction through a small choice of one of the equation parameters. This then gives us a unique solution for each value of that parameter, which varies smoothly with the parameter. Often, we need to simplify the ODE in order to find a contraction, removing terms that do not uniformly shrink with the relevant parameters. For this, we use normal forms. 11 2.2 Resonant terms and normal forms Often when dealing with a system of ODEs, we would like to transform them into a simpler form. The real difficulty is understanding what terms can be removed, which cannot, and what transformations will achieve this. The study of normal forms provides the answers to these questions. The reduction to normal forms is achieved through power series expansions in the distance from an equilibrium position. These transformations do not always converge, though in C ∞ these transformations can generally be made rigorous. Poincaré developed the formal theory that allows us to simplify a non-resonant vector field into a linear field at a singularity. The vector fields he studied are formal power series and the transformations are formal diffeomorphisms. Consider the vector valued power series v(x) = Ax + · · · , x ∈ Rn where A has eigenvalues λ1 , . . . , λn . A resonance is defined between eigenvalues as a relation of the form λi = X kj λj , j with each kj ≥ 0 and P kj ∈ Z kj ≥ 2. A non-resonant vector field has no resonances between any of its eigenvalues. Poincaré argued formally that the equation ẋ = Ax + · · · can be reduced to ẏ = Ay when the original vector field in x is non-resonant. More generally, any non-resonant term can be removed while resonant terms remain. 12 Consider the example 1 2 u 2 0 0 u uv w v = 0 1 0 v + u2 + vw + h.o.t.. w 0 0 −1 w vw2 t By using normal forms, this can be reduced to 1 0 0 u u 0 2 v = 0 1 0 v + u2 + h.o.t. 0 0 −1 w w vw2 t near (0, 0, 0). Only the resonant terms remain. These are essentially the terms that grow or decay at the same rate as the linear terms in each component, and they cannot be separated from the linear flow. The normal form coordinates we use are only valid locally around the equilibrium. We will use them to understand the dynamics very near to different fixed points, but we will need to resort to different techniques to understand how the evolution occurs between the equilibria. 2.3 Stable manifold theorem In general, very little progress can be made on the study of nonlinear ODEs. However, if we can find fixed points to these nonlinear systems, then we can sometimes make a great deal of progress near the fixed points by linearizing the equations. Generally, we want to require the fixed point is hyperbolic such that the nonlinear contributions become arbitrarily small as compared to the linear flow at the equilibrium. We then find stable and unstable manifolds very near to the stable and unstable directions of the linear flow. The stable manifold theorem, as quoted from [18, Theorem 1.3.2], makes this claim more 13 precise. Theorem 1. Suppose that ẋ = f (x) has a hyperbolic fixed point x̄. Then there exist local s (x̄), W u (x̄), of the same dimensions n , n as those of stable and unstable manifolds Wloc s u loc the eigenspaces E s , E u of the linearized system ẏ = Df (x̄)y, y ∈ Rn , s (x̄), W u (x̄) are as smooth as the function f . and tangent to E s , E u at x̄. Wloc loc These stable and unstable manifolds are only given locally by this result. They can be extended globally by flowing the manifold out for infinite negative and positive time. This result also tells us nothing when we have a nonzero eigenvalue λ with Re λ = 0 in the linear system. More can be said in this situation, but it is not needed for our work. Fixed points are the most basic invariant manifolds, but they are dynamically boring. However, the stable and unstable manifolds of these fixed points can create very interesting dynamics. They can intersect with other stable and unstable manifolds for the same and different equilibria. It is this very phenomena that gives rise to the radially symmetric solutions we study. 2.4 Roughness theorem for exponential dichotomies In itself a very useful theorem, proving the persistence of exponential dichotomies under various perturbations provides a nice example of many of the techniques employed in this thesis. Exponential dichotomies for the system d u = A(ξ; λ)u dξ with u ∈ Cn are defined in [32, Definition 3.1]. (4.1) 14 Definition 4.1. Let I = R+ , R− or R, and fix λ∗ ∈ C. We say that eq. (4.1) with λ = λ∗ fixed, has an exponential dichotomy on I if constants K > 0 and κs < 0 < κu exist as well as a family of projections P (ξ), defined and continuous for ξ ∈ I, such that the following is true for ξ, ζ ∈ I. • With Φs (ξ, ζ) := Φ(ξ, ζ)P (ζ) where Φ(ξ, ζ) is the evolution operator of equation (4.1), we have s (ξ−ζ) |Φs (ξ, ζ)| ≤ Keκ , ξ ≥ ζ, ξ, ζ ∈ I. ξ ≤ ζ, ξ, ζ ∈ I. • Define Φu (ξ, ζ) := Φ(ξ, ζ)(id − P (ζ))), then u (ξ−ζ) |Φu (ξ, ζ)| ≤ Keκ , • The projections commute with the evolution, Φ(ξ, ζ)P (ζ) = P (ξ)Φ(ξ, ζ), so that Φs (ξ, ζ)u0 ∈ R(P (ξ)), ξ ≥ ζ, ξ, ζ ∈ I Φu (ξ, ζ)u0 ∈ N(P (ξ)), ξ ≤ ζ, ξ, ζ ∈ I. Now we consider a small perturbation of the linear system ut = Au + B(t)u (4.2) where we assume A is hyperbolic and supt |B(t)| < ε. Assuming A is a constant matrix is not necessary, we only need that the equation ut = A(t)u possesses exponential dichotomies itself. Also define the spectral projections P0s and P0u onto the set of eigenvalues of A with negative and positive real part respectively. Then there exist positive constants η, K > 0 such that for t ≥ 0 we can write down the estimates At s −At u e P0 + e P0 ≤ Ke−ηt . Theorem 2. For a sufficiently small ε, the perturbed system (4.2) possesses an exponential 15 dichotomy with constants K̃ and 0 < γ < η for some δ > 0. Proof. Banach’s fixed point theorem can be applied to an appropriate integral equation to construct the perturbed exponential dichotomies. These are denoted by Φs (t, s) and Φu (t, s). The integral equation comes from the variation of constants formula, Z t Z t Φs (t, s) = eA(t−s) P0s + eA(t−τ ) P0s B(τ )Φs (τ, s)dτ + eA(t−τ ) P0u B(τ )Φs (τ, s)dτ s ∞ Z s A(t−τ ) s u e P0 B(τ )Φ (τ, s)dτ, t ≥ s ≥ t0 − t0 Z t Z t Φu (t, s) = eA(t−s) P0u + eA(t−τ ) P0u B(τ )Φu (τ, s)dτ + eA(t−τ ) P0s B(τ )Φu (τ, s)dτ s t0 Z ∞ + eA(t−τ ) P0u B(τ )Φs (τ, s)dτ, s ≥ t ≥ t0 . s This equation can be found in the literature, for example [32, equation (3.8)]. The right hand side is a contraction in the space χs = {Φs : Φs (t, s) defined and continuous for t ≥ s ≥ t0 } χu = {Φu : Φu (t, s) defined and continuous for s ≥ t ≥ t0 } with norms kΦs ks = t≥s≥t0 kΦ ku = s≥t≥t0 u sup eγ(t−s) |Φs (t, s)| sup eγ(s−t) |Φu (t, s)|. We have now established a new exponential dichotomy, given by Φs,u , for the perturbed system. This completes the proof of the theorem. In the proof of the above theorem, the use of the contraction mapping principle on a fixed point equation in function space was crucial to establishing the claim. The fixed point equation came directly from the variation of constants formula, and the fixed point 16 was the function of interest. In a similar way, we will often use the variation of constants formula to write down an equation for our nonlinear flows in a perturbative regime. These will generally be solved by finding a contraction in an appropriate space. Unlike the above proof, we often need the explicit estimates on the solution provided by the variation of constants formula. 2.5 Turing bifurcations A Turing bifurcation is a diffusion-driven instability. When this was first proposed in [37], it was a radical idea because diffusion was believed to only stabilize systems. Here, through a simple calculation, we will provide an explicit example where the introduction of diffusion drives an otherwise stable system into an unstable configuration. This section closely follows the discussion in [26, Chapter 2] and requires a linear stability analysis. Consider the following reaction-diffusion system for u, v ∈ R: ut = γf (u, v) + ∆u vt = γg(u, v) + d∆v with γ > 0 < d, d 6= 1 and a homogeneous steady state f (u0 , v0 ) = g(u0 , v0 ) = 0. First, this system must be stable without the diffusive terms. The coordinates ũ = u − u0 and ṽ = v − v0 are more useful for the following calculations. To perform the linear stability analysis we start with ũ fu (u0 , v0 ) fv (u0 , v0 ) ũ =γ =: γA~u. ṽ gu (u0 , v0 ) gv (u0 , v0 ) ṽ By plugging in the ansatz ũ λt ∝e , ṽ 17 we arrive at the eigenvalue problem fu (u0 , v0 ) fv (u0 , v0 ) γ − λI = λ2 − γλ(fu + gv ) + γ 2 (fu gv − gu fv ) = 0. gu (u0 , v0 ) gv (u0 , v0 ) The eigenvalues are given by 1p 2 1 γ (fu + gv )2 − 4γ 2 (fu gv − fv gu ). λ± = γ(fu + gv ) ± 2 2 This gives us the conditions fu + gv < 0 and fu gv − fv gu > 0 to guarantee a stable system (the real part of the eigenvalues must be negative). Using the coordinates ũ = u−u0 and ṽ = v − v0 , we now show that adding the diffusive term causes an instability. The linearized equation around (0, 0) is ũt fu (u0 , v0 ) fv (u0 , v0 ) ũ 1 0 ∆ũ =γ + . ṽt gu (u0 , v0 ) gv (u0 , v0 ) ṽ 0 d ∆ṽ Define 1 0 D= . 0 d Now we want to study solutions to this equation of the form ~u(t) ∝ X eλt eikx k where the k are Fourier wavenumbers. The eigenvalues, λ, can then be found as roots of the characteristic polynomial |λI − γA + Dk 2 | = 0. After taking the determinant, we find the following expression for the eigenvalues 0 = λ2 + λ k 2 (1 + d) − γ(fu + gv ) + h(k 2 ) with h(k 2 ) = dk 4 − γ(dfu + gv )k 2 + γ 2 |A|. 18 If we can find a k 6= 0 such that a solution to this eigenvalue problem has a positive real part, then this system is now unstable due to the addition of a diffusive term. The k = 0 case is assumed stable as it is the case without diffusion. An instability will occur if h(k 2 ) < 0 for some choice of k. After a bit of work, this tells us two more conditions that will ensure an instability in the full system with diffusion: dfu + gv > 0, and (dfu + gv )2 > fu gv − fv gu . 4d Turing instabilities, through the mechanism just described, create patterns in reactiondiffusion systems. They are often applied to any pattern forming system simply because they are the best understood model that can produce such structures. Unfortunately, many of these situations cannot be shown, or have not been shown to fit into such a framework. In the work below, we use the Swift–Hohenberg equation because it is a normal form for such bifurcations, and because it is often studied. It is important, however, to remember that other models may exist that better serve to explain many experimental patterns. Chapter Three Numerical Exploration 20 The contents of this chapter were published under the title Snaking of radial solutions of the multi-dimensional Swift–Hohenberg equation: a numerical study in Physica D; see [24]. We are interested in the formation and parameter dependence of localized stationary radial solutions of the variational Swift–Hohenberg equation ut = −(1 + ∆)2 u − µu + νu2 − u3 , x ∈ Rn . (0.1) This equation was first derived by Swift and Hohenberg [36] to describe the effects of random thermal fluctuations on fluid convection just below onset. As shown for instance in [34], the steady Swift–Hohenberg equation is also the normal-form equation for smallamplitude radial solutions at Turing bifurcations in reaction-diffusion systems. More generally, the Swift–Hohenberg equation serves as a paradigm for bistable pattern-forming systems: it exhibits a plethora of interesting localized and non-localized patterns that have also been found in many other biological and physical systems [12, 16, 19, 28, 30]. Our interest is in localized radial steady-state solutions of (0.1). Part of our motivation stems from the observation made in [23, 3] that localized stripe, hexagon and rhomboid patches emerge from localized radial solutions via symmetry-breaking bifurcations. In addition to their relevance to such patterned patches, radial solutions are of interest in their own right in many physical systems, and we refer to [6, 13, 20] for references to systems that admit localized patterns of the shape discussed below. We now discuss equation (0.1) in more detail. Throughout this chapter, we take ν > 0 as the case ν < 0 is obtained upon replacing u by −u. Unless stated otherwise, all computations presented below are, in fact, done for ν = 1.6. The background state u = 0 is stable for µ > 0 and destabilizes in a Turing bifurcation at µ = 0. The Turing bifurcation gives rise to spatially periodic stationary patterns with period near 2π, which we refer to p as rolls. Rolls bifurcate into the region µ > 0 for ν > ν∗ = 27/38 ≈ 0.84 and into the region µ < 0 otherwise. For fixed ν > ν∗ , rolls are initially unstable, but when continued towards increasing µ, they undergo a fold bifurcation for sufficiently large µ at which they 21 stabilize. They then return back as stable patterns towards decreasing µ and finally cross µ = 0 with positive amplitude. As mentioned above, we focus on localized stationary radial solutions u(x, t) = u(|x|) of the Swift–Hohenberg equation. Such patterns satisfy the equation ∂r2 n−1 + ∂r + 1 r 2 u = −µu + νu2 − u3 , r ∈ R+ (0.2) with the boundary conditions ur (0) = urrr (0) = 0 and limr→∞ u(r) = 0, where r := |x|. To measure and represent the spatial width of localized radial patterns, we use their oneand two-dimensional L2 -norms given, respectively, by kuk2L2x := Z ∞ 0 |u(x)| dx, 2 kuk2L2r := Z 0 ∞ |u(r)|2 r dr. In (0.2), we can clearly consider n as a continuous parameter and examine the dependence of localized patterns on the continuous dimension parameter n. We are particularly interested in solution profiles u(r) that exist for µ > 0 and are, in an appropriate sense, composed of the stable roll structures that we discussed above. In one space dimension, these radial profiles resemble stable rolls with a localized envelope superimposed on them as illustrated in Figure 3.1, so that they can be thought of as localized rolls. In the planar case, the radial profiles we are interested in appear as localized target patterns; see Figure 3.2. We now summarize some of the known results about localized radial structures in dimension n = 1, 2, 3 for µ > 0. When n = 1, equation (0.2) is reversible and Hamiltonian, and much is known about localized radial patterns and their bifurcation diagrams [4, 5, 6, 7, 8, 11, 40]. Localized roll structures, which we refer to as pulses, exist for ν > ν∗ . Symmetric pulses that are invariant under x 7→ −x snake: their bifurcation branch, obtained by plotting the width of the roll plateau as measured by their L2x -norm against the parameter µ, resembles a vertical sinusoidal curve; see Figure 3.1. As we move up along the branch, pulses broaden as new rolls are added on either end at every other fold. As shown in Figure 3.1, there are 22 u 1 3 1 �u�2L2 x x 4 2 2 3 4 5 0.18 0.2 5 µ Figure 3.1: The center panel contains the bifurcation diagram of 1D localized pulses. The symmetric profiles that correspond to parameters on the light-colored curve have a maximum at r = 0 as shown in panels (1), (2), and (5), while the symmetric profiles corresponding to the dark-colored branch have a minimum at r = 0 as illustrated in panel (3). As we move up on each branch, a pair of new rolls is added to the solution profile at every other fold bifurcation. The two different branches discussed above are connected by ladder branches that correspond to asymmetric profiles as indicated in panels (3)-(5). These asymmetric structures bifurcate at pitchfork bifurcations near each fold from the symmetric pulses. two branches of symmetric pulses with either a positive maximum or a negative minimum at x = 0, and these branches are connected by horizontal ladder branches that correspond to asymmetric localized roll patterns. Among the other known solutions are symmetric 2-pulses, which are bound states of two individual well-separated localized roll structures. Two-pulses exist along figure-eight isolas that lie inside the regions formed by two consecutive ladder branches and the two snaking curves that connect them [38, 39]. More precisely, symmetric 2-pulses exist along a two-parameter family of isolas that are parametrized by (s, `), where ` ∈ N denotes the number of rolls in each of the two individual localized roll structures that make up the 2-pulse, and s ∈ N is the number of small-amplitude oscillations near u = 0 in between the two individual pulses [21]. Thus, s can be thought of as a measure of the separation width, while ` represents the L2x -norm of the 2-pulse. In particular, a countably infinite number of 2-pulses are expected to exist for each fixed value of their L2x -norm, and these 2-pulses are distinguished from each other by the increasing separation distance between the two individual pulses. In two dimensions, several different kinds of localized radial patterns were recently found in [22]. First, for each ν > 0, spots bifurcate from µ = 0 into µ > 0. As illustrated in Figure 3.2, these spots resemble J0 Bessel functions near r = 0, and they have an initial 23 �u�2L2r u(r) Spot Spot r u(r) µ �u�2L2r Ring Ring r µ Figure 3.2: Shown are profiles, representative color plots, and bifurcation branches of localized planar spot A solutions in the top row and of the two localized planar ring solutions in the bottom row. Profiles and color plots correspond to solutions at (µ, ν) = (0.005, 1.6). [Reproduced from [22]]. amplitude of order √ µ for small µ. From now on, we refer to these structures as spot A solutions. In addition to these spots, two ring solutions emerge from µ = 0 for each fixed √ ν > ν∗ . These solutions have an overall sech-like shape with a maximum of order µ that √ occurs at r ' 1/ µ. For ν > ν∗ , spot A and the two rings appear to snake as can be seen in Figure 3.2. All of these solutions were proved to exist for 0 < µ 1 in [22]. In three dimensions, numerical evidence for the existence of spots was presented in [22]; their existence near onset is proved in [25]. In contrast to the planar case, 3D spots do not appear to snake: instead, the L2r -norm along branches of localized spots stays bounded. Our goal in this chapter is to understand the change in the behavior of spots and rings when the dimension switches from two to three, and to investigate how the 1D, 2D, and 3D structures described above are related to each other. To elucidate the different behaviors of profiles and branches as n varies, we treat n as a continuous parameter and use numerical continuation techniques to follow spots and rings from n = 2 upwards to n = 3 and downwards to n = 1. In particular, the focus of this chapter is on numerical computations, though we will outline some possible avenues for analysis and rigorous proofs in §3.5 below. We now briefly summarize our results. 24 Spot A Spot B Spot B u u r Spot A r Figure 3.3: The profiles of spots A and B with (µ, ν) = (0.005, 1.6) are compared in the left panel, while an enlarged plot of spot A is shown separately in the right panel. Note that spot B resembles an inverted spot A but with a much larger amplitude. The zeros of both profiles appear to align well for r 1. First, we discovered a second family of planar 2D spots, from now on referred to as spot B, which seem to exist only for ν > ν∗ . In contrast to the spot A structures, spot B solutions have a negative minimum at r = 0 as shown in Figure 3.3. In addition, they 3 1 appear to scale like µ 8 as µ → 0 and are therefore not captured by the µ 2 -scaling used in the analysis of spot A solutions in [22]. Second, when we follow spots A and B and the two ring structures down in dimension to n = 1, we find that spot A and B become, respectively, the symmetric 1D pulses with a maximum and a minimum at r = 0 that we discussed above. The two rings, however, turn into symmetric 1D 2-pulses. Recall that symmetric 2-pulses exist along a two-parameter family of isolas, and the mechanism for the production of isolated branches from two connected ring snaking curves turns out to be quite complicated. Our numerical continuation results show that each ring curve folds over onto itself several times in a complicated manner and then pinches off a number of 2-pulse isolas. On the other hand, we also found 2-pulse isolas that are not connected to the ring branches upon increasing n but instead shrink to a point and disappear. Our third result concerns the snaking structure of spots A and B for 2 ≤ n ≤ 3, which turns out to be equally complicated. Recall that indefinite snaking was predicted in [22] from the numerical computations presented there. It turns out that the computations in [22] were stopped at a value of the L2r -norm that was not large enough to reveal the 25 more complicated bifurcation structure that we report on here. Indeed, as we follow spot A up on its bifurcation curve, the curve eventually turns around, and the L2r -norm of the spots begins to decrease again. At this point, the profile of the underlying pattern transforms from a spot to the profile of one of the two rings. Similarly, spot B broadens for a while, but eventually transforms into the second ring and follows the ring bifurcation curve downwards towards decreasing L2r -norm. In particular, spots and rings are pairwise connected in parameter space. Above these two connected curves lies a family of stacked isolas of localized structures, which also terminates for a large enough value of the L2r norm. Above these stacked isolas, we found a connected U -shaped solution curve that seems to extend up to infinite L2r -norm. Both of the two branches of this curve snake and the associated profiles cycle through spot A and B solutions. These branches seem to continue indefinitely towards increasing L2r -norm, but the width of the snaking regions in the µ-direction decreases. We also gain insight into how the snaking curves above and below the isolas depend on the parameter µ and will discuss this further in §3.2. We proceed as follows. Section 3.1 describes the numerical techniques used. Section 3.2 details the bifurcation structures for n = 2. In §3.3, the changes of the bifurcation structure are explored when n is increased from two to three, while §3.4 discusses how these structures change when n is decreased from two to one. Section 3.5 presents conclusions and open problems. 3.1 Numerical algorithms For the sake of clarity, we briefly outline the numerical protocols used in the exploration of the snaking diagrams. To continue localized radial profiles, we numerically solved 26 boundary-value problems that are based on the first-order system u u1 u2 d u1 = dr u u3 2 u3 −(1 + µ)u + νu2 − u3 − 2( n−1 r u1 + u2 ) + (n−1)(n−3) u1 (r r2 − u2 ) − 2(n−1) u3 r (1.1) on the interval (0, L) together with the Neumann boundary conditions u1 (0) = 0, u3 (0) = 0, u1 (L) = 0, u3 (L) = 0 (1.2) at r = 0, L. Unless stated differently, we used ν = 1.6 in all computations. We employed auto07p [14] to continue solutions of (1.1)-(10.1) in the parameter µ. Computing the connected snaking branches of symmetric 1D pulses and planar spots and rings is then straightforward. To ensure that the results do not depend on the value of L and to prevent boundary effects, we checked for each computation that the computed patterns are sufficiently small near the boundary and, in addition, repeated these computations for significantly larger values of L (typically at least doubling L). In the rest of this section, we outline the changes that are necessary to continue asymmetric pulses and to find isolas of symmetric 2-pulses, planar spots and planar rings. Finding isolas: There are several types of isolas that appear in our calculations, and it requires different techniques to find them. When an isola lies above a snaking segment, we must move around in parameter space in order to find the isola. Fortunately, the bifurcation structure provides an easy solution. As n is decreased, the height of the connected snaking curve is found to increase. Thus, we initially continue a solution in n for fixed µ towards an appropriate smaller value of n, and then fix this value of n and follow the snaking curve in µ towards increasing L2r -norm. Afterwards, we fix the parameter µ and continue in n towards increasing n until we reach its original value. If we continued high enough in the L2r -norm in the second step, the final solution will lie on an isola, which we can now trace 27 out by continuing in µ for fixed n. When we continue ring structures from dimension two to dimension one, isolas of 2-pulses are pinched off the bifurcation curves. To find these isolas, we continue a large number of solutions with starting data in a single period of the snaking structure towards decreasing n. Computing asymmetric 1D pulses: The computation of asymmetric pulses for n = 1, when the system (1.1) is autonomous, requires an additional phase condition to fix the location of the localized pattern somewhere inside the interval (0, L). We use the usual integral phase constraint Z 0 L old uold 1 (x)(u(x) − u (x)) dx, (1.3) where uold refers to the solution evaluated at a previous continuation step. In order to solve the phase constraint, we add the term cu1 to the last component in (1.1), so that c can be thought of as a wave speed: theoretically, c should vanish identically during continuation; in practice we found that c is typically of order 10−12 and certainly never exceeds 10−6 . To find starting data, we break the pitchfork bifurcation through which the asymmetric states appear, which will allow us to obtain asymmetric pulses by continuing the known symmetric pulses. To break the reflection symmetry r 7→ −r present for n = 1, we add the term δ sin r to the fourth component of (1.1). Thus, to find asymmetric pulses, we solve (1.1) with the expression (0, 0, 0, cu1 + δ sin r)t added to its right-hand side, together with (10.1)-(1.3). We start with a symmetric 1D pulse away from the pitchfork bifurcation and continue initially in δ up to a fixed small value, typically near δ = 0.05. Afterwards, we continue in µ for fixed δ until we encounter a fold bifurcation. Once we have passed the fold bifurcation, we continue in δ for fixed µ until δ becomes zero. The resulting structure is then the desired asymmetric profile on a ladder branch, which can be validated by continuing again in µ. During the above computations, we allow c to vary, although, as explained above, its value will stay close to zero. 28 3.2 Localized 2D states In this section, we focus on the bifurcation diagram of spots and rings for equation (0.2) with n = 2. We emphasize that other localized structures may exist but these are not p considered here. We fix ν = 1.6 and note that ν exceeds the critical value ν∗ = 27/38 ≈ 0.84 below which rings do not exist. We consider exclusively the regime µ > 0, where u = 0 is stable for (0.1). As already mentioned, the existence of three solution branches associated with smallamplitude spot A structures and two ring patterns was proved in [22] for 0 < µ 1 in the regions ν > 0 for spot A and ν > ν∗ for rings. The numerical evidence presented in [22] indicated that these branches begin to snake indefinitely as in the one-dimensional case. Indeed, the three solution branches were continued in µ a significant distance away from the origin, and convincing snaking was seen with the associated folds approaching two vertical asymptotes. As in the one-dimensional case, additional localized rolls are added at every other fold along the branch near the tail of these localized structures. It turns out, however, that this picture changes drastically when a large enough number of localized rolls has been added to the underlying pattern or, in other words, when the L2r -norm has become sufficiently large. In Figure 3.4(i), we present computations that indicate that the spot A branch and one of the ring branches are connected in parameter space. In other words, if we continue spot A solutions towards increasing L2r -norm, then the branch will reach a maximal L2r -norm near which the spot A profiles transform into rings, and the branch will then continue downwards towards decreasing L2r -norm along the ring branch. At the maximal L2 -norm, the underlying profile consists of around 20 rolls. Figure 3.4(ii) shows the results of a similar computation, where we continued the second ring along its bifurcation branch. At the top of the solution branch shown in Figure 3.4(ii), the ring profiles transform into spot-like profiles, with the maximal amplitudes occurring near the core at r = 0, and the associated branch descends towards lower values of the L2r -norm. This second spot (spot B) has not been observed before, and we will comment on 29 !u!2L2r (i) secondary snaking structure !u!2L2r µ (ii) µ Figure 3.4: Shown are the connected bifurcation curve of spot A and one of the ring solutions in panel (i) and the bifurcation branch of spot B and the second ring solution in panel (ii). In the upper right corner of panel (i), the branch oscillates between three folds aligned approximately at µ ≈ 0.18, 0.19, and 0.21, and we refer to the part of the branch that oscillates between the two rightmost folds as the secondary snaking structure. Note that the vertical L2r -axes in panels (i)-(ii) are scaled differently: in particular, the spot A branch reaches a larger value of the L2r -norm. The solution profiles at the points labelled (a)-(d) are shown in Figure 3.5. u(r) u(r) r r u(r) u(r) r r Figure 3.5: Panels (a)-(d) contain the solution profiles of spots and rings at the parameter values labelled (a)-(d) on the branches shown in Figure 3.4. As the spot and ring branches are traversed towards increasing L2r -norm, additional rolls are added at the right tail of the localized profiles. The maximal (minimal) amplitude of spot A (spot B) always occurs at r = 0 along the branch. For rings, u(0; µ) oscillates between positive and negative values as we move from one leftmost fold to the next on the branch; new rolls are created only at the tail but not near r = 0. We refer to the movies at http://www.dam.brown.edu/people/mccalla/SpotAmovie.mpg and http://www.dam.brown.edu/people/mccalla/SpotBmovie.mpg for further details on the behavior of spots and rings. 30 (i) �u�2L2r (ii) �u�2L2r secondary! snaking! structure Spot B µ µ Figure 3.6: Panel (i) shows in blue the connected snaking branch of the spot B and ring B solutions from Figure 3.4(ii) together with a stack of isolas, plotted in red and alternately in dashed and solid, along which profiles resemble those of spot B and ring B. Panel (ii) contains the spot A curve (in dark cyan) and the spot B branch (in blue) from Figure 3.4 together with the stacked isolas (in red) from panel (i). Note that the isolas align well with the secondary snaking structure visible near the top of the spot A branch, indicating that that they pinch off from the spot A branch as n is changed. its properties in more detail below. From now on, we refer to the ring structures connected to spot A and B as the ring A and ring B patterns, respectively. Representative profiles of spots and rings are shown in Figure 3.5. As indicated in Figure 3.4(i), a secondary snaking structure is visible near the upperright part of the spot A branch, where the branch oscillates between three distinct limits rather than two: we refer to the part of the branch that oscillates between the two rightmost folds as the secondary snaking structure. Its presence appears to be related to a stacked family of isolated branches of spot B and ring B structures that fill the region in between the secondary snaking structure we just discussed and the spot B branch from Figure 3.4(ii). These isolas along with the spot B branch are shown in Figure 3.6(i) and together with both spot branches in Figure 3.6(ii). Above the spot A branch, we found a second family of 31 �u�2L2r µ Figure 3.7: Shown is the first isola (in green) of a second family of stacked isolas that appears above the spot A branch (plotted in dark cyan). �u�2L2r (i) µ �u�2L2r (ii) µ Figure 3.8: The lower parts of both panels contain the connected snaking branch of spot A and ring A (in dark cyan) from Figure 3.4. Above this branch, we found a family of stacked isolas (plotted in green) that include the isola shown in Figure 3.7. The stack of isolas extends only up to a value of the L2r -norm at which the profiles consist of approximately 38 rolls. Above this value, we found a single connected solution curve (drawn in brown) that consists of two intertwined branches that both snake, seemingly indefinitely. For clarity, we show only one of the two intertwined branches in the upper part of panel (ii). Solution profiles along the upper snaking curve can be found in the accompanying http://www.dam.brown.edu/people/mccalla/SpotABmovie.mpg. 32 stacked isolas. The first of these isolas is shown in Figure 3.7, while some of the remaining isolas are presented in Figure 3.8(i). Along each isola, the solution profile changes in an intricate way between the four spot and ring profiles. The second family of isolas ends at a value of the L2r -norm that corresponds to profiles that contain around 38 rolls. Above this value, we found another connected solution branch that consists of two intertwined arms, each of which snakes as indicated in Figure 3.8. Note that both of these arms oscillate back and forth between folds that align themselves along four distinct curves. As we follow either vertical branch of the snaking curve up, the amplitude of the pattern near the core at r = 0 oscillates up and down between the maximum of spot A and the minimum of spot B. These oscillations create new rolls near the core as we move up on the branch, while no new rolls are formed near the right tails of the localized structures: this is in sharp contrast to the situation along the lower spot A and B branches or the situation for n = 1. Note that Figure 3.8 also indicates that the width of the top snaking branches decreases as we move up on the branch. Figure 3.9 contains log-log plots of the L2r -norm against the difference of the µ values at which folds occur from the Maxwell point µ = 0.2 at which the fully nonlinear 1D roll patterns with vanishing Hamiltonian have zero energy1 . These results suggests that the width shrinks to zero as the L2r -norm goes to infinity. Note though that we do not know whether the top branch continues upwards indefinitely or whether it, too, ends at a finite value of the L2r -norm. Finally, we comment in more detail on the planar spot B solutions that we encountered. Recall that their profiles are shown in Figure 3.3. These spots differ in various ways from the spot A patterns found in [22]. First, spot B resembles the Bessel function −J0 near its core, and its amplitude is therefore negative near r = 0. More importantly, Figure 3.10 3 shows that the supremum norm of spot B appears to scale like µ 8 as µ approaches zero, so 1 that spot B is significantly larger than spot A, whose amplitude scales with µ 2 . Another significant difference is that spot A was proved to bifurcate at µ = 0 from the trivial 1 Equation (0.1) is variational for n ≥ 1, and (0.2) is Hamiltonian for n = 1; see [4, 5] and references therein for details. 33 -5 -6 -6 -7 4 5 4 5 (i) log(−u(0)) log u(0) Figure 3.9: The two panels show log-log plots of the two leftmost and two rightmost folds of the high snaking branch shown in Figure 3.8, indicating that the snaking branch converges algebraically to the Maxwell point µ = 0.2 of 1D rolls. 0.49 log µ − 0.0071 (ii) 0.374 log µ + 0.7178 log µ log µ 1 Figure 3.10: Panel (i), reproduced from [22], indicates that the amplitude of spot A scales as µ 2 as µ approaches zero. As shown in panel (ii), the amplitude of spot B appears to scale approximately like µ0.374 . background state u = 0 for each fixed ν > 0 [22]. In contrast, as shown in Figure 3.11, p we were not able to continue spot B below ν = ν∗ = 27/38 ≈ 0.84, which is the value at which rings cease to exist. Thus, we believe that the bifurcation mechanism that leads to the existence of spot B solutions depends crucially on the far field even though their profile envelopes appear to decay to zero monotonically in r. 3.3 The connection between 2D and 3D branches The bifurcation diagram for 2 ≤ n ≤ 3 is similar to the 2D case, except that the height and width of the isolas and the snaking branches decrease significantly as n is increased. 34 µ = 0.1 µ = 0.05 µ = 0.01 µ = 0.005 µ = 0.001 µ = 0.0005 µ = 0.0001 µ = 0.00005 !u!2L2r ν ν= � 27 38 ln(µ) ν Figure 3.11: To delineate the existence region of spot B, we continued spot B in the parameter ν for several fixed values of µ and visualize the resulting solution branches in two different ways: in the left panel, we plot ν versus the squared L2r -norm (the values of µ decrease from right to left), while p the right panel shows log µ versus ν. Note that the solution branches stay above the critical value ν = 27/38 and that the L2r -norm of the associated profiles goes to infinity as ν approaches the lower end of each branch. To illustrate these behaviors, we show in Figure 3.12 the lower snaking branches of the two spot-ring pairs and, in Figure 3.13, the upper snaking branch of the two spots for different values of n. We do not show our computations of the family of stacked isolas between the spot A and spot B branches or of the second family of isolas that exist between the lower and upper snaking branches shown in Figure 3.12 and 3.13, respectively. These isolas look qualitatively similar to those for n = 2, but they are narrower and there are fewer of them as the height of the overall bifurcation diagram decreases. In §3.2, we found that the width of the upper snaking branch for n = 2 shrinks as we move up along the branch. Figure 3.13 indicates furthermore that the overall width of these branches decreases as n increases. In addition, the two arms of the upper snaking branch that overlap significantly for n = 2 become separate for n = 3. Even though both of these arms lie to the left of the Maxwell point µ = 0.2, fitting the folds using a loglog plot indicates that the spine of these branches aligns itself with a curve of the form µ = 0.2 − Ckuk−1.38 for some constant C > 0. L2 r Note that we used the L2r -norm in Figures 3.12 and 3.13. It might be more natural to use the n-dependent norm kuk2L2n := Z 0 ∞ |u(r)|2 rn−1 dr, (3.1) 35 �u�2L2r (i) Spot A �u�2L2r �u�2L2r (ii) Spot B �u�2L2r µ µ µ µ Figure 3.12: The bifurcation curves of spot A and spot B solutions are presented in panels (i) and (ii), respectively, for different values of the dimension parameter n. The insets show the branches for n = 3 in more detail. (i) (ii) �u�2L2r �u�2L2r µ µ Figure 3.13: Panel (i) contains the upper snaking branches of spots for n = 2 (in brown), n = 2.3 (in cyan), and n = 3 (in black). Panel (ii) contains the two arms of the snaking branch for n = 3 to illustrate that they do not overlap. or appropriate scalar multiples thereof, but since using this norm did not reveal any features not already visible in the L2r -norm, we decided to use the latter. 3.4 The connection between 1D and 2D branches In this section, we investigate to which of the localized 1D pulses the planar states connect when we decrease n. Thus, we start with the planar spots and rings that we found in §3.2 and continue them in n towards n = 1. Note that the spot and ring branches are connected at the top for some large value of the L2r -norm and that this value increases as n decreases. Furthermore, the associated profiles change from spots to rings only near the 36 �u�2L2x µ Figure 3.14: The two curves plotted in cyan diamonds correspond to the limits at n = 1 of the lower planar spot A and spot B branches when continued in n. The profiles along these branches for n = 1 coincide with the 1D pulses shown in Figure 3.1. The solid figure-eight isolas plotted in red arise when we continue the two ring branches from n = 2 down to n = 1 using the methods outlined in §3.1. The profiles along each isola are symmetric 1D 2-pulses. µ µ n = 1.2 �u�2L2r n=1 µ �u�2L2x n = 1.2 n = 1.29 n = 1.3 µ n = 1.3 n = 1.302 µ µ µ Figure 3.15: The left panel contains the ring A branch for different values of n plotted in the planar L2r -norm. The curve for n = 1.2 is connected but clearly shows structures that will pinch off to become individual isolas for smaller values of n. These isolas continue to form and pinch off as the dimension is decreased further, thus leading to isolas of 2-pulses with a given L2r -norm and an arbitrary separation between the pulses. The right panel shows the ring A branch for n = 1.2 and n = 1.3 but now plotted in the one-dimensional L2x -norm. Note that the curve for n = 1.2 appears to cover an entire family of what will later become separate 2-pulse isolas at n = 1. top of their respective branches, and we can therefore distinguish these two branches easily and continue them separately towards n = 1. Figure 3.14 contains the solution branches at n = 1 that we obtain when we continue the planar spots and rings in n from two to one dimensions. As expected, the planar spot A and spot B patterns connect to the symmetric 1D localized roll structures shown in Figure 3.1 that have, respectively, a maximum or a minimum at r = 0. The situation 37 �u�2L2x u(r) µ 0 25 50 75 100 r 0 25 50 75 100 125 Figure 3.16: The left panel contains four isolas at n = 1 that are found from the two planar ring branches through continuation in n. The right panel contains the solution profiles at the topmost intersection of these isolas with the line µ = 0.195: the profiles in panels (a)-(b) and B come from ring A, while the profiles in panels (c)-(d) arise from ring B. Since these profiles were computed with Neumann boundary conditions at r = 0, they can be reflected across r = 0 and therefore correspond to 2-pulses. for rings is more complicated, and the limiting set we obtain at n = 1 when continuing each of the two ring branches towards decreasing n is actually a set of isolated branches that correspond to symmetric 1D 2-pulses. The mechanism that leads from an initially connected bifurcation curve at n = 2 to a family of isolas at n = 1 is elucidated in Figure 3.15. As we decrease n, each ring curve becomes entangled with itself and begins to pinch off isolas. In the left panel of Figure 3.15, we show three isolas for n = 1 that are formed from the ring A branch and correspond to 2-pulses with similar L2r -norm. Observe that the 2-pulse branches look quite different in the right panel of Figure 3.15, where we plot them in the one-dimensional L2x -norm. We remark that plotting solution branches in the n-dependent norm from (3.1) actually obfuscates the relation between the branches for n > 1 and their limits at n = 1. The reason is that the solutions change near r = 0 during continuation, and varying the power of r in the norm during continuation can hide or amplify the effect of these changes. Thus, solution branches appear better represented by using a fixed norm. Before we address the relation between planar rings and symmetric 1D 2-pulses in more detail, we briefly summarize some of the results for 2-pulse isolas from [21] as these will be useful in the forthcoming discussion. As shown in [21], for each pair (s, `) of sufficiently large integers, there are four isolas along which symmetric 1D 2-pulses exist: the parameter 38 u(r) k=1 s=1 ks=3 =3 k=2 s=2 r r r Figure 3.17: The profiles shown here at n = 1 were found through continuation from rings. Due to the Neumann conditions imposed at r = 0, these solutions correspond to symmetric 2-pulses with different separation distances represented by the number s of small oscillations near r = 0. (i) upper! fold (ii) upper folds upper! fold lower folds lower! fold lower! fold Figure 3.18: When we continue an asymmetric 1D pulse that is centered some distance away from x = 0 in n, we obtain the isolas in panel (i) which shrink and eventually disappear. Panel (ii) contains continuation results in (µ, n) of the two upper and lower folds along the isolas. As n increases, the lower folds disappear in a cusp, thus making the isola more circular, while the collision of the remaining upper folds corresponds to the point at which the isola disappears. ` represents the number of large-amplitude rolls in each of the two pulses that make up the 2-pulse, while s is the number of small-amplitude oscillations in between the two 1-pulses. Thus, s measures the separation distance of the two pulses, while ` represents the width of each pulse. For each of the four 2-pulses that exist for a given pair (s, `), let u(r) be its profile and denote by j the quadrant in which the pair (u(0), urr (0)) lies, then the integer j ∈ {1, . . . , 4} characterizes the 2-pulse uniquely among the four 2-pulses. In other words, the jth 2-pulse u(r) has (u(0), urr (0)) in the jth quadrant, where j ∈ {1, . . . , 4}. In Figure 3.16, we plot the four isolas that belong to the same pair (s, `) in the left panel and the associated symmetric 2-pulse profiles in panels (a)-(d), which correspond respectively to j = 1, . . . , 4. The 2-pulses in panels (a)-(b) arise when we continue ring A towards n = 1, while the 2-pulses in panels (c)-(d) come from the planar ring B. This is 39 consistent with the preceding discussion as the rolls contained in the two planar rings differ by a phase shift of half their period (in the limit µ → 0, the rings are given by the Bessel function ±J0 ). Figure 3.17 shows the profiles of 2-pulses for different values of s that we obtained by following rings towards dimension one. Note that we cannot be sure whether all isolas of symmetric 2-pulses are obtained from continuing the two planar ring patterns to dimension one. In fact, it is hard to envision that each ring branch folds up on itself infinitely often to generate a countably infinite number of isolas for each given ring width ` as this would require the ring structures to move away from r = 0 to generate 2-pulses for all possible separation distances s. To test whether there are isolas that do not connect to the ring branch, we computed one of the asymmetric localized 1D pulses that exist for n = 1 along the ladder branches shown in Figure 3.1. We then placed this pulse at position r = r0 inside the interval [0, r1 ], where 1 r0 r1 are chosen so that the profile is close to zero for r = 0 and r = r1 . Afterwards, we continued this profile in n using Neumann boundary conditions. The choice of Neumann conditions guarantees that we can view the resulting profile as a 2-pulse for n = 1 and a ring solution for n > 1. Furthermore, the choice of r0 guarantees that the 2-pulse corresponds to a large value of the separation parameter s. The resulting bifurcation diagrams for three different values of n are shown in Figure 3.18. Thus, it appears as if these 2-pulse isolas shrink to a point and disappear without connecting back to one of the ring branches. We did not investigate systematically which of the 2-pulse isolas shrink to zero in the same fashion but believe that this happens for all 2-pulses with larger values of s. 3.5 Discussion The numerical explorations indicate how the localized radial patterns that exist in dimensions one to three are related to each other when the dimension parameter n is treated as a continuous variable. In particular, we found that planar spots connect to symmetric 1D pulses, while planar rings become the symmetric 1D 2-pulses. We also resolved the 40 apparent discrepancy between snaking in 2D and non-snaking in 3D that was reported in [22]. Our results show that neither planar nor 3D spots snake; instead, the bifurcation diagram is similar in both cases and consists of branches that snake over a long but finite interval which are followed by stacked isolas for sufficiently large values of the L2 -norm of the underlying patterns. We also found a new family of localized radial structures of the planar Swift–Hohenberg equation that we have referred to as spot B solutions. These spots do not seem to obey 3 1 the expected µ 2 -scaling when µ goes to zero and instead seem to scale like µ 8 . They also p appear to exist only for values of ν above the critical value ν∗ = 27/38. The analytical techniques used in [22] to prove the existence of planar rings and spot A states can also be utilized to investigate the existence of spot B solutions: a preliminary formal analysis in chapter 4 corroborates that spot B solutions exist only for ν > ν∗ and predicts an 3 amplitude scaling µ 8 ; making this formal study rigorous is the goal in chapter 4. While our numerical computations give a quite detailed picture of the planar bifurcation diagram, we do not understand the complicated structure of different connected solution branches that alternate with families of stacked isolas that they revealed. Perhaps the best approach to gain a theoretical understanding of these diagrams is to carry out a perturbation analysis of symmetric 1D pulses in the continuous bifurcation parameter n near dimension one. Indeed, the mechanism that leads to the one-dimensional bifurcation structure shown in Figure 3.1 is well understood, and the recent dynamical-systems investigation in [4] may allow us to carry out a perturbation analysis of h i2 ∂r2 + ∂r + 1 u = −µu + νu2 − u3 , r r ∈ R+ (5.1) in := n − 1 near = 0. We remark though that such a perturbation analysis may turn out to be difficult, given the complexity of the bifurcation diagrams for n > 1. In the remainder of this section, we briefly outline a formal argument that one could utilize when attempting to understand how radial spots of (5.1) for 0 < 1 emerge from 41 symmetric 1D pulses. First, we remark that the perturbation from = 0 is, despite its appearance, a regular perturbation as the singularity at r = 0 for > 0 can be resolved by choosing logarithmic variables near r = 0. Thus, we expect that each given symmetric 1D pulse persists for sufficiently small positive > 0 provided we stay away from pulses that undergo fold or pitchfork bifurcations. Thus, the key issue is to understand pulses whose L2 -norm is large as we cannot guarantee uniformity of the persistence interval in for such pulses (nor do we anticipate uniformity given the complex bifurcation structure we expect to find for n > 1). To discuss the persistence of such solutions, observe that (5.1) is Hamiltonian at = 0 with Hamiltonian given by H(u) = ur urrr − u2rr (1 + µ)u2 νu3 u4 + u2r + − + . 2 2 3 4 To gain an initial understanding into the behavior of localized solutions of (5.1) for > 0, it is natural to compute the change of the Hamiltonian along such a solution. Assuming that u(r; ) is a family of solutions of (5.1) that is bounded in ≥ 0 and satisfies u(r; ) → 0 as r → ∞ uniformly in , we find via a straightforward calculation that H(u(∞; )) − H(u(0; )) = Z 0 ∞ d H(u(r; )) dr = −2 dr Z 0 ∞ [urrr + ur ] ur dr + O(2 ). r Next, assume that u resembles a spatially periodic roll pattern with wavenumber κ for r ∈ (1, R) with R 1. Arguing now formally by assuming that these rolls are of the form cos κr and proceeding as above, we find that H(u(R; )) − H(u(0; )) ≈ −2 Z 1 R [urrr + ur ] ur dr ≈ κ2 (κ2 − 1) log R. r Thus, as R increases, the wavenumber κ can stay within a bounded interval only if = 0 or else κ approaches unity like 1/ log R. On the other hand, [5, Figure 12] shows that 1D front solutions that connect a roll pattern with wavenumber κ to the trivial state u = 0 exist only when κ = κc for a certain κc < 1. Hence, if we fix 0 < 1, then the wavenumber of rolls inside an extended localized pattern needs to change along the spatial profile from κ = 1 to κ = κc , where the solution can return to u = 0. This suggests that an analysis of 42 spots for dimensions near one needs to account for roll patterns for wavenumbers κ in an entire interval [κc , 1]. Chapter Four Existence Near Onset for Spot B in 2D 44 The existence of the ring solutions and spot A for the planar radially symmetric Swift– Hohenberg equation was rigorously shown in [22]. In chapter 3, spot B was found numerically to exist in two and three dimensions. See figure 4.1 for a reminder of spot A and spot B. This chapter will be devoted to rigorously showing the existence of spot B in two dimensions. Recall the Swift–Hohenberg equation is ut = −(1 + ∆)2 u − µu + f (u), ~x ∈ Rn , (0.1) with f (u) = νu2 − κu3 + O(u4 ). We are interested in stationary radially symmetric solutions that decay to 0 as r → ∞. We only study the case where 0 < µ 1; small amplitude solutions bifurcate from the trivial state at µ = 0. The term ν > 0 is always taken as positive; the leading order analysis for ν < 0 is identical upon taking u → −u. From here on, we will ignore the O(u4 ) terms as they do not alter the following analysis. These solutions then satisfy 1 (∂r2 + ∂r + 1)2 u1 = −µu1 + νu21 − κu31 . r (0.2) The main theorem in this section follows. Theorem 3. Fix ν > 0 and κ ∈ R such that c03 := 3κ 19ν 2 − < 0, 4 18 there exists a µ0 > 0 such that equation (0.1) has a stationary localized radial solution u(r) for every µ ∈ (0, µ0 ). These solutions remain bounded near the trivial state u = 0 and for 3 √ a fixed r0 , they asymptotically appear as u(r) = −βµ 8 J0 (r) + O( µ) for µ → 0 uniformly in 0 ≤ r ≤ r0 . Note β > 0. 45 Spot A Spot B Spot B u r Figure 4.1: Spot A versus spot B. 4.1 Geometry of the rings and spots The Swift–Hohenberg equation can be rewritten as a four dimensional first order system. It can be made autonomous by defining the additional variable α = 1 r which satisfies αr = −α2 . Consider the profile of localized stationary radial solutions as the orbit of an ordinary differential equation in the radial variable r. The solutions will be understood in terms of three different coordinate charts: the core manifold, the transition chart, and the rescaling chart. The rings and spot A are already understood in these various charts, and spot B turns out to be a piecewise connection of one of the rings with spot A. See figure 4.2 for an illustration. All solutions that remain smooth and bounded in an interval [0, r0 ] can be understood using regular perturbation theory with Bessel functions; the resulting set is referred to as the core manifold and forms a two-dimensional manifold. Unfortunately, the core manifold does not capture the dynamics from the far field. For this, we need both the transition chart and rescaling chart. The rescaling chart tracks the exponential growth and decay of solutions as r → ∞. 46 Core Transition Rescaling Ring Spot B Spot A Figure 4.2: The pictured schematic represents the core manifold, the transition region, and the rescaling chart. The core manifold, at top left, is a two-dimensional manifold that captures the smooth bounded solutions within the interval [0, r0 ] for a fixed but finite r0 > 0. The transition chart, at bottom left, captures the algebraic growth and decay of solutions. The rescaling chart, at right, captures the exponential decay of solutions as the radius goes to ∞. The blue solid curves represent spot A and the ring on the left and right respectively. The dashed red line represents spot B. Equation (0.2) exhibits equilibria corresponding to the exponential growth and decay rates s (µ) of these equilibria in the rescaling chart. The goal is to track the stable manifolds W∞ back until r = r0 and then seek intersections with the core manifold. These equilibria are hyperbolic and the accompanying stable manifolds are two-dimensional. As r → ∞, the system becomes the Ginzburg–Landau equation. The ring solution lies near a known solution of the Ginzburg–Landau equation, and spot B lies near the same solution in the far field. Using this information, the stable manifold of the appropriate far-field equilibrium can be tracked into the transition chart. Spot B only exists for c03 < 0 because the rings only exist for c03 < 0 and spot B is a ring glued to spot A. The transition chart coordinates track the algebraic decay rates of the solutions. Roughly speaking, the coordinates are of the form v = r uur : equilibria of v correspond to algebraic decay rates for u, and a solution of the form u(r) = rα maps to the equilibrium solution v(r) = α. In this chart, there are equilibria at ± 12 corresponding to Bessel function solutions of Swift–Hohenberg at µ = 0. The ring solutions pass from the core manifold to the positive equilibrium and then out to the far field. Spot A goes through the negative equilibrium on its way from the core manifold to the far field. In Figure 4.2, an illustration of the spot and ring solutions is shown in the context of the three different regions. 47 Spot B passes by both equilibria. Starting with the section of the stable manifold in the transition chart near to the equilibrium at 21 , we integrate in backwards time around the fixed point. A diffeomorphism is then used to pass from this equilibrium to − 12 . The section is then propagated around the last fixed point and matched to the core manifold. In this way, spot B is shown to exist. Many of the preliminary steps in this proof were drawn directly from [22]. We will begin by reviewing the appropriate sections from that paper. Afterwards, we will present a formal calculation for the existence of spot B that predicts the proper scaling of the amplitude and provides intuition for the proof. 4.2 Life at the core Equation (0.1) can be rewritten as the following system: (∂ 2 + 1 ∂r + 1)u1 = u2 r r (2.1) (∂r2 + 1 ∂r + 1)u2 = −µu1 + νu21 − κu31 . r This can be recast as a first order system Ur = AU + F(U, µ), 0 0 0 0 A= −1 1 0 −1 1 0 − 1r 0 0 1 , 0 1 −r (2.2) 0 0 F(U, µ) = . 0 2 3 −µu1 + νu1 − κu1 For any fixed but finite r0 > 0, the small radially symmetric solutions of equation (2.2) that are bounded and smooth in an interval [0, r0 ] were characterized in [22] and are referred to as the core region. System (2.2), when linearized about U = 0 with µ = 0 to give Ur = AU , has a set of four linearly independent solutions Vi . These are in terms of the 48 Bessel functions Jk and Yk of the first and second kind. Quoting [22, Equation (2.4)], V1 (r) = √ V2 (r) = √ V3 (r) = √ V4 (r) = √ 2π(J0 (r), 0, −J1 (r), 0)T , 2π(rJ1 (r), 2J0 (r), rJ0 (r), −2J1 (r))T , 2π(Y0 (r), 0, −Y1 (r), 0)T , 2π(rY1 (r), 2Y0 (r), rY0 (r), −2Y1 (r))T . The core manifold is a two-dimensional manifold in R4 for fixed r0 , and we denote the projection onto the space spanned by V1,2 (r0 ) as P−cu (r0 ). P−cu (r0 ) has the span of V3,4 (r0 ) as its null space. We may now recall the following lemma. Lemma 2.1 ([22, Lemma 1]). Fix n = 2 and r0 > 0, then there are constants δ0 , δ1 > 0 so that the set W−cu (µ) of solutions U (r) of (2.2) for which sup0≤r≤r0 |U (r)| < δ0 is, for |µ| < δ0 , a smooth two-dimensional manifold. Furthermore, U ∈ W−cu (µ) with |P−cu (r0 )U (r0 )| < δ1 if and only if ˜ + |d| ˜ 2) U (r0 ) = d˜1 V1 (r0 ) + d˜2 V2 (r0 ) + V3 (r0 )Or0 (|µ||d| 1 2 3 2 ˜ ˜ ˜ ˜ +V4 (r0 ) √ + o(1) ν d1 + Or0 (|µ||d| + |d1 | + |d2 | ) . 3 (2.3) ˜ < δ1 , where the right-hand side in (2.3) depends for some d˜ = (d˜1 , d˜2 ) ∈ R2 with |d| ˜ µ), and o(1) is the Landau symbol in r0 as r0 → ∞. smoothly on (d, −1 The o(1) estimate can be improved to O(r0 2 ). This both simplifies our notation and the final step in the proof of theorem 3. Lemma 2.2. Equation (2.3) can be replaced by ˜ + |d| ˜ 2) U (r0 ) = d˜1 V1 (r0 ) + d˜2 V2 (r0 ) + V3 (r0 )Or0 (|µ||d| 1 − 12 2 3 2 ˜ ˜ ˜ ˜ +V4 (r0 ) √ + O(r0 ) ν d1 + Or0 (|µ||d| + |d1 | + |d2 | ) . 3 (2.4) Proof. In the proof of [22, Lemma 1], the quadratic coefficient in d˜1 in front of V4 (r0 ) is 49 found by evaluating the integral πν 4 Z r0 0 πν sJ0 (s) ds = 4 3 Z 0 ∞ sJ0 (s) ds + o(1) . 3 The o(1) term accounts for the integral of the tail. The J0 Bessel function, as r → ∞, has q 3 2 the expansion πr cos(r − π4 ) + O(r− 2 ): see [1, (9.1.10), (9.1.11) and section 9.2] or [22, Table 1]. We will use this asymptotic form to calculate the contribution of the tail integral Z ∞ sJ0 (s)3 ds = r0 = Z ∞ s r0 √ Z 2 2 r 3 π 2 cos(s − ) + O(s− 2 ) πs 4 !3 ds 3 1 π √ cos3 (s − ) + O(s− 2 )ds 4 s r π "0 r ! r !! r ! r !#∞ r √ 2s 2s 6s 6s 1 1 √ 3 3 S = +C +S −C +O π π π π r0 π 6 r0 r 1 . = O r0 = ∞ 3 2 The functions S(s) and C(s) are the Fresnel sine and cosine integrals, respectively, and the asymptotic form for the Fresnel functions is found from [27, Chapters 7.5, 7.12]. 4.3 Normal forms We start from equation (2.2) to understand the solution for large r of the stationary radial Swift–Hohenberg equation. We can treat this as an autonomous equation by adding the variable α = 1 r. Here we are interested in the far field which implies 0 < α 1. This variable is governed by the additional differential equation αr = −α2 . Equation (2.2) can 50 then be recast as u1 u3 u2 u4 d . u = −u1 − αu3 + u2 3 dr u4 −u2 − αu4 − µu1 + νu21 − κu31 −α2 α (3.1) Following [15, 22, 34], we use the normal form coordinates 1 0 0 2i U = à + B̃ + c.c. i 1 0 −2 (3.2) which are equivalent to à 1 2u1 − i(2u3 + u4 ) = , 4 B̃ −u4 − iu2 U = (u1 , u2 , u3 , u4 )T . (3.3) The equation, transformed into these coordinates, becomes α α à + B̃ + ï + O (|µ| + |Ã| + |B̃|)(|Ã| + |B̃|) 2 2 α α¯ = i− B̃ − B̃ + O (|µ| + |Ã| + |B̃|)(|Ã| + |B̃|) 2 2 Ãr = B̃r i− (3.4) αr = −α2 . Let α0 := r0−1 and ε2 := µ. We can now use theory developed in [34] to transform this equation into a more useful form. Lemma 3.1 ([22, Lemma 2]). Fix 0 < m < ∞, then there is a change of coordinates A à −iφ(r) [1 + T (α)] + O (|ε|2 + |Ã| + |B̃|)(|Ã| + |B̃|) =e B B̃ (3.5) 51 such that (3.4) becomes α Ar = − A + B + RA (A, B, α, ε) 2 α ε2 Br = − B + A + c03 |A|2 A + RB (A, B, α, ε) 2 4 (3.6) αr = −α2 . The transformation (3.5) is polynomial in (A, B, α), smooth in ε. T (α) = O(α) is linear and upper triangular for each α. The function φ(r) satisfies φr = 1 + O(ε2 + |α|3 + |A|2 ), The constant is c03 = 3κ 4 − 19ν 2 18 . φ(0) = 0. (3.7) The remainder terms are RA (A, B, α, ε) 2 X = O |Aj B 3−j | + |α|3 |A| + |α|2 |B| + (|A| + |B|)5 + |ε|2 |α|m (|A| + |B|) (3.8) j=0 RB (A, B, α, µ) 1 X = O |Aj B 3−j | + |α|3 |B| + |ε|2 (|ε|2 + |α|3 + |A|2 )|A| + (|A| + |B|)5 + |ε|2 |α|m |B| . j=0 The core manifold must be transformed into the (A, B)-coordinates given by (3.5) before we can map them into our transition chart coordinates and match them to the far-field. The core manifold is mapped into (A, B) in [22, Equation (3.23)] 2 2 2 A W−cu (ε)|α=α0 : = ei[−π/4+O(α0 )+Oα0 (|ε| +|d| )] (3.9) B q √ ˜ −1 ˜ 2 2 ˜ ˜ α0 d1 [1 + O(α0 )] − α0 d2 [i + O(α0 )] + Oα0 (|ε| |d| + |d| ) × . √ √ ˜ √ √ ˜2 2 2 3 ˜ + |d˜2 | + |d˜1 | ) − α0 d2 [i + O(α0 )] − [1/ 3 + O( α0 )]ν α0 d1 + Oα0 (|ε| |d| 52 Summary: We have now collected all the necessary results from [22]. From this, we have the normal form coordinates for the equations in the far-field and the core manifold in these coordinates. We still must transform everything into our rescaling and transition charts and match the far-field to the core manifold. The next section will establish the rescaling coordinates between the far-field and matching regimes. 4.4 The rescaling chart The rescaling coordinates are A ε −αA/2 + B z = = ε εA √ = ε= µ A2 = z2 ε2 α2 = (4.1) α ε s = εr. The Swift–Hohenberg equation expressed as (3.6), in these coordinates, becomes the system Ar B̃ 1 = + RA ε ε ε 1 = εA2 z2 + RA ε 1 = ε A2 z 2 + 2 R A ε ∂r A2 = 53 and ∂r z2 = = " # 1 B̃r B̃ − 2 Ar ε A A " 2 # 1 α4 A + 14 µA + c03 |A|2 A − αB̃ + αRA + RB B̃ − 2 (B̃ + RA ) ε A A εα22 1 RB c0 z2 + α2 RA + 2 + ε + 3 |εA2 |2 − εα2 z2 − εz22 + 4 4 ε εA2 ε A2 1 + α22 1 z 2 + α2 0 2 2 = ε RA + 3 RB . + c3 |A2 | − α2 z2 − z2 + 2 4 ε A2 ε A2 = We want to change our evolution variable r into s = εr. The calculations for the remainders are straightforward and tedious: α2 RA ε2 A2 , ε22 A2 z2 + , α2 ε2 , ε2 = (4.2) 2 2 X = O |Aj B 3−j | + |α|3 |A| + |α|2 |B| + (|A| + |B|)5 + |ε|2 |α|m (|A| + |B|) j=0 = |A2 |O |ε2 |4 + |ε2 |3+m |α2 |m and α2 RB ε2 A2 , ε22 A2 z2 + , α2 ε2 , ε2 = (4.3) 2 1 X = O |Aj B 3−j | + |α|3 |B| + |ε|2 (|ε|2 + |α|3 + |A|2 )|A| + (|A| + |B|)5 + |ε|2 |α|m |B| j=0 = |A2 |O |ε2 |5 + |ε2 |4+m |α2 |m . The remainders in the rescaling chart are all found to be of order O(|ε2 |2 ). We arrive at the equation in the rescaling chart ∂s A2 = A2 z2 + O(|ε2 |2 ) ∂ s z2 = 1+ 4 α22 ∂ s ε2 = 0 ∂s α2 = −α22 . + c03 |A2 |2 − α2 z2 − z22 + O(|ε2 |2 ) (4.4) 54 Summary: s (ε) back in time we need to convert it into the transition Once we follow W∞ chart coordinates (5.1). The next section derives equation (3.6) in the transition chart s (ε) back in time up to the matching point α = α . coordinates so that we may follow W∞ 0 4.5 The transition chart We transform (3.6) into coordinates (A1 , z1 ) such that z1 is approximately the algebraic decay rate of the solutions. Let z and τ be defined by z = − α2 + B A and eτ = r. Then the matching coordinates are A1 = z1 = ε1 = A α z 1 B =− + α 2 αA ε α α1 = α. These satisfy the equations ∂ r A1 = Ar Aαr − 2 α α 1 = αA1 + αA1 z1 + RA α 1 = α A1 + A1 z1 + 2 RA α and Br BAr Bαr − − 2 2 αA αA α A α ε2 B − α2 A + B + RA − 2 B + 4 A + c03 |A|2 A + RB B = − + 2 αA αA A 2 1 z + ε2 1 1 1 1 2 = + αc03 |A1 |2 − α z1 + + α z1 + − RA + 2 RB 4α 2 2 αA1 α A1 " # 1 z1 + 1 ε2 1 = α −z12 + + 1 + c03 |A1 |2 − 2 2 RA + 3 RB 4 4 α A1 α A1 ∂r z1 = (5.1) 55 where RA and RB are now evaluated at (A, B) = (A, α2 A1 z1 + α2 A 1 2 ). We also need the r derivatives of ε1 and α1 which are ∂r ε1 = − εαr = ε = αε1 α2 ∂r α1 = −αα1 . This finally gives the equation in the transition chart ∂τ A1 = A1 [1 + z1 ] + ∂τ z1 = −z12 + 1 RA α2 (5.2) z1 + 1 1 + ε21 1 + c03 |A1 |2 − 2 2 RA + 3 RB 4 α A1 α A1 ∂τ ε1 = ε1 ∂τ α1 = −α1 . We still need to find RA and RB in terms of (A1 , z1 ). It will be important to show the remainders in (5.2) vanish when A1 = 0, ε1 = 0 and z1 = − 21 . Introduce the coordinate z− = z 1 + 1 2 then, abbreviating the algebraic details, z− 1 2 RA (α1 A1 , α1 A1 z1 + , α1 , α12 ε21 ) = α2 A1 2 2 X z− = O |Aj B 3−j | + |α|3 |A| + |α|2 |B| + (|A| + |B|)5 + |ε|2 |α|m (|A| + |B|) α 2 A1 j=0 = 2 X z− O |(α1 A1 )j (α12 z− A1 )3−j | + |α1 |3 |α1 A1 | + |α1 |2 |α12 z− A1 | α12 A1 j=0 ! +(|α1 A1 | + |α12 z− A1 |)5 + |α12 ε21 ||α1 |m (|α1 A1 | + |α12 z− A1 |) z− O |α1 |4 |A1 | + |ε21 ||α1 |m+3 |A1 | 2 α 1 A1 = z− O |α1 |2 + |ε21 ||α1 |m+1 = z− O |α1 |2 = 56 and 1 α3 A = 1 RB ((α1 A1 ), (α12 A1 z− ), α1 , α12 ε21 ) = (5.3) 1 X 1 O |Aj B 3−j | + |α|3 |B| + |ε|2 (|ε|2 + |α|3 + |A|2 )|A| α 3 A1 j=0 ! +(|A| + |B|)5 + |µ||α|m |B| = 1 X 1 O |(α1 A1 )j (α12 A1 z− )3−j | + |α1 |3 |α12 A1 z− | α13 A1 j=0 +|ε1 α1 |2 (|ε1 α1 |2 + |α1 |3 + |α1 A1 |2 )|α1 A1 | +(|α1 A1 | + |α12 A1 z− |)5 + |ε1 α1 |2 |α1 |m |α12 A1 z− | ! 1 5 5 5 2 5 2 m+4 O |α | |A | + |α | |A ||z | + |ε | |α | |A | + |ε ||α | |A z | 1 1 1 1 − 1 1 1 1 1 − 1 α13 A1 = |α1 |2 O |A1 |4 + |z− | + |ε1 |2 . (5.4) = Reiterating the transition chart equation with the remainders included gives ∂τ A1 = A1 1 + z1 + O(|α1 |2 ) ∂τ z1 = −z12 + 1+ 4 ε21 ∂ τ ε1 = ε1 0 2 2 + c3 |A1 | + |α1 | O |A1 | + z1 + 1 + |ε1 | 2 (5.5) ∂τ α1 = −α1 . These are the coordinates we will use for the matching, so it is important to express the core manifold in terms of them. The computation is straightforward and results in W−cu (ε)|α=α0 : (5.6) 2 2 2 A1 = ei[−π/4+O(α0 )+Oα0 (|ε| +|d| )] 1 −2 −3/2 ˜ 2 ˜ 2 ˜ ˜ × α0 d1 [1 + O(α0 )] − α0 d2 [i + O(α0 )] + Oα0 (|ε| |d| + |d| ) √ √ ˜ + |d˜2 |2 + |d˜1 |3 ) 1 −d˜2 [i + O(α0 )] − [1/ 3 + O( α0 )]ν d˜21 + Oα0 (|ε|2 |d| z1 = − + . ˜ + |d| ˜ 2) 2 α0 d˜1 [1 + O(α0 )] − d˜2 [i + O(α0 )] + Oα (|ε|2 |d| 0 57 Summary: Most of the computations will be done in these coordinates. As we already have the core manifold at the matching point r = r0 , it remains to understand the props (ε) around the equilibria at z = ± 1 . These computations will comprise agation of W∞ 1 2 most of the analysis in this chapter. The transition and rescaling charts are related by the following transformations: A1 = 4.6 A2 , α2 z1 = z2 , α2 ε1 = 1 , α2 α1 = ε2 α2 = εα2 . The fixed points Because we will be examining the flow around the two equilibria in the matching regime, we want to use coordinates where the fixed points are located at the origin. The equation in the transition chart is ∂τ A1 = A1 1 + z1 + O |α1 |2 ∂τ z 1 = −z12 ∂ τ ε1 = ε1 1 + ε21 1 0 2 2 + + c3 |A1 | + |α1 | O |A1 | + |z1 + | + |ε1 | 4 2 ∂τ α1 = −α1 . We find two fixed points given by p± := (A1 , z1 , ε1 , α1 ) = (0, ± 21 + O(ε1 ), 0, 0). We can better understand the dynamics around the fixed points by changing to appropriate coordinates. First around the equilibrium p+ using z+ = z1 − P+ : ∂ τ A+ ∂τ z+ 1 2 gives 3 2 = A+ + z+ + O |α+ | 2 ε2 2 = −z+ − z+ + + + c03 |A+ |2 + O |α+ |2 4 ∂τ ε+ = ε+ ∂τ α+ = −α+ . (6.1) 58 Repeating this procedure with the coordinates z− = z1 + 1 2 (around the p− fixed point) leads to P− : ∂τ A− ∂τ z − 1 2 = A− + z− + O |α− | 2 ε2 2 = z− − z− + − + c03 |A− |2 + |α− |2 O (|A− | + |z− | + |ε− |) 4 (6.2) ∂τ ε− = ε− ∂τ α− = −α− . Remark 6.1. The same proof of existence could be attempted for u ∈ R3 but the equilibrium corresponding to P− loses hyperbolicity in the A1 direction. The coefficient of the A− term changes from 1 2 to 0. From this, we would expect logarithmic terms to arise when we integrate around this equilibrium. This could cause problems for the final matching. Summary: The P+ coordinates move the origin to (A1 , z1 ) = (0, 21 ), while P− centers it at (A1 , z1 ) = (0, − 21 ). In P+ , the origin is a hyperbolic equilibrium. The linearized system has decaying eigenvector along the z+ direction and growing eigenvector in the A+ direction. On the other hand, P− exhibits growth along both directions. See figure 4.7 for a picture. 4.7 The formal argument in two dimensions 3 Our goal is to find a simple argument that explains how the scaling of µ 8 arises. In this section, we will proceed formally and loosely: we will immediately drop all higher order terms and set every allowable constant to one. Additionally we will restrict ourselves to real variables, the normal form coordinates. Then, for c03 < 0 and ν > ν∗ , there is a ring solution connecting the far field to the equilibrium p+ . The intersection of the unstable s (µ) is transverse in the real coordinates. This enables us to follow manifold of p+ and W∞ s (µ) that lies near (A , z ) = (−η, 0) = (−1, 0) until τ = a piece of W∞ + + 1 r0 = 1. This is 59 Core Transition Rescaling Ring Spot B Spot A s cu Figure 4.3: The solid blue curve for the ring lies in the transverse intersection of W∞ and W− . We can s find starting data near (A+ , z+ ) = (−η, 0) in W∞ . done first by moving from the section defined by ε1 = ε01 = 1 into the section z+ = − 21 using the linearized flow around p+ . We then switch to the linearized flow in P− to track all the way back in time to our matching point. Finally we find the intersections between s (µ) at r = 1. Consider the equation in the the core manifold and this segment of W∞ 0 transition chart ∂τ A1 = A1 [1 + z1 ] + O |α1 |2 |A1 | ∂τ z1 = −z12 + ∂τ ε1 = ε1 1 ε21 + + c03 |A1 |2 + O |α1 |2 4 4 ∂τ α1 = −α1 . As mentioned at the start of the section, we will work with the simplified equations ∂τ A+ = 3 A+ 2 (7.1) ∂τ z+ = −z+ around p+ and ∂τ A− = 1 A− 2 ∂τ z − = z − (7.2) 60 around p− . We actually desire a more general argument than this would provide, so in the charts P+ and P− , respectively, we will study the equations ∂τ A+ = aA+ (7.3) ∂τ z+ = −z+ and ∂τ A− = bA− (7.4) ∂τ z − = z − . Recall equation (5.6) for the core manifold in the transition chart coordinates W−cu (µ)|r=r0 : 2 2 A1 = ei[−π/4+O(1/r0 )+Or0 (|µ|+|d| )] h√ i 3/2 ˜ + |d| ˜ 2) × r0 d˜1 [1 + O(r0−1 )] − r0 d˜2 [i + O(r0−1 )] + Or0 (|µ||d| √ √ ˜ + |d˜2 |2 + |d˜1 |3 ) 1 −d˜2 [i + O(r0−1 )] − [1/ 3 + O( α0 )]ν d˜21 + Or0 (|µ||d| , z1 = − + ˜ + |d| ˜ 2) 2 d˜1 /r0 [1 + O(r−1 )] − d˜2 [i + O(r−1 )] + Or (|µ||d| 0 0 0 which we reduce to W−cu (µ)|r=1 : A1 = d˜1 − id˜2 1 −id˜2 − d˜21 z1 = − + . 2 d˜1 − id˜2 By setting d˜2 = 0, we can write down an estimate for the core at the matching point A1 = d˜1 (7.5) 1 z1 = − − d˜1 . 2 We need to know how long to follow the flow until we land at the matching point r = 1. 61 In other words, starting from the assumption at τ = 0 that ε1 = 1 and letting −T be the time such that α1 (−T ) = 1 we find T = − ln ε. s is given by (A , z ) = (−1, −y) with y > 0. Equation Assume the stable manifold W∞ + + (7.3) admits solutions A+ = −eaτ , z+ = −ye−τ , τ ≤ 0. 1 From this, we can deduce the time τ+ when z+ (τ+ ) = − 21 = −ε 4 y and extract τ+ = ln y where we have again changed an unimportant constant to one. Plugging this in, we find A1 (τ+ ) = −eaτ+ = −y a z1 (τ+ ) = 0. Now we have the amount of time we need to integrate around P− ε τ− = ln ε − ln y = ln . y Solving equation (7.4) from (A− (0), z− (0)) = (−y a , 12 ) for time τ− = ln yε gives A− = −y a ebτ− = −y a z− = eτ− = εb = −y a−b εb yb ε . y Now recall (7.5) to find A1 = d˜1 = −y a−b εb z− = −d˜1 = y a−b εb . 62 Matching to the core in the z− component and solving tells us 1−b y a−b+1 = ε1−b ⇒ y = ε a−b+1 . Equivalently, the final amplitude of the J0 Bessel function in the core is given by d˜1 = −y a−b εb = −εb ε a−b+1 (1−b) = ε a−b+1 a−b a a = µ 2(a−b+1) For spot B, a = 3 2 (7.6) 3 3 and b = 21 . This results in the scaling d˜1 = ε 4 = µ 8 . The above s with Wcu . calculation has additionally told us the initial offset necessary to match W∞ − 1−b 1 Note that y = ε a−b+1 ⇒ y = ε 4 for our system. We will assume an offset of this form in 1 the proof of the theorem: y = ε 4 x. The eigenvalues around both fixed points p± contribute to the final scaling. In order to understand the scaling of similar systems in general, we would need to have a detailed understanding of the eigenvalues and the flow around all the fixed points. Summary: 3 Notice that the unusual scaling of µ 8 , the negative value at the origin, and the existence region coupled to the ring all arise naturally from the formal argument. The main difficulties for a rigorous proof involve integrating the full equations with nonlinearities around the different fixed points. 4.8 The flow around the equilibria in the transition chart In this section, we establish several lemmas needed to prove the main theorem. They govern how the flow moves between and around the different equilibria. With these lemmas in hand, we can propagate the solutions backwards from r = ∞ to r = r0 and then match 63 with the core manifold to complete the proof. 4.8.1 Transversality and the ring: The formal argument relies on the existence of an orbit between fixed points in the transition and rescaling charts. The orbit should exist in the plane defined by α1 = 0 and ε2 = 0 as illustrated in Figure 4.4. Its existence follows from the work of [33]: in appropriate coordinates, the equation in the invariant plane matches the equation studied by Scheel. Consider the coordinates (a, b) from [22] that are related to our coordinates (A, B) via A= so that b = as + a 2s . √ µa, B = µb, s r=√ µ In these coordinates the far field equation, ignoring remainders, becomes ass = − as a a + 2 + + c03 |a|2 a. s 4s 4 (8.1) We want to understand the implications of [22, Lemma 4] for our equation in the different charts. Lemma 8.1 ([22, Lemma 4]). Assume that c03 < 0, then, for each integer n ≥ 0, equation (8.1) has a bounded nontrivial real solution a(s) = qn (s) that has precisely n simple zeros √ for s ∈ (0, ∞) and satisfies qn (s) = O( s) as s → 0 and (qn , qn0 )(s) → 0 exponentially as s → ∞. Furthermore, the linearization of (8.1) about qn (s) does not have a nontrivial real-valued solution that is bounded uniformly on R+ . If c03 > 0, then the only bounded solution of (8.1) on R+ is a(s) ≡ 0. We can translate this into the following result for our system. Lemma 8.2. For ε = 0 and c03 < 0, there is a locally unique solution connecting the equilibrium (A1 , z1 ) = (0, 21 ) in the transition chart to the equilibrium (A2 , z2 ) = (0, − 21 ) in the rescaling chart. Also, in the real subspace, which is invariant for α1 = 0, the unstable man- 64 α1 ε2 1 (A1 , z1 ) = (0, ) 2 Invariant Plane: (α1 = ε2 = 0) 1 (A2 , z2 ) = (0, − ) 2 Figure 4.4: A cartoon of the orbit. We are looking for a ring solution that lies in the invariant plane where α1 = 0 and ε2 = 0. The remainder terms in both the transition and rescaling chart then drop out and we are left with (8.1). ifold of (A1 , z1 ) = (0, 12 ) transversely intersects the stable manifold of (A2 , z2 ) = (0, − 21 ). In R2 this structure persists up to O(ε) corrections for all sufficiently small 0 < ε 1. Proof. First, the charts, and consequently solutions of (5.5), (4.4) and (8.1), are related by the transformations A2 = a, A1 = as, 1 b + , 2s a 1 sb z1 = − + , 2 a z2 = − s = α2−1 ε1 = s, (8.2) ε α1 = . s Recall the transition chart, ∂τ A1 = A1 [1 + z1 ] + A1 O |α1 |2 ∂τ z1 = −z12 + ∂τ ε1 = ε1 1 ε21 + + c03 |A1 |2 + O |α1 |2 4 4 ∂τ α1 = −α1 , and note that the remainder terms disappear when α1 = 0. We are then firmly in the situation of Lemma 8.1. From the lemma, we know the asymptotic form of qn (s) as s → 0. 65 Also recall b = as + a 2s √ √ and as s → 0 then qn (s) = c1qn s + o( s). Note c1qn 6= 0: for all sufficiently small values of c1qn > 0, all solutions of this form are strictly positive away from the origin [33]. The only exception for c1 = 0 is the trivial solution. Looking at the asymptotics in the transition chart, 3 3 A1 (s) = as = c1qn s 2 + o(s 2 ) → 0 as s → 0 1 sb 1 s a z1 (s) = − + =− + as + 2 a 2 a 2s 1 sas sas 1 =− + = + 2 a 2 a √ √ 1 cq s + o( s) 1 √ → as s → 0, = 1n √ 2 2cqn s + o( s) we see that the solution associated with qn via (8.2) is in the unstable manifold of the equilibrium (A1 , z1 ) = (0, 21 ). The rescaling chart is given by ∂s A2 = A2 z2 + O(|ε2 |2 ) ∂s z2 = 1 1 2 + α + c03 |A2 |2 − α2 z2 − z22 + O(|ε2 |2 ) 4 4 2 ∂s ε2 = 0 ∂s α2 = −α22 . Again, the remainders disappear when confined to the invariant plane given by α1 = ε2 = 0. To understand the solution in the rescaling chart, we must use as s → ∞ then qn (s) = −s/2 −s/2 cs e √s + o( e √s ) with cs 6= 0 and o(1) → 0 as s → ∞. See Lemma 2.1 in the appendix. The limit in the far field as s → ∞ is e−s/2 e−s/2 A2 (s) = a = cs √ + o( √ ) → 0 as s → ∞ s s 1 b 1 1 a 1 as 1 as z2 (s) = − + = − + as + =− + + = 2s a 2s a 2s 2s a 2s a e−s/2 1 e−s/2 1 1 √ √ −cs 2 s 1 + s + o( s ) − 1 + s + o(1) 1 = = 2 → − as s → ∞. −s/2 −s/2 e√ e√ 1 + o(1) 2 cs s + o( s ) 66 The solutions qn belong in the stable manifold of the equilibrium (A2 , z2 ) = (0, − 21 ). This means the solution qn lies in the intersection of the stable and unstable manifolds of the equilibria discussed above. This intersection is transverse in R2 by lemma 8.1 and locally unique because there are no real-valued nontrivial uniformly bounded solutions to the linearized equation around the solution qn . Everything depends smoothly on ε and the intersection of the manifolds is transverse thus this all persists with small perturbations, though we expect O(ε) changes to the equilibria and solutions. Remark 8.3. The real heteroclinic orbit arising from qn can be rotated by an arbitrary phase, eiγ for any γ ∈ R, to produce a one parameter family of heteroclinic connections in s . The phase component eiγ only multiplies the A component in the stable manifold W∞ 1 s as it cancels in the z component. We will need to choose γ to solve the matching W∞ 1 s . equation between the core manifold and W∞ Summary: This crucially provides our starting data in the transition chart coordinates. Additionally, it provides us with a range of starting values near to the P+ equilibrium for an appropriately small choice of µ. We need this freedom to find the intersections between the core and far-field. 4.8.2 P+ : Recalling the formal analysis, spot B looks like a ring solution and a spot A solution glued together somewhere in the transition chart. We first must understand the flow around the equilibrium at z1 = 1 2 near the solution predicted by [33]. Refer to figure 4.5 for a picture. This corresponds to pulling the ring back from the far field and looking at it in the transition regime. After moving the p+ fixed point to the origin through the coordinates 67 P + A1 p+ ε1! = "ε01 1 ,0 2 z1 −η −δ+ s W∞ ε1 s Figure 4.5: The equilibrium at (z1 , A1 ) = ( 21 , 0) is pictured. A segment of W∞ is drawn in red. By choosing the appropriate point on this manifold, the matching equations are solved. We need to understand how this segment is affected by the flow between and around the two equilibria. z+ := z1 − 1 2 we arrive at the equation 3 2 ∂ τ A+ = A+ + z+ + O |α+ | 2 ε2 2 ∂τ z+ = −z+ − z+ + + + c03 |A+ |2 + O |α+ |2 4 (8.3) ∂τ ε+ = ε+ ∂τ α+ = −α+ . Several low-order non-resonant terms in the z+ equation cause difficulties when we try to solve. We have to transform them away. We will do this calculation in two steps: first 2 ) terms and then we will we explicitly write down transformations to eliminate the O(α+ flatten out the unstable manifold. 68 Lemma 8.4. Equation (8.3) can be brought into the form 3 ∂τ A+ = A+ + O (|A+ | + |ẑ+ | + |ε+ | + |α+ |) 2 i hε + 2 2 2 + α+ ĝ2 + A+ c03 A+ + α+ ĝ3 ∂τ ẑ+ = −ẑ+ 1 + ẑ+ + α+ ĝ1 + ε+ 4 (8.4) ∂ τ ε+ = ε+ ∂τ α+ = −α+ with ĝ1 = ĝ1 (A+ , ẑ+ , ε+ , α+ ) and ĝ2,3 = ĝ2,3 (A+ , ε+ , α+ ) through a smooth transformation of the form ẑ+ = z+ + f (α+ ) where f (α+ ) = O(|α+ |2 ). 2 ). Proof. We want a coordinate change of the form ẑ+ = z+ + f (α+ ) with f (α+ ) = O(α+ Upon taking the derivative of this expression, we find ∂τ ẑ+ = ∂τ z+ + f 0 (α+ )∂τ α+ 2 =: g(A+ , ẑ+ , ε+ , α+ ) + α+ g̃(α+ , f (α+ )) − α+ f 0 (α+ ) with g(0, 0, 0, α+ ) = 0 for all α+ . We have decomposed the original equation into terms that are functions solely of α+ and those that are not. The terms that are solely functions 2 ) because this is the lowest order that appeared in the original of α+ must be at least O(α+ equation, and our transformation is also of this order. We want to remove them through our choice of f . In this equation, both g and g̃ depend on f (α+ ) which is still unknown. We can determine them by substituting z+ = ẑ+ − f (α+ ) into the right hand side of equation 2 ) terms means solving (8.3). Removing O(α+ 2 α+ g̃(α+ , f (α+ )) = α+ f 0 (α+ ) f 0 (α+ ) = α+ g̃(α+ , f (α+ )) which is an ODE for f with initial condition f (0) = 0. This IVP has a unique solution, 2 ) as claimed. This second statement follows because f (α+ ), and additionally f (α+ ) = O(α+ f 0 (α+ ) = O(α) which is clearly seen from the ODE and f (0) = 0. This f defines our desired 69 transformation. The equation for ẑ+ becomes ∂τ ẑ+ = g(A+ , ẑ+ , ε+ , α+ ) 2 = −ẑ+ − ẑ+ + ε2+ 2 + c03 |A+ |2 + α+ [A+ ĝ1 + ẑ+ ĝ2 + ε+ ĝ3 ] 4 with ĝi (A+ , ẑ+ , ε+ , α+ ) well-defined and smooth. Remember no terms that solely depend 2 [A + ẑ + ε ]) contributions immediately come on α+ appear in g by definition, but O(α+ + + + 2 ) in the original equation. In a slightly different form this is from the O(α+ i hε + 2 2 2 + α+ ĝ2 + A+ c03 A+ + α+ ĝ3 ∂τ ẑ+ = −ẑ+ 1 + ẑ+ + α+ ĝ1 + ε+ 4 with ĝ1 = ĝ1 (A+ , ẑ+ , ε+ , α+ ) and ĝ2,3 = ĝ2,3 (A+ , ε+ , α+ ). As we have changed only the z+ coordinate, and it only appears in higher order terms in the A+ component, we are now done. Our next step is to flatten the unstable manifold of equation (8.4). We want to show this can be done through another transformation solely on the ẑ term. Lemma 8.5. There exists a smooth coordinate change z̃+ = ẑ+ +G(A+ , ε+ ), with G(A+ , ε+ ) = O(|A+ |2 + |ε+ |2 ), in which equation (8.4) becomes ∂ τ A+ = A+ 3 + O (|A+ | + |ε+ | + |α+ | + |z̃+ |) 2 (8.5) ∂τ z̃+ = −z̃+ [1 + O (|A+ | + |ε+ | + |α+ | + |z̃+ |)] ∂ τ ε+ = ε+ ∂τ α+ = −α+ . Proof. We want to flatten out the unstable manifold in order to make (z+ , α+ ) = 0 invariant. We need to show that we can achieve this through transformations only on the 70 ẑ+ term. It follows from (8.4) that the unstable manifold can be written as 0 α+ g1 (A+ , ε+ ) u W = (A+ , ẑ+ , ε+ , α+ ) : = = ẑ+ g2 (A+ , ε+ ) g2 (A+ , ε+ ) because α+ = 0 is invariant which requires Wu ⊂ {α+ = 0}. Here g2 (A+ , ε+ ) = O(|A+ |2 + |ε+ |2 ). Using the coordinate transformation z̃+ = ẑ+ − g2 (A+ , ε+ ) leads to Wu = {(A+ , z̃+ , ε+ , α+ ) : (z̃+ , α+ ) = 0}, which is invariant, and in turn ∂τ z̃+ = ∂τ ẑ+ − ∂A+ g2 (A+ , ε+ )∂τ A+ − ∂ε+ g2 (A+ , ε+ )∂τ ε+ = −z̃+ (1 + O(|A+ | + |ε+ | + |α+ | + |z̃+ |)). This last equality occurs because we made the unstable manifold invariant and no terms of the form g(α+ ) appear from the transformation. The G from above is given by g2 . We are left with system (8.5) after only performing transformations on the z+ term. Set s (ε) is transverse to the α1 = 0, we know that in the real subspace the stable manifold W∞ unstable manifold of p+ emerging from the equilibrium (A1 , z1 ) = (0, 21 + O(ε)) by lemma 8.2. In our flattened coordinates, the unstable direction near the equilibrium p+ becomes s (ε) is a curve transverse to the A+ direction. This then means the stable manifold W∞ the A+ direction but near the equilibrium. By remark 8.3, a whole one parameter family s (ε) can be found by multiplying with eiγ . This proves the following of curves lying in W∞ lemma. 0 Lemma 8.6. Fix ε > 0, ε01 > 0 and define the incoming section Σ+ 1 around p+ as ε+ = ε1 . s (ε) ∩ Σ+ as a We can parameterize the stable manifold intersected with this section W∞ 1 71 0 and γ by A (0) = −eiγ η(ε0 ) + O(|z̃ 0 | + |ε|) and z̃ (0) = −z̃ 0 with 0 < η function of z̃+ + + + + 1 0. and 0 ≤ z̃+ Note the expansion 3 3 η(ε01 ) = c01 (ε01 ) 2 + o((ε01 ) 2 ) (8.6) for some constant c01 > 0: see corollary 3.1 in the appendix for a proof. We want to find τ+ > 0 such that initial data in Σ+ 1 , the incoming section around p+ , lands in the outgoing section Σ+ 2 around p+ given by z̃+ (−τ+ ) = −δ+ . The value of τ+ 0 . This is found by solving equation (8.5). When we solve, we only want will depend on z̃+ to integrate in the decaying direction for the A+ and z̃+ components. This is easy to do because we are solving a boundary value problem and know what each component should be in the relevant sections. Lemma 8.7. Consider equation (8.5) where we have flattened the unstable manifold of 3 1 p+ . There exists ε0 > 0 and x0 > 0 such that for all 0 < ε < ε0 , ε 4 < x < x0 ε− 8 , and initial data of the form 0 A+ (0) = −eiγ η + O(|z̃+ | + |ε|) 1 z̃+ (0) = −xε 4 ε+ (0) = ε01 α+ (0) = ε , ε01 which lies in the stable manifold of (A2 , z2 ) = (0, − 21 + O(ε)), we can solve backward up to 1 4 0 the unique time τ+ (ε, x, η, δ+ ) = − ln( xε δ+ ) + O δ+ + η + ε1 + α0 for which z̃+ (−τ+ ) = 72 −δ+ . This solution is i 3 3 1 + O δ + η + ε0 + α + 0 1 A+ (−τ+ ) = −e η + O(ε ) x 2 ε 8 3 δ+2 h z̃+ (−τ+ ) = −δ+ ε+ (−τ+ ) = 1 4 iγ 1 4 xε ε01 δ+ 1 + O δ+ + η + ε01 + α0 3 α+ (−τ+ ) = ε 4 δ+ 1 + O δ+ + η + ε01 + α0 . 0 xε1 Additionally note that τ+ depends on the value of z̃+ (0). Proof. Solutions of the fixed point equation Rτ 3 0 A+ (τ ) = −eiγ η + O(|z̃+ | + |ε|) e 0 [ 2 +O(A+ +ε+ +α+ +z̃+ )]dσ − z̃+ (τ ) = −δe (8.7) Rτ −τ+ [1+O(A+ +ε+ +α+ +z̃+ )]dσ for functions A+ (τ ) and z̃+ (τ ) are equivalent to solutions of equation (8.5). To prove the theorem, we need to show solutions exist for a range of τ , and then we need to show we can find a τ+ which gives solutions with our desired initial data. We are solving a boundary value problem where our boundary terms lie in the sections ε+ = ε01 and z̃+ = −δ+ . We want to solve from the intersection of the first section, ε+ = ε01 , with the stable manifold of (A2 , z2 ) = (0, − 21 + O(ε)) backwards in time into the second section, z̃+ = −δ+ . Two of the components are explicitly solvable, ε+ and α+ , as functions of τ : ε+ (τ ) = εo1 eτ , α+ (τ ) = This shows |α+ (τ )| ≤ 1 r0 α+ ε+ = ε ε −τ 1 e ≤ . 0 r0 ε1 for τ ≥ −T with −T defined as the time where we match to the core manifold. Define −τ+ as the value of τ when we land in the second section, then 0 −τ+ ≥ ln εr . ε0 1 73 Define the spaces XA = {A+ : A+ (τ ) ∈ C defined and continuous for − τ+ ≤ τ ≤ 0} Xz = {z̃+ : z̃+ (τ ) ∈ C defined and continuous for − τ+ ≤ τ ≤ 0} , equipped with the weighted norms kϕk+ A = sup −τ+ ≤τ ≤0 kϕk+ z = −τ+ ≤τ ≤0 sup 3 e− 2 τ |ϕ(τ )| e−(τ+ −τ ) |ϕ(τ )|. We want to show that the right hand side of the fixed point equation (8.7) is a contraction mapping in these norms. The uniform contraction mapping theorem then gives a smooth solution to these equations. We can streamline the following calculations by noting a few important, but straightforward, properties of these norms. First, we can find pointwise estimates by 3 τ + |φ(τ )| ≤ e 2 τ kφk+ A and |φ(τ )| ≤ e kφkz . Second there is a constant C such that we can estimate the integrals sup Z τ −τ+ ≤τ ≤0 0 |φ(σ)|dσ ≤ Ckφk+ A and sup Z τ −τ+ ≤τ ≤0 −τ+ |φ(σ)|dσ ≤ Ckφk+ z uniformly in τ+ and φ. We start by defining a map F : XA ⊕ Xz 7→ XA ⊕ Xz via the right hand side of equation (8.7) so that Rτ 3 0 [FA (A+ , z̃+ )] (τ ) = −eiγ η + O(|z̃+ | + |ε|) e 0 [ 2 +O(A+ +ε+ +α+ +z̃+ )]dσ [Fz (A+ , z̃+ )] (τ ) = −δ+ e − Rτ −τ+ [1+O(A+ +ε+ +α+ +z̃+ )]dσ 74 Looking at this map in our norms and using the properties above gives kFA (A+ , z̃+ )k+ A = sup −τ+ ≤τ ≤0 Rτ 3 3 0 e− 2 τ − eiγ η + O(|z̃+ | + |ε|) e 0 [ 2 +O(A+ +ε+ +α+ +z̃+ )]dσ ε τ+ + + iγ 0 0 ≤ − e η + O(|z̃+ | + |ε|) 1 + O kA+ kA + ε1 + 0 e + kz̃+ kz ε1 and R − τ [1+O(A+ +ε+ +α+ +z̃+ )]dσ e−(−τ+ −τ ) − δ+ e −τ+ −τ+ ≤τ ≤0 ε τ+ + + 0 ≤ − δ+ 1 + O kA+ kA + ε1 + 0 e + kz̃+ kz . ε1 kFz (A+ , z̃+ )k+ z = sup From the formal analysis, we expect τ+ = O | 14 ln ε| . Additionally, we can choose η(ε01 ) arbitrarily small by varying ε01 : see equation (8.6). For a fixed value of ε, we could find a very large displacement in the z+ direction from this choice of η. Luckily, we also expect 1 the displacement in the z+ direction to vary with ε 4 so by choosing a small enough value s (ε) that for the parameter ε, we can again find some segment of the stable manifold W∞ lies in the valid range for our coordinates. The right hand sides can be made small through an appropriate choice of ε and η. There are two important points to notice. Our solution operator is a well defined continuous operator, and it is Lipschitz with constant less than 1 2. To see this, we take the derivatives of FA and Fz : kDA FA (A+ , z̃+ )(Ã)kA + Z R τ [ 3 +O(Ã+ε+ +α+ +z̃+ )]dσ τ √ iγ 0 ≤ −e η + O(|z̃+ | + |ε|) e 0 2 ÃO( α0 )dσ 0 A ε 0 0 τ+ kÃk+ ≤ C − eiγ η + O(|z̃+ | + |ε|) 1 + O kA+ k+ + kz̃+ k+ z A + ε1 + 0 e A ε1 75 and + Z τ Rτ √ − −τ [1+O(A+ +ε+ +α+ +z̃)]dσ + kDz Fz (A+ , z̃+ )(z̃)kz ≤ δ+ e z̃O( α0 )dσ −τ+ z ε τ+ 0 kz̃k+ + kz̃+ k+ ≤ C − δ+ 1 + O kA+ k+ z z, A + ε1 + 0 e ε1 + where the terms multiplying kÃk+ A and kz̃kz give the size of the derivative and can be made small by the previous calculation. We can conclude that for each large enough choice of τ+ 1 we have a unique solution to the boundary value problem (8.7). 0 , η, δ) that gives our desired initial data must still be found. For The value of τ+ (ε, z̃+ 1 0 = −xε 4 where x is this, we need to know the solution. The formal analysis implies z̃+ some positive quantity bounded away from zero. In backwards time, A+ should decay and z̃+ should grow. We estimate the solution as 0 A+ (τ ) = −eiγ η + O(|z̃+ | + |ε|) e − z̃+ (τ ) = −δ+ e Rτ −τ+ Rτ » 0 » „ 1+O δ+ eσ +εo1 eσ + 3 +O 2 „ δ+ eσ +εo1 eσ + 1 ε −σ e +δ+ e 2 (−τ+ −σ) ε0 1 1 ε −σ e +δ+ e 2 (−τ+ −σ) ε0 1 «– dσ «– dσ . 1 Noting that ε 4 is the leading order term over ε, we can integrate this into h i 3 1 A+ (τ ) = −eiγ η + O(ε 4 ) e 2 τ eh1 (τ ) z̃+ (τ ) = −δ+ e−τ+ e−τ eh2 (τ ) (τ ) with |h1 (τ )| = O(δ+ + η + ε01 + α0 ), |h2 (τ )| = O(δ+ + η + ε01 + α0 ), and | dhdτ2 + | = O(δ+ + η + ε01 + α0 ). Furthermore z̃+ (0) = −δ+ e−τ+ eh2 (0) which enables us to calculate τ+ . Equating our expected initial z+ from the formal analysis 76 to this expression we find 1 0 z̃+ = xε 4 = δ+ e−τ+ eh2 (0) 1 e τ+ xε 4 = =⇒ δ+ eh2 (0) 1 −τ+ = ln( 1 xε 4 xε 4 ) − h2 (0) = ln( ) + O δ+ + η + ε01 + α0 . δ+ δ+ We now have a value for τ+ which both enables us to solve equation (8.5) and gives the right initial data. Now it remains to write down the value of the equation in the outgoing section Σ+ 2 so we can continue this calculation in a different chart. In the end, our solution in Σ+ 2 is h i A+ (τ+ ) = −eiγ η + O(ε ) 1 4 h 1 xε 4 δ+ !3 2 3 eh1 (τ+ )− 2 h2 (0) i 3 3 eh1 (τ+ )− 32 h2 (0) = −e η + O(ε ) x 2 ε 8 3 δ+2 1 4 iγ z̃+ (τ+ ) = −δ+ 1 ε+ (τ+ ) = εo1 eτ+ = εo1 xε 4 δ+ eh2 (0) 3 ε ε 4 δ+ eh2 (0) α+ (τ+ ) = 0 e−τ+ = 0 , x ε1 ε1 which completes the proof. We can invert the z̃+ coordinate to find h i 3 3 1 + O(δ + η + ε0 + α ) 1 + 0 1 iγ A+ (τ+ ) = −e η + O(ε 4 ) x 2 ε 8 3 δ+2 1 z+ (τ+ ) = −δ+ + O(ε 2 ) 1 ε+ (τ+ ) = ε01 xε 4 1 + O(δ+ + η + ε01 + α0 ) δ+ 3 α+ (τ+ ) = ε 4 δ+ 1 + O(δ+ + η + ε01 + α0 ) . 0 ε1 x 77 We are now concerned with transporting the solutions into the valid range for the normal form around the other fixed point. Summary: We have introduced many new constants, and would like to recall their 3 1 expected values here: 0 < ε < ε0 , ε 4 < x < x0 ε− 8 , |h1 (τ )| = O(δ+ + η + ε01 + α0 ), |h2 (τ )| = 3 3 (τ ) O(δ+ + η + ε01 + α0 ), and | dhdτ2 + | = O(δ+ + η + ε01 + α0 ). Note η(ε01 ) = c01 (ε01 ) 2 + o((ε01 ) 2 ). 4.8.3 Connecting P+ to P− We establish a diffeomorphism Π(A1 , z1 , α1 , ε1 ) to understand how the solutions vary be− tween Σ+ 2 given by z̃+ = −δ+ and the third section Σ1 which is approximately given by z1 ≈ − 21 + δ− . The constants δ+ and δ− must be chosen small enough that they are within the range of the normal form coordinates around p+ and p− , respectively. Rather than define our third section as the flat plane z1 = − 12 + δ− for some choice of δ− , we define it as δ− the section that arises by integrating backwards for a fixed time τ0 = − ln (1−δδ++)(1−δ −) from the incoming section. We need to choose δ− small enough such that the solutions we are interested in (those for small enough ε) land in the valid range of the normal form calculations around (0, − 12 ). The diffeomorphism comes from the flow in our original transition chart coordinates: ∂τ A1 = A1 1 + z1 + O |α1 |2 ∂τ z1 = −z12 + ∂τ ε1 = ε1 1 ε21 + + c03 |A1 |2 + O |α1 |2 4 4 ∂τ α1 = −α1 . We are avoiding any fixed points of the flow, so the function defined by flowing initial data in the section z̃+ = −δ+ for time τ0 is well-defined, smooth and invertible: it is a diffeomorphism. When A1 = α1 = ε1 = 0, the z1 axis is invariant. We expand around 78 P+ A1 z1 ! (0, 0) " 1 ,0 2 −η s W∞ s W∞ s Figure 4.6: In this cartoon, we display the expected motion of W∞ from the positive equilibrium backwards in time towards the negative equilibrium. this axis to control the solution growth in the other directions as we move between the sections. The equation on the z1 axis reduces to 1 ∂τ z1 = −z12 + , 4 and this is easily solved as z1 (τ ) = eτ − ec . 2 (eτ + ec ) Next we want the time scale to traverse from z1 = z1 (0) = 1 2 1 2 − δ+ to z1 = − 12 + δ− . Assuming − δ+ we can find the constant in the above equation. We find after some algebra c = ln δ+ 1 − δ+ , and that the time to the next section is given by τ0 = − ln δ+ δ− (1 − δ+ )(1 − δ− ) . 79 This motivates our integration time above. The constants δ± do not depend on ε. We can write down the following expansion for our diffeomorphism around the invariant axis: ag1 (a, z, ε, α) 1 − + δ + zg (a, z, ε, α) − 2 1 2 Π(a, − δ+ + z, ε, α) = 2 εg (a, z, ε, α) 3 αg4 (a, z, ε, α) a(c1 + O(|a| + |z| + |ε| + |α|)) − 1 + δ− + z(c2 + O(|a| + |z| + |ε| + |α|)) 2 = . ε(c + O(|a| + |z| + |ε| + |α|)) 3 α(c4 + O(|a| + |z| + |ε| + |α|)) 3 From corollary 3.2 in the appendix, c1 = δ+2 p δ− (1 + O(δ+ + δ− ))). In the end, we inte- grate for a finite time to move between the normal forms. This does not change the order in ε of any of our solutions, and for small offsets just appears as a scalar change in the distance to the invariant z1 axis. Summary: For small enough ε, the diffeomorphism above will not send any nonzero component to zero or change the sign. It also does not alter the order in ε of any of the components. The constants c1,2,3,4 > 0 are positive. 80 P− z1 P+ A1 ! " 1 − ,0 2 ! (0, 0) s W∞ s W∞ " 1 ,0 2 s W∞ s Figure 4.7: Here the full motion of W∞ between P− (on the left) and P+ (on the right) is shown with the expected linear growth rates around and between the equilibria. Sections 4.8.2, 4.8.3 and 4.8.4 are equivalent to stepping across this picture from right to left. 4.8.4 P− : We are left with initial data in our final coordinate chart: we need to propagate this into the section α1 = 1 r0 and match it to the core manifold. Our goal is to solve 1 2 + z− + O |α− | ∂τ A− = A− 2 ε2 2 ∂ τ z − = z− − z− + − + c03 |A− |2 + |α− |2 O (|A− | + |z− | + |ε− |) 4 (8.8) ∂ τ ε− = ε− ∂τ α− = −α− backwards in time until the matching point r = r0 . The initial data for A− and z− is given by 3 p 1 2 A− (0) = −e η δ+ δ− (1 + O(δ+ + δ− )) + O(|ε| 4 ) iγ h i x 32 ε 38 1 × 1 + O(δ+ + η + ε01 + α0 + |ε| 4 ) 3 δ+2 (8.9) 81 and 1 z− (0) = δ− + O(|ε| 2 ) (8.10) with the constants given in the previous section. Recall this was for a sufficiently small 3 1 0 < ε, 0 < ε01 , ε 4 < x < x0 ε− 8 and 0 < δ± as needed in the proof of Lemma 8.7. We know the time from the initial data in the first chart to the matching time, and the amount of time we have solved between the first three sections to arrive at our present equation. From this information we can easily find how long we should integrate in backwards time: τ− = T − τ+ − τ0 = − ln εr0 + ln ε01 1 4 xε δ+ ! (8.11) + ln δ+ δ− (1 − δ+ )(1 − δ− ) + O δ+ + η + ε01 + α0 3 α0 ε01 xδ− = − ln ε + ln + O δ+ + η + ε01 + α0 . 4 (1 − δ+ )(1 − δ− ) This gives us the initial data in the α− and ε− components: 1 ε− (0) = ε 4 xε01 3 1 δ− 0 eO(δ+ +η+ε1 +α0 ) =: cε ε 4 (1 − δ+ )(1 − δ− ) (8.12) 3 ε 4 (1 − δ+ )(1 − δ− ) O(δ+ +η+ε01 +α0 ) e =: cα ε 4 . α− (0) = 0 δ− xε1 Equation (8.8), unfortunately, is not easily solved for this amount of time. We again circumvent this problem by using normal form coordinates. Lemma 8.8. Using the C ∞ transformation Ã− = A− (1 + O(A− + z− + ε− + α− )) α2 α3 2 z̃− = z− + O(Aα−1 z− ε − ) + O α− (A− + z− + ε− ) (8.13) 82 with P i=1,..,4 αi ≥ 2, equation (8.8) can be recast as 1 2 ∂τ Ã− = Ã− + O |α− | (z̃− + ε− + Ã− ) 2 2 ∂τ z̃− = z̃− + c03 |Ã− |2 + O |α− |2 (Ã2− ε− z̃− + z̃− + Ã4− + ε2− ) (8.14) ∂ τ ε− = ε− ∂τ α− = −α− . Proof. By the Poincare-Dulac theorem (see [2, pgs. 181-184]), we can find a formal transformation to remove non-resonant terms up to arbitrary order. The result in [9, Theorem of Equivalence] states that a C ∞ transformation exists which carries one C ∞ vector field onto another if there is a formal transformation between their corresponding Taylor expansions. For this equation, resonant terms in the A− component are of the m εl αn+m+l for any non-negative n, m and l, while resonant terms of the z form A2n+1 z− − − − − n A2m εl αn+m+l−1 with n + m + l ≥ 1. Also notice that for a component are of the form z− − − − m εn αp then ∂ f (A , z , ε , α ) = O Al z m εn αp + term f (A− , z− , ε− , α− ) = O Al− z− τ − − − − − − − − − − m−1 n p 2 O Al− z− ε− α− ε2− + A2− + α− . Equation (8.14) does not contain several leading order resonant terms. These are resonant terms where α− appears as a linear multiplier. These will not appear from the transformations used to remove the non-resonant terms. The leading order non-resonant terms we want to transform away do not depend on α− . Also, 2 as the leading order. Taking the in the derivatives of A− , z− and ε− , we only see α− derivative of these polynomial terms additionally cannot lower the order in α− as was mentioned above. The addition of any term linear in α− would need to be immediately cancelled by the same term because all other components only change α− at quadratic order. All higher order resonant terms are captured by equation (8.14). For these leading order terms, explicit transformations can be written down to verify this, but they are not sufficient for the claims on the remainders. Now we need to discuss the remainders in equation (8.8) and their effect on the form of the transformation (8.13). The remainders in equation (8.8) are |α− |2 O (|A− | + |z− | + |ε− |). 83 As a result, the set (A− , z− , ε− ) = (0, 0, 0) is invariant. The existence of a C ∞ transformation between our two equations is now guaranteed, but we must still show that it is of the form shown in equation (8.13). Specifically, we want to show that if A− = 0 is invariant, and Ã− = 0 is invariant, then our transformation also leaves A− = 0 invariant. The transformation between these two vector fields is given by [9, Lemma 3.1] wherein a sequence, σi , of transformations is defined which limit to the appropriate coordinate transformation. Denoting our initial and final vector fields as T and T̃ respectively, the sequence is defined in various parts of the space with one of the following transformations: σ1 p = p, T̃ T −1 p, or T̃ −1 T p and σk p = σk−1 p, T̃ σk−1 T −1 p, or T̃ −1 σk−1 T p. First consider σ1 . When it is the identity, it certainly respects the invariance property above. In the other two cases, it is a combination of functions which leave the appropriate set invariant (as the inverse of these vector fields does). This means σ1 has the desired property. The argument for each σk is the same inductively, only now with combinations of three functions that respect the invariance. Then, as each element in the sequence has the desired property, so does their limit and that establishes the desired property. A similar argument can be used about the invariant set (A− , z− , ε− ) = (0, 0, 0) to prove the previous claim, completing the proof. We now transform the initial data into our new coordinates: p 1 Ã− (0) = −e η δ+ δ− (1 + O(δ+ + δ− )) + O(|ε| 4 ) iγ 3 2 h i x 32 ε 83 1 × 1 + O(δ+ + δ− + η + ε01 + α0 + |ε| 4 ) 3 δ+2 1 2 z̃− (0) = δ− + O(δ− + ε 4 ). In the chart P+ we had to solve a boundary value problem in order to land in the second section. To do this, we used a variation of constants formula and showed it was a contraction in an appropriate space. Here we are no longer solving a boundary value 84 problem, but the rest of the computation is very similar. We only integrate in one direction in time for both components here, and we know how long to integrate before we begin. α ε01 xδ− Lemma 8.9. Let τ− = − 43 ln ε + ln (1−δ0+ )(1−δ + O δ+ + η + ε01 + α0 as in equation −) f 0 | ≤ z̃ f and (8.11). There exists a Ãf− > 0, z̃− > 0 and ε0 > 0 such that for |Ã0− | ≤ Ãf− , |z̃− − 0 , ε0 , α0 ) ε ≤ ε0 we can solve equation (8.14) with initial data (Ã− , z̃− , ε− , α− ) = (Ã0− , z̃− − − backwards in time up to time τ for each −τ− ≤ τ ≤ 0. In the section α1 = 1 r0 the solution is 3√ 3√ 7√ 3 0 Ã− (−τ− ) =ε 4 F −1 Â01 + O ε 2 F Â0− ẑ− + ε 4 F Â0− + ε 2 (Â0− )2 F " # 13 3 3 3 −1 0 −1 0 0 2 0 2 0 z̃− (−τ− ) =ε 4 F ẑ1 + ln(cτ ε)F c3 ε 2 |Â− | + O ε 4 (Â− ) ẑ− 4 3 0 2 ) + ε3 (Â0− )4 + ε2 , + O ε 2 (ẑ− where F = F(δ− , δ+ , ε01 , x, r0 ) := 0 α0 ε01 xδ− eO(δ+ +η+ε1 +α0 ) . (1 − δ+ )(1 − δ− ) Proof. This can again be proved by finding a contraction for the variation of constants operator in an appropriate weighted space. The ε dependence of each term in the equation is important when we try to solve. We handle this dependence by factoring ε out before 3 we solve. Using the coordinates Ã− = ε 8 Â− and z̃− = ẑ− , equation (8.14) is rewritten as 3 1 ∂τ Â− = Â− + O |α− |2 Â− (ẑ− + ε− + ε 8 Â− ) 2 3 3 3 2 2 ∂τ ẑ− =ẑ− + c03 ε 4 |Â− |2 + O ε 4 Â2− α− ε− ẑ− + |α− |2 (ẑ− + ε 2 Â4− + ε2− ) ∂τ ε− =ε− ∂τ α− = − α− . We track the solutions from the previous section into the final section at α1 = 1 r0 . First 85 we explicitly solve the equations for α− and ε− as 1 ε− (τ ) = cε ε 4 eτ 3 α− (τ ) = cα ε 4 e−τ ≤ 1 r0 with cα and cε defined as in equation (8.12). The expected decay rates are different around this equilibrium than in P+ ; the norms must reflect this. We use the same spaces as before, XA and Xz , equipped with the new norms kϕk− A := kϕk− z := 1 sup e− 2 σ |ϕ(σ)| τ− ≤σ≤0 sup e−σ |ϕ(σ)| τ− ≤σ≤0 to solve the equations. We want to write down several estimates to speed our calculations. First, we have pointwise estimates 1 τ − |φ(τ )| ≤ e 2 τ kφk− A and |φ(τ )| ≤ e kφkz and there is a constant C, independent of τ− , such that we have the integral estimates Z sup τ τ− ≤τ ≤0 0 |φ(σ)|dσ ≤ Ckφk− A and sup Z τ− ≤τ ≤0 0 τ |φ(σ)|dσ ≤ Ckφk− z. Our approach here is the same as was done around P+ . Restarting from a variation of constants formula, we want the solution to Z 1 3 e 2 (τ −σ) O |α− |2 Â− (ẑ− + ε− + ε 8 Â− ) dσ 0 " # Z τ 3 3 3 2 2 + ε 2 Â4− + ε2− ) + ε 4 Â2− α− ε− ẑ− dσ. ẑ− (τ ) =eτ ẑ10 + eτ −σ c03 ε 4 |Â− |2 + O |α− |2 (ẑ− Â− (τ ) =e 1 τ 2 Â01 + τ 0 Taking these expressions one at a time, we look at their right hand sides in the norms 86 defined above. Starting with the Â− component and recalling the estimate τ ≤ 3 4 ln cτ ε kÂ− kA = ≤ ≤ ≤ ≤ Z t 1 3 sup e |e Â0− | + sup e e 2 (t−σ) O |α− |2 Â− (ẑ− + ε− + ε 8 Â− ) dσ 0 τ ≤t≤0 τ ≤t≤0 Z t 1 1 3 1 e− 2 σ O |α− |2 kÂ− kA e 2 σ (eσ kẑ− kz + ε− + ε 8 e 2 σ kÂ− kA ) dσ |Â0− | + sup τ ≤t≤0 0 Z t O c2α ε 23 e−2σ kÂ− kA (eσ kẑ− kz + cε ε 14 eσ + ε 38 e 12 σ kÂ− kA ) dσ |Â0− | + sup τ ≤t≤0 0 3 7 15 3 |Â0− | + O c2α ε 2 e−τ kÂ− kA kẑ− kz + c2α cε ε 4 e−τ kÂ− kA + c2α ε 8 e− 2 τ kÂ− k2A 3 3 |Â0− | + O c2α ε 4 kÂ− kA kẑ− kz + c2α cε εkÂ− kA + c2α ε 4 kÂ− k2A . − 12 t 1 t 2 − 21 t Now the ẑ− component kẑ− kz = ≤ ≤ ≤ ≤ 0 sup e−t |et ẑ− | τ ≤t≤0 Z t 3 3 3 2 2 et−σ c03 ε 4 |Â− |2 + O |α− |2 (ẑ− + sup e−t ε− ẑ− dσ + ε 2 Â4− + ε2− ) + ε 4 Â2− α− 0 τ ≤t≤0 Z t 3 0 |ẑ− | + sup e−σ c03 ε 4 eσ kÂ− k2A τ ≤t≤0 0 3 3 +O |α− |2 (e2σ kẑ− k2z + ε 2 e2σ kÂ− k4A + ε2− + ε 4 ε− e2σ kÂ− k2A kẑ− kz ) dσ Z t 3 0 c03 ε 4 kÂ− k2A |ẑ− | + sup τ ≤t≤0 0 3 1 2 23 −σ 2 −σ 4 2 −σ 2 +O cα ε (e kẑ− kz + ε 2 e kÂ− kA + cε ε 2 e + cε εkÂ− kA kẑ− kz dσ 3 0 |ẑ− | + c03 ε 4 kÂ− k2A |τ | 3 5 2 23 −τ 2 −τ 4 2 21 −τ 2 2 2 2 +O cα ε (e kẑ− kz + ε e kÂ− kA + cε ε e ) + |τ |cα cε ε kÂ− kA kẑ− kz 3 3 3 0 |ẑ− | + c03 ε 4 kÂ− k2A | ln ε + ln cτ | 4 4 3 9 5 2 2 4 2 45 2 2 4 4 2 +O cα (ε kẑ− kz + ε kÂ− kA + cε ε ) + | ln ε|cα cε ε kÂ− kA kẑ− kz 87 We also need to look at the derivatives of the right hand side, but this comes from the same calculation and can similarly be made small. From this calculation, we see a contraction for sufficiently small ε. We can solve these equations. 1 We also want good estimates for the solution and we find these by using Â− (τ ) = Â0− e 2 τ , 1 3 0 eτ , ε = c ε 4 eσ and α = c ε 4 e−σ : ẑ− (τ ) = ẑ− − ε − α Â− (τ ) = e 1 τ 2 Â01 + Z τ e 1 (τ −σ) 2 0 " 3 8 O |α− | Â− (ẑ− + ε− + ε Â− ) 2 # dσ 3 1 7 1 15 0 − 12 τ = e 2 τ Â01 + O ε 2 Â0− ẑ− e + ε 4 Â0− e− 2 τ + ε 8 (Â0− )2 e−τ # " Z τ 3 3 3 2 2 ε− ẑ− dσ + ε 2 Â4− + ε2− ) + ε 4 Â2− α− ẑ− (τ ) = eτ ẑ10 + eτ −σ c03 ε 4 |Â− |2 + O |α− |2 (ẑ− 0 = eτ ẑ10 + τe τ " 3 c03 ε 4 |Â0− |2 +O ε 5 2 0 (Â0− )2 ẑ− # 3 0 2 + O ε 2 (ẑ− ) + ε3 (Â0− )4 + ε2 . Transforming back into (Ã− , z̃− ) gives 15 3 1 17 9 1 0 − 21 τ Ã− (τ ) = ε 8 e 2 τ Â01 + O ε 8 Â0− ẑ− e + ε 8 Â0− e− 2 τ + ε 4 (Â0− )2 e−τ " # 5 3 τ 0 τ 0 34 0 2 0 2 0 0 2 z̃− (τ ) = e ẑ1 + τ e c3 ε |Â− | + O ε 2 (Â− ) ẑ− + O ε 2 (ẑ− ) + ε3 (Â0− )4 + ε2 . Then at the endpoint, using 3 α0 ε01 xδ− τ− = − ln ε + ln + O δ+ + η + ε01 + α0 , 4 (1 − δ+ )(1 − δ− ) and 3 eτ− = ε− 4 3 0 α0 ε01 xδ− eO(δ+ +η+ε1 +α0 ) =: ε− 4 F(δ− , δ+ , ε01 , x, α0 ), (1 − δ+ )(1 − δ− ) 88 we find 15 3 17 1 9 1 0 21 τ− Ã− (−τ− ) = ε 8 e− 2 τ− Â01 + O ε 8 Â0− ẑ− e + ε 8 Â0− e 2 τ− + ε 4 (Â0− )2 eτ− 3√ 3√ 7√ 3 0 = ε 4 F −1 Â01 + O ε 2 F Â0− ẑ− + ε 4 F Â0− + ε 2 (Â0− )2 F " # 3 5 3 0 0 2 z̃− (−τ− ) =e−τ− ẑ10 − τ− e−τ− c03 ε 4 |Â0− |2 + O ε 2 (Â0− )2 ẑ− + O ε 2 (ẑ− ) + ε3 (Â0− )4 + ε2 # " 13 3 3 0 =ε F −1 ẑ10 + ln(cτ ε)F −1 c03 ε 2 |Â0− |2 + O ε 4 (Â0− )2 ẑ− 4 3 0 2 + O ε 2 (ẑ− ) + ε3 (Â0− )4 + ε2 . 3 4 Our initial data is then transformed into 3 p 1 2 4 = −e η δ+ δ− (1 + O(δ+ + δ− ))) + O(|ε| ) Â01 iγ h × 1 + O(δ+ + δ− + η + 1 ε01 i x 32 + α0 + |ε| ) 3 δ+2 1 4 2 ẑ10 = δ− + O(δ− + |ε| 4 ), and our final data is given by 3 p 1 2 A1 (−τ− ) = −ε e η δ+ δ− (1 + O(δ+ + δ− ))) + O(|ε| 4 ) (8.15) h i 1 x × 1 + O(δ+ + δ− + η + ε01 + α0 + |ε| 4 ) 3 p δ+2 α0 ε01 δ− 3 1 1 1 2 z1 (−τ− ) = − + ε 4 δ− + O(δ− + |ε| 4 ) 1 + O δ+ + δ− + η + ε01 + α0 0 2 α0 xε1 δ− 2 h i x 32 3 1 3 + ln(cτ ε)C1−1 r0 x2 c03 ε 2 −c1 eiγ η + ηO(δ+ + η + ε01 + α0 + |ε| 4 ) 3 4 δ+2 3 4 iγ 89 with C1 = = F α0 x 0 ε01 δ− eO(δ+ +η+ε1 +α0 ) (1 − δ+ )(1 − δ− ) = ε01 δ− 1 + O δ+ + δ− + η + ε01 + α0 Summary: Note that C1 > 0. A plethora of constants have been introduced throughout the chapter, and relevant information about them has been collected in the summaries at the end of each section. Most of these have been defined to be positive, and greater than zero for sufficiently small ε. Everything is now in place to find the intersections between the core and far-field. 4.9 Proof of the main theorem: matching the core To complete the proof of theorem 3, we must solve for intersections between the core manifold and equation (8.15) in terms of the variables (d˜1 , d˜2 ), x and γ as functions of ε. s (ε) can be parameterized by η, γ and x as By applying the lemmas from section 4.8, W∞ 3 1 in equation (8.15). Recall that x0 > 0 and ε 4 < x < x0 ε− 8 . The core manifold, in the appropriate matching coordinates, is given by (5.6) 2 2 2 A1 = ei[−π/4+O(α0 )+Oα0 (|ε| +|d| )] 1 −2 −3/2 ˜ 2 ˜ 2 ˜ ˜ × α0 d1 [1 + O(α0 )] − α0 d2 [i + O(α0 )] + Oα0 (|ε| |d| + |d| ) √ √ ˜ + |d˜2 |2 + |d˜1 |3 ) 1 −d˜2 [i + O(α0 )] − [1/ 3 + O( α0 )]ν d˜21 + Oα0 (|ε|2 |d| z1 = − + . ˜ + |d| ˜ 2) 2 α0 d˜1 [1 + O(α0 )] − d˜2 [i + O(α0 )] + Oα (|ε|2 |d| 0 90 The system of equations we must solve is h i x 3 1 −ε 4 eiγ η 1 + O(δ+ + δ− + η + ε01 + α0 + |ε| 4 ) p 0 ε1 2 2 (9.1) 2 = ei[−π/4+O(α0 )+Oα0 (|ε| +|d| )] i h ˜ + |d| ˜ 2) × d˜1 [1 + O(α0 )] − α0−1 d˜2 [i + O(α0 )] + Oα0 (|ε|2 |d| (9.2) and 3 ε4 = 1 1 0 4 1 + O δ + δ + η + ε + α + |ε| + − 0 1 α0 xε01 √ √ ˜ + |d˜2 |2 + |d˜1 |3 ) −d˜2 [i + O(α0 )] − [1/ 3 + O( α0 )]ν d˜2 + Oα (|ε|2 |d| 1 0 ˜ + |d| ˜ 2) α0 d˜1 [1 + O(α0 )] − d˜2 [i + O(α0 )] + Oα0 (|ε|2 |d| (9.3) . ˜ 2 ) we can remove the phase First, by redefining γ̂ = γ − π/4 + O(r0−2 ) + Or0 (|ε|2 + |d| in (9.1). We also want to rescale as (d˜1 , d˜2 ) = (ε 4 dˆ1 , ε 2 dˆ2 ). We obtain the system 3 3 and 0 = 3 ˆ + |ε| 34 |d| ˆ 2) 0 = dˆ1 [1 + O(α0 )] − ε 4 α0−1 dˆ2 [i + O(α0 )] + Oα0 (|ε|2 |d| h i xη 1 0 4 + (cos(γ̂) + i sin(γ̂)) 1 + O(δ+ + δ− + η + ε1 + α0 + |ε| ) p 0 ε1 1 + O δ+ + δ− + η + ε01 + α0 + |ε| 1 4 " # ˆ2 3 d 2 2 ˆ ) ˆ + |ε| 4 |d| dˆ1 − iε + Oα0 (|ε| |d| α0 3 4 √ √ +xε01 dˆ2 [i + O(α0 )] + [1/ 3 + O( α0 )]νxε01 dˆ21 5 ˆ + |ε| 23 |dˆ2 |2 + |ε| 34 |dˆ1 |3 ). +xε01 Oα0 (|ε| 4 |d| Our goal will be to apply the Implicit Function Theorem to solve this system. By equating 91 the real and imaginary parts of the two equations, we arrive at the following system: 3 ˆ + |ε| 43 |d| ˆ 2) 0 = dˆ1 [1 + O(α0 )] + dˆ2 O(ε 4 ) + Oα0 (|ε|2 |d| h i xη 1 + cos(γ̂) 1 + O(δ+ + δ− + η + ε01 + α0 + |ε| 4 ) p 0 ε1 3 3 ˆ + |ε| 4 |d| ˆ 2) 0 = dˆ1 O(α0 ) − ε 4 α0−1 dˆ2 [1 + O(α0 )] + Oα0 (|ε|2 |d| h i xη 1 + sin(γ̂) 1 + O(δ+ + δ− + η + ε01 + α0 + |ε| 4 ) p 0 ε1 1 0 0 = dˆ1 O δ+ + δ− + η + ε1 + α0 + |ε| 4 # " 3 dˆ 3 1 2 ˆ + |ε| 4 |d| ˆ 2) − 1 + O δ+ + δ− + η + ε01 + α0 + |ε| 4 + Oα0 (|ε|2 |d| ε4 α0 5 ˆ + |ε| 32 |dˆ2 |2 + |ε| 34 |dˆ1 |3 ) +xε01 dˆ2 [1 + O(α0 )] + xε01 Oα0 (|ε| 4 |d| h i 1 3 0 2 2 ˆ ˆ ˆ 4 4 0 = 1 + O δ+ + δ− + η + ε1 + α0 + |ε| d1 + Oα0 (|ε| |d| + |ε| |d| ) 1 dˆ2 O δ+ + δ− + η + ε01 + α0 + |ε| 4 α0 √ √ 0ˆ +xε d2 O(α0 ) + [1/ 3 + O( α0 )]νxε0 dˆ2 3 +ε 4 1 1 1 ˆ + |ε| |dˆ2 |2 + |ε| |dˆ1 |3 ). +xε01 Oα0 (|ε| |d| 5 4 3 2 3 4 For ε = 0 and fixed α0 , ε01 , η and δ± , this becomes xη 0 = dˆ1 [1 + O(α0 )] + cos(γ̂) 1 + O(δ+ + δ− + η + ε01 + α0 ) p 0 ε1 xη 0 = dˆ1 O(α0 ) + sin(γ̂) 1 + O(δ+ + δ− + η + ε01 + α0 ) p 0 ε1 0 = dˆ1 O δ+ + δ− + η + ε0 + α0 + xε0 dˆ2 [1 + O(α0 )] 1 0 = 1 1 + O δ+ + δ− + η + ε01 + α0 3 3 √ √ dˆ1 + xε01 dˆ2 O(α0 ) + [1/ 3 + O( α0 )]νxε01 dˆ21 . Recalling η(ε01 ) = c01 (ε01 ) 2 + o((ε01 ) 2 ), while simultaneously substituting in the new param- 92 eter y = xε01 , simplifies this significantly into 0 = dˆ1 [1 + O(α0 )] + cos(γ̂) 1 + O(δ+ + δ− + ε01 + α0 ) c01 y 0 = dˆ1 O(α0 ) + sin(γ̂) 1 + O(δ+ + δ− + ε01 + α0 ) c01 y 0 = dˆ1 O δ+ + δ− + ε01 + α0 + y dˆ2 [1 + O(α0 )] √ √ 0 = 1 + O δ+ + δ− + ε01 + α0 dˆ1 + y dˆ2 O(α0 ) + [1/ 3 + O( α0 )]νy dˆ21 . (9.4) Letting the constants α0 , δ+ , δ− , ε01 → 0 with c01 > 0 a constant, we obtain 0 = dˆ1 + cos(γ̂)c01 y (9.5) 0 = sin(γ̂)c01 y 0 = y dˆ2 √ 0 = dˆ1 + [1/ 3]νy dˆ21 . A straightforward calculation shows this system is satisfied by y = s√ 3 =: y0 νc01 (9.6) γ̂ = 0 dˆ1 s √ c01 3 = − ν dˆ2 = 0. Call this solution ~y0 . The matrix of partial derivatives for the first three components of (9.5) is fdˆ1 ,γ̂,dˆ2 (~y0 ) = 0 0 1 0 c0 y 0 0 1 0 0 y0 with determinant c01 y02 . There exists a K > 0 such that the solution persists for α0 , δ+ , δ− , ε01 < 93 K. We can repeat this procedure for the full equation to find a solution with estimates s√ 1 √ 3 [1 + O(δ+ + δ− + ε01 + α0 + |ε| 4 )] 0 νc1 1 √ γ̂ = O(δ+ + δ− + ε01 + α0 + |ε| 4 ) s √ 1 √ c01 3 dˆ1 = − [1 + O(δ+ + δ− + ε01 + α0 + |ε| 4 )] ν 1 √ ˆ d2 = O(δ+ + δ− + ε01 + α0 + |ε| 4 ). y = (9.7) Transforming back into the original variables d˜1 and d˜2 , we find d˜1 d˜2 s √ 1 √ c01 3 [1 + O(δ+ + δ− + ε01 + α0 + |µ| 8 )] = −µ ν 3 1 √ = µ 4 O(δ+ + δ− + ε01 + α0 + |µ| 8 ) 3 8 (9.8) completing the proof of theorem 3. 4.10 The breakdown of monotonicity As a consequence of the above proof, we expect the envelope of a spot B solution to be non-monotone. The transition chart captures the algebraic decay rate of the envelope of the solution, so a positive value of z1 corresponds to growth and a negative value to decay. Spot A decreases monotonically from the core out to infinity: in our coordinates, the negative z1 component never changes sign. The ring, on the other hand, grows and then decreases which corresponds to a positive z1 coordinate in the transition chart, but it eventually connects to the exponentially decaying far field. Because spot B passes near to the equilibrium z1 = 1 2 and the ring solution, there is a region where it should have a positive algebraic growth rate. Outside of this region it is near spot A, and should be decaying. By looking at the spot B profile as µ is reduced towards zero, we should be able to see each of these regions. The amplitude of the oscillations should start large and decrease, then there should be a short stretch where the amplitude grows, but this should 94 µ ∈ [0.04, 0.14] |u(r̃)| Decreasing µ r̃ Figure 4.8: The radial rescaling is done with γ = 1/2 corresponding to r̃ = εr. The absolute values of several solutions are plotted for various values of µ ∈ [0.04, 0.14]. Note the change in amplitude of the extrema as µ is varied: the amplitude equation can be visualized through these values. For the largest value of µ, the maximum amplitude is not at the origin. This is an example of the expected breakdown of monotonicity, though we are far from the valid range of our calculations. As µ is decreased, the second largest extremum shrinks while third largest extremum grows. When µ is made small enough, we expect a shift in which is largest. This can be seen in figure 4.9. |u(r̃)| µ = 0.0059 r̃ Figure 4.9: This plot is similar to figure 4.8, only for a value of µ = 0.0059. As is highlighted by the dashed line, there is now a non-monotonicity in the amplitude with respect to r. cease and the amplitudes should decay towards zero. To see this behavior, we first rescale the calculation with the variables r̃ = µγ r. The 95 equation becomes d u = u1 dr̃ d u1 = u2 dr̃ d u2 = u3 dr̃ d n−1 u3 = µ−4γ −(1 + µ)u + νu2 − u3 − 2µ−2γ ( u1 + u2 ) dr̃ r̃ (n − 1)(n − 3) u1 2(n − 1) + ( − u2 ) − u3 r̃2 r̃ r̃ on the interval (0, L) together with the Neumann boundary conditions u1 (0) = 0, u3 (0) = 0, u1 (L) = 0, u3 (L) = 0 (10.1) at r̃ = 0, L. We use the same methods as discussed in chapter 3 to solve it. Normally as the bifurcation parameter is reduced towards zero the support of the solution starts to fill the calculation interval and hits the boundary. At this point, the calculation is no longer valid. Using the unscaled equation would require repeatedly extending the calculation interval then continuing down in µ. By rescaling, the solution stays fixed in the interval and we can easily follow the solution down to small values of µ. The results of these computations are displayed in figures 4.8 and 4.9. They provide another nice validation of our proof. 4.11 Discussion Spot B is essentially spot A and a ring glued together. The maximum amplitude for both 3 √ spot A and the rings scales as µ; this is much smaller than the µ 8 scaling we see for spot B. It is surprising that spot B exhibits this large scaling considering its construction. It also obscured its existence in the work of [22]. The method of proof used in [22] to construct the spot and ring solutions relied on rescaling the equations with the expected amplitude. This scaling can usually be found log(−u(0)) 96 0.102 log µ − 0.171 0.25 log µ + 0.89 log | log µ| − 0.87 Spot B log(µ) Figure 4.10: The amplitude of spot B in three dimensions is shown as a function of µ. The data is insufficient to fully acquire the scaling: both a linear scaling and a linear scaling multiplied by a logarithmic 1 term adequately fit the data. From the formal analysis, we expect the scaling to look like µ 4 (log µ)α for some unknown constant α. through computer experimentation if the solutions have already been numerically found. The spot B solutions do not appear as solutions in those equations, and their existence was unexpected. It would be useful to understand the mechanisms that produce spot B in general and develop a theory to help predict where these odd solutions might arise. We would like to find this solely from formal calculations. The three-dimensional Swift–Hohenberg equation also exhibits spot B solutions, and a similar approach as above should capture them. As mentioned in section 4.6, we lose hyperbolicity around the p+ fixed point which would be an added difficulty in the proof. This, however, provides another example of a piecewise constructed solution with an odd scaling. In the 3D case, the scaling is hard to determine numerically; this is seen in figure 4.10. The Swift–Hohenberg equation for every intermediate value of the dimension n provides further examples. For arbitrary dimension n, the P+ and P− charts are given by z+ = z1 − 1 + n−1 2 and 97 z− = z 1 + n−1 2 respectively. The equations in these charts are 5−n A+ 2 ∂ τ A+ = ∂τ z+ = −z+ and 3−n A− 2 ∂ τ A− = ∂τ z− = z− . From equation (7.6) in the formal analysis, we expect the scaling 5−n d˜1 = µ 2(a−b+1) = µ 8 a 1 for general dimension n. In three dimensions, this would predict a scaling of µ 4 . This is unfortunately a critical value for the formal analysis as b = 0, and logarithmic terms will be important in the leading order analysis. The formal argument in section 4.7 implies the final scaling depends on the eigenvalues around both equilibria in a nontrivial fashion. This is not obviously the case from considering the numerical results. We can extract the scaling extremely well even when the profiles do not exhibit any obvious non-monotonicity. From this, we expect that z1 remains negative throughout the transition region and thus would not expect the p+ equilibrium to crucially affect the scaling. This is clearly not the case, though. In the 3D case, the formal argument is invalid because of the loss of hyperbolicity. It is likely that the scaling observed actually involves logarithmic terms multiplying the decay rate, but this is difficult to see numerically and requires involving higher order terms analytically. Chapter Five Conclusion 99 5.1 Summary of main results In this thesis, I have explored the bifurcation structure of several radially symmetric stationary families of solutions to the Swift–Hohenberg equation for various spatial dimensions. I discovered that the bifurcation curves of these solutions for dimension greater than one do not appear to snake. Instead, a convoluted family of curves and bifurcations is seen that connects the well understood one dimensional snaking case to the planar and spherical cases. The two-pulses, living on isolas in one dimension, become rings in higher dimensions. The isolas connect with the continuous snaking curves of one-pulses and eventually form snaking regions in higher dimensions. This inquiry exposed the existence of a hitherto unknown family of spot solutions, spot B. The maximum value of these solutions 3 expressed an unexpected scaling, µ 8 , with respect to the bifurcation parameter as µ → 0. I rigorously proved the existence of spot B in two dimensions using a novel proof which recovers the unexpected scaling of the spot solution. This proof relies on the existence of spot A and the rings as was established in [22]. The scaling and existence region of spot B arises naturally from the proof, and I expect similar solutions can be found in other systems. I have also proved the existence of spot A in three dimensions, using the same analytic framework as in [22]. This is not included in this thesis as I expect a simpler and cleaner proof can be constructed using the techniques from the spot B proof. The 3D rings should also be straightforward to establish in this setting, but additional difficulties may arise for proving the existence of spot B in three dimensions. 5.2 Open questions Very little has been proven rigorously about snaking and the bifurcation curves seen in chapter 3. Due to the vast number of bifurcations, and the sensitivity to the phase of the solutions with respect to the radial variable, it seems difficult to make any progress. 100 Many questions relating to the stability of the spots and rings are completely unanswered. Not even the linearized spectrum of the planar and spherical solutions has been computed. Unstable eigenvalues are expected to exist, such as the instability that causes hexagons to bifurcate from spots, but it would be interesting to know how many unstable eigenvalues there were, and what they correspond to. The easiest first step would be a numerical study of the radially symmetric spectrum. The radially symmetric problem is one dimensional and many techniques exist to study it both numerically and analytically. It is more interesting to study the stability with respect to arbitrary perturbations. This also could be done numerically, or rigorously using the following approach. In order to study the spectrum analytically, we would let us denote either a radially symmetric ring or spot solution which has already been shown to exist. By considering an ansatz of the form u(r, φ, t) = us (r) + u(r)eλt+ilφ , we would then plug this into the governing equations and solve for each λ(l). A similar analysis using the coordinates from the spot B proof should illuminate the spectrum for arbitrary perturbations. The previous analytic approach in [22] required rescaling the solutions with their expected amplitude as a function of µ. As a result, it failed to reveal the spot B solutions. I expect similarly constructed solutions may be missed in other settings because of their odd scalings. A formal framework wherein these scalings could be found a priori would be extremely useful in the general existence theory, however this question remains completely open. Studying several other systems might help illuminate these issues further: these systems are the 3D Swift–Hohenberg equation, the forced complex Ginzburg–Landau equation, and a coupled Turing–Hopf system. Multi-pulses, such as hexagon patches, are both interesting and ubiquitous. Because the solutions are well localized, there is hope that multi-pulses can be constructed by gluing spots and rings together. In three dimensions, it would be interesting to find spots concentrated on a crystal lattice structure. This work would require a detailed analysis 101 in the tails of these spot solutions. A general approach to produce multi-pulses in two dimensions has been introduced in [41]. The study of pattern formation in the Swift–Hohenberg equation poses many interesting and important questions. The results apply to a host of planar problems where Turing bifurcations are the source of the interesting dynamics. The Swift–Hohenberg equation is also one of the simplest model systems to study snaking, and could illuminate a great deal about the general pattern forming process. Appendix A Proofs for Asymptotics 103 A.1 The general approach Lemma 1.1. Fix a real constant a > 0. Consider the equation u0 = u (a + O(u + f (x))) . (1.1) Let the function f (s) satisfy the condition Z −∞ 0 |f (s)|ds ≤ C. There is a δ > 0 such that for any initial data |u0 | < δ the equation (1.1) has solution u(x) = ceax (1 + O() + o(eax )) (1.2) for x ∈ [−∞, 0]. Proof. Let x ∈ (−∞, 0], then equation (1.1) is satisfied by the solution to the fixed point equation u(x) = u(0)e Rx 0 (1.3) a+O(u(s)+f (s))ds = u(0)eax e Rx 0 O(u(s)+f (s))ds . Defining the function v(x) := e−ax u(x), this then becomes v(x) = v(0)e Rx 0 O(eas v(s)+f (s))ds . We want to solve for smooth and bounded v on the interval (−∞, 0] with the uniform norm 104 kvk∞ : v(x) = v(0)e ≤ v(0)e Rx 0 Rx 0 O(eas v(s)+f (s))ds O(eas kvk∞ +f (s))ds ≤ v(0)eO(kvk∞ )+ Rx 0 O(f (s))ds . In the uniform norm, kvk∞ ≤ v(0)eO(kvk∞ )+O() which has a solution by the contraction mapping principle with the estimate v(x) = v(0) (1 + O(kvk∞ + )) = v(0) + o(v(0)) Transforming back into u(x) finishes the proof of the lemma. A.2 Decay estimates for the ring in the far field Recall equation (8.1) from chapter 4 as a a + 2 + + c03 |a|2 a. s 4s 4 (2.1) as a a + 2+ s 4s 4 (2.2) e−s/2 es/2 a(s) = cs √ + cu √ . s s (2.3) ass = − The simplified equation ass = − has solutions 105 s , we must have c 6= 0 and c = 0. Our goal is Clearly, for a nontrivial solution in W∞ s u to show equation (2.1) has decaying solutions that are to leading order the same as the decaying solutions of (2.3). We already know this equation has a nontrivial exponentially decaying solution with exponentially decaying derivative by [22, Lemma 4]. Lemma 2.1. There is a constant cs 6= 0 such that e−s/2 e−s/2 qn (s) = cs √ + o( √ ) s s as s → ∞. Proof. The simplified equation (2.2) has an exponential dichotomy. By projecting onto the stable part, as we already know the solution and its derivative decay exponentially, we have the original exponential dichotomy with a small perturbation. The solution then persists to leading order. We could also proceed more directly and recast this as a first q s order system using the approach from [22, Chapter 4]. In the variables â = µ A and b̂ = √ s µ B this becomes âs = b̂ b̂s = â + c03 |â|2 â. 4 By using weighted norms that track the expected decay rate − 12 for the stable solution, the c03 |â|2 â term can be seen as a small perturbation. Using the variation of constants formula to express the solution, we can use the contraction mapping principle to find our solutions. 106 A.3 A.3.1 Corollaries for different f (s) Corollary 1: η(ε01 ) Corollary 3.1. Our initial data in the transition chart (η(ε01 ), ε01 ) satisfies 3 3 η(ε01 ) = c01 (ε01 ) 2 + o((ε01 ) 2 ) for a constant c01 > 0. s (ε) transversely intersects the unstable manifold of p , we can find Proof. Because W∞ + this relation from the solution to equation (8.5) when α+ = z̃+ = 0 and using ε+ = cε eτ : ∂ τ A+ = A+ 3 τ + O (|A+ | + e ) . 2 By lemma 1.1, this has solution 3 3 A+ (τ ) = c01 ε+2 + o(ε+2 ). s (ε) near its transverse intersection with the In order for us to choose initial data in W∞ unstable manifold of p+ , it must respect this scaling. A.3.2 Corollary 2: c1 (δ− , δ+ ) Corollary 3.2. There exists 0 such that for ε < 0 the constant c1 (δ− , δ+ ) is given by 3 c1 (δ− , δ+ ) = δ+2 p δ− (1 + O(δ+ + δ− ))) . Proof. Here, we just want to solve equation (5.5) from initial data of order (A1 , z1 , α1 , ε1 ) = 3 1 3 1 O(ε 8 ), −δ+ + O(ε 2 ), O(ε 4 ), O(ε 4 ) . We can find a solution to this from a proof similar 107 to that for lemma 1.1. Recall c = ln τ0 = − ln δ+ 1 − δ+ δ+ δ− (1 − δ+ )(1 − δ− ) and for (A1 , α1 , ε1 ) = (0, 0, 0) z1 (τ ) = eτ − ec 2(eτ + ec ) for τ ∈ [−τ0 , 0]. Consider then, with τ0 > 0, c1 Z 1 es − ec 2 = Exp 1+ + O(|ε| )ds 2 (es + ec ) 0 ! −τ0 1 1 + τ0 O(|ε| 2 ) = e−τ0 Exp − s + log (−2(es + ec )) 2 0 −τ0 1 e−τ0 + ec τ0 O(|ε| 12 ) = e − 2 τ0 e 1 + ec s 1 δ+ δ− δ+ δ− = δ+ + eτ0 O(|ε| 2 ) (1 − δ+ )(1 − δ− ) 1 − δ− 3p = δ+2 δ− (1 + O(δ+ + δ− ))) . Bibliography [1] M. Abramowitz and I. A. Stegun. Handbook of mathematical functions with formulas, graphs, and mathematical tables. New York: Dover, 1972. [2] V. I. Arnol0 d. Geometrical methods in the theory of ordinary differential equations, volume 250 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Science]. Springer-Verlag, New York, 1983. Translated from the Russian by Joseph Szücs, Translation edited by Mark Levi. [3] D. Avitabile, D. J. B. Lloyd, J. Burke, E. Knobloch, and B. Sandstede. To snake or not to snake in the planar Swift–Hohenberg equation. SIAM J. Appl. Dynam. Syst., 9:704–733, 2010. [4] M. Beck, J. Knobloch, D. J. B. Lloyd, B. Sandstede, and T. Wagenknecht. Snakes, ladders, and isolas of localized patterns. SIAM J. Math. Anal., 41(3):936–972, 2009. [5] J. Burke and E. Knobloch. Localized states in the generalized Swift-Hohenberg equation. Phys. Rev. E, 73(5):056211, 2006. [6] J. Burke and E. Knobloch. Homoclinic snaking: structure and stability. Chaos, 17(3):037102, 2007. [7] J. Burke and E. Knobloch. Snakes and ladders: localized states in the Swift-Hohenberg equation. Phys. Lett. A, 360(6):681–688, 2007. [8] S. Chapman and G. Kozyreff. Exponential asymptotics of localised patterns and snaking bifurcation diagrams. Phys. D, 238(3):319 – 354, 2009. [9] K.-T. Chen. Equivalence and decomposition of vector fields about an elementary critical point. Am. J. Math., 85(4):pp. 693–722, 1963. [10] C. Chicone. Ordinary differential equations with applications, volume 34 of Texts in Applied Mathematics. Springer-Verlag, New York, 1999. [11] P. Coullet, C. Riera, and C. Tresser. Stable static localized structures in one dimension. Phys. Rev. Lett., 84(14):3069–3072, 2000. [12] M. Cross and P. Hohenberg. Pattern formation outside of equilibrium. Rev. Mod. Phys., 65(3):851–1112, 1993. [13] J. H. P. Dawes. The emergence of a coherent structure for coherent structures: localized states in nonlinear systems. Phil. Trans. R. Soc. A, 368(1924):3519–3534, 2010. 108 109 [14] E. J. Doedel and B. Oldeman. auto07p: continuation and bifurcation software for ordinary differential equations. Technical report, Concordia University, 2009. [15] C. Elphick, E. Tirapegui, M. E. Brachet, P. Coullet, and G. Iooss. A simple global characterization for normal forms of singular vector fields. Phys. D, 29(1-2):95–127, 1987. [16] P. C. Fife. Pattern formation in gradient systems. In B. Fiedler, editor, Handbook of Dynamical Systems 2, pages 679–719. North-Holland, Amsterdam, 2002. [17] C. Gollwitzer, I. Rehberg, and R. Richter. Via hexagons to squares in ferrofluids: experiments on hysteretic surface transformations under variation of the normal magnetic field. J. Phys.: Condensed Matter, 18(38):S2643, 2006. [18] J. Guckenheimer and P. Holmes. Nonlinear oscillations, dynamical systems, and bifurcations of vector fields, volume 42 of Applied Mathematical Sciences. Springer-Verlag, New York, 1983. [19] R. B. Hoyle. Pattern Formation. Cambridge University Press, Cambridge, 2006. [20] E. Knobloch. Spatially localized structures in dissipative systems: open problems. Nonlinearity, 21(4):T45–T60, 2008. [21] J. Knobloch, D. Lloyd, B. Sandstede, and T. Wagenknecht. Isolas of 2-pulse solutions in homoclinic snaking scenarios. J. Dynam. Diff. Eqns., 23:93–114, 2011. 10.1007/s10884-010-9195-9. [22] D. J. B. Lloyd and B. Sandstede. Localized radial solutions of the Swift–Hohenberg equation. Nonlinearity, 22(2):485–524, 2009. [23] D. J. B. Lloyd, B. Sandstede, D. Avitabile, and A. R. Champneys. Localized hexagon patterns of the planar Swift–Hohenberg equation. SIAM J. Appl. Dynam. Syst., 7:1049–1100, 2008. [24] S. McCalla and B. Sandstede. Snaking of radial solutions of the multi-dimensional Swift–Hohenberg equation: A numerical study. Phys. D, 239(16):1581 – 1592, 2010. [25] S. McCalla and B. Sandstede. Existence of 2D and 3D spots in the stationary radial Swift-Hohenberg equation. In preparation, 2011. [26] J. D. Murray. Mathematical biology. II, volume 18 of Interdisciplinary Applied Mathematics. Springer-Verlag, New York, third edition, 2003. Spatial models and biomedical applications. [27] NIST. http://dlmf.nist.gov/7.5, accessed on 15 March 2011. [28] L. M. Pismen. Patterns and Interfaces in Dissipative Dynamics. Springer Verlag, Berlin, 2006. [29] H.-G. Purwins. http://www.uni-muenster.de/Physik.AP/Purwins/DC/indexen.html, accessed on 14 October 2010. [30] M. I. Rabinovich, A. B. Ezersky, and P. D. Weidman. The Dynamics of Patterns. World Scientific Publishing, River Edge, 2000. [31] R. Richter and I. V. Barashenkov. Two-dimensional solitons on the surface of magnetic fluids. Phys. Rev. Lett., 94(18):184503, May 2005. 110 [32] B. Sandstede. Stability of travelling waves. In Handbook of dynamical systems, Vol. 2, pages 983–1055. North-Holland, Amsterdam, 2002. [33] A. Scheel. Subcritical bifurcation to infinitely many rotating waves. J. Math. Anal. Appl., 215(1):252–261, 1997. [34] A. Scheel. Radially symmetric patterns of reaction-diffusion systems. Mem. Amer. Math. Soc., 165(786), 2003. [35] E. Sheffer, H. Yizhaq, E. Gilad, M. Shachak, and E. Meron. Why do plants in resource-deprived environments form rings? Ecological Complexity, 4(4):192 – 200, 2007. [36] J. Swift and P. C. Hohenberg. Hydrodynamic fluctuations at the convective instability. Phys. Rev. A, 15(1):319–328, 1977. [37] A. M. Turing. The chemical basis of morphogenesis. Phil. Trans. R. Soc. London B, 237(641):37–72, 1952. [38] G. H. M. van der Heijden, A. R. Champneys, and J. M. T. Thompson. Spatially complex localisation in twisted elastic rods constrained to a cylinder. Internat. J. Solids Structures, 39:1863–1883, 2002. [39] M. K. Wadee, C. D. Coman, and A. P. Bassom. Solitary wave interaction phenomena in the strut buckling model incorporating restabilisation. Phys. D, 163(1-2):26–48, 2002. [40] P. D. Woods and A. R. Champneys. Heteroclinic tangles and homoclinic snaking in the unfolding of a degenerate reversible Hamiltonian-Hopf bifurcation. Phys. D, 129(3-4):147–170, 1999. [41] S. Zelik and A. Mielke. Multi-pulse evolution and space-time chaos in dissipative systems. Mem. Amer. Math. Soc., 198(925):vi+97, 2009.