MODULATIONAL STABILITY OF MULTI-PULSES WITHIN THE FUNCTIONALIZED CAHN-HILLIARD GRADIENT FLOW By Hayriye G¨u¸ckır C¸ akır A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Mathematics – Doctor of Philosophy 2019 ABSTRACT MODULATIONAL STABILITY OF MULTI-PULSES WITHIN THE FUNCTIONALIZED CAHN-HILLIARD GRADIENT FLOW By Hayriye G¨u¸ckır C¸ akır The Functionalized Cahn-Hilliard (FCH) energy is a model describing the interfacial energy in a phase separated mixture of amphiphilic molecules and a solvent. On a bounded domain in R, the Euler-Lagrange equation for the mass constrained Functionalized Cahn- Hilliard(FCH) free energy with zero functionalization terms is derived and a large family of multi-pulse critical points is constructed. We show that the FCH energy with no functiona- lization terms subject to a mass constraint has global minimizers over a variety of admissible sets. We introduce a multi-pulse ansatz as the extensions of the periodic multi-pulse critical points to R and establish the H2-coercivity of the second variation of the energy about multi-pulse ansatz. Modulational stability and the dynamic evolution of the multi-pulse ansatz with respect to the Π0-gradient flow are also addressed. ACKNOWLEDGMENTS I would like to express my deepest appreciation and gratitude to my advisor Dr. Keith Promislow for his continuous personal and professional support throughout my years at MSU. I would like to thank him for his contributions to my knowledge on partial differential equations and applied mathematics with several courses he taught and useful discussions during our regular meetings in spite of his busy schedule. The completion of this dissertation would not be possible without his guidance, extensive knowledge and endless patience. I would also like to thank my committee members Dr. Jeffrey Schenker, Dr. Russell Schwab and Dr. Zhengfang Zhou for their helpful feedback and valuable time. No words would be enough to express my gratitude to my parents and my sisters, Hale and Esra, for their belief in me and unconditional support to pursue my dreams. I am deeply indebted to my husband, Firat. Thank you for sharing my dreams and walking this journey with me with joy for the last 14 years. Last but not least, I am grateful to my children, Miran and Nehir, who have given me happiness and power keeping me motivated to complete this work. iii TABLE OF CONTENTS LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 Functionalized Cahn-Hilliard Free Energy . . . . . . . . . . . . . . . . . . . . 1.2 Description of the Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3 Main Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v 1 1 4 8 Chapter 2 The Euler-Lagrange Equation . . . . . . . . . . . . . . . . . . . . 2.1 Derivation of the Euler-Lagrange Equation over Various Admissible Sets . . 2.2 The Euler-Lagrange Equation with a small mass constraint . . . . . . . . . . 2.2.1 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.2 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.4 2.2.5 Energy values at the critical points . . . . . . . . . . . . . . . . . . . 11 12 24 24 27 Solutions to the Euler-Lagrange Equation with a small mass constraint 29 Some Remarks 32 33 Chapter 3 Existence Of the Minimizers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1 Existence of the Minimizers Introduction: n-pulses Chapter 4 Modulational Stability of n-pulses . . . . . . . . . . . . . . . . . . 4.1 H2-coercivity of the second variation of I . . . . . . . . . . . . . . . . . . . . 4.1.1 . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1.2 H2-coercivity of the second variation of I about n-pulse ansatz . . . . 4.1.3 H2-coercivity of L2 n . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Modulational Stability of n-Pulses . . . . . . . . . . . . . . . . . . . . . . . . 4.3 Pulse Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 37 46 46 47 49 59 65 77 87 APPENDIX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 iv LIST OF FIGURES Figure 1.1: When the hydrocarbons in different shapes introduced into a solvent, the solvent particles create a cavity to avoid the solute(left)[Wiebe et al., 2012]. The simulation depicts packing of amphiphilic molecule at interface between external solvent molecules and internal solvent(right). . . . . . . . . . . . Figure 1.2: Figure 1.2a is the phase plane for the ODE (1.2.6) which demonstrates the orbit (solid line) homoclinic to b− attaining its maxima at UM and the periodic solutions (dotted lines) that are the solutions to the boundary value problem(1.2.5). Figure 1.2b is an example of an n = 4 pulse solution to (1.2.5) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (cid:16) L2 (cid:17) ∈ W s((cid:126)b) where (cid:126)b = (b, 0, 0, 0) mimic the behavior of the Figure 2.1: Figure 2.1a represents a phase plane for the dynamical system given in (2.2.2) extended to R for λ = 0. Figure 2.1b is a phase plane for the dynamical system when λ (cid:54)= 0. The boundary conditions (cid:126)u(0) ∈ W u((cid:126)b) and (cid:126)u whole line system. The distance between the fixed points of the dynamical systems is |b − b−| = ε while |u(0) − b| = O(e−κ L2 ε ) (cid:28) ε. However, for the clarity of the graph both distances are depicted in similar lengths. . . ε 2 7 30 36 Figure 2.2: The reduced energy, I(Φn), versus mass constraint M . The blue lines demonstrate the energy values at the critical points, Φn, for n = 0, 1 2, 2. The red lines represent the infimum of the energy values over all the blue lines. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.1: Figure (a) is a depiction for the spectrum of Ln(p). In the descending order λ0 > 0 = λ1 > λ2 > ... (red disks) are the eigenvalues of L1(p) = L. Ln(p) has n associated eigenvalues(black crosses) to each localized eigenvalue of L1(p) such that |λk−λk,j|j=1,...,n decays exponentially with growing pulse separation. Figure (b) demonstrates the spectrum for L2 2 , 1, 3 n(p). 53 v Chapter 1 Introduction 1.1 Functionalized Cahn-Hilliard Free Energy Amphiphilic molecules are chemical compounds consisting of a hydrophobic group and a hydrophilic group, such as lipids and surfactants. When a molecule with an amphiphilic structure is introduced to a solvent(water), the hydrophobic group may alter the structure of the solvent which causes an increase in the free energy of the system. As a respond to this change the system minimizes the contact between the hydrophobic group with formation of bilayer interfaces and pores. Network formations differ from single layer interfaces that occur in binary metals and other purely hydrophobic blends. While single layer interfaces separate two distinct phases from each other, bilayers separate one phase by thin sheets of other phase. The Cahn-Hilliard(CH) free energy has been used broadly to model single layer interfaces in hydrophobic blends. In 1958, the CH free energy was proposed as (cid:90) Ω E(u) = |(∇u)|2 + W (u)dx, ε2 2 (1.1.1) in [Cahn and Hilliard, 1958] to describe the free energy of an interface occurred by a phase separation in a binary mixture due to spinodal decomposition. Here on a fixed domain Ω ⊂ Rn, u ∈ H1(Ω) is the volume fraction of one of the components in the binary mixture, 1 W : R → R is the free energy density of the mixture and ε is the thickness of the interface. Figure 1.1: When the hydrocarbons in different shapes introduced into a solvent, the solvent particles create a cavity to avoid the solute(left)[Wiebe et al., 2012]. The simulation depicts packing of amphiphilic molecule at interface between external solvent molecules and internal solvent(right). Since CH was introduced, its minimizers, minimizers subject to a constraint and critical points of the CH have been broadly studied. Although a double well with two unequal depth minima is generic, W (u) is assumed to be a smooth double-well potential with two equal depth minima at b±,i.e, W (b−) = W (b+) and a maxima at b− < b0 < b+ in most studies. This assumption on the form of the potential does not affect the following minimization (cid:27) (u − b−)dx = c (cid:90) Ω , (1.1.2) problem with a mass constraint (cid:26)(cid:90) min Ω |(∇u)|2 + W (u)dx : u ∈ H1(Ω), ε2 2 2 Bdx.doi.org/10.1021/jp209088u|J.Phys.Chem.CXXXX,XXX,000–000TheJournalofPhysicalChemistryCARTICLE2.MODELThepartialmolarvolume,VX,ofasolute,X,inagivensolvent,S,isthederivativeoftheoverallvolume,V,withrespecttothenumberofmolesofsolute,nX.VX¼∂V∂nXP,Tð2ÞFordilutesolutions,thiscanbereplacedwithafinite-differenceexpressionVX¼VðnS,nXÞVðnS,0ÞnXð3ÞThatis,VXisdefinedbythedifferencebetweenthevolumeofsolutionandthevolumeofpuresolventdividedbynX.Weproposetouseconstant-pressuremoleculardynamics(MD)simulationstoobtainthisdifferenceforasinglesolutemoleculeX(nX=1/NA;NAistheAvogadroconstant)andanN-particlesolventN3S(nS=N/NA).ThevolumeofasingleparticleV(X)isthengivenbyVðXÞ¼VðXþN3SÞVðN3SÞð4ÞandVX=NA3V(X).Thismethodmimicstheexperimentalprocedureofdeterminingpartialmolarvolumesofstablecom-poundsandisequallyapplicabletocalculationofpartialmolarvolumesofshort-livedTSspecies,forwhichsuchexperimentalprocedureisnotfeasible.Figure1showsfourhydrocarbonsimmersedinamodelsolvent.Ineachcase,thesolventtrajectorydemonstratesaclearpatternofavoidanceduetotheshort-rangesolventsoluterepulsion,whichresultsintheformationofacavityaroundthesolute.Theincrementalincreaseinthevolumeoftheoverallsystemduetocavityformationisdescribedbyeq4.Thesizeandtheshapeofthecavitydependonthegeometryofthesolute,thestrengthandtypeofthesolutesolventinteractions,aswellastemperatureandpressure.Volumefluctuationsinaconstant-pressureMDrunarequitesignificantandexceedthevalueofanincrementalvolumeincreaseduetoasinglemolecule.(SeeFigure2.)However,theaveragevolumefluctuatesmuchless,andthestandarderrorofthecumulativeaveragedecreasesasthesquarerootofthelengthofthetrajectory.AscanbeseeninFigure3,theerrorscanbereducedtoanacceptablelevelifanMDrunissufficientlylong.3.COMPUTATIONALDETAILSMDcalculationswereperformedusingtheGROMACSpackage9forasystemof256solventmoleculesinacubicboxwithperiodicboundaryconditions.Thesystemwasmaintainedataconstantpressureof0.1MPaandaconstanttemperaturematchingexperimentalconditionsusingBerendsentemperatureandpressurecoupling.10TheMDtrajectorieswereobtainedusingleapfrogintegrationwith1fstimestepwithinteractioncutoffradiusFigure1.Hydrocarbonsimmersedinamodelsolvent(clockwisefromthetopleftcorner):hexane,cyclopentadiene,benzene,andtoluene.Thesolventtrajectoryisrepresentedbyanoverlayofsolventconfigurationsacquiredatdifferentinstantsoftime.Solventparticles(white)avoidthesolute,thusformingacavityofthematchingsizeandshape.Figure2.InstantaneousandaverageMDvolumesofasystemof256cyclopentadienemolecules.Theincrementalcontributionfromasinglemoleculeestimatedbytheexperimentalmolarvolumeof82cm3/molisca.0.14nm3.Largeamplitudefluctuationsoftheinstantaneousvolumearesomewhatstabilizedbyaveragingover10psintervals.Furtherimprovementisreachedbyusingacumulativeaverage.Chem.Soc.Rev.Thisjournalis©TheRoyalSocietyofChemistry2018atomisticforcefieldsaremostsuitableforsimulatingtheself-assemblyofbio-inspiredmolecules.Comparativestudiesaresparseinthiscontext.AnexhaustivesetofMDsimulationsonthefoldingoftetra-andpentapeptides,performedusingCHARMM,AMBER,aswellasOPLS,showedclearinconsistenciesbetweentheseforcefields.26Asimilarobservationwasmadeinastudyonthedimerizationpropensityofaminoacidsidechainresidues.27Packingmotifs.Keepingtheaboveconsiderationsinmind,AAsimulationshavecontributedalottowardourunderstandingofthepackingmotifsofsupramolecularassemblies.Typicalsystemsetupsrangefromsmall-scalestudiesonmonomer–monomerinteractions,tolargerscaleself-assemblysimulationsandthestudyofpre-formedaggregates.Atthesmallestscale,i.e.monomersanddimers,AAsimulationsaremostlyusedtoimproveorvalidatetheunderlyingforcefield.Forinstance,Bejagametal.developedanAAmodelfor1,3,5-benzenetricarboxamide(BTA)usingQMcalculationsofgasphasedimersasareference.28TheenergyandstructureoftheBTAdimerattheQMlevelcouldbecapturedbytheAAmodel.TheAAmodelwasthenusedtostudydimerizationinnonane,andadimerfreeenergyof13kcalmol1wasobtained.Attachmenttoapre-constructedfibrewasfoundtobemorefavourable,demonstratingcoopera-tivityoftheprocess.Anotherniceexampleofforcefielddevel-opmentandvalidationistheworkofSasselietal.29TheauthorsdescribeasystematicparameterizationoftheFmocmoietyusingCHARMM,basedonQMdata(monomerinsolvent,dimer,torsionalpotentials)aswellasexperimentaloctanol/waterparti-tioningcoefficients.Validationofthemodelcamefromself-assemblyofdifferenttypesofFmoc-peptidesinwater,oneformingfibres,theothersphericalaggregates,bothinagree-mentwithexperiment.Inadditiontoforcefielddevelopment,smallscalesimulationscanbeveryusefultodetectpossiblebindingandstackingmodesofthebuildingblocks.30However,thequestionalwaysremainswhetherthesemodesarerepresen-tativefortheself-assembledstate.Initialself-assemblypathways.Theessenceofsupramolecularsystemsisthattheyformviaself-assembly.TocapturethisprocessinAAsimulationsisnottrivial,giventhelimitsintimeandlengthscalesthatcanbeachieved.Thetimescalesexploredareusuallylimitedtothenanosecondrange,withsystemsizestypicallycontainingafew100monomers.MostoftheAAself-assemblysimulationsreportedtodatearethereforenotleadingtowell-definedstructures,butoftenresultintheformationofsmallirregularaggregates.Despitethiscaveat,self-assemblysimulationsareimportantforshowingpossiblestackingmotifsinsidetheaggregateandrevealingthedrivingforcesoftheinitialgrowthprocess.Quiteanumberofstudiesdealwiththeself-assemblyofshortpeptides31–35orpeptideconjugates.36–38Alreadysuchsimplesystemsgiverisetoarichbehaviourintermsoftheinitialself-assemblykinetics,exem-plifiedbytheformationofunexpectedalpha-helicalinter-mediatestructuresofbeta-fibrilformingpeptides,31andthestrongeffectofpeptiderigidityontheorderinsidetheaggre-gatesformed.33Tofurthercomplicatetheissue,thedrivingforcesappearhighlysystem-dependent.Thep–p-stackingwasfoundtobeimportantindrivingtheinitialself-assemblyofpeptide–drugconjugatesintofibres,36whereasfibreformationintheabsenceofaromaticresidueswasobservedforothersmallpeptides.32OneofthemostinterestingAAstudiesonself-assemblyhasbeenreportedbyGarzonietal.39TheystudiedthemechanismofgrowthofBTAfibrilsinaqueousconditions.Basedonsimulationsofself-assembledshortnon-equilibriumaggregatesofdifferentlength,estimatesofthemonomer‘freeenergy’instacksofdifferentlengthcouldbeobtained.Theresultsindicateacooperativemechanismforsupramolecularpolymergrowth,whereacriticalsizemustbereachedintheaggregatesbeforeemergenceandamplificationoforderintotheexperimentallyobservedfibers.DetailedanalysisofthesimulationdatasuggeststhatH-bondingisamajorsourceofthisstabilizationenergy.Theworkprovidesevidenceforthekeydrivingforceofhydrogenbondinginenhancingthepersistencyofmonomerstackingandamplifyingtheleveloforderintothegrowingsupramolecularpolymer.Studiesintomoleculardrivingforcescanevenhavepracticalimpactinpopularculturesuchasavant-gardecuisine,asthe‘spherification’ofliquidfoodonthenanoscalewasfoundtobedrivenbycalcium-inducedaggregationofalginate.40Aniceexampleofthedifficultythatsimulationsattheall-atomlevelfacetoobtainequilibratedself-assembledstructuresisprovidedbytheworkofHaverkortetal.41Theyperformedextensiveself-assemblysimulationsofamphi-pseudoisocyanine,adyethatformshighlyordered,single-walledcylindricalJ-aggregatesinexperiment,seeFig.3.Thesimulatedstructure,althoughinthecorrectmorphology,containedahighamountofdisorder,despitemultiplecyclesofsimulatedannealing.Note,however,thatthepreciselevelofdisorderisnotknown,asthisisdifficulttoassessexperimentally.Itisconceivablethattheidealizedstructuresoftenusedtoillustratethepackingofmoleculesinsidefibresareunrealisticandoverlyordered.Idealizedstartingstructures.Thelackofknowledgeaboutthepackingdetailsofthesupramolecularaggregatesisoneofthemainreasonstoresorttoall-atomsimulations.Theideaofthisso-called‘‘top-down’’simulationapproachisthat,startingfromFig.3Self-assemblyofamphiphiliccyaninedyesintotubestructure.Thedyemoleculeisamphi-pseudoisocyanine,acyaninedyewithtwohydro-phobictails(CH3andC18H37).Thedyemoleculesaredepictedwiththetailspurple;thenitrogenatomsaredarkblue;thearomaticcarbonsarepink;thelinkerbetweenthearomaticringsislightblue;andhydrogensconnectedtoaromaticcarbonsarewhite.Perchloratecounterionsareshowningreen.Forthewatermolecules,oxygenatomsareredandhydrogenatomsarewhite.Figureadaptedwithpermissionfromref.41.Copyright2013AmericanChemicalSociety.ReviewArticleChemSocRevOpen Access Article. Published on 24 April 2018. Downloaded on 25/04/2018 02:04:47. This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.View Article Online since this problem is equivalent to minimizing (cid:90) (cid:90) ˜E(u) = = ε2 2 ε2 2 Ω Ω (cid:90) Ω (u − b−)dx |(∇u)|2 + W (u)dx − λ |(∇u)|2 + W (u)dx − cλ, (1.1.3) where λ is a Lagrange multiplier and so cλ is only a fixed quantity. The functional ˜E is the same as E with the well W replaced by W (u) + λu. An appropriate choice of λ will render W (u) + λu a double well with equal depth minima. The critical points of the CH are the solutions to δE δu = 0 where δE δu = ε2∆u − W(cid:48)(u), (1.1.4) is the variational derivative with respect to L2 inner-product. Changing variables into the inner variables z = z ε , at the leading order the critical point equation becomes z u − W(cid:48)(u) = λ. ∂2 (1.1.5) From the phase-plane analysis, the critical points of the CH subject to a mass constraint are heteroclinic orbits or single layer interfaces which are the co-dimension 1 interfaces between two distinct phases. Further, the single layer interfaces are the minimizers of the CH subject to a mass constraint. The Γ-convergence as ε → 0 for the single layers structures of CH free energy to a scaled surface area was established in [Modica, 1987] and [Sternberg, 1988]. A model for the network formation in amphiphilic mixtures motivated by small-angle X-ray scattering (SAXS) data was introduced in [Gompper and Schick, 1990] and [Teubner and Strey, 1987] adding a higher order term in the CH free energy. Based on these models, The FCH free energy was developed in [Promislow and Wetton, 2009] and in 3 [Gavish et al., 2011] as a mathematical model to describe the interfacial energy in a phase- separated mixture with amphiphilic molecules. F (u) = ε2∆u − W(cid:48)(u) (cid:17)2 − εp (cid:18) η1ε2 2 (cid:19) |∇u|2 + η2W (u) dx, (1.1.6) (cid:16) (cid:90) 1 2 Ω where ε (cid:28) 1 is the ratio of amphiphilic molecule length to domain size and η1 > 0,η2 ∈ R. For p = 1, the FCH corresponds to the strong functionalization while for p = 2 it is a model for the weak functionalization. It is generic to assume that W (u) is a double-well with two unequal depth minima at b±. Further, due to the quadratic term in FCH it is not possible to rewrite the energy in terms of an equal-depth double-well. We also assume that α± := W(cid:48)(cid:48)(b±) > 0. As a consequence, the dominant term in the FCH energy is the square of the first L2 variational derivative of CH energy. Indeed the FCH energy can be viewed as measuring L2 distance to the critical points of the associated CH energy. When η1 = 0 and η2 = 0, all critical points of CH free energy are the global minimizers of the FCH free energy since = 0 asserts that F (u) = 0. In [Promislow and Zhang, 2013], the existence δE δu of global minimizers was established over a variety of admissible function space for a general class of high order energies such as FCH free energy. Further, the authors showed that the the critical points of CH are the critical points of the higher order energies. 1.2 Description of the Problem Let a small amount of a polymer(soap) be added to a solvent(water) in a container with impermeable walls and allow the system reach an equilibrium. Suppose that the mass of the soap, m, scales with ε with a relation m = εM where ε > 0 is a small parameter and 4 M = O(1). We adapt the FCH free energy introduced in the previous section to model the free energy of this system. For fixed L2 > 0, independent of ε, and Ω = [0, L2] ⊂ R, the FCH free energy describing the free energy of the soap-water mixture is (cid:90) L2 0 (cid:16) 1 2 (cid:17)2 I(u) = ε2∂2 xu − W(cid:48)(u) dx, (1.2.1) subject to the mass constraint (cid:90) L2 0 m := (u − b−)dx = εM, (1.2.2) where M ∈ [0, M∗] ⊂ R for fixed M∗ and u satisfying non-flux boundary conditions. Here the density function u ∈ H2(Ω) map Ω into R+ and the potential W (u) is an unequal double well with two minima at b± for which W (b−) = 0 > W (b+) and a maxima at bM where b− < bM < b+. For simplicity, we prefer converting our problem from macroscopic to misroscopic level by converting spacial variables into inner variables. Introducing the inner variable z = x ε in (1.2.1), the inner scaling of the FCH free energy takes the form (cid:90) L2 ε (cid:16) (cid:17)2 z u − W(cid:48)(u) ∂2 1 2 dz, (1.2.3) (u − b−)dz = M. (1.2.4) I (u) = subject to the mass constraint 0 (cid:90) L2 ε 0 In this thesis, we aim constructing a special class of critical points of the inner scaling of the FCH free energy subject to the mass constraint (1.2.4) as the possible minimizers of the free 5 energy. Apparently, I(u) ≥ 0 for all u ∈ H2(Ω). When I(u) is free of any constraints, the solutions to  ∂2 z u = W(cid:48)(u), ∂zu(0) = 0, ∂zu( L2 ε ) = 0, (1.2.5) are global minimizers of I since they return I(u) = 0. To establish the existence of these solutions we write the (1.2.5) as a non-linear system of first order differential equations v(cid:48) 1 = v2 v(cid:48) 2 = W(cid:48)(v1) (1.2.6) and analyze the associated phase-plane diagram. The dynamical system (1.2.6) has two saddle points at (b±, 0) and a center at the equilibrium point (bM , 0).(See Figure 1.2.) A homoclinic solution is an orbit connecting a saddle point to itself. From the first integral of (1.2.5) 1 2 (∂zu)2 = W (u) + Cn, (1.2.7) for any constant Cn. For the choice Cn = 0, we deduce that we have an orbit starting at b−, hitting the uz = 0 axis at UM for which W (u) = 0 and joining back to b− by reversibility. Thus, there exists a homoclinic solution, φh, converging to b− as z → ∞. The homoclinic solutions do not satisfy the Neumann boundary conditions, having exponentially small derivatives at z = 0 and z = L2 ε . We can construct periodic solutions of (1.2.5) which do satisfy the boundary conditions. The solutions satisfying the boundary conditions of (1.2.5), φn, are the periodic solutions at the center starting at a point between b− and UM on the axis uz = 0 and ending at a point on the same axis when z = L2 ε making n = k+1 2 , 6 k ∈ N tours. (See Figure 1.2.) By adjusting the value of the minimum of φn we adjust the period, which we can tune to be an integral multiple of so that the zero derivative points lie at z = 0 and z = L2 ε . Translating the n-pulse periodic L2 ε gives an exact solution of (1.2.5). (a) Phase plane diagram (b) An illustration for φ4 Figure 1.2: Figure 1.2a is the phase plane for the ODE (1.2.6) which demonstrates the orbit (solid line) homoclinic to b− attaining its maxima at UM and the periodic solutions (dotted lines) that are the solutions to the boundary value problem(1.2.5). Figure 1.2b is an example of an n = 4 pulse solution to (1.2.5) Upon this analysis, among many possible critical points of the mass constrained inner scaling of the FCH free energy we desire to obtain a special class of those which are the expansions of the minimizers of the unconstrained inner scaling of the FCH free energy, n-periodic pulses. Further, motivated by [Promislow, 2002] we survey the modulational stability and dynamical evolution of the n-pulse structure of the inner scaling of the FCH energy. For this purpose, we introduce the mass-preserving projection gradient flow of the FCH energy given in (1.2.3) ut = −Π0 δI δu (u), u(z, 0) = U0(z), 7 (1.2.8) uzuUMb−UMb−φnzL2 where the zero-mass projection, Π0, is given as Π0f := f −(cid:104)f(cid:105)Ω with (cid:104)f(cid:105)Ω := 1|Ω| (cid:90) Ω f (x)dx and observe that any critical point of the. As the zero-mass projection gradient of the FCH energy evolves the mass of the initial data is preserved, (cid:104)u(cid:105)Ω = 0, d dt and the FCH energy given in (1.2.3), I, decreases, I(u) ≤ 0. d dt (1.2.9) (1.2.10) The mass-preserving gradient flow of the CH free energy modeling a phase separation process in a binary mixture was analyzed in [Rubinstein and Sternberg, 1992]. 1.3 Main Results In Chapter 2, the main goal is to construct a class of the critical points of the inner scaling of the FCH free energy subject to the mass constraint over the space of functions u ∈ H2(Ω) satisfying the no-flux boundary conditions. In this purpose, we derive the Euler-Lagrange(E- L) equation of the inner scaling of the FCH,  z − W(cid:48)(cid:48)(u)(cid:1)(cid:0)∂2 z u − W(cid:48)(u)(cid:1) = λε1, (cid:0)∂2 (cid:17) (cid:16) L2 = 0, ∂zu(0) = 0, ∂zu ∂3 z u(0) = 0, ∂3 z u ε (cid:17) (cid:16) L2 ε = 0, (1.3.1) 8 over the admissible space A1 := {u ∈ H2 0, (cid:18)(cid:20) (cid:21)(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:90) L2 ε 0 L2 ε (u − b−)dz = M, uz(0) = uz (cid:18) L2 (cid:19) ε = 0}. (1.3.2) Motivating the corresponding scaling of the Langrange multiplier λε = ελ to the scaling of the mass constraint, we use an asymptotic expansion to solve the E-L equation and obtain a class of the critical points of the mass-constrained FCH energy I. Another result we present in this chapter is that the mass constraint value has a significant impact on the minimizers belonging to the class of the critical points we construct. In Chapter 3, utilizing the classical tools from Calculus of Variations we obtain the existence of the global minimizers of the FCH energy, I, subject to the mass-constraint (1.2.4) over the admissible space A1. In Chapter 4, we introduce n-pulse ansatz as the corrected extensions of the n-pulse solutions of the E-L equation to the whole line R. We establish the H2-coercivity of the second variation of the FCH energy I about n-pulse ansatz and further the H2-coercivity of the second variation of the FCH energy I about periodic multi-pulses φn. With an application of the result on modulational stability of the steady-state solutions of the gradient system in [Promislow, 2002], we demonstrate that n-pulse ansatz, the steady-state solutions of the mass-preserving projection gradient of I, is stable in the modulational sense. The evolution equations of the pulse-locations and the background are derived as ¯λ(cid:48) = 0, p(cid:48) i = −α 3/2− (cid:16) e−√ α−(pi+1−pi) − e−√ α−(pi−pi−1)(cid:17) + O(δ2). (1.3.3) With an analysis of the evolution equations of the pulse locations we conclude that the 9 stationary solutions, equally spaced n-pulses(periodic) are spectrally stable. More significantly, for a given initial data in an ε2-neighborhood of the n-pulse ansatz we recover the pulse dynamics to O(δ2) where δ is exponentially small in ε. Moreover, the solutions remain within a O(δ) neighborhood in H2 of the periodic n-pulses(equally spaced). 10 Chapter 2 The Euler-Lagrange Equation In the calculus of variations, the Euler-Lagrange equations are used to construct the critical points of a functional. In this section, we will derive the Euler-Lagrange equation for the problem (cid:90) L2 ε (cid:16) 1 2 min u∈A (cid:17)2 (cid:90) L2 ε z u − W(cid:48) (u) ∂2 dz subject to (u − b−)dz = M, (2.0.1) 0 where the choice of the space of admissible functions A ⊂ H2(cid:16)(cid:104) 0 (cid:105)(cid:17) 0, L2 ε will be addressed later. As described earlier, the problem is based on an experiment during which a polymer is being added in to a solvent to form a dispersion in a fixed container with an impermeable boundary and then the system is allowed to relax to reach its equilibrium. To model this, we consider u ∈ H2(cid:16)(cid:104) (cid:105)(cid:17) 0, L2 ε satisfying the mass constraint and the Neumann boundary conditions due to impermeable membrane. With this description, we define the space of the admissible functions A1 := {u ∈ H2 (cid:18)(cid:20) 0, L2 ε (cid:21)(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:90) L2 ε 0 (u − b−)dz = M, uz(0) = uz (cid:18) L2 (cid:19) ε = 0}. (2.0.2) 11 spaces and A2 = L2 ε ε 0 (cid:21)(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:90) L2 (cid:21)(cid:19)(cid:12)(cid:12)(cid:90) L2 ε 0 (2.0.3) (cid:27) ) = 0 . (2.0.4) L2 ε Additionally, the free energy given (1.2.3) is well-defined on the following natural admissible (cid:26) A0 := u ∈ H2 (cid:18)(cid:20) 0, (cid:27) , (u − b−)dz = M (cid:26) u ∈ H2 (cid:21)(cid:19) (cid:18)(cid:20) 0, L2 ε (cid:18)(cid:20) 0, L2 ε ∩ H1 0 (u − b−)dz = M, uz(0) = uz( It may be easily observed that A2 ⊂ A1 ⊂ A0. We will construct the Euler-Lagrange equation over all these spaces and discuss the necessary boundary conditions. In the sequel, we will further see that the value of the mass constraint has a considerable impact on the form of the actual minimizer(s). 2.1 Derivation of the Euler-Lagrange Equation over Various Admissible Sets The Euler-Lagrange equation characterizes the smooth critical points of the free energy functional I(·). Recall that we consider u ∈ H2 satisfying the mass constraint udz = M ε (cid:90) L2 and the boundary conditions uz(0) = uz = 0 in the problem and our main purpose is 0 to construct the Euler-Lagrange equation for the critical points satisfying these conditions in this section. We first consider the largest space of admissible functions (cid:26) A0 := u ∈ H2 (cid:18)(cid:20) 0, L2 ε (cid:21)(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:90) L2 ε 0 (cid:27) . (u − b−)dz = M (2.1.1) 12 (cid:17) (cid:16) L2 ε Let u ∈ A0 be any critical point of the energy I(u) subject to the mass constraint and form a curve u + τ v ∈ A0 for τ ∈ R and v ∈ A(cid:48) 0 where (cid:26) (cid:18)(cid:20) A(cid:48) 0 = v ∈ H2 0, L2 ε (cid:21)(cid:19)(cid:12)(cid:12)v ⊥ 1 (cid:27) (cid:105)(cid:17) 0, L2 ε , (2.1.2) is the tangent plane to A0. The orthogonality condition is seen to be required by observing that u + τ v ∈ A0 holds if and only if v ∈ H2(cid:16)(cid:104) u + τ vdz = 0, this and d dτ (cid:90) L2 ε 0 second requirement implies 0 = d dτ = (cid:90) L2 (cid:90) L2 0 ε ε (u + τ v) dz, vdz, (2.1.3) and we deduce that v ⊥ 1. 0 We denote by i(τ ) the evaluation of I on the curve u + τ v i(τ ) := I(u + τ v). (2.1.4) For the following calculations, let us assume that i(τ ) is differentiable at τ = 0 which will be established in Theorem 2.1.2 later. Assuming that u is a critical point of I, i has zero derivative at τ = 0, i.e. i(cid:48)(0) = 0. We formally calculate the variation of I ε (cid:90) L2 (cid:90) L2 (cid:90) L2 0 0 ε ε 0 i(cid:48)(τ ) = = = d dτ (cid:16) (cid:17)2 z (u + τ v) − W(cid:48)(u + τ v) ∂2 (cid:17)(cid:16) z (u + τ v) − W(cid:48)(u + τ v) ∂2 z v −(cid:16) (cid:17) z (u + τ v) − W(cid:48)(u + τ v) ∂2 ∂2 1 2 (cid:16) (cid:16) dz, z v − W(cid:48)(cid:48)(u + τ v)v ∂2 (cid:17) dz, z (u + τ v) − W(cid:48)(u + τ v) ∂2 (cid:17) W(cid:48)(cid:48)(u + τ v)vdz. (2.1.5) 13 Let τ = 0 and since i(cid:48)(0) = 0 ε (cid:16) (cid:90) L2 0 = i(cid:48)(0) = (cid:17) z u − W(cid:48)(u) ∂2 Further, if the critical point u ∈ H4(cid:16)(cid:104) (cid:48). 0 for all v ∈ A0 ∂2 z v −(cid:16) (cid:105)(cid:17) 0, L2 ε (cid:17) z u − W(cid:48)(u) ∂2 W(cid:48)(cid:48)(u)vdz, (2.1.6) we may twice integrate by parts the second term in the integrand and obtain 0 := Gu(v) = ε 0 (cid:90) L2 (cid:124) (cid:16) (cid:124) −∂z (cid:16) (cid:16) z u − W(cid:48)(u) ∂2 ∂2 z (cid:17) z u − W(cid:48)(u) ∂2 v (cid:17) z u − W(cid:48)(u) ∂2 (cid:17) z u − W(cid:48)(u) ∂2 (cid:17) −(cid:16) (cid:123)(cid:122) (cid:12)(cid:12)(cid:12) L2 (cid:16) (cid:123)(cid:122) + A 0 ε B (cid:17) W(cid:48)(cid:48)(u) (cid:12)(cid:12)(cid:12) L2 (cid:125) 0 ε , ∂zv + vdz (cid:125) (2.1.7) for all v ∈ A(cid:48) 0. The map v (cid:55)→ Gu(v) is the weak formulation of the variational derivative of the free energy I(u). The equality in (2.1.7) holds for all v ∈ A0 but different subspaces A(cid:48) 0 afford information on different terms in the RHS of (2.1.7). In particular, if we choose v from the subspace S(cid:48) 0 then the boundary terms B are 0 0 = {v ∈ C∞ 0, c L2 ε (cid:16)(cid:104) (cid:105)(cid:17)(cid:12)(cid:12)(cid:12)v ⊥ 1} ⊂ A(cid:48) (cid:17)(cid:16) z − W(cid:48)(cid:48)(u) ∂2 (cid:17) z (u) − W(cid:48)(u) ∂2 (cid:105)(cid:17) we infer that and we deduce that (cid:90) L2 ε 0 for all v ∈ S(cid:48) 0. By the density of S(cid:48) (cid:16) 0 in L2(cid:16)(cid:104) vdz = 0, (2.1.8) 0 in A(cid:48) (cid:16) 0, L2 ε (cid:17)(cid:16) z − W(cid:48)(cid:48)(u) ∂2 (cid:17) ⊥ A(cid:48) 0, z u − W(cid:48)(u) ∂2 14 (2.1.9) (cid:26) v ∈ H2(cid:16)(cid:104) since A(cid:48) 0 = 0, L2 ε (cid:16) (cid:27) (cid:105)(cid:17)(cid:12)(cid:12)v ⊥ 1 we see that A(cid:48)⊥ 0 = span{1} and hence (cid:17) z u − W(cid:48)(u) ∂2 = λε1, (cid:17)(cid:16) z − W(cid:48)(cid:48)(u) ∂2 (2.1.10) 0 (v(0), v( for some Lagrange multiplier λε. Consequently, the relation (2.1.8) holds for all v ∈ A(cid:48) even for those which are not in S(cid:48) conclude that B = 0 to preserve the equality in (2.1.7). Since the trace map v ∈ A(cid:48) 0. This information implies that A = 0 in (2.1.7) and we 0 (cid:55)→ ε )) ∈ R4 is onto, we may choose v1, v2, v3, v4 for which this trace L2 map yields the canonical basis e1, e2, e3, e4. These choices of v show that the critical point u from the admissible space A0 satisfy the following boundary conditions, L2 ε ), ∂zv(0), ∂zv( z u − W(cid:48)(u)∂zu(cid:1)(cid:12)(cid:12) z u − W(cid:48)(u)∂zu(cid:1)(cid:12)(cid:12)z=0= 0, (cid:0)∂3 (cid:0)∂3 z u − W (u)(cid:1)(cid:12)(cid:12) z u − W (u)(cid:1)(cid:12)(cid:12)z=0= 0, (cid:0)∂2 (cid:0)∂2  Proposition 2.1.1. Any critical point of the problem (2.0.1) over A0 that lies in H4(cid:16)(cid:104) We summarize the results obtained so far in the following proposition. (2.1.11) z= L2 ε z= L2 ε = 0, = 0. (cid:105)(cid:17) 0, L2 ε satisfies  (cid:0)∂2 z − W(cid:48)(cid:48)(u)(cid:1)(cid:0)∂2 z u − W(cid:48)(u)(cid:1) = λε1, z u − W(cid:48)(u)∂zu(cid:1)(cid:12)(cid:12)z=0 = 0, z u − W(cid:48)(u)∂zu(cid:1)(cid:12)(cid:12) (cid:0)∂3 (cid:0)∂3 z u − W (u)(cid:1)(cid:12)(cid:12)z=0= 0, z u − W (u)(cid:1)(cid:12)(cid:12) (cid:0)∂2 (cid:0)∂2 = 0. z= L2 ε = 0, z= L2 ε (2.1.12) Another admissible space, which is actually our main focus of interest for the minimization 15 problem, is given by (cid:26) (cid:18)(cid:20) (cid:21)(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:90) L2 ε 0 L2 ε A1 = u ∈ H2 0, (u − b−)dz = M, uz(0) = uz (cid:18) L2 (cid:19) ε (cid:27) = 0 , (2.1.13) where Neumann-boundary conditions uz(0) = uz = 0 emulate no-flux boundary (cid:17) (cid:16) L2 ε conditions due to the impermeable boundary. Following the same procedure we adopted as deriving the Euler-Lagrange equation satisfied by any u ∈ A0, we achieve (2.1.8) for all v ∈ A(cid:48) 1 where (cid:26) (cid:18)(cid:20) A(cid:48) 1 = v ∈ H2 0, L2 ε (cid:21)(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)v ⊥ 1, vz(0) = vz (cid:18) L2 (cid:19) ε (cid:27) = 0 , (2.1.14) is the tangent plane to the admissible space A1. Taking this into consideration and inserting the boundary conditions uz(0) = uz (cid:17) (cid:16) L2 ε = 0 imposed in the admissible space A1 in (2.1.7) we obtain (cid:17) z (u) − W(cid:48)(u) ∂2 0 = ∂z (cid:16) v (cid:12)(cid:12)(cid:12) L2 ε 0 , (2.1.15) A1 ∩ H4(cid:16)(cid:104) (cid:105)(cid:17) for all v ∈ A(cid:48) 1. Similar to the previous case, we conclude that any critical point u ∈ L2 ε satisfies 0,  z − W(cid:48)(cid:48)(u)(cid:1)(cid:0)∂2 z u − W(cid:48)(u)(cid:1) = λε1, (cid:0)∂2 (cid:17) (cid:16) L2 = 0, ∂zu(0) = 0, ∂zu ∂3 z u(0) = 0, ∂3 z u ε (cid:17) (cid:16) L2 ε = 0. (2.1.16) 16 The last natural admissible space we discuss in this section is (cid:26) (cid:18)(cid:20) (cid:21)(cid:19) (cid:18)(cid:20) A2 = u ∈ H2 0, L2 ε ∩ H1 0 0, L2 ε (cid:21)(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:90) L2 ε 0 (u − b−)dz = M, uz(0) = uz (cid:18) L2 (cid:19) ε (cid:27) = 0 , (2.1.17) which has the tangent plane (cid:26) (cid:18)(cid:20) A(cid:48) 2 = v ∈ H2 0, L2 ε (cid:21)(cid:19) (cid:18)(cid:20) 0, (cid:21)(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)v ⊥ 1, vz(0) = vz (cid:18) L2 (cid:19) ε L2 ε ∩ H1 0 (cid:27) = 0 . (2.1.18) In a similar manner to the previous two cases, it can be easily demonstrated that any minimizer u ∈ A2 ∩ H4([0, L2 ε ]) solves the Euler-Lagrange equations  z u − W(cid:48)(u)(cid:1) = λε1, z − W(cid:48)(cid:48)(u)(cid:1)(cid:0)∂2 (cid:0)∂2 (cid:16) L2 (cid:17) = 0, ∂zu(0) = 0, ∂zu u(0) = 0, u ε (cid:17) (cid:16) L2 ε = 0, (2.1.19) and no new boundary conditions arise. From now on, for simplicity of notation we let A := A1. Our earlier construction of the Euler-Lagrange equation assumed that i(τ ) is differentiable at τ = 0. The following theorem provides a justification for this fact. Theorem 2.1.2. Consider I(u) given (1.2.3). Then, for any u, v ∈ H2(cid:16)(cid:104) (cid:105)(cid:17) 0, L2 ε i(τ ) = I(u + τ v), (2.1.20) is differentiable at τ = 0. 17 Proof. Take any u, v ∈ H2(cid:16)(cid:104) (cid:105)(cid:17) 0, L2 ε and set i(τ ) = I(u + τ v). (2.1.21) First, i(τ ) is finite for all τ since u ∈ H2(cid:16)(cid:104) (cid:105)(cid:17) 0, L2 ε and the square of L2 functions lie in L1. Let τ (cid:54)= 0. Evaluate the difference quotient for a.e. z (cid:90) L2 ε 0 = i(τ ) − i(0) τ (cid:0)∂2 z (u + τ v) − W(cid:48) (u + τ v)(cid:1)2 1 2 dz − 1 2 (cid:0)∂2 z u − W(cid:48) (u)(cid:1)2 τ dz. (2.1.22) Inserting the Taylor series expansion of W(cid:48)(u + τ v) with integral remainder, W(cid:48)(u + τ v) = W(cid:48)(u) + τ W(cid:48)(cid:48)(u)v + W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds, (2.1.23) in to (2.1.22), and after some simplifications we find that the quotient reduces to (cid:90) u+τ v u i(τ ) − i(0) τ = 1 τ (cid:90) L2 ε 0 ε 0 1 τ − 1 2 (cid:90) L2 −(cid:16) (cid:90) L2 − τ ε = = (cid:90) u+τ v (cid:19)2 W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds 1 2 (cid:18) z v − W(cid:48)(u) − τ W(cid:48)(cid:48)(u)v − (cid:16) (cid:17)2 ∂2 z u + τ ∂2 z u − W(cid:48) (u) (cid:16) (cid:17) (cid:17)(cid:16) ∂2 z v − W(cid:48)(cid:48)(u)v z u − W(cid:48)(u) (cid:17)(cid:18)(cid:90) u+τ v ∂2 ∂2 z u − W(cid:48)(u) (cid:17)(cid:18)(cid:90) u+τ v (cid:16) ∂2 z v − W(cid:48)(cid:48)(u)v ∂2 dz, + u τ u u (cid:18)(cid:90) u+τ v W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds (cid:19) + τ 2(cid:16) z v − W(cid:48)(cid:48)(u)v W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds (cid:19) ∂2 u W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds dz+ + + (cid:19)2 (cid:17)2 dz, (2.1.24) Lτ (z)dz, 0 18 (cid:19)2 (cid:17) (cid:18)(cid:90) u+τ v u + W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds (cid:19) W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds + τ 2(cid:16) (cid:19)(cid:21) W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds . z v − W(cid:48)(cid:48)(u)v ∂2 + (cid:17)2 + (2.1.25) . For the proof we first apply the triangle inequality to the where we have introduced (cid:34) Lτ := 1 τ τ (cid:17)(cid:16) (cid:16) z v − W(cid:48)(cid:48)(u)v z u − W(cid:48)(u) (cid:17)(cid:18)(cid:90) u+τ v ∂2 ∂2 −(cid:16) z u − W(cid:48)(u) (cid:17)(cid:18)(cid:90) u+τ v (cid:16) ∂2 z v − W(cid:48)(cid:48)(u)v ∂2 (cid:105)(cid:17) −τ u u L2 ε We claim that Lτ ∈ L1(cid:16)(cid:104) (cid:90) L2 (cid:90) L2 integrand and obtain 0, ε ε |Lτ (z)|dz = 1 τ 0 0 + + (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)τ (cid:16) (cid:17)(cid:16) z u − W(cid:48)(u) (cid:18)(cid:90) u+τ v ∂2 −(cid:16) z u − W(cid:48)(u) + τ 2(cid:16) ∂2 (cid:16) z v − W(cid:48)(cid:48)(u)v ∂2 z v − W(cid:48)(cid:48)(u)v ∂2 − τ (cid:17) (cid:19)2 W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds z v − W(cid:48)(cid:48)(u)v ∂2 (cid:17)(cid:18)(cid:90) u+τ v (cid:17)2 (cid:17)(cid:18)(cid:90) u+τ v + u u u + (cid:19) W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)dz. 19 + (cid:19)2(cid:12)(cid:12)(cid:12)(cid:12) (cid:125) + W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds 1 τ Taking inside the integral and by the triangle inequality we have 0 0 u 1 τ :=C + + :=A :=B τ u (cid:90) L2 ε (cid:90) L2 ε |Lτ (z)|dz ≤ z v − W(cid:48)(cid:48)(u)v ∂2 (cid:123)(cid:122) (cid:17)(cid:18)(cid:90) u+τ v (cid:123)(cid:122) (cid:17)2(cid:12)(cid:12)(cid:12)(cid:12) (cid:125) z v − W(cid:48)(cid:48)(u)v ∂2 (cid:17)(cid:18) 1 (cid:17)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:16) (cid:17)(cid:16) (cid:124) (cid:125) (cid:123)(cid:122) z u − W(cid:48)(u) ∂2 (cid:12)(cid:12)(cid:12)(cid:12)τ (cid:18) 1 (cid:90) u+τ v (cid:124) W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds (cid:12)(cid:12)(cid:12)(cid:12)(cid:16) (cid:124) z u − W(cid:48)(u) ∂2 (cid:12)(cid:12)(cid:12)(cid:12)τ (cid:16) (cid:124) (cid:12)(cid:12)(cid:12)(cid:12)(cid:16) (cid:124) (cid:105)(cid:17) A, B, C, D and E are in L1(cid:16)(cid:104) Adz ≤(cid:13)(cid:13)(cid:13)∂2 (cid:90) u+τ v (cid:123)(cid:122) z v − W(cid:48)(cid:48)(u)v ∂2 z u − W(cid:48) (u) (cid:90) L2 ε (cid:13)(cid:13)(cid:13)∂2 0, L2 ε (cid:123)(cid:122) (cid:13)(cid:13)(cid:13)L2 τ u :=D :=E + + + 0 By the Lebesgue Dominated Convergence Theorem, it suffices to show that each of the terms . For the first term, by the Holder’s inequality W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds dz. + (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) (cid:125) (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) (cid:125) (2.1.26) (2.1.27) (2.1.28) (cid:13)(cid:13)(cid:13)L2 , z v − W(cid:48)(cid:48) (u) v , (cid:17) +(cid:13)(cid:13)W(cid:48)(cid:48)(u)v(cid:13)(cid:13)L2 (cid:17) +(cid:13)(cid:13)W(cid:48)(cid:48)(u)(cid:13)(cid:13)L∞ (cid:107)v(cid:107)L2 (cid:13)(cid:13)(cid:13)W (n)(s) . Then, from the triangle inequality we have (cid:90) L2 ε 0 Adz ≤(cid:16)(cid:13)(cid:13)(cid:13)∂2 ≤(cid:16)(cid:13)(cid:13)(cid:13)∂2 z u z u (cid:13)(cid:13)(cid:13)L2 (cid:13)(cid:13)(cid:13)L2 +(cid:13)(cid:13)W(cid:48)(u)(cid:13)(cid:13)L2 +(cid:13)(cid:13)W(cid:48)(u)(cid:13)(cid:13)L2 (cid:17)(cid:16)(cid:13)(cid:13)(cid:13)∂2 (cid:17)(cid:16)(cid:13)(cid:13)(cid:13)∂2 z v z v (cid:13)(cid:13)(cid:13)L2 (cid:13)(cid:13)(cid:13)L2 (cid:13)(cid:13)(cid:13)L∞ ≤ α1 for all s ∈(cid:2)−(cid:107)u(cid:107)L∞ − τ (cid:107)v(cid:107)L∞ ,(cid:107)u(cid:107)L∞ + τ (cid:107)v(cid:107)L∞(cid:3). Since u ∈ H2 implies (cid:107)u(cid:107)L∞ is bounded, By the smoothness of W , for each n there exists an α1 > 0 such that 20 we conclude that A ∈ L1(cid:16)(cid:104) (cid:105)(cid:17) 0, L2 ε that there exists a constant α1 such that for each u, v ∈ H2(cid:16)(cid:104) (cid:90) L2 Ddz ≤ α1τ. ε (cid:105)(cid:17) 0, L2 ε . The same arguments show (2.1.29) 0 On the other hand, bounding B, C and E requires estimates on the integral remainder term. From the smoothness of W and the compact range of u and u + τ v, there exists a α2 > 0 such that |W(cid:48)(cid:48)(cid:48)(s)| ≤ α2 for s running over u to u + τ v. Then, (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 1 τ (cid:90) u+τ v u W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds (cid:90) u+τ v (cid:12)(cid:12)u + τ v − s(cid:12)(cid:12)ds. (2.1.30) By the change of variables t = u + τ v − s we obtain (cid:90) u+τ v u (cid:12)(cid:12)u + τ v − s(cid:12)(cid:12)ds = |t|dt ≤ τ 2 2 |v2|. (2.1.31) τ v u τ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ α2 (cid:90) 0 Inserting this in (2.1.30) we have (cid:90) u+τ v u W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ α2τ 2 (cid:12)(cid:12)v2(cid:12)(cid:12). (2.1.32) With these estimates we deduce that Bdz ≤ W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds (cid:13)(cid:13)(cid:13)(cid:13)2 , L2 , (2.1.33) τ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 1 (cid:90) L2 ε 0 (cid:13)(cid:13)(cid:13)(cid:13)(cid:90) u+τ v (cid:13)(cid:13)(cid:13)v2(cid:13)(cid:13)(cid:13)2 α2 2 2 (cid:107)v(cid:107)4 α2 u L2 L4 . ≤ τ 4 4 τ 4 4 = 21 Here we controlled the L4 norm by H1 and L∞ norms using the estimate (cid:90) L2 ε 0 (cid:107)v(cid:107)4 L4 = v4dz ≤ (cid:107)u(cid:107)2 L∞ (cid:107)u(cid:107)2 L2 ≤ (cid:107)u(cid:107)2 H1 (cid:107)u(cid:107)2 L∞ , (2.1.34) and further L∞ norm by H1 norm (cid:12)(cid:12)(cid:12)(cid:12)(cid:90) L2 ε z (cid:12)(cid:12)(cid:12)(cid:12) ≤ 2(cid:107)v(cid:107)L2(cid:107)∂zv(cid:107)L2 ≤ (cid:107)v(cid:107)2 ∂x(v2)dx (cid:107)v(cid:107)2 L∞ = sup z∈Ω we obtain L2 + (cid:107)∂zv(cid:107)2 L2 ≤ (cid:107)v(cid:107)2 H1 , (2.1.35) (cid:107)v(cid:107)4 L4 ≤ (cid:107)v(cid:107)4 H1 ≤ C (cid:107)v(cid:107)4 H2 . Utilizing this estimate in (2.1.33), we have (cid:90) L2 ε 0 Bdz ≤ τ 4 4 2 (cid:107)v(cid:107)4 α2 H2 . With the same arguments we obtain upper bounds for C and E such that W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)dz, (cid:90) L2 ε (cid:90) L2 ε (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:16) (cid:90) L2 ε z u − W(cid:48)(u) ∂2 u τ (cid:17)(cid:18) 1 (cid:90) u+τ v (cid:17)(cid:12)(cid:12)(cid:12)(cid:12)dz, z u − W(cid:48) (u) (cid:13)(cid:13)(cid:13)L2 +(cid:13)(cid:13)W(cid:48)(u)(cid:13)(cid:13)L2 ∂2 (cid:13)(cid:13)(cid:13)H2 (cid:17) + α1 z u , z u (cid:12)(cid:12)(cid:12)(cid:12)v2(cid:16) (cid:16)(cid:13)(cid:13)(cid:13)∂2 (cid:16)(cid:13)(cid:13)(cid:13)∂2 0 (cid:107)v(cid:107)2 L4 (cid:107)v(cid:107)2 H2 (cid:17) , Cdz = 0 0 ≤ τ α2 2 ≤ τ α2 2 ≤ τ α2 2 (2.1.36) (2.1.37) (2.1.38) 22 and(cid:90) L2 ε 0 (cid:90) L2 ε 0 Edz = ≤ τ 2α2 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)τ (cid:16) (cid:90) L2 ε 0 (cid:12)(cid:12)(cid:12)(cid:12)v2(cid:16) z v − W(cid:48)(u)v ∂2 (cid:17)(cid:18) 1 τ (cid:90) u+τ v (cid:17)(cid:12)(cid:12)(cid:12)(cid:12)dz, u z v − W(cid:48) (u) v ∂2 W(cid:48)(cid:48)(cid:48)(s) (u + τ v − s) ds (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)dz, (2.1.39) ≤ (1 + α1) τ 2α2 (cid:107)v(cid:107)3 H2 . Considering all these estimates for A, B, C, D and E and u, v ∈ H2(cid:16)(cid:104) 2 (cid:105)(cid:17) (cid:17)(cid:16)(cid:13)(cid:13)∂2 0, L2 ε z v(cid:13)(cid:13)L2 +(cid:13)(cid:13)W(cid:48)(cid:48)(u)(cid:13)(cid:13)L∞ (cid:107)v(cid:107)L2 we conclude (cid:17) ∈ (cid:16)(cid:13)(cid:13)∂2 z u(cid:13)(cid:13)L2 +(cid:13)(cid:13)W(cid:48)(u)(cid:13)(cid:13)L2 (cid:90) L2 (cid:105)(cid:17) 0 ε as τ → 0. |Lτ (z)|dz ≤ γ + O(τ ), 0, L2 ε that there exists a constant γ := R such that which implies that Lτ ∈ L1(cid:16)(cid:104) (cid:17)(cid:16) Lτ →(cid:16) z u − W(cid:48)(u) ∂2 (2.1.40) Indeed our estimates applied to (2.1.25) show that (cid:17) z v − W(cid:48)(cid:48)(u)v ∂2 τ → 0 as f or a.e z, (2.1.41) Applying the Dominated convergence theorem, from (2.1.40) and (2.1.41) we show that the limits i(cid:48)(0) = lim τ→0 i(τ ) − i(0) τ (cid:90) L2 (cid:16) 0 ε = lim τ→0 (cid:90) L2 ε = 0 Lτ dz, z u − W(cid:48)(u) ∂2 exist. This proves that i is differentiable at τ = 0. 23 (cid:17)(cid:16) (cid:17) z − W(cid:48)(cid:48)(u) ∂2 vdz, (2.1.42) 2.2 The Euler-Lagrange Equation with a small mass constraint Recall that the inner scaling of the Functionalized Cahn-Hilliard free energy with no functionalization terms is (cid:90) L2 ε 0 1 2 (cid:16) (cid:17)2 z u − W(cid:48)(u) ∂2 dz. I (u) = In this section, we address the dependence of the minima of the functional I upon the mass M , (cid:90) L2 ε 0 (u − b−)dz = M, (2.2.1) by analyzing an asymptotic expansion of the Euler-Lagrange equation (2.1.10). We construct a class of solutions of the E-L equation which corresponds to a scaling of the Lagrange multiplier, λε = ελ, arising from the ε-dependence of the mass constraint, m. Following this information, specifically, we seek solutions to the perturbed E-L equation, 2.2.1 Notation Introduce the operator (cid:16) (cid:17)(cid:16) z − W(cid:48)(cid:48)(Φ) ∂2 (cid:17) z Φ − W(cid:48)(Φ) ∂2 = λε. L := ∂2 z − W(cid:48)(cid:48)(φh), 24 (2.2.2) (2.2.3) that is the linearization of (1.2.5) about φh, and define the operator Ln(p) := ∂2 z − W(cid:48)(cid:48) (un) , acting on H2 (R) where un is the extension φn to R defined as n(cid:88) (cid:0)z − pj (cid:1) + b−, (2.2.4) (2.2.5) un := φh j=1 where φh := φh − b− and p = (p1, p2, ..., pn)t ∈ Rn is the vector of pulse locations. The admissible set of pulse locations is given by P = {p ∈ Rn : pi < pi+1 i = 1, ..., n and ∆p ≥ l}, for (2.2.6) |pi − pj| and l > 0 is sufficiently large. Let Bj denote the functions where ∆p = min i(cid:54)=j L−j1 ∈ L∞(R) for j = 1, 2 that are the solutions to LjBj = 1, and orthogonal to the kernel of L. Actually, the function Bj is in the form Specifically, B1 takes the form Bj = ˆBj + Bj,∞. B1 = ˆB1 − 1 α− , 25 (2.2.7) (2.2.8) (2.2.9) where α− is the coercivity value of W (u) at b− and ˆB1 is the solution to the L ˆB1 = 1 − W(cid:48)(cid:48)(φh) α− . (2.2.10) Since φh → b− at an exponential rate as z → ∞, the RHS of (2.2.10) is in L2(R) and even about z = 0, hence orthogonal to φ(cid:48) h ∈kerL. The existence of B2 is established similarly. Further, we truncate Bj to have compact support on the bounded interval [0, L2 ε ] and introduce the functions so that n(cid:88) i=1 Bj,n := ˆBj(z − pi) + Bj,∞, Lj n,bBj,n = 1 + o(δ2) (2.2.11) (2.2.12) where Ln,b represents the restriction of the operator Ln(p) to the bounded domain [0, Remark 2.2.1. The kernel of the operator L := ∂2 about the homoclinic solution φh is spanned by φ(cid:48) h. L2 ε ]. z − W(cid:48)(cid:48) (φh), the linearization of (1.2.6) Indeed, the equation (1.2.6) has a translational symmetry on R since φh(z + γ) also solves the equation (1.2.6). If we insert φh(z + γ) in the ODE (1.2.6) and take its derivative with respect to γ we have Lφh (cid:48) = 0, (2.2.13) and conclude that kerL = span{φ(cid:48) h}. Since L is a second order Sturm-Liouville operator acting on an unbounded domain it has real-valued simple eigenvalues that can be written in a (cid:48) is an eigenfunction of L corresponding strictly decreasing order. Since kerL = span{φh (cid:48)}, φh 26 to the eigenvalue λ = 0. Since the eigenfunctions of a Sturm-Liouville operator, {ψj}, has j (cid:48) has one node, it is the second largest eigenfunction, ψ1, corresponding simple zeros and φh to the second eigenvalue λ1 = 0. Then, there exists a ground-state eigenfunction, ψ0 corresponding to λ0 > 0 and the remainder of the spectrum is real and O(1) distance to the left of 0 by Weyl’s essential spectrum theorem. Remark 2.2.2. By standard perturbation theory, the properties of the point spectrum of Ln(p) presented in Lemma 4.1.4 carry over up to exponentially small terms to the operator on the large bounded domain, Ln,b(p). (See Section 9.6 in [Kapitula and Promislow, 2013].) For the purposes of exposition we do not distinguish between the operator Ln(p) acting upon the whole line or Ln,b(p) acting upon the large bounded domain when inverting the operator, except where doing so is essential to the argument. 2.2.2 Motivation Before we proceed to the asymptotic expansion analysis of the E-L equation, we would like to motivate ε-dependence of the Lagrange multiplier, λε, due to the ε-dependence of the total mass, m. Assume uε is a solution to equation (2.2.2), or to z u − W(cid:48)(cid:48)(cid:48)(u)(∂zu)2 +(cid:0)W(cid:48)(cid:48)(u)W(cid:48)(u) − λε (cid:1) = 0, (2.2.14) z u − 2W(cid:48)(u)∂2 ∂4 written explicitly. Rather than dealing with possible different types of solutions to (2.2.2), we focus on the construction of multipulse solutions as the possible minima of I. For this purpose, we simplify the problem taking ε → 0 which extends the domain to R+ and the considering its 27 even extension to R−. Further, keeping the mass constraint fixed (cid:90) R (u0 − b−)dz = M, (2.2.15) we deduce that u0 → b− as z → ±∞. Consistent with u0 → b− as z → ±∞, the exponential dichotomies of (2.2.2) on R imply z (u0 − b−) → 0 for k = 1, 2, 3 exponentially as ∂k z → ±∞, (2.2.16) (See Section 2.1.4 in [Kapitula and Promislow, 2013]). In other words, there exist constants c, κ > 0 such that where (cid:126)b− = (b−, 0, 0, 0), (cid:126)u0(·) = (cid:0)u0(·), u(cid:48) |(cid:126)u0 − (cid:126)b−| ≤ ce−κ|z|. 0 (·)(cid:1). Returning to (2.2.14), setting (2.2.17) 0(·), u(cid:48)(cid:48) 0(·), u(cid:48)(cid:48)(cid:48) ε = 0 and taking z → ±∞ by (2.2.16) we obtain the equality W(cid:48)(b−)W(cid:48)(cid:48)(b−) = λ0. (2.2.18) Since W(cid:48)(b−) = 0 we conclude that λ0 = 0. For the finite domain problem associated to (2.2.2), namely when ε (cid:54)= 0 but small, we choose boundary conditions that best approximate the whole line problem. In particular, we assume u becomes asymptotically close to b which solves W(cid:48)(b)W(cid:48)(cid:48)(b) = λε, (2.2.19) 28 and impose the exponential dichotomies like boundary conditions (cid:126)u(0) ∈ Wu((cid:126)b) and (cid:126)u ∈ Ws((cid:126)b), where (cid:126)b = (b, 0, 0, 0), (cid:126)u(·) = (cid:0)u(·), u(cid:48)(·), u(cid:48)(cid:48)(·), u(cid:48)(cid:48)(cid:48)(·)(cid:1) and Wu((cid:126)b) and Ws((cid:126)b) are the (cid:18) L2 (cid:19) ε unstable and stable manifolds of (cid:126)b, respectively. This yields a finite domain problem that best approximates the whole-line problem. This assumption is consistent with the mass constraint, away from the pulses scale like (cid:90) L2 ε 0 M = (u − b−)dz ≈ L2 ε (b − b−) = O(1), (2.2.20) and we deduce that b = b− + O(ε). From (2.2.19) we deduce that λε =(cid:0)W(cid:48)(cid:48)(b−)(cid:1))2(b − b−) = O(ε). (2.2.21) To fix notation, we write b = b− + εb1 and λε = ελ for b1 = O(1) and λ = O(1). 2.2.3 Solutions to the Euler-Lagrange Equation with a small mass constraint In the light of the comments about our motivation discussed in Section 2.2.2, our main focus of interest is constructing a class of solutions to (2.2.2) when λ = λ. We are specifically interested in an asymptotic expansion of these solutions which has an expansion form Φ = ϕ0 + εϕ1 + O(ε2). (2.2.22) 29 (a) Phase plane for when λ = 0 (b) Phase plane when λ (cid:54)= 0 Figure 2.1: Figure 2.1a represents a phase plane for the dynamical system given in (2.2.2) extended to R for λ = 0. Figure 2.1b is a phase plane for the dynamical system when λ (cid:54)= 0. The boundary conditions (cid:126)u(0) ∈ W u((cid:126)b) and (cid:126)u mimic the behavior of the whole line system. The distance between the fixed points of the dynamical systems is |b− b−| = ε while |u(0)− b| = O(e−κ L2 ε ) (cid:28) ε. However, for the clarity of the graph both distances are depicted in similar lengths. (cid:17) ∈ W s((cid:126)b) where (cid:126)b = (b, 0, 0, 0) (cid:16) L2 ε (cid:16) (cid:16) (cid:16) ελ = = = To find ϕ0 and ϕ1 we first insert (2.2.22) in (2.2.2). (cid:17)(cid:16) (cid:17) z Φ − W(cid:48) (Φ) z − W(cid:48)(cid:48) (Φ) (cid:17)(cid:17)(cid:16) ∂2 ∂2 ϕ0 + εϕ1 + O(ε2) ∂2 z − W(cid:48)(cid:48)(cid:16) (cid:16) z −(cid:0)W(cid:48)(cid:48) (ϕ0) + εϕ1W(cid:48)(cid:48)(cid:48) (ϕ0+)(cid:1) + O(ε2) ∂2 z ∂2 , (cid:17)(cid:16) ϕ0 + εϕ1 + O(ε2) (cid:17) −(cid:0)W(cid:48) (ϕ0) + εϕ1W(cid:48)(cid:48) (ϕ0)(cid:1) + O(ε2) ∂2 z ϕ0 + ε∂2 . (cid:17) − W(cid:48)(cid:16) (cid:17)(cid:17) , ϕ0 + εϕ1 + O(ε2) z ϕ1+ (2.2.23) Matching the terms with the same order of ε, we obtain O(1) and O(ε) equations (cid:16) (cid:17)(cid:16) (cid:17) z ϕ0 − W(cid:48) (ϕ0) ∂2 z − W(cid:48)(cid:48) (ϕ0) ∂2 = 0, (2.2.24) 30 (cid:16) z − W(cid:48)(cid:48) (ϕ0) ∂2 (cid:17)(cid:16) z ϕ1 − ϕ1W(cid:48)(cid:48) (ϕ0) ∂2 (cid:17) (cid:16) (cid:17) z ϕ0 − W(cid:48) (ϕ0) ∂2 + ϕ1W(cid:48)(cid:48)(cid:48) (ϕ0) = λ, (2.2.25) As our base profile we choose the solutions φn for n ∈ { k+1 2 : k ∈ N} to the problem (1.2.5) as a special classes of solutions of equation (2.2.24). Inserting ϕ0 = φn in (2.2.25), λ = (cid:16) (cid:16) (cid:17)(cid:16) z − W(cid:48)(cid:48) (φn) (cid:17)(cid:16) ∂2 z − W(cid:48)(cid:48) (φn) ∂2 = = L2 nϕ1, (cid:17) z ϕ1 − ϕ1W(cid:48)(cid:48) (φn) (cid:17) ∂2 z − W(cid:48)(cid:48) (φn) ∂2 ϕ1, , (2.2.26) where we have denoted the linearization of (1.2.6) about the periodic solution φn by For each n = 1, 2, 3, the equation Ln := ∂2 z − W(cid:48)(cid:48)(φn). L2 nϕ1 = λ, (2.2.27) (2.2.28) has a solution ϕ1 ∈ L2 if and only if λ ⊥ kerLn by the Fredholm Alternative. Indeed, λ is orthogonal to kerLn because we have (cid:90) R (cid:48)dz = 0, λφn (2.2.29) since kerLn = span{φn (cid:48)} and φn (cid:48) is odd about z = 0. Then, the solution to (2.2.28) can be written as ϕ1 = λB2,n, (2.2.30) 31 where B2,n solves L2 nB2,n = 1. (2.2.31) Therefore, the periodic solutions to the Euler-Lagrange equation (2.1.10) in other words the critical points of I subject to the mass constraint (1.2.4) have the asymptotic expansion Φn = φn + ελB2,n + O(cid:16) ε2(cid:17) . (2.2.32) 2.2.4 Some Remarks Recall that the equation (1.2.5) has solutions, φn, which are global minimizers of the free energy (1.2.3) when I(u) is not subject to any constraints. These have background state, b = b−. In Section 2.2.2 we motivate the construction of the solutions with background. b = bn where bn − b− = O(ε) solves W(cid:48)(bn)W(cid:48)(cid:48)(bn) = λε. (2.2.33) Since W(cid:48)(b−) = 0 it follows λε is also O(ε). Let τ (bn) be the period of the orbit φn. The periodic solutions Φn to the Euler-Lagrange equation (2.1.10) which satisfy τ (bn) = fit precisely n periods of the orbit into the interval solve the boundary conditions. Moreover, the period scales like τ (bn) = O(cid:16) and is monotonically decreasing as b−b− increases, achieving a minima at the center equilibrium. and may be translated to exactly (cid:16) 1 (cid:17)(cid:17) bn−b− ln L2 nε can , however we further restrict the size (cid:105) (cid:104) 0, L2 ε ετ (bn) = O(cid:16) 1 L2 ε|lnε| (cid:17) 32 These considerations suggest n = of bn so that n = O(1) Remark 2.2.3. If τ (b) (cid:54)= L2 nε then we state without proof that by adjusting the associated φ for which (φ, φ(cid:48)) passes through (b, 0), we may arrive at a solution of (1.2.5) that satisfies ε ) = O(e− L2 the boundary conditions. Indeed by translating φ we may achieve φ(cid:48)(0), φ(cid:48)( ε ). L2 The corrections to φ are exponentially small and have no impact on the value of the energy or total mass to the order considered here. In particular, we define (cid:22) L2 τ (b)ε (cid:23) − 1 2 n := . (2.2.34) we proceed formally with the construction of Φn ignoring the issue of exponentially small boundary mismatch. 2.2.5 Energy values at the critical points Inserting the critical points (2.2.32) in (1.2.3) and in the mass constraint (1.2.4) we calculate their energy values and their mass constraint M to determine which has minimum energy at prescribed mass. 0 (cid:90) L2 (cid:90) L2 (cid:90) L2 (cid:90) L2 (cid:90) L2 0 0 0 ε 0 ε ε ε ε ∂2 z (cid:17)2 dz, φn + ελB2,n + O(ε2) (cid:16) z Φn − W(cid:48) (Φn) ∂2 (cid:16) (cid:16) ε2λ2(cid:16) ε2λ2(cid:0)LnB2,n ε2λ2(cid:0)LnB2,n zB2,n − W(cid:48)(cid:48) (φn)B2,n ∂2 (cid:1)2 + O(ε3)dz, (cid:1)(cid:0)LnB2,n 1 2 1 2 1 2 1 2 1 2 I (Φn) = = = = = (cid:1) + O(ε3)dz. (cid:17)(cid:17)2 φn + ελB2,n + O(ε2) dz, (cid:17) − W(cid:48)(cid:16) (cid:17)2 + O(ε3)dz, (2.2.35) 33 Since Ln’s are self-adjoint and from (2.2.31) we obtain the reduced energy (cid:90) L2 (cid:90) L2 0 ε ε 0 1 2 1 2 I (Φn) = = ε2λ2(cid:16)L2 nB2,n (cid:17)B2,n + O(ε3)dz, ε2λ2B2,n + O(ε3)dz. Let ¯B2,n denote the mass for B2,n that is (cid:90) L2 B2,n := ε B2,ndz, 0 = α−2− L2 ε + O(1), and Mn be the mass for the critical point (2.2.32) (cid:90) L2 (cid:90) L2 0 ε ε 0 Mn : = = Φndz (cid:2)(φn − b−) + ελB2,n (cid:3) dz. (2.2.36) (2.2.37) (2.2.38) Substituting the mass (2.2.37), we have Mn = L2α−2− λ + (cid:90) L2 ε 0 (φn − b−)dz. (2.2.39) For the sake of convenience, we write the mass of φn in terms of the mass of φh, (φn − b−)dz = n (φh − b−)dz, (2.2.40) (cid:90) L2 ε 0 (cid:90) L2 ε 0 = nMh, 34 where Mh is the mass for the homoclinic solution φh, namely (cid:90) L2 ε 0 Mh := (φh − b−)dz. We insert (2.2.40) in (2.2.38) and obtain Mn = nMh + λα−2− L2. (2.2.41) (2.2.42) Setting Mn = M and solving for λ gives an expression for λ in terms of the mass constraint M , λ = α2− L2 (M − nMh) . (2.2.43) Utilizing this formula of λ in (2.2.36), we calculate the reduced energy critical point of I in (cid:90) L2 ε terms of the mass constraint M I (Φn) = = = ε2λ2B2,ndz, 1 2 0 ε2λ2B2,n, 1 2 εα2− 2L2 (M − nMh)2 . (2.2.44) We minimize (2.2.44) for n to find the n-pulses with the minimal free energy. Note that when M = nMh, Φn is a global minimizer. Since I(Φn) is a discrete function of n, the closest value of n to M Mh is the minima of I (Φn). We conclude that the inner scaling of FCH free energy, I(u), attains its minima, over the n pulse solutions we have constructed, at Φn for which n is closest to M Mh . From Figure 2.2 , it can be observed that among the n-pulse profiles we have constructed 35 Figure 2.2: The reduced energy, I(Φn), versus mass constraint M . The blue lines demonstrate the energy values at the critical points, Φn, for n = 0, 1 2 , 2. The red lines represent the infimum of the energy values over all the blue lines. 2 , 1, 3 the one with minimal energy sensitively depend on the mass constraint. In the sequel, we show the existence of a global minima and that integral values of n the associated n-pulse Φn is a local minima of I. 36 01/2mhmh3/2mh2mhM*MI(n)I(0)(M)I(1/2)(M)I(1)(M)I(3/2)(M)I(2)(M)1/23/23/22BAGH--←izµ← Chapter 3 Existence Of the Minimizers 3.1 Existence of the Minimizers In this section, we use classical tools from direct methods of Calculus of Variations to establish the existence of a global minimizer the energy functional (1.2.3) subject to the mass constraint (1.2.4) following the procedures in [Promislow and Zhang, 2013]. Recall that the space of admissible functions is given by (cid:26) (cid:18)(cid:20) (cid:21)(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:90) L2 ε 0 0, L2 ε A = u ∈ H2 (u − b−)dz = M, uz(0) = 0, uz (cid:18) L2 (cid:19) ε (cid:27) = 0 . We consider a general form for the double-well potential W (u). In addition to the assumptions earlier on, we suppose that W (u) is convex at infinity and satisfies some growth conditions and b− = 0 for the sake of easiness in the calculations. Specifically, there exist p > 1, c, c− > 0, u0 > 0 and β > 1 sufficiently large such that  W(cid:48)(u) ≤ c|u|p, W(cid:48)(u) ≤ c−, |u| ≥ 1, |u| < 1, W(cid:48)(cid:48)(u) > β, |u| > u0. 37 (3.1.1) (3.1.2) In the following lemma, we establish an estimate for H1 norm of u ∈ A which is utilized in the proof of the H2-coercivity of the energy in Theorem 3.1.2. For simplicity in the proof of existence of a global minima, for fixed ε > 0 and L2 ∈ R we introduce scaled Cahn-Hilliard free energy functional on a large bounded domain 0, (cid:104) (cid:105) ∈ R L2 ε (cid:90) L2 ε 0 1 2 E = (∂zu)2 + W (u)dz. (3.1.3) The variational derivative of E with respect to L2 inner product is = −∂2 z u + W(cid:48)(u), δE δu and note that the FCH free energy, I, can be written in terms of δE δu (3.1.4) (3.1.5) I(u) = = (cid:19)2 dz, ε . δu δu 1 2 0 1 2 (cid:18) δE (cid:90) L2 (cid:13)(cid:13)(cid:13)(cid:13)δE (cid:13)(cid:13)(cid:13)(cid:13)2 (cid:13)(cid:13)(cid:13)(cid:13)2 (cid:13)(cid:13)(cid:13)(cid:13) δE L2 δu L2 Lemma 3.1.1. Fix ε > 0. There exists a constant C > 0 such that (cid:107)u(cid:107)2 H1 ≤ + C, (3.1.6) for all u ∈ A. Proof. Multiplying δE δu by u and integrating by parts, we obtain (cid:90) L2 ε 0 (cid:90) L2 ε 0 (cid:16)−∂2 (cid:17) z u + W(cid:48)(u) u dz = (cid:90) L2 ε 0 u δE δu dz = 38 (∂zu)2 + uW(cid:48)(u)dz. (3.1.7) So, (cid:90) L2 ε 0 (cid:90) L2 ε 0 (∂zu)2dz = − uW(cid:48)(u)dz. u δE δu (3.1.8) Applying Young’s inequality to the first term on the right and adding (cid:107)u(cid:107)2 L2 to both sides, we have (cid:90) L2 ε 0 (cid:107)u(cid:107)2 H1 ≤ (cid:18) δE (cid:19)2 − uW(cid:48)(u) + u2dz, δu u2 − uW(cid:48)(u)dz. = I(u) + u2 2 1 2 ε + (cid:90) L2 (cid:90) L2 0 3 2 ε We need to get an upper bound for h(z)dz where 0 h(z) := u2 − uW(cid:48)(u). 3 2 From the assumption (3.1.2) there exists β and u0 such that W(cid:48)(u) > β (u − u0) + W(cid:48)(u0) for u > u0, W(cid:48)(u) < β(u0 + u) + W(cid:48)(−u0) for u < −u0. Multiplying (3.1.11) and (3.1.12) by −u and adding 3 2 u2 , we obtain (3.1.9) (3.1.10) (3.1.11) (3.1.12) (cid:19) − β (cid:19) (cid:18)3 (cid:18)3 2 2 h(u) < u2 + βu0u − uW(cid:48)(u0) for u > u0, (3.1.13) h(u) < − β u2 − βu0u − uW(cid:48)(−u0) for u < −u0. (3.1.14) The inequalities (3.1.13) and (3.1.14) imply that on R/[−u0, u0] h(u) is bounded above for sufficiently large β because it is bounded by a function whose dominant term is quadratic 39 with a negative coefficient for sufficiently large β. Further, h(u) is bounded on [−u0, u0] since it is continuous. Hence, h(u) is bounded above for sufficiently large β, in other words there exists a constant C > 0 independent of ε such that (cid:90) L2 ε 0 h(z)dz < C. (3.1.15) From this and (3.1.9) , the bound we aimed in (3.1.6) has been achieved. Theorem 3.1.2. Fix ε > 0 and L2, independent of . The energy functional given in (1.2.3) subject to the mass constraint (1.2.4) has a global minimizer over the admissible set A. In other words, there exists at least one u ∈ A satisfying I(w) ≥ I(u), (3.1.16) for all w ∈ A. Proof. Since I(·) is well-defined on A and bounded below by 0, we define m := infw∈AI(w) ≥ 0. The key step in the proof of the existence of a minimizer is to show the coercivity of I(u) over the admissible set A. More specifically we show that there exist constants β and γ such that (I(u)) 3 2 ≤ −γ + β(cid:107)u(cid:107)H2. (3.1.17) To establish this bound we pursue the preliminary step of bounding the variational derivative 40 of E, given in (3.1.3). By the relation obtained from (3.1.4) ∂2 z u = δE δu − W(cid:48)(u). Taking L2 norm of both sides and applying the triangle inequality, we obtain (cid:13)(cid:13)(cid:13)∂2 z u (cid:13)(cid:13)(cid:13)L2 δu (cid:13)(cid:13)(cid:13)(cid:13)δE (cid:13)(cid:13)(cid:13)(cid:13)δE (cid:13)(cid:13)(cid:13)(cid:13)δE δu δu , − W(cid:48)(u) (cid:13)(cid:13)(cid:13)(cid:13)L2 +(cid:13)(cid:13)W(cid:48)(u)(cid:13)(cid:13)L2 , (cid:13)(cid:13)(cid:13)(cid:13)L2 (cid:13)(cid:13)(cid:13)(cid:13)L2 +(cid:13)(cid:13)W(cid:48)(u)(cid:13)(cid:13)L2 , 2 +(cid:13)(cid:13)W(cid:48)(u)(cid:13)(cid:13)L2 . 1 = ≤ ≤ = (2I(u)) and by the second relation in (3.1.5), (cid:13)(cid:13)(cid:13)∂2 z u (cid:13)(cid:13)(cid:13)L2 Then, we add (cid:107)u(cid:107)H1 on both sides of (3.1.20) and obtain (cid:107)u(cid:107)H2 ≤ (2I(u)) 2 +(cid:13)(cid:13)W(cid:48)(u)(cid:13)(cid:13)L2 + (cid:107)u(cid:107)H1 . 1 Here using the bound for H1 norm of u constructed in Lemma 3.1.1, we have (cid:107)u(cid:107)H2 ≤ (2I(u)) ≤ (2I(u)) 1 2 +(cid:13)(cid:13)W(cid:48)(u)(cid:13)(cid:13)L2 + (cid:107)u(cid:107)H1 , (cid:32)(cid:13)(cid:13)(cid:13)(cid:13)δE (cid:13)(cid:13)(cid:13)(cid:13)2 2 +(cid:13)(cid:13)W(cid:48)(u)(cid:13)(cid:13)L2 + δu 1 L2 (cid:33) 1 2 + C , and by the second relation in (2.1.24), we obtain (cid:107)u(cid:107)H2 ≤ (2I(u)) 2 +(cid:13)(cid:13)W(cid:48)(u)(cid:13)(cid:13)L2 + (2I(u) + C) 1 1 2 . 41 (3.1.18) (3.1.19) (3.1.20) (3.1.21) (3.1.22) (3.1.23) Further, by the assumptions (3.1.1) we have (cid:13)(cid:13)W(cid:48)(u)(cid:13)(cid:13)2 L2 = c2 (cid:90) (cid:90) L2 ε |u|>1 (cid:90) |u|2pdz + c2− (cid:90) |u|2pdz + c2− ≤ c2 0 |u|2dz, |u|<1 |u|2dz, (3.1.24) |u|<1 ≤ c2 (cid:107)u(cid:107)2p ≤ c2 (cid:107)u(cid:107)2p L2p + c2− L2 H1 + c2− L2 ε ε . Taking square root of both sides of (3.1.24) and by Lemma 3.1.1, we deduce that there exists a c > 0 such that (cid:13)(cid:13)W(cid:48)(u)(cid:13)(cid:13)L2 ≤ c (cid:18) (cid:18) 1 + (cid:13)(cid:13)(cid:13)(cid:13)δE δu (cid:13)(cid:13)(cid:13)(cid:13)p (cid:19) (cid:19) = c 1 + (2I(u)) , , (3.1.25) L2 p 2 for p > 1. Inserting (3.1.25) in (3.1.23) we conclude that there exist constants α, γ > 0 such that (cid:107)u(cid:107)H2 ≤ α + γ (I(u)) p 2 , (3.1.26) for p > 1 and this provides the H2 coercivity of I(u). The other essential part of the proof is showing the weak lower semi-continuity of the energy. Since the energy functional is bounded below, there exists a minimizing sequence {uk}∞ k=1 ∈ A and so, I(uk) → m. (3.1.27) From the H2 coercivity of I(u), the sequence {uk} is bounded in H2 and there exists a 42 subsequence {ujk } and ¯u ∈ H2 such that (cid:42) ¯u weakly in H2, ujk (3.1.28) )n → ¯u in H1 since H1 ⊂⊂ H2. Before the and further, there exists a subsequence (ujk proof of weak lower semi-continuity of I(u), we need to verify that such ¯u resides in A. First (cid:42) ∂z ¯u weakly in L2 and ∂2 (cid:42) ∂2 z ¯u in L2. From integration by parts, z ujk observe that ∂zujk for any w ∈ H2(cid:16)(cid:104) 0, for any v ∈ C∞(cid:16)(cid:104) 0, we have (cid:17) (cid:16) ∂2 z w vdz = − (cid:90) L2 ε 0 ∂zw∂zv + (∂zw) v(cid:12)(cid:12) L2 ε 0 , (3.1.29) . We substitute w = ujk in (3.1.31).Since ujk ∈ A the boundary L2 ε (cid:105)(cid:17) (cid:90) L2 (cid:105)(cid:17) 0 ε L2 ε ε (cid:90) L2 (cid:105)(cid:17) 0 term is 0 and hence the equation (3.1.31) becomes (cid:90) L2 ε (cid:16) (cid:17) ∂2 z ujk vdz = − ε (cid:90) L2 (cid:105)(cid:17) 0 0, L2 ε 0 On the other hand, we substitute w = ¯u ∈ H2(cid:16)(cid:104) (cid:90) L2 (cid:17) (cid:16) ∂2 z ¯u vdz = − ε 0 for all v ∈ C∞(cid:16)(cid:104) 0, L2 ε ∂zujk ∂zv. (3.1.30) in (3.1.31) and obtain ∂z ¯u∂zv + (∂z ¯u) v(cid:12)(cid:12) L2 ε 0 , (3.1.31) . Consequently, since ∂zujk (cid:42) ∂z ¯u weakly in L2 and ∂2 z ujk (cid:42) ∂2 z ¯u in L2 and comparing the non-zero terms in (3.1.30) and (3.1.31) we conclude that (∂z ¯u) v(cid:12)(cid:12) L2 ε 0 = 0, 43 (3.1.32) for all v ∈ C∞(cid:16)(cid:104) (cid:105)(cid:17) 0, L2 ε . This proves that ∂z ¯u(0) = ∂z ¯u( ε ) and hence, ¯u ∈ A. L2 For the weak lower-semi continuity of I(u), we utilize the weak convergence of the first variation of F, (cid:16) δF δu Dujk , ujk , z (cid:17) δF δu (cid:42) which is established in Lemma 3.1.3. (D¯u, ¯u, z) weakly in L2, (3.1.33) Since weak convergence is lower semi continuous, (cid:90) L2 ε (cid:18) δF 0 δu (cid:16) Dujk , ujk , z (cid:17)(cid:19)2 dz ≥ (cid:90) L2 ε 0 (cid:18) δF δu (cid:19)2 (D¯u, ¯u, z) dz. lim inf Considering ¯u ∈ A, it follows that m ≤ I(¯u) ≤ lim inf I(ukj ) = m. Thus, the energy, I, attains its minima at ¯u. (3.1.34) (3.1.35) (cid:42) ¯u weakly in H2 and strongly in L2, then the variational derivative Lemma 3.1.3. If ujk of F, given in (3.1.3), converges δF δu (cid:105)(cid:17) . 0, L2 ε weakly in L2(cid:16)(cid:104) W(cid:48)(cid:16) ujk Proof. We already know that ∂2 (cid:17) → W(cid:48)(¯u) in L2(cid:16)(cid:104) z ujk L2 ε 0, uniformly bounded there exists ξjk (cid:16) (cid:17) δF δu (cid:42) Dujk , ujk , z (D¯u, ¯u, z) , (3.1.36) (cid:42) ∂2 z ¯u in L2(cid:16)(cid:104) (cid:105)(cid:17) (cid:105)(cid:17) ∈ L∞(cid:16)(cid:104) 0, L2 ε (cid:105)(cid:17) (cid:13)(cid:13)(cid:13)ξjk 0, L2 ε with 44 . By the mean value theorem and since ¯u, ujk and it suffices to show that (cid:13)(cid:13)(cid:13)L∞ uniformly bounded, are independent of ε, such that (cid:12)(cid:12)W(cid:48)(cid:16) ujk Squaring both sides of (3.1.37) and integrating over 0, (cid:13)(cid:13)(cid:13)W(cid:48)(cid:16) ujk (cid:13)(cid:13)(cid:13)L2 (cid:17) − W(cid:48)(¯u) weakly in H2(cid:16)(cid:104) which converges to zero since W(cid:48)(cid:48)(ξjk in L2(cid:16)(cid:104) (cid:105)(cid:17) and also, ujk (cid:105)(cid:17) L2 ε 0, 0, . L2 ε L2 ε − ¯u|. we obtain )||ujk (cid:105) (cid:17) − W(cid:48)(¯u)(cid:12)(cid:12) ≤ |W(cid:48)(cid:48)(ξjk (cid:104) ≤(cid:13)(cid:13)(cid:13)W(cid:48)(cid:48)(ξjk ≤(cid:13)(cid:13)(cid:13)W(cid:48)(cid:48)(ξjk ) is uniformly bounded in L∞(cid:16)(cid:104) → ¯u in L2(cid:16)(cid:104) . Thus, W(cid:48)(ujk (cid:13)(cid:13)(cid:13)L2 (cid:13)(cid:13)(cid:13)ujk (cid:13)(cid:13)(cid:13)L∞ − ¯u (cid:107)L2, (cid:105)(cid:17) − ¯u) )(ujk 0, L2 ε ) , (3.1.37) (3.1.38) (cid:105)(cid:17) 0, and ujk L2 (cid:42) ¯u ε ) converges W(cid:48)(¯u) 45 Chapter 4 Modulational Stability of n-pulses In this chapter, our main interest is the dynamical stability of a manifold of n-pulses given as the graph of an n-pulse ansatz, Φn(z, p, ¯λ). These are related to the periodic multi-pulse solutions, {Φn|n ∈ N}, constructed in Chapter 2 as a class of critical points to the free energy I. The H2-coercivity of the second variation of I about Φn(z, p, ¯λ), modulational stability and dynamic evolution of the n-pulse ansatz with respect to the Π0-gradient flow are addressed. 4.1 H 2-coercivity of the second variation of I In this section, we prove the H2-coercivity of the second variation of the free energy I about n-pulse ansatz, representation theorem δ2I δu2 (Φn(z, p), ¯λ). The second variation is defined from the Riesz (cid:12)(cid:12)(cid:12)τ =0 d2 dτ 2 I(u + τ v) (cid:68) δ2I δu2 (u)v, v (cid:69) , L2 = (4.1.1) for any v ∈ H2(R). Definition 4.1.1. : Let D be a subspace of Hilbert space H. A linear operator A : D → H satisfying (cid:104)Au, u(cid:105) ≥ µ(cid:107)u(cid:107)2 , ∀u ∈ D (4.1.2) 46 for some µ > 0 is called a coercive operator and µ is called the coercivity constant. 4.1.1 Introduction: n-pulses Recall that n-pulse ansatz defined on R is given in (2.2.5) as n(cid:88) (cid:0)z − pj (cid:1) + b−, φh (4.1.3) un := j=1 where φh := φh − b− and p = (p1, p2, ..., pn)t ∈ Rn is the vector of pulse locations. The admissible set of pulse locations is given by P = {p ∈ Rn : pi < pi+1 i = 1, ..., n and ∆p ≥ l}, for (4.1.4) where ∆p = min i(cid:54)=j |pi−pj| and l > 0 is sufficiently large so that the exponential terms e−√ α−l arising in the calculations are negligible. We extend un to be defined on all of R, and add a correction term that reduces the size of residual. Recalling the definition of B2,n given in (2.2.31), we introduce the corrected extension by Φn(z, p, ¯λ) := un + δ¯λB2,n, (4.1.5) and let Mn = {Φn(p, ¯λ)|p ∈ P} be the n-dimensional manifold formed by these solutions. Let Xn(p) represent the tangent plane to the manifold Mn, Xn(p) = span{ ∂Φn(p) : p ∈ ∂pi P ⊂ Rn} = span{φ(cid:48) h(z − pi)}n i=1. Recall that in (2.2.4) we introduced the operator Ln(p) := ∂2 z − W(cid:48)(cid:48) (un) , 47 (4.1.6) acting on H2 (R). The second variation of the free energy I about n-pulse ansatz is (cid:16) L := z − W(cid:48)(cid:48)(Φn) ∂2 (cid:17)2 −(cid:16) (cid:17) z Φn − W(cid:48)(Φn) ∂2 W(cid:48)(cid:48)(cid:48)(Φn). (4.1.7) Taylor expanding (4.1.7) about un up to O(δ2) terms we obtain L = (cid:17)2 −(cid:16) (cid:16)Ln − δ¯λW(cid:48)(cid:48)(cid:48)(un)B2,n + O(δ2) (cid:17)(cid:16) − δ¯λW(cid:48)(cid:48)(un)B2,n + O(δ2) + O(δ2) (cid:17) . z un − W(cid:48)(un) + δ¯λ∂2 ∂2 W(cid:48)(cid:48)(cid:48)(un) + δ¯λW (4)(un)B2,n+ zB2,n (4.1.8) (4.1.9) . Recalling that LnB2,n = B1,n we obtain (cid:16)Ln − δ¯λW(cid:48)(cid:48)(cid:48)(un)B2,n + O(δ2) (cid:17)2 −(cid:16) L = z un − W(cid:48)(un)+ ∂2 (cid:17) W(cid:48)(cid:48)(cid:48)(un) + δ¯λW (4)(un)B2,n + O(δ2) (cid:17)(cid:16) + δ¯λB1,n + O(δ2) (cid:105) for i = 1, . . . , n we have un = φi +O(e−√ z un − W(cid:48)(un) about φh(z − pi) and obtain ∂2 From the definition of un, partitioning the domain into sub-intervals about pulse locations as α−(cid:96)) on each sub-interval. We z un − W(cid:48)(un) = −W(cid:48)(cid:48)(φh(z − Taylor expand ∂2 pi+pi+1 (cid:104) pi−1+pi 2 2 , pi))e−√ √ α−(cid:96) + O(e−2 α−(cid:96)) on Summing up over all sub-intervals provides ∂2 Here we set δ = e−√ α−(cid:96). α−(cid:96) which will be our scaling value of the background parameter through i=1 W(cid:48)(cid:48)(φh(z − pi))e−√ since φh solves the equation (1.2.6). (cid:104) pi−1+pi 2 pi+pi+1 , (cid:105) z un − W(cid:48)(un) = − n(cid:80) 2 the rest of this thesis. 48 We insert this expansion into (4.1.9) and calculate the inner product (cid:17) W(cid:48)(cid:48)(cid:48)(un)+ (4.1.10) (cid:104)Lv, v(cid:105) = (cid:68)(cid:18)(cid:16)Ln − δ¯λW(cid:48)(cid:48)(cid:48)(un)B2,n + O(δ2) + −(cid:16)−δ n(cid:88) (cid:19) (cid:17)2 W(cid:48)(cid:48)(φh(z − pi)) + δ¯λB1,n (cid:69) i=1 + + O(δ2) v, v , for any v ∈ H2. Here we must be careful when expanding the quadratic term and simplifying further. Precisely, we have (cid:104)Lv, v(cid:105) = (cid:68)(cid:16)L2 nv − δ¯λLn(W(cid:48)(cid:48)(cid:48)(un)B2,nv) − δ¯λW(cid:48)(cid:48)(cid:48)(un)B2,nLnv + −(cid:16)−δ W(cid:48)(cid:48)(cid:48)(un)v+ (cid:69) W(cid:48)(cid:48)(φh(z − pi)) + δ¯λB1,n n(cid:88) (cid:17) i=1 v, v + O(δ2 (cid:107)v(cid:107)2 H2). (cid:17) + (4.1.11) δ2I Further, for the proof of H2-coercivity of L and later to verify H2-coercivity of δu2 (φn) := L2 n we establish and utilize H2-coercivity of the second variation of the inner scaling of FCH free energy about n-pulse ansatz at the leading order so it is worth noting that (cid:104)Lv, v(cid:105) = (cid:104)L2 nv, v(cid:105) + O(δ (cid:107)v(cid:107)2 H2). (4.1.12) 4.1.2 H 2-coercivity of the second variation of I about n-pulse ansatz For the stability analysis of the n-pulse ansatz purposes, we establish the H2-coercivity of the second variation of I about n-pulse ansatz given in (4.1.10). Theorem 4.1.2. Consider the inner scaling of FCH free energy, I, given in (1.2.3) and 49 n-pulse ansatz given in (4.1.5). Then, the bilinear form induced by L given in (4.1.10) is coercive, i.e, there exists a µ > 0 independent of ε such that (cid:104)Lv, v(cid:105) ≥ µ(cid:107)v(cid:107)H2(R), (4.1.13) for all v ∈ X⊥ n (p). The essential step in the proof of this theorem is the H2-coercivity of L2 n over X⊥ Before we present the proof of Theorem 4.1.2 we establish the H2-coercivity of L2 X⊥ n (p). n (p). n over Theorem 4.1.3. Consider the operator Ln given in (2.2.4) on H2 (R) which is the linearization of (1.2.6) about n-pulses, un, given in (2.2.5). Then, the bilinear form induced by L2 n is coercive, i.e, there exists a ˜µ > 0 independent of ε such that (cid:104)L2 nv, v(cid:105) ≥ ˜µ(cid:107)v(cid:107)H2(R), for all v ∈ X⊥ n (p). δ2I δu2 (un) = (4.1.14) The H2-coercivity of the operator L2 n(p) arises from the L2-coercivity of L2 n(p). We prove the L2-coercivity of L2 n(p) in Lemma 4.1.4. Lemma 4.1.4. There exists a ˜µ > 0 such that (cid:104)L2 n(p)v, v(cid:105) ≥ ˜µ(cid:107)v(cid:107)2 L2(R) , (4.1.15) for all v ∈ X⊥ n (p). 50 Proof. It suffices to show that there exists a ˜µ > 0 such that (cid:104)(L2 n(p) − ˜µ)v, v(cid:105) ≥ 0, (4.1.16) for all v ∈ X⊥ n (p). Ln 2 (p) − ˜µ : H2(R) ⊂ L2(R) → H−2(R) is a self-adjoint operator on L2(R). Let b[v, v] := (cid:104)(L2 n(p) − ˜µ)v, v(cid:105), (4.1.17) be the bilinear form associated to Ln induced by the constrained bilinear form defined by the restriction of b[v, v] to X⊥ (p) in the L2 inner product. We first find the operator 2 n (p). We introduce the orthogonal projection P : H2(R) → Xn(p) and Π := I − P which has the range X⊥ n (p). The bilinear form constrained to X⊥ n (p) induces the constrained operator (L2 n(p) − ˜µ)Π := Π(L2 n(p) − ˜µ)Π : H2(R) ∩ X⊥ n (p) → ΠH−2(R). (4.1.18) spectra of Π(L2 The operator Π(L2 n(p) − ˜µ)Π is self adjoint, so its spectrum is real-valued. If the point n(p) − ˜µ)Π is strictly positive for some values of ˜µ then we obtain (4.1.31). n(p) − ˜µ)Π is strictly positive. By It remains to show that the point spectra of Π(L2 Proposition 5.3.1 in [Kapitula and Promislow, 2013], number of the negative eigenvalues of the constrained operator is given by the difference between the number of the negative eigenvalues of the operator and the constrained matrix, D(˜µ),namely, n((L2 n(p) − ˜µ)Π) = n(L2 n(p) − ˜µ) − n(D(˜µ)), (4.1.19) 51 where n represents the count of the negative eigenvalues and D(˜µ) is given as Dij(˜µ) := (cid:104)φ(cid:48)(z − pi), (L2 n(p) − ˜µ)−1φ(cid:48)(z − pj)(cid:105). (4.1.20) We first calculate the number of negative eigenvalues of the operator L2 n(p) − ˜µ by directly examining the spectrum of the operator. Since u1 = φh(z − p1) and φh is translation invariant, the linearized operator L about φh and L1(p) are the same operators. Hence, we have that σ(L) = σ(L1(p)). On the other hand, the essential spectrum of L1 is σess(L1) = {−k2 − α− : k ∈ R}. (4.1.21) Further, σess(Ln(p)) = σess(L1(p)) by the classical Weyl essential spectrum theorem since the operators have the same limiting states, limz→±∞un. (See Chapter 3 in [Kapitula and Promislow, 2013].) On the other hand, to each point spectrum λk of L1(p), there are associated n eigenvalues of Ln(p), {λk,j}n |λk − λk,j| decays exponentially with growing pulse j=1 such that max j=1,...,n separation.(See [Sandstede, 1998].) Since Ln(p) is a self-adjoint operator, by the spectral mapping theorem σ(L2 n(p)) = (σ(Ln(p)))2. Recall that in Remark 2.2.1 we present that the eigenvalues of Ln has an order λ0 < 0 = λ1 < λ2 < . . . . By the choice of ˜µ > 0, the eigenvalues of L2 n(p) associated to the eigenvalue of L1(p) at λ1 = 0 are shifted to the left by ˜µ(See Figure 4.1). Choosing n(p) − ˜µ are ˜µ > 0 but less than the minimum of λ2 positive except n eigenvalues of Ln(p) associated to λ1 = 0, {λ1,j}n j=1. 0 and λ2 2 we see that the eigenvalues of L2 We conclude that n(L2 n(p) − ˜µ) = n. 52 (4.1.22) (a) Spectrum of Ln(p) (b) Spectrum of L2 n(p) Figure 4.1: Figure (a) is a depiction for the spectrum of Ln(p). In the descending order λ0 > 0 = λ1 > λ2 > ... (red disks) are the eigenvalues of L1(p) = L. Ln(p) has n associated eigenvalues(black crosses) to each localized eigenvalue of L1(p) such that |λk − λk,j|j=1,...,n decays exponentially with growing pulse separation. Figure (b) demonstrates the spectrum for L2 n(p). The next step is calculating the number of negative eigenvalues of the constrained matrix, n(D(˜µ)). The eigenfunctions of L2 n(p)− ˜µ corresponding to the eigenvalues {λ2 (cid:96)=1 are in h(z−pj) up to exponentially small terms. (See [Sandstede, 2001].) the form ψ1,i =(cid:80)n j=1 βijφ(cid:48) 1,i}n Using the definition of D we have the identity n(cid:88) j=1 Dij(˜µ)βij = (cid:104)ψ1,i, (L2 n(p) − ˜µ)−1φ(cid:48) h(z − pi)(cid:105). (4.1.23) Since inverse of a self-adjoint operator is also self-adjoint, transposing (L2 n(p) − ˜µ)−1 onto (4.1.24) ψ1,i provides n(cid:88) j=1 (cid:68) ψ1,i λ1,i − ˜µ (cid:69) . , φ(cid:48) h(z − pi) Dij(˜µ)βij = 53 σ(Ln)σess(Ln)λ2λ1=0λ0σ(L2n)σess(L2n)λ2λ1=0λ0222 Then, inserting the formula of ψ1,i we obtain j=1 βijφ(cid:48) h(z − pj) λ1,i − ˜µ (cid:69) h(z − pi) , φ(cid:48) (cid:68)(cid:80)n n(cid:88) j=1 Dij(˜µ)βij = + exp. small terms (4.1.25) = βii λ1,i − ˜µ + exp. small terms. Let A be the matrix of the coefficients βij of eigenfunction ψ1,i which is comprised of the vector of columns Bi = (βi1, βi2, . . . , βin)T . Then, by (4.1.25) we have DBi = 1 λ1,i − ˜µ Bi, (4.1.26) up to exponentially small terms. Hence, we conclude that a value of ˜µ ∈ R for which 1 λ1,i − ˜µ ∈ σ(D) and there exists min i=1,...,n λ2,i > ˜µ > max i=1,··· ,n λ1,i ∀ i = 1, . . . , n, (4.1.27) that guarantees n(D(˜µ)) = n. (4.1.28) If we insert (4.1.22) and (4.1.28) in (4.1.19), we obtain the desired result as n((L2 n(p) − ˜µ)Π) = 0, (4.1.29) which implies the L2 coercivity of L2 n(p), i.e. there exists a ˜µ > 0 independent of domain size such that (cid:104)L2 n(p)v, v(cid:105) ≥ ˜µ(cid:107)v(cid:107)2 L2 , (4.1.30) 54 for all v ∈ X⊥ n (p). Corollary 4.1.5. There exists a ˜µ1 > 0, independent of the domain size, such that (cid:104)L2 n(p)v, v(cid:105) ≥ ˜µ1 (cid:107)v(cid:107)2 H2 , (4.1.31) for all v ∈ X⊥ n (p). Proof. We already proved the L2-coercivity of the bilinear form induced by L2 n(p) in Lemma 4.1.4. We utilize this result to establish the H2-coercivity of the bilinear form induced by L2 n(p). By the L2-coercivity of L2 n, there exists a constant ˜µ > 0 such that (cid:104)Lnv,Lnv(cid:105) ≥ ˜µ(cid:107)v(cid:107)2 L2(R) , (4.1.32) for all v ∈ Xn ⊥(p). Expanding the inner product, we fix θ ∈ (0, 1) and write (cid:104)Lnv,Lnv(cid:105) = θ(cid:104)Lnv,Lnv(cid:105) + (1 − θ)(cid:104)Lnv,Lnv(cid:105), ∂2 z v + P ∂2 z v v + Qv2dz + (1 − θ)˜µ(cid:107)v(cid:107)2 L2 , (4.1.33) where P := −2W(cid:48)(cid:48)(un) and Q := (W (un))2. Applying the Holder’s inequality to the term with P ,(cid:13)(cid:13)v∂2 z v(cid:13)(cid:13)L1 (cid:16) (cid:17) we obtain (cid:17)2 2v(cid:13)(cid:13)L2 R (cid:90) ≥ θ (cid:16) (cid:13)(cid:13)∂z (cid:18)(cid:13)(cid:13)(cid:13)∂2 (cid:13)(cid:13)(cid:13)2 (cid:13)(cid:13)(cid:13)2 (cid:13)(cid:13)(cid:13)∂2 ≤ (cid:107)v(cid:107)L2 z v z v L2 (cid:104)Lnv,Lnv(cid:105) ≥ θ ≥ θ − (cid:107)P(cid:107)L∞ (cid:107)v∂zv(cid:107)L1 − (cid:107)Q(cid:107)L∞ (cid:107)v(cid:107)2 L2 , L2 L2 − θ (cid:107)P(cid:107)L∞ (cid:107)∂zv(cid:107)L2 (cid:107)v(cid:107)L2 + ((1 − θ)˜µ − θ (cid:107)Q(cid:107)L∞)(cid:107)v(cid:107)2 + (1 − θ)˜µ(cid:107)v(cid:107)2 L2 . (cid:19) (4.1.34) 55 We apply Young’s inequality to the second term on the second line of (4.1.34), (cid:19) L2 + ((1 − θ)˜µ − θ (cid:107)Q(cid:107)L∞)(cid:107)v(cid:107)2 L2 , (cid:107)P(cid:107)2 (cid:17)(cid:111) (cid:107)v(cid:107)2 L∞ L2 . (4.1.35) independent of domain size we . (4.1.36) (cid:13)(cid:13)(cid:13)2 (cid:13)(cid:13)(cid:13)∂2 (cid:18) L∞ (cid:107)v(cid:107)2 (1 − θ)˜µ − θ (cid:107)Q(cid:107)L∞ − θ 2 − θ 2 (cid:107)P(cid:107)2 z v L2 − θ 2 + (cid:104)Lnv,Lnv(cid:105) ≥ θ θ 2 = L2 z v (cid:13)(cid:13)(cid:13)2 (cid:13)(cid:13)(cid:13)∂2 (cid:13)(cid:13)(cid:13)2 (cid:13)(cid:13)(cid:13)∂2 (cid:16) (cid:110) θ z v L2 , 2 Choosing θ∗ = min (1 − θ)˜µ − θ (cid:107)Q(cid:107)L∞ − θ 2 (cid:107)P(cid:107)2 L∞ obtain (cid:104)Lnv,Lnv(cid:105) ≥ θ∗ (cid:19) + (cid:107)v(cid:107)2 L2 (cid:18)(cid:13)(cid:13)(cid:13)∂2 z v (cid:13)(cid:13)(cid:13)2 L2 = θ∗ (cid:107)v(cid:107)2 H2 . Proof of Theorem 4.1.2. We show the H2-coercivity of the second variation of I about n-pulse ansatz using bilinear form given in (4.1.11). (cid:68)Lv, v (cid:69) = (cid:68)(cid:18) + −(cid:16)−δ n(cid:88) (cid:69) − δ¯λ (cid:68)Ln(W(cid:48)(cid:48)(cid:48)(un)B2,nv), v n(cid:88) (cid:17) nv − δ¯λLn(W(cid:48)(cid:48)(cid:48)(un)B2,nv) − δ¯λW(cid:48)(cid:48)(cid:48)(un)B2,nLnv L2 (cid:17) (cid:69) W(cid:48)(cid:48)(cid:48)(un)v, v (cid:69) − δ¯λ (cid:68) W(cid:48)(cid:48)(cid:48)(un)B2,nLnv, v (cid:69) (cid:69) − −δ¯λ (cid:68) W(cid:48)(cid:48)(cid:48)(un)B1,nv, v W(cid:48)(cid:48)(φh(z − pi)) + δ¯λB1,n nv, v W(cid:48)(cid:48)(cid:48)(un) W(cid:48)(cid:48)(φh(z − pi))v, v − δ i=1 + (cid:104)Lv, v(cid:105) ≥(cid:68)L2 (cid:68) (4.1.37) + O(δ2 (cid:107)v(cid:107)2 H2) (cid:69) + + (4.1.38) i=1 + O(δ2 (cid:107)v(cid:107)2 H2). From Corollary 4.1.5, there exists a ˜µ1 > 0 independent of ε such that (cid:104)L2 nv, v(cid:105) ≥ ˜µ1 (cid:107)v(cid:107)2 H2 56 and hence (cid:68)Ln(W(cid:48)(cid:48)(cid:48)(un)B2,nv), v H2 − δ¯λ n(cid:88) W(cid:48)(cid:48)(cid:48)(un) W(cid:48)(cid:48)(φh(z − pi))v, v (cid:69) − δ¯λ (cid:68) W(cid:48)(cid:48)(cid:48)(un)B2,nLnv, v (cid:69) (cid:69) − δ¯λ (cid:68) W(cid:48)(cid:48)(cid:48)(un)B1,nv, v + (cid:69) + (4.1.39) (cid:104)Lv, v(cid:105) ≥ ˜µ1 (cid:107)v(cid:107)2 (cid:68) − δ i=1 + O(δ2 (cid:107)v(cid:107)2 H2). By the smoothness of the functions W over bounded functions and the smoothness of B1,n and B2,n and by the definition of Ln we get the following estimates for the inner products in (4.1.39). Since Ln is self-adjoint, transposing that onto v we have (cid:68)Ln(W(cid:48)(cid:48)(cid:48)(un)B2,nv), v (cid:69) = W(cid:48)(cid:48)(cid:48)(un)B2,nv,Lnv (cid:68) (cid:69) =(cid:13)(cid:13)W(cid:48)(cid:48)(cid:48)(un)B2,nvLnv(cid:13)(cid:13)L1 (cid:13)(cid:13)L∞(cid:13)(cid:13)vLnv(cid:13)(cid:13)L1 ≤(cid:13)(cid:13)W(cid:48)(cid:48)(cid:48)(un)(cid:13)(cid:13)L∞(cid:13)(cid:13)B2,n (4.1.40) (4.1.41) Then, we apply the Holder’s inequality to the term with L1 norm (cid:68)Ln(W(cid:48)(cid:48)(cid:48)(un)B2,nv), v (cid:69) ≤(cid:13)(cid:13)W(cid:48)(cid:48)(cid:48)(un)(cid:13)(cid:13)L∞(cid:13)(cid:13)B2,n (cid:13)(cid:13)L∞(cid:13)(cid:13)Lnv(cid:13)(cid:13)L2 (cid:107)v(cid:107)L2 . Here we use the fact that there exists a constant c > 0 such that(cid:13)(cid:13)Lnf(cid:13)(cid:13)L2 ≤ c(cid:107)f(cid:107)H2 for all f ∈ H2(R) and get (cid:68)Ln(W(cid:48)(cid:48)(cid:48)(un)B2,nv), v (cid:69) ≤ c(cid:13)(cid:13)W(cid:48)(cid:48)(cid:48)(un)(cid:13)(cid:13)L∞(cid:13)(cid:13)B2,n (cid:13)(cid:13)L∞ (cid:107)v(cid:107)H2 (cid:107)v(cid:107)L2 ≤ θ1 (cid:107)v(cid:107)2 H2 , (4.1.42) for some θ1 > 0. With a similar calculation, we get an upper bound for the second inner 57 product,(cid:68) (cid:69) W(cid:48)(cid:48)(cid:48)(un)B2,nLnv, v = (cid:13)(cid:13)(cid:13)W(cid:48)(cid:48)(cid:48)(un)B2,nLnv2(cid:13)(cid:13)(cid:13)L1 (cid:13)(cid:13)(cid:13)Lnv2(cid:13)(cid:13)(cid:13)L1 (cid:13)(cid:13)L∞ ≤(cid:13)(cid:13)W(cid:48)(cid:48)(cid:48)(un)(cid:13)(cid:13)L∞(cid:13)(cid:13)B2,n (cid:13)(cid:13)L∞(cid:13)(cid:13)Lnv(cid:13)(cid:13)L2 (cid:107)v(cid:107)L2 ≤(cid:13)(cid:13)W(cid:48)(cid:48)(cid:48)(un)(cid:13)(cid:13)L∞(cid:13)(cid:13)B2,n (cid:13)(cid:13)L∞(cid:13)(cid:13)Lnv(cid:13)(cid:13)2 ≤(cid:13)(cid:13)W(cid:48)(cid:48)(cid:48)(un)(cid:13)(cid:13)L∞(cid:13)(cid:13)B2,n H2 ≤ θ2 (cid:107)v(cid:107)2 H2 , (4.1.43) for some θ2 > 0. For the remaining two inner products, there exist constants θ3 > 0 and θ4 > 0 such that and (cid:68) W(cid:48)(cid:48)(cid:48)(un) (cid:68) n(cid:88) i=1 W(cid:48)(cid:48)(φh(z − pi)v, v (cid:69) ≤ θ3 (cid:107)v(cid:107)2 H2 , W(cid:48)(cid:48)(cid:48)(un)B1,nv, v (cid:69) ≤ θ4 (cid:107)v(cid:107)2 H2 . (4.1.44) (4.1.45) Inserting the estimates (4.1.42), (4.1.43), (4.1.44) and (4.1.45) in (4.1.39) provides us (cid:104)Lv, v(cid:105) ≥ ˜µ1 (cid:107)v(cid:107)2 H2 − δ¯λθ1 (cid:107)v(cid:107)2 − δθ3 (cid:107)v(cid:107)2 H2 − δ¯λθ4 (cid:107)v(cid:107)2 H2 − δ¯λθ2 (cid:107)v(cid:107)2 H2 + H2 + O(δ2 (cid:107)v(cid:107)2 H2). (4.1.46) Choosing the ˜µ2 = max{¯λθ1, ¯λθ2, θ3, ¯λθ4}, we obtain the H2-coercivity of δ2I δu2 (Φn) as (cid:104)Lv, v(cid:105) ≥ ˜µ1 (cid:107)v(cid:107)2 H2 − δ ˜µ2 (cid:107)v(cid:107)2 H2 . (4.1.47) We also show the H2-coercivity of the L2 n(p) restricted to the bounded domain [0, L2 ε ] 58 that is an essential step in the verification of the assumption (H3) in Section 4.2. Corollary 4.1.6. Let the operator Ln,b(p) acting on H2(cid:16)(cid:104) (cid:105)(cid:17) 0, L2 ε be the restriction of the Ln(p) given in (2.2.4) to the large bounded domain [0, by L2 n,b(p) is coercive, i.e, there exists a µ∗ > 0 independent of ε such that L2 ε ]. Then, the bilinear form induced (cid:104)L2 n,b(p)v, v(cid:105) ≥ µ∗(cid:107)v(cid:107) (cid:18)(cid:20) H2 0, L2 ε (cid:21)(cid:19), (4.1.48) for all v ∈ Y (cid:48) n(p). Proof. Let ˜v ∈ Y (cid:48) n(p) have a compact support. Let v ∈ H2(R) be the extension of ˜v to R so that we have (cid:104)L2 n,b(p)v, v(cid:105) = (cid:104)L2 n(p)v, v(cid:105), (4.1.49) for all v ∈ X⊥ n (p). By Corollary 4.1.5, there exists a ˜µ1 such that (cid:104)L2 n,b(p)v, v(cid:105) = (cid:104)L2 n(p)v, v(cid:105) ≥ ˜µ1(cid:107)v(cid:107)H2(R), n (p). Since H2(cid:16)(cid:104) 0, L2 ε (cid:105)(cid:17) with compact support is dense in H2(cid:16)(cid:104) for all v ∈ X⊥ (4.1.50) (cid:105)(cid:17) 0, L2 ε the result follows. 4.1.3 H 2-coercivity of L2 n For the stability of the periodic multi-pulse solutions, Φn, given in (2.2.37) of the free energy I, the H2-coercivity of the second variation of I about Φn at leading order is established utilizing the H2-coercivity of the operator L2 n. See Appendix for the derivation of second variation of I about Φn at leading order. 59 Theorem 4.1.7. Consider the operator Ln given in (2.2.27) on H2(cid:16)(cid:104) (cid:105)(cid:17) 0, L2 ε which is the linearization of (1.2.6) about periodic multi-pulse solutions, φn, at the leading order. Then, n is coercive, i.e, there exists a µ∗ > 0 independent the bilinear form induced by δ2I δu2 (un) = L2 of ε such that for all v ∈ X⊥ n (p). nv, v(cid:105) ≥ µ∗(cid:107)v(cid:107)H2(R), (cid:104)L2 (4.1.51) Lemma 4.1.8. There exist smooth functions (p, η) : H2 → RN × H2 satisfying p(0) = p∗ such that for all (cid:107)v(cid:107)H2 = O(1) sufficiently small Φn + v = Φn(z, p, ¯λ) + η(v), (4.1.52) where η(v) ∈ X⊥ n (p). Further, there exists α > 0 constant such that (cid:107)p(v) − p∗(cid:107)L2 ≤ α(cid:107)v(cid:107)H2, (4.1.53) where v ∈ X⊥ n (p). Proof. Introduce F = (F1, ..., Fn)t where Fi(v, p) := (cid:104)v + Φn(p) − Φn(p), φ(cid:48) n(z − pi)(cid:105), (4.1.54) for i = 1, ..., n and (cid:104)·,·(cid:105) is L2 inner product. Fi = 0 for each i since η(v) = v+Φn(¯p)−Φn(p) ∈ X⊥ n (p). Indeed, Fi for each i attains its minima for some values of v on the compact set of p ∈ P ⊂ Rn. Assume that F attains its minima at p(v0) and so, p = p(v0) is one solution to F (v, p) = 0 for v = v0. We apply the Implicit Function Theorem to F (v, p). The ij th 60 entry of the gradient ∇pF is (cid:18) ∇pF (cid:12)(cid:12)(cid:12)(v0,p(v0) (cid:19) = ij = h(z − pi)(cid:105)(cid:17)(cid:12)(cid:12)(cid:12)(v0,p(v0)) , i = j, i (cid:54)= j,   (cid:16)−(cid:104)∂Φn(p) , φ(cid:48) ∂pi h(z − pi)(cid:105)+ + (cid:104)v + Φn − Φn(p),−φ(cid:48)(cid:48) , φ(cid:48) h(z − pi(v(0)), φ(cid:48) h(z − pi)(cid:105)(cid:12)(cid:12)(cid:12)(v0,p(v0)) h(z − pi(v0))(cid:105)(cid:17) + (cid:104)h,−φ(cid:48)(cid:48) (cid:104) ∂Φn(p) ∂pj (cid:16)−(cid:104)φ(cid:48) , h(z − pi(v0))(cid:105)+ , 0, i = j, i (cid:54)= j, (4.1.55) where h := v0 + Φn − Φn(p(v0)) is small with (cid:107)h(cid:107)H2 = O(δ). By a proper choice of far apart pulse locations p ∈ P, φ(cid:48)(z − pi) satisfy  κ for 0 for (φ(cid:48) h(z − pi), φ(cid:48) h(z − pj)2 = i = j, i (cid:54)= j, (4.1.56) for constant κ > 0. From this relation, we write the gradient as (cid:12)(cid:12)(cid:12)(v0,p(v0) ∇pF = −κI + O(δ). (4.1.57) We see that ∇pF (cid:12)(cid:12)(cid:12)(v0,p(v0) has non-zero determinant, and from this the Implicit Function Theorem implies that there exists a neighborhood (v0, p(v0)) and a unique function p(v) such that F (p(v), v) = 0. Further, by the Implicit Function Theorem p is smooth since Fi 61 is smooth. Thus, the Taylor expansion of p about v = 0 yields p(v) = p(0) + (cid:104) δp δv (0), v(cid:105)H2 + O((cid:107)v(cid:107)2 H2). (4.1.58) From this, we obtain |p(v) − p(0)| ≤ (cid:13)(cid:13)(cid:13)(cid:13) δp δv (cid:13)(cid:13)(cid:13)(cid:13)H−2 (0) for a constant θ > 0. (cid:107)v(cid:107)2 H2 + c((cid:107)v(cid:107)2 H2) ≤ θ(cid:107)v(cid:107)2 H2, (4.1.59) Proof of Theorem 4.1.7. We prove the H2-coercivity of L2 L2 n(p). Expanding the bilinear form induced by L2 by L2 n(p) and utilizing H2 coercivity of the bilinear form induced by L2 n(p) from Corollary n(p) using H2-coercivity of n(p) in terms of the bilinear form induced 4.1.5, we obtain a lower bound for the first term in the expansion, (cid:104)L2 n(p)v, v(cid:105) = (cid:104)L2 n(p)v, v(cid:105) + (cid:104)(L2 H2 + (cid:104)(L2 n(p) − L2 n(p) − L2 n(p))v, v(cid:105), n(p))v, v(cid:105), ≥ ˜µ1 (cid:107)v(cid:107)2 (4.1.60) for all v ∈ X⊥ n (p) with (cid:107)v(cid:107)H2 (cid:28) 1. Hence, it suffices to show that the remaining term in the expansion (4.1.60) is small. We attack this term splitting it into (cid:104)(L2 n(p) − L2 n(p))v, v(cid:105) = (cid:104)(L2 n(p) − L2 n(p∗))v, v(cid:105) + (cid:104)(L2 n(p∗) − L2 n(p))v, v(cid:105). (4.1.61) 62 Let u∗ n := un(z, p∗). (cid:17) n(p∗) − L2 n(p) (cid:16)L2 Define v, (cid:17)2 z − W(cid:48)(cid:48)(un) ∂2 n) − 2∂zv∂z(W(cid:48)(cid:48)(u∗ z vW(cid:48)(cid:48)(u∗ n)(cid:1)2 n)) +(cid:0)W(cid:48)(cid:48)(u∗ z vW(cid:48)(cid:48)(un) − 2∂zv∂z(W(cid:48)(cid:48)(un)) +(cid:0)W(cid:48)(cid:48)(un)(cid:1)2 (cid:0)W(cid:48)(cid:48)(un)(cid:1) , (4.1.62) (cid:17) (cid:17) v v , . v = (cid:16) (cid:16) −(cid:16) = z z z − W(cid:48)(cid:48)(u∗ ∂2 n) z v − ∂2 ∂4 z v − ∂2 ∂4 (cid:17)2 v −(cid:16) n)(cid:1) v − 2∂2 (cid:0)W(cid:48)(cid:48)(u∗ (cid:0)W(cid:48)(cid:48)(un)(cid:1) v − 2∂2 A0(z, p) :=(cid:0)W(cid:48)(cid:48)(un)(cid:1)2 − ∂2 (cid:0)W(cid:48)(cid:48)(un)(cid:1) , A1(z, p) := −2∂z A2(z, p) := −2W(cid:48)(cid:48)(un), z (4.1.63) (4.1.64) which are smooth functions in z. Then, (cid:16)L2 n(p∗) − L2 n(p) (cid:17) v = (A0(z, p∗) − A0(z, p)) v, + (A1(z, p∗)∂zv − A1(z, p)) ∂zv, (cid:17) (cid:16) z v − A2(z, p) A2(z, p∗)∂2 + ∂2 z v, and so, by the Mean Value Theorem for several variable functions, there exist ξ0, ξ1, ξ2 ∈ [p(0), p(v)] so that (cid:10)(cid:16)L2 n(p∗) − L2 n(p) (cid:17) v, v(cid:11) ≤ (cid:90) L2 ε 0 (A0(z, p) − A0(z, p∗)) v2 + (A1(z, p) − A1(z, p∗)) v∂zv + (A2(z, p) − A0(z, p)) v∂2 (cid:90) L2 z vdz, ε |∇pA0(ξ0)||p(v) − p∗|v2 + |∇pA1(ξ1)||p(v) − p∗|v∂zv ≤ 0 + |∇pA2(ξ2)||p(v) − p∗|v∂2 z vdz, 63 ε (cid:90) L2 ≤(cid:13)(cid:13)∇pA0(ξ0)(cid:13)(cid:13)L∞ |p(v) − p∗|v2dz+ (cid:90) L2 +(cid:13)(cid:13)∇pA1(ξ1)(cid:13)(cid:13)L∞ (cid:90) L2 +(cid:13)(cid:13)∇pA2(ξ2)(cid:13)(cid:13)L∞ |p(v) − p∗|v∂2 |p(v) − p∗|v2dz+ 0 0 ε ε z vdz. 0 Utilizing the bound for |p(v) − p∗| in Lemma 4.1.8, (cid:10)(cid:16)L2 n(p) − L2 (cid:17) n(p∗) v, v(cid:11) ≤ α(cid:107)v(cid:107)H2 (cid:90) L2 ε 0 v2dz (cid:18)(cid:13)(cid:13)∇pA0(ξ0)(cid:13)(cid:13)L∞ (cid:90) L2 (cid:90) L2 +(cid:13)(cid:13)∇pA1(ξ1)(cid:13)(cid:13)L∞ +(cid:13)(cid:13)∇pA2(ξ2)(cid:13)(cid:13)L∞ 0 ε ε 0 v∂zvdz+ (cid:19) v∂2 z vdz . {(cid:13)(cid:13)∇pAj(ξj)(cid:13)(cid:13)L∞}. (cid:17) v, v(cid:11) ≤ Cα(cid:107)v(cid:107)H2 n(p∗) Now let C = max 1≤j≤3 (cid:10)(cid:16)L2 n(p) − L2 (4.1.65) (4.1.66) On the other hand, a similar calculation gives a bound for the second term in (4.1.60) (cid:104)(L2 n(p) − L2 n(p∗))v, v(cid:105) = ≤ (cid:16) (cid:90) L2 (cid:90) L2 0 ε ε 0 (cid:18) (cid:107)v(cid:107)2 H2 + (cid:107)∂zv(cid:107)L2 (cid:107)v(cid:107)L2 + (cid:107)∂zv(cid:107)L2 (cid:107)v(cid:107)L2 (cid:19) , ≤ C1(cid:107)v(cid:107)3 H2. (4.1.67) (cid:17)2 −(cid:16) (cid:17)2 z − W(cid:48)(cid:48) (φn) ∂2 ( z − W(cid:48)(cid:48) (u∗ ∂2 n) )v2dz, (A0(z) − A0(z, p∗)) v2 + (A1(z) − A1(z, p∗)) v∂zv, + (A2(z) − A0(z, p∗)) v∂2 ≤ C2 (cid:107)v(cid:107)2 H2 . z vdz (4.1.68) 64 Hence, inserting (4.1.67) and (4.1.68) in (4.1.61), (cid:104)(L2 n(p) − L2 n(p))v, v(cid:105) ≥ −C1 (cid:107)v(cid:107)3 H2 − C2 (cid:107)v(cid:107)2 H2 , and (4.1.69) in (4.1.60) we obtain the H2-coercivity (cid:104)L2 n(p)v, v(cid:105) ≥ ˜µ1 (cid:107)v(cid:107)2 ≥ α(cid:107)v(cid:107)2 H2 + (cid:104)(L2 H2 − C1 (cid:107)v(cid:107)3 n(p) − L2 n(p))v, v(cid:105), H2 . (4.1.69) (4.1.70) 4.2 Modulational Stability of n-Pulses We establish the modulational stability for n-pulse ansatz, Φn, with respect to the Π0 gradient flow of the inner scaling of the FCH energy given (1.2.3) with an application of Theorem 2.1 in [Promislow, 2002] where modulational stability of manifolds of quasi- stationary solutions to dispersive equations is established. Introduce the Π0-gradient flow of the inner scaling of the FCH free energy, I, given (1.2.3) ut = −Π0 δI δu (u), (4.2.1) where δI δu is the first variational derivative of I with respect to L2 inner product and Π0 is the mass preserving L2 projection Π0f := f − ε L2 65 (cid:90) L2 ε 0 f (z)dz. (4.2.2) Consider the family of multi-pulse critical points {Φn : n ∈ N} of the free energy I and the n-dimensional manifold Mn = {Φn(p, ¯λ)|p ∈ P} where P ⊂ Rn defined in (4.1.4). For the modulational stability of the n-pulses {Φn : n ∈ N}, we apply renormalization techniques from [Promislow, 2002]. We are interested in the evolution of the solutions which lie in a neighborhood of the manifold Mn which consists of the n-pulse solutions Φn(p, ¯λ) for p ∈ P. Introduce u(z, t) = Φn(z, p, ¯λ) + w(z). (4.2.3) We reduce the dynamics of (4.2.1) near the manifold Mn to a weakly non-linear flow which is predominantly controlled by the terms that are linear in their deviation, w, of U from Mn, and non-linear in the pulse evolution p = p(t) about Φn(p, ¯λ). Taylor expanding the flow (4.2.1) about Φn(p, ¯λ) we obtain − Π0 δI δu (u(z, t)) = −Π0 (Φn(p, ¯λ)) − Π0L2 n(p)w + N (w), (4.2.4) (cid:18) δI δu (cid:19) where Ln(p) given (2.2.4) is the linearization about un and N represents the non-linear terms. The linearization about un will be weakly time dependent through the slow evolution of the pulse positions, p, and the background state ¯λ. The n + 1 parameters form the coordinates of the slow n-pulse manifold, one of which is determined by the mass constraint. In Theorem 2.1 of [Promislow, 2002], there are some assumptions on the linearized operator and the manifold of the steady-state solutions to the gradient flow. Here we present those assumptions adapting to the linearized operator −Π0L2 n(p) and the manifold Mn but defer the verification of each assumption to the proof of Theorem 4.2.1. (H0) The manifold Mn is quasi-steady, i.e, for δ > 0, the scaling of ¯λ, there exists M > 0 66 such that (cid:13)(cid:13)(cid:13)(cid:13)−Π0 δI δu (cid:13)(cid:13)(cid:13)(cid:13)H2 (Φn(p, ¯λ)) ≤ M δ. (4.2.5) (H1) The spectrum of each operator −Π0L2 n(p) consists of a stable part σs ⊂ {λ|λ ≤ −ks} for some ks > 0 and a slow part σ0 ⊂ {λ||λ| ≤ c0e− k0 ε } for some c0, k0 > 0. (H2) Each operator −Π0L2 n(p) generates a C0 semigroup Sp which satisfies (cid:13)(cid:13)Sp(t)u(cid:13)(cid:13)H2 ≤ M e−kst (cid:107)u(cid:107)H2 , n (p)∩ H2(cid:16)(cid:104) (cid:105)(cid:17) 0, L2 ε , where X⊥ n (p) is perpendicular to the tangent (4.2.6) for all t ≥ 0, u ∈ Y (cid:48) n := X⊥ plane Xn(p) of Φn(p, ¯λ). Let Yp represent the slow space of the linearized operator −Π0L2 n(p), the n+1 dimensional space associated with small eigenvalues of −Π0L2 n(p). Recall that Xn(p) = span{ ∂Φn(p) ∂pi (cid:110){φ(cid:48) plane to the manifold Mn. Introduce Xn+1(p) = span (cid:18)(cid:20) : p ∈ P ⊂ Rn} = span{φ(cid:48) h(z − pi)}n h(z − pi)}n (cid:21)(cid:19) i=1 ∪ {B2,n}(cid:111) and i=1 is the tangent Y (cid:48) n+1(p) := X⊥ n+1(p) ∩ H2 0, L2 ε , (4.2.7) where X⊥ n+1(p) is the orthogonal space to Xn+1(p). (H3) Yp is well-approximated by Xn+1(p). In [Promislow, 2002], this assumption is utilized to establish the coercivity of the linearized operator. Instead, here we will establish the H2-coercivity of Π0L2 of L2 n+1(p) which follows from the H2-coercivity n(p). Recall that H2-coercivity of L2 n(p) over the space Y (cid:48) n,b(p) over Y (cid:48) n,b(p) over Y (cid:48) n(p) was established in Corollary 4.1.6 using Corollary 4.1.5. We assume that the adjoint of the elements in Yp and Xn+1(p) satisfy the following 67 normalization condition 1 0 (cid:104)ζi, ζ † j(cid:105) = for i = j, for i (cid:54)= j. With this condition, the adjoint of Xn+1(p) is defined as † n+1(p):=Xn(p)∪span{1}. X (H4) The normalized adjoint eigenvectors {ψ † 1, ..., ψ † n+1} of the space Yp satisfy (cid:16)(cid:13)(cid:13)(cid:13)ψ † i (p) (cid:13)(cid:13)(cid:13)H2 max i=1,...,n p∈P for some M . (cid:13)(cid:13)(cid:13)∇2 + † i (p) pψ (cid:17) ≤ M, (cid:13)(cid:13)(cid:13)H2 (4.2.8) (4.2.9) (4.2.10) Theorem 4.2.1. Fix a pulse separation value (cid:96) = O(ε−1) > 0 in P ⊂ Rn. Then, there exists a manifold Mn = {Φn(p, ¯λ)|p ∈ P} satisfying the hypothesis (H0)-(H4) for some constants M and k and there exist ε0, M0 for ε ∈ [0, ε0] such that for all initial data u(z, t0) = u0(z) within ε2-neighborhood in H2-norm of Mn whose mass lies within δ-neighborhood of the mass of Φn(p, ¯λ), the solution u of (4.2.1) can be decomposed as u(z) = Φn(z, p(t), ¯λ) + w(z, t), (4.2.11) where the deviation w ∈ Y (cid:48) n+1(p(t)) satisfies (cid:107)w(·, t)(cid:107)H2 ≤ M0(ε2e−ks(t−t0) + δ) f or t ≥ t0. (4.2.12) The pulse locations p(t) = (p1, . . . , pn)t may be chosen to lie on a smooth curve in P. After 68 an initial transient T i ∼ 1| ln(ε)|, that is, for t > t0 + T i, the evolution of the pulse locations is governed to leading order by the closed system (cid:68) − Π0 p(cid:48) i = δI δu (Φn(p, ¯λ)), 1(cid:13)(cid:13)φ(cid:48) (cid:13)(cid:13)2 h L2 φ(cid:48) h(z − pi) (cid:69) + O(δ2) f or t ≥ t0 + T i, (4.2.13) for i = 1, . . . , n. Proof. Here we verify that the Mn and the linearized operator −Π0L2 n(p) satisfy the hypothesis (H0)-(H4) and a direct application of Theorem 2.1 from [Promislow, 2002] provides the result. (H0) We prove that the manifold Mn is quasi-steady, namely, (cid:13)(cid:13)(cid:13)(cid:13)−Π0 δI δu (cid:13)(cid:13)(cid:13)(cid:13)H2 (Φn(p, ¯λ)) ≤ M δ, for some M > 0. Recall that Φn(p, ¯λ) = un + δ¯λB2,n where un =(cid:80)n j=1 φh with φh = φh − b−. The Taylor series expansion of δI (4.2.14) (cid:0)z − pj (cid:1) + b− δI δu (Φn(p, ¯λ)) = = = (cid:16) (cid:16) (cid:16) δu about un provides (cid:17)(cid:16) (cid:17) z Φn(p) − W(cid:48)(Φn(p)) z − W(cid:48)(cid:48)(Φn(p)) (cid:17)(cid:16) ∂2 ∂2 z − W(cid:48)(cid:48)(un + δ¯λB2,n (cid:17)(cid:16) ∂2 z − W(cid:48)(cid:48)(un) − δ¯λB2,nW(cid:48)(cid:48)(cid:48)(un) z (B2,n)+ z un + δ¯λ∂2 ∂2 ∂2 δ2¯λ2(B2,n)2W(cid:48)(cid:48)(cid:48)(un) − W(cid:48)(un) − δ¯λB2,nW(cid:48)(cid:48)(un)) − 1 (cid:105) (cid:17) z (un + δ¯λB2,n) − W(cid:48)(un + δ¯λB2,n) ∂2 (cid:17) 2 . (cid:104) pi−1+pi Since the tail-tail interaction of the adjacent pulses dominates the value of un, on each we can write un = φi−1 + φi + φi+1 where φi = φh(z − pi) and δu(Φn(p, ¯λ)) about φi we window φi = φi − b−. Letting φ∆i := φi−1 + φi+1 and Taylor expanding δI pi+pi+1 2 2 , (4.2.15) 69 obtain δI δu (Φn(p, ¯λ)) = i=1 ∂2 (cid:1)(cid:17)(cid:16) (cid:16) (cid:17)(cid:16) (cid:17) z − W(cid:48)(cid:48)(Φn(p, ¯λ)) z Φn(p, ¯λ) − W(cid:48)(Φn(p, ¯λ)) (cid:16) n(cid:88) z − W(cid:48)(cid:48)(cid:0)φi + φ∆i + δλB2,n ∂2 ∂2 (cid:17) − W(cid:48)(cid:16) z − W(cid:48)(cid:48)(φi) − W(cid:48)(cid:48)(cid:48)(φi)(cid:0)φ∆i + δ¯λB2,n)(cid:1)(cid:17)(cid:16) (cid:16) n(cid:88) zB2,n − W(cid:48)(φi) − W(cid:48)(cid:48)(φi)(cid:0)φ∆i + δ¯λB2,n φi + φ∆i + δλB2,n + δ¯λ∂2 ∂2 i=1 = = z (φi + φ∆i + δλB2,n) ∂2 z φi + ∂2 ∂2 z φ∆i+ (cid:1)(cid:17) . (4.2.16) Letting Li = ∂2 written as δI δu (Φn(p, ¯λ)) = i = W(cid:48)(φi), δI z − W(cid:48)(cid:48)(φi) and taking into account that φ(cid:48)(cid:48) (cid:16)Li − W(cid:48)(cid:48)(cid:48)(φi) (cid:16) n(cid:88) (cid:16)Li − W(cid:48)(cid:48)(cid:48)(φi)φ∆i n(cid:88) (cid:17)Liφ∆i + δ¯λ φ∆i + δ¯λB2,n i=1 = (cid:17)(cid:17)(cid:16)Liφ∆i + δ¯λLiB2,n (cid:16)Li − W(cid:48)(cid:48)(cid:48)(φi)φ∆i (cid:17) (cid:17)LiB2,n+ δu(Φn(p, ¯λ)) can be (4.2.17) i=1 − δ¯λW(cid:48)(cid:48)(cid:48)(φi)B2,nLiφ∆i + O(δ2). Applying −Π0 onto (4.2.17) and recalling the definition of B2,n given in (2.2.12) we rearrange 70 the terms −Π0 δI δu (Φn(p, ¯λ)) = − n(cid:88) Π0 i=1 = − n(cid:88) i=1 Π0 (cid:17)LiB2,n+ (cid:18)(cid:16)Li − W(cid:48)(cid:48)(cid:48)(φi)φ∆i (cid:16)L2 (cid:17)Liφ∆i + δ¯λ (cid:16)Li − W(cid:48)(cid:48)(cid:48)(φi)φ∆i (cid:19) − δ¯λW(cid:48)(cid:48)(cid:48)(φi)B2,nLiφ∆i + O(δ2) (cid:19) − δ¯λW(cid:48)(cid:48)(cid:48)(φi)B2,nLiφ∆i + O(δ2) i φ∆i − W(cid:48)(cid:48)(cid:48)(φi)φ∆iLiφ∆i + δ¯λ − δ¯λW(cid:48)(cid:48)(cid:48)(φi)φ∆iB1,n+ . Since Π0 onto any constant is 0, Π0δ¯λ = 0 and hence, −Π0 δI δu (Φn(p, ¯λ)) = n(cid:88) (cid:16)−Π0L2 i=1 + δ¯λΠ0 (cid:0)W(cid:48)(cid:48)(cid:48)(φi)φ∆iLiφ∆i (cid:1)(cid:17) (cid:0)W(cid:48)(cid:48)(cid:48)(φi)B2,nLiφ∆i i φ∆i + Π0 + O(δ2). (4.2.18) (cid:1) + δ¯λΠ0 (cid:0)W(cid:48)(cid:48)(cid:48)(φi)φ∆iB1,n (cid:1) + Then, we calculate the H2-norm of (4.3.5). By the triangle inequality we have (cid:13)(cid:13)(cid:13)(cid:13)−Π0 δI δu (Φn(p, ¯λ)) (cid:13)(cid:13)(cid:13)(cid:13)H2 i φ∆i (cid:18)(cid:13)(cid:13)(cid:13)−Π0L2 ≤ n(cid:88) + δ¯λ(cid:13)(cid:13)Π0 + δ¯λ(cid:13)(cid:13)Π0 (cid:13)(cid:13)(cid:13)H2 +(cid:13)(cid:13)Π0 (cid:0)W(cid:48)(cid:48)(cid:48)(φi)φ∆iLiφ∆i (cid:1)(cid:13)(cid:13)H2 + (cid:0)W(cid:48)(cid:48)(cid:48)(φi)φ∆iB1,n (cid:1)(cid:13)(cid:13)H2 (cid:0)W(cid:48)(cid:48)(cid:48)(φi)B2,nLiφ∆i + O(δ2). (cid:19) i=1 (4.2.19) (cid:1)(cid:13)(cid:13)H2 + (4.2.20) Since Π0 is a H2-orthogonal projection, for any function u ∈ H2 we have the estimate (cid:107)Π0u(cid:107)H2 ≤ (cid:107)u(cid:107)H2 , (4.2.21) 71 (cid:13)(cid:13)H2 + (4.2.22) i=1 i φ∆i (cid:13)(cid:13)(cid:13)H2 +(cid:13)(cid:13)W(cid:48)(cid:48)(cid:48)(φi)φ∆iLiφ∆i (cid:19) (cid:18)(cid:13)(cid:13)(cid:13)L2 ≤ n(cid:88) + δ¯λ(cid:13)(cid:13)(cid:0)W(cid:48)(cid:48)(cid:48)(φi)φ∆iB1,n (cid:1)(cid:13)(cid:13)H2 + (cid:1)(cid:13)(cid:13)H2 + δ¯λ(cid:13)(cid:13)(cid:0)W(cid:48)(cid:48)(cid:48)(φi)B2,nLiφ∆i (cid:13)(cid:13)(cid:13)W (k)(un) + O(δ2). and hence the (4.2.20) becomes (cid:13)(cid:13)(cid:13)(cid:13)−Π0 δI δu (Φn(p, ¯λ)) (cid:13)(cid:13)(cid:13)(cid:13)H2 We know that the functions Bj,n ∈ H2 for j = 1, 2 and by the smoothness of W , for k ≤ 4 (cid:13)(cid:13)(cid:13)L∞ ≤ α1. Here the value of α1 depends upon there exists an α1 > 0 such that the uniform bound on (cid:107)un(cid:107)L∞. Utilizing all these facts we get upper bound for the H2 norm of -Π0 δI δu(Φn(p)), (cid:13)(cid:13)(cid:13)(cid:13)−Π0 δI δu (Φn(p, ¯λ)) (cid:13)(cid:13)(cid:13)(cid:13)H2 +(cid:13)(cid:13)W(cid:48)(cid:48)(cid:48)(φi)(cid:13)(cid:13)L∞ (cid:107)φ∆iLiφ∆i(cid:107)H2 + i=1 i φ∆i (cid:13)(cid:13)(cid:13)H2 (cid:18)(cid:13)(cid:13)(cid:13)L2 ≤ n(cid:88) + δ¯λ(cid:13)(cid:13)W(cid:48)(cid:48)(cid:48)(φi)(cid:13)(cid:13)L∞(cid:13)(cid:13)(cid:0)φ∆iB1,n (cid:1)(cid:13)(cid:13)H2 + + δ¯λ(cid:13)(cid:13)W(cid:48)(cid:48)(cid:48)(φi)(cid:13)(cid:13)L∞(cid:13)(cid:13)(cid:0)B2,nLiφ∆i (cid:1)(cid:13)(cid:13)H2 (cid:18)(cid:13)(cid:13)(cid:13)L2 ≤ n(cid:88) + α1 (cid:107)φ∆iLiφ∆i(cid:107)H2 + i φ∆i (cid:19) + O(δ2) (cid:13)(cid:13)(cid:13)H2 (cid:13)(cid:13)(cid:0)φ∆iB1,n (cid:1)(cid:13)(cid:13)H2 + (cid:1)(cid:13)(cid:13)H2 (cid:13)(cid:13)(cid:0)B2,nLiφ∆i i=1 + δ¯λα1 + δ¯λα1 (cid:19) + O(δ2). (4.2.23) 72 Here we use the fact that the product of two H2 functions lies in H2(See Appendix.), (cid:13)(cid:13)(cid:13)(cid:13)−Π0 δI δu (Φn(p, ¯λ)) (cid:13)(cid:13)(cid:13)(cid:13)H2 (cid:18)(cid:13)(cid:13)(cid:13)L2 ≤ n(cid:88) i=1 i φ∆i (cid:13)(cid:13)(cid:13)H2 (cid:13)(cid:13)B1,n (cid:13)(cid:13)H2 + (cid:13)(cid:13)H2 (cid:107)Liφ∆i(cid:107)H2 (cid:13)(cid:13)B2,n + δ¯λα1 (cid:107)φ∆i(cid:107)H2 + δ¯λα1 + α1 (cid:107)φ∆i(cid:107)H2 (cid:107)Liφ∆i(cid:107)H2 + (cid:19) + O(δ2). (4.2.24) H2+2k for some c1 > 0 and obtain ≤ c1 (cid:107)f(cid:107) (cid:18) (cid:107)φ∆i(cid:107)H6 + α1 (cid:107)φ∆i(cid:107)H2 (cid:107)Liφ∆i(cid:107)H2 + (cid:13)(cid:13)H2 + (cid:13)(cid:13)B1,n (cid:19) (cid:13)(cid:13)H2 (cid:107)φ∆i(cid:107)H2 + δ¯λα1 (cid:107)φ∆i(cid:107)H2 + δ¯λα1 (cid:13)(cid:13)B2,n (4.2.25) + O(δ2). We also use the fact that (cid:13)(cid:13)(cid:13)(cid:13)−Π0 δI δu (Φn(p, ¯λ)) nf (cid:13)(cid:13)(cid:13)Lk (cid:13)(cid:13)(cid:13)(cid:13)H2 (cid:13)(cid:13)(cid:13)H2 ≤ n(cid:88) i=1 (cid:104) pi−1+pi 2 pi+pi+1 2 , On the window (cid:105) , we write φ∆i = φmaxe−√ (cid:104) pi−1+pi pi+pi+1 , 2 2 (cid:105) for k = 1, 2, . . . on M > 0 such that where φmax is the amplitude of the pulse. Using this form of φ∆i, we obtain (cid:107)φ∆i(cid:107) α−(z−pi−1) + φmaxe−√ α−(pi+1−z), (4.2.26) Hk = O(δ) . Hence, we conclude that there exists a constant (cid:13)(cid:13)(cid:13)(cid:13)−Π0 δI δu (cid:13)(cid:13)(cid:13)(cid:13)H2 (Φn(p, ¯λ)) ≤ M δ. (4.2.27) (H1) The spectrum of each operator −Π0L2 n(p) consists of a stable part σs and a slow part σ0. The spectrum of L2 n(p) has been examined in detail in the proof of Lemma (4.1.4). 73 It can be easily seen that σ(−Π0L2 n(p)) = σ(−Π0L2 n(p)Π0) ∪ {0} and further σ(−Π0L2 n(p)Π0) = σs ∪ σ0, (4.2.28) where σs ⊂ {λ|λ ≤ −ks} for some ks > 0 and σ0 ⊂ {λ||λ| ≤ c0e− k0 ε } for some k0, c0 > 0 and σ0 consists of n + 1 eigenvalues. (H2) Each operator −Π0L2 n(p) generates a C0 semigroup Sp which satisfies (cid:13)(cid:13)Sp(t)u(cid:13)(cid:13)H2 ≤ M e−kst (cid:107)u(cid:107)H2 , (4.2.29) for all t ≥ 0, u ∈ X⊥ n+1(p). This estimate on the semigroup is a result of Pr¨uss-Gearhart Theoremwhich states that the boundedness of a C0-semigroup generated by an operator is from boundedness of the resolvent on the right half plane(See [Gearhart, 1978] and [Pr¨uss, 1984]). To complete the verification, we need to show the boundedness of the resolvent of the constrained operator over the space X⊥ constrained to the space X⊥ Pp : H2(R) → Xn+1(p). Introduce the constrained operator Πp(−Π0L2 generating the constrained semigroup onto X⊥ of −Π0L2 n+1(p) because the semigroup is n+1(p). Let Πp = I − Pp be the orthogonal projection where n(p)Π0)Πp that is n(p)Π0, it is easy to see that , σ(Πp(−Π0L2 n+1(p). Since we already obtained the spectrum n(p)Π0)) \ σ0 n(p)Π0)Πp) = σ((−Π0L2 and hence σ(Πp(−Π0L2 self-adjoint operator and so, ζ ∈ R(Πp(−Π0L2 n(p)Π0)Πp) ⊂ {λ|λ ≤ −ks}. On the other hand, Πp(−Π0L2 n(p)Π0)Πp) = {z ∈ C|(Πp(−Π0L2 n(p)Π0)Πp is a n(p)Π0)Πp − 74 ζ) is bijective} and λ ∈ σ(Πp(−Π0L2 n(p)Π0)Πp), (cid:13)(cid:13)(cid:13)(cid:13)(cid:16) (cid:17)−1(cid:13)(cid:13)(cid:13)(cid:13) ≤ sup λ Πp(−Π0L2 n(p)Π0)Πp − ζ |λ − ζ|−1, (4.2.30) (See [Kato, 1976]). (H3) As mentioned earlier, here we establish the H2-coercivity of Π0L2 X⊥ n+1(p) which is a result of the H2-coercivity of L2 (4.1.5) that there exists a ˜µ1 > 0, independent of domain size, such that n(p) over X⊥ n(p) over the space n (p). Recall from Corollary (cid:104)L2 n(p)v, v(cid:105) ≥ ˜µ1 (cid:107)v(cid:107)2 H2 , for all v ∈ X⊥ Let v ∈ X⊥ n (p). n (p). Then, w = Π0v ∈ X⊥ n+1(p) ⊂ X⊥ n (p) and (cid:104)Π0L2 n(p)w, w(cid:105) = (cid:104)Π0L2 n(p)Π0v, Π0v(cid:105) = (cid:104)L2 = (cid:104)L2 = (cid:104)L2 0v(cid:105) n(p)Π0v, Π2 n(p)Π0v, Π0v(cid:105) n(p)w, w(cid:105) ≥ ˜µ1 (cid:107)w(cid:107)2 H2 , which proves the H2-coercivity of Π0L2 n(p) over the space X⊥ n+1(p). (H4) The normalized adjoint eigenvectors {ψ1, ..., ψn+1} of the space Xn+1(p) which are same with the normalized eigenvectors of the space since L2 n(p) is self-adjoint satisfy (cid:16)(cid:107)ψi(p)(cid:107)H2 + max i=1,...,n p∈P (cid:13)(cid:13)(cid:13)∇2 (cid:17) ≤ M, (cid:13)(cid:13)(cid:13)H2 pψi(p) (4.2.31) 75 for some M since φ(cid:48) h(z − pi) ∈ Cn for some n. Analyticity of the eigenvectors of an unbounded, self-adjoint operator with compact resolvent is proved in details in [Kriegl et al., 2011]. Following the verification of these conditions required in the main theorem in [Promislow, 2002], we apply the theorem to problem 4.2.1 and obtain the desired result. 76 4.3 Pulse Dynamics In this section, we study the ODE’s given in (4.2.13) to understand the evolution of an initial data given in a neighborhood of the manifold Mn. Reformulating the evolution equations of pulse locations given in [Promislow, 2002] we obtain n + 1 evolution equations, n equations for the pulse locations and an equation for the background parameter ¯λ, i =(cid:10)−Π0 p(cid:48) δI δu (Φn(p, ¯λ)), h(z − pi)(cid:11) + O(δ2), φ(cid:48) 1(cid:13)(cid:13)φ(cid:48) (cid:13)(cid:13)2 h L2 for i = 1, . . . , n and the evolution of background parameter ¯λ(cid:48) =(cid:10)−Π0 δI δu (Φn(p, ¯λ)), 1(cid:11), (4.3.1) (4.3.2) where 1 is the adjoint eigenfunction corresponding to B2,n which is one of the n+1 eigenfunctions of −Π0 δu(Φn(p, ¯λ)). δI Recall that the n-pulse ansatz is given as Φn(p, ¯λ)) = un + δλB2,n with un defined in (2.2.5). Note that the mass of an n-pulse configuration is M = nMh + b−L2 ε + δ¯λ L2 εα2− , (4.3.3) (cid:90) L2 ε 0 (φh − b−)dz is the mass of the homoclinic solution. where Mh = Remark 4.3.1. We assume that n is a fixed number, independent of ε, and |pi+1 − pi| ≥ (cid:96) for all i = 1, . . . , n. These choices provide us δ := e−√ α−(cid:96) (cid:28) εp for all p. 77 Let the initial mass of the polymer added to the solvent be M0 = nMh + b−L2 ε + ¯M0, (4.3.4) where ¯M0 is the excess mass remaining after the formation of n-pulses. Evaluating the inner product given in (4.3.1) and (4.3.2) we would like to construct explicit ODE’s for the pulse locations and the background state ¯λ to study their evolutions. Inserting the expression obtained in (4.2.17) in the inner product in (4.3.1) we have (cid:28) p(cid:48) i = − Π0 (cid:18)(cid:16)Li − W(cid:48)(cid:48)(cid:48)(φi)φ∆i (cid:17)Liφ∆i + δ¯λ (cid:16)Li − W(cid:48)(cid:48)(cid:48)(φi)φ∆i (cid:17)LiB2,n+ (cid:19) (cid:29) (cid:13)(cid:13)2 1(cid:13)(cid:13)φ(cid:48) − δ¯λW(cid:48)(cid:48)(cid:48)(φi)B2,nLiφ∆i + O(δ2) φ(cid:48) , . i i L2 (4.3.5) Remark 4.3.2. Since Π0 is self-adjoint, we project that onto φ(cid:48) φ(cid:48) i + O(δ). Hence, the higher order terms in p(cid:48) i are O(δ2). i and obtain that Π0φi (cid:48) = p(cid:48) i = − (cid:28) = − (cid:28)(cid:16)Li − W(cid:48)(cid:48)(cid:48)(φi)φ∆i (cid:17)Liφ∆i + δ¯λ − δ¯λW(cid:48)(cid:48)(cid:48)(φi)B2,nLiφ∆i, i φ∆i − W(cid:48)(cid:48)(cid:48)(φi)φ∆iLiφ∆i + δ¯λL2 L2 − δ¯λW(cid:48)(cid:48)(cid:48)(φi)B2,nLiφ∆i, (cid:16)Li − W(cid:48)(cid:48)(cid:48)(φi)φ∆i (cid:17)LiB2,n+ (cid:29) 1(cid:13)(cid:13)φ(cid:48) (cid:13)(cid:13)2 (cid:29) iB2,n − δ¯λW(cid:48)(cid:48)(cid:48)(φi)φ∆iLiB2,n+ + O(δ2) Π0φ(cid:48) L2 i i + O(δ2). 1(cid:13)(cid:13)φ(cid:48) (cid:13)(cid:13)2 i L2 φ(cid:48) i (4.3.6) 78 (cid:28) Here since L2 i is self-adjoint and φ(cid:48) i ∈ ker(Li), we have (cid:104)L2 i (·), φ(cid:48) i(cid:105) = 0. − W(cid:48)(cid:48)(cid:48)(φi)φ∆iLiφ∆i − δ¯λW(cid:48)(cid:48)(cid:48)(φi)φ∆iLiB2,n − δ¯λW(cid:48)(cid:48)(cid:48)(φi)B2,nLiφ∆i, p(cid:48) i = − + O(δe−√ (cid:28) W(cid:48)(cid:48)(cid:48)(φi)φ∆iLiφ∆i, φ(cid:48) (cid:28) W(cid:48)(cid:48)(cid:48)(φi)(φ∆iLiB2,n + B2,nLiφ∆i), α−(cid:96)) (cid:29) + δ¯λ = i (cid:13)(cid:13)2 1(cid:13)(cid:13)φ(cid:48) i L2 (cid:29) φ(cid:48) i L2 1(cid:13)(cid:13)φ(cid:48) (cid:13)(cid:13)2 (cid:29) i φ(cid:48) i + O(δ2). (4.3.7) Transposing W(cid:48)(cid:48)(cid:48)(φi) to the right of the inner products and writing W(cid:48)(cid:48)(cid:48)(φi)φ(cid:48) we obtain p(cid:48) i = (cid:13)(cid:13)2 1(cid:13)(cid:13)φ(cid:48) i L2 (cid:28) φ∆iLiφ∆i,(cid:0)W(cid:48)(cid:48)(φi)(cid:1) (cid:124) (cid:123)(cid:122) z A (cid:29) (cid:125) +δ¯λ (cid:13)(cid:13)2 1(cid:13)(cid:13)φ(cid:48) i L2 (cid:28) (cid:124) + O(δ2). φ∆iLiB2,n + B2,nLiφ∆i,(cid:0)W(cid:48)(cid:48)(φi)(cid:1) i = (W(cid:48)(cid:48)(φi))z, (cid:29) (cid:125) z (cid:123)(cid:122) B For the calculation of A we use the explicit formula of Li, (cid:68) (cid:68) (cid:69) φ∆iLiφ∆i,(cid:0)W(cid:48)(cid:48)(φi)(cid:1) (cid:69) −(cid:68) ∆i,(cid:0)W(cid:48)(cid:48)(φi)(cid:1) φ∆iφ(cid:48)(cid:48) z z A = = ∆i,(cid:0)W(cid:48)(cid:48)(φi)(cid:1) z W(cid:48)(cid:48)(φi)φ2 (4.3.8) (4.3.9) (cid:69) . (cid:16) 1 Taking W(cid:48)(cid:48)(φi) to the other side in the second inner product and observing that W(cid:48)(cid:48)(φi)(W(cid:48)(cid:48)(φi))z = (cid:0)W(cid:48)(cid:48)(φi)(cid:1)2(cid:17) , we obtain 2 z (cid:68) ∆i,(cid:0)W(cid:48)(cid:48)(φi)(cid:1) φ∆iφ(cid:48)(cid:48) z (cid:69) −(cid:68) φ2 ∆i, (cid:16)(cid:0)W(cid:48)(cid:48)(φi)(cid:1)2(cid:17) (cid:69) . z 1 2 (4.3.10) A = To evaluate these inner products, we write φi = b− + φmaxe−√ α−|z−pi| for all i = 1,··· , n 79 where φmax represents the maximum value of a single pulse in the n-pulse configuration, and α−(pi+1−z). we have φ∆i = φmaxe−√ on the window pi+pi+1 2 (cid:105) , 2 (cid:104) pi−1+pi (cid:68) (φmaxe−√ −(cid:68) φmaxe−√ A = α− α−(z−pi−1) + φmaxe−√ α−(z−pi−1) + φmaxe−√ α−(z−pi−1)+φmaxe−√ (cid:69) α−(pi+1−z))2,(cid:0)W(cid:48)(cid:48)(φi)(cid:1) (cid:16)(cid:0)W(cid:48)(cid:48)(φi)(cid:1)2(cid:17) (cid:69) α−(pi+1−z), + z . 1 2 z We integrate by parts the terms in the last two lines of (4.3.11) and obtain √ α−(cid:10)φ2 α−(cid:10)φ2 maxe−2 √ maxe−2 √ A = −2α− √ + α−(pi+1−z) − φ2 α−(pi+1−z) − φ2 maxe−2 √ maxe−2 √ α−(z−pi−1), W(cid:48)(cid:48)(φi)(cid:11)+ α−(z−pi−1),(cid:0)W(cid:48)(cid:48)(φi)(cid:1)2(cid:11). Further, writing the explicit formula for φi and Taylor expanding W(cid:48)(cid:48)(φi) about b− we obtain (4.3.11) (4.3.12) (4.3.13) (cid:68) A = − 2α− √ + √ = −2α− α− (cid:68) √ α− + α−(z−pi−1), α−(z−pi−1), α− √ + √ √ α− (cid:68) α−(pi+1−z) − φ2 √ maxe−2 φ2 W(cid:48)(cid:48)(b− + φmaxe−√ (cid:68) maxe−2 (cid:16) φ2 W(cid:48)(cid:48)(b− + φmaxe−√ √ α−(pi+1−z) − φ2 maxe−2 φ2 W(cid:48)(cid:48)(b−) + W(cid:48)(cid:48)(cid:48)(b−)φmaxe−√ √ √ maxe−2 maxe−2 (cid:16) φ2 W(cid:48)(cid:48)(b−) + W(cid:48)(cid:48)(cid:48)(b−)φmaxe−√ √ maxe−2 (cid:69) α−|z−pi|) maxe−2 α−(pi+1−z) − φ2 (cid:17)2(cid:69) α−|z−pi|) √ α−(z−pi−1), maxe−2 α−|z−pi|(cid:69) α−|z−pi|(cid:17)2(cid:69) α−(pi+1−z) − φ2 α−(z−pi−1), + . 80 Recall that W(cid:48)(cid:48)(b−) = α− and define γ− := W(cid:48)(cid:48)(cid:48)(b−). Rearranging (4.3.13) provides A = −2α− √ α−(cid:10)φ2 α−(cid:10)φ2 √ + √ α−(pi+1−z) − φ2 maxe−2 α− + γ−φmaxe−√ √ α−(pi+1−z) − φ2 maxe−2 α2− + 2α−γ−φmaxe−√ α−(z−pi−1), √ √ maxe−2 α−|z−pi|(cid:11)+ maxe−2 α−|z−pi| + γ2−φ2 √ α−(z−pi−1), maxe−2 (4.3.14) α−|z−pi|(cid:11). √ √ α− to the right side of For further simplification we move the constants −2α− α− and the inner product. A =(cid:10)φ2 √ maxe−2 +(cid:10)φ2 maxe−2 √ α−α2− + 2 − 2α2− √ α−(z−pi−1), √ α−(pi+1−z) − φ2 maxe−2 √ √ α−γ−φmaxe−√ α− − 2α− α−(pi+1−z) − φ2 maxe−2 √ α−α−γ−φmaxe−√ α−|z−pi| + α−(z−pi−1), √ √ α−|z−pi|(cid:11)+ (4.3.15) α−|z−pi|(cid:11). √ maxe−2 α−γ2−φ2 Adding up these two inner products and grouping or canceling common terms we obtain √ maxe−2 √ − α2− √ A =(cid:10)φ2 =(cid:10)e−2 (cid:124) +(cid:10)e−2 (cid:124) √ α−φ2 √ α− + γ2− α−(pi+1−z) − e−2 √ I √ α−(pi+1−z) − e−2 α−(pi+1−z) − φ2 α−(z−pi−1), α−|z−pi|(cid:11) √ maxe−2 √ maxe−2 (cid:123)(cid:122) α−(z−pi−1),−φ2 α−(z−pi−1), γ2− √ (cid:123)(cid:122) II maxα2− √ α−(cid:11) (cid:125) maxe−2 α−φ4 (cid:104) pi−1+pi 2 pi+pi+1 2 , (cid:105) (4.3.16) α−|z−pi|(cid:11) (cid:125) . √ . Let mi = pi−1+pi 2 and We evaluate these inner products on the interval 81 mi+1 = pi+pi+1 2 . We start by calculating the first inner product I, mi+1(cid:90) √ (e−2 √ I = −α2− α−φ2 max √ α−(pi+1−z) − e−2 α−(z−pi−1))dz = −α2−φ2 max 2 = −α2−φ2 max 2 + e−√ mi √ √ α−(pi+1−z) + e−2 (e−2 (cid:16) α−(pi+1−pi) − e−√ e−√ α−(pi+1+pi−2pi−1) − e−√ α−(z−pi−1)) α−(2pi+1−pi−pi−1) α−(pi−pi−1)(cid:17) . (cid:12)(cid:12)(cid:12)(cid:12)mi+1 mi Similarly, we calculate the second inner product and obtain √ II = γ2− α−φ4 α−(pi+1−z) − e−2 √ α−(z−pi−1))e−2 √ α−|z−pi|)dz mi+1(cid:90) √ (e−2 α−(pi+1−pi−1)+ 4 √ max α−(pi+1−pi)(cid:17) (cid:16)1 mi √ (e−2 α−(pi+1−pi) − e−2 √ (pi+1 − pi)e−2 (cid:16)1 α−(pi−pi−1)+ (pi − pi−1)e−2 2 √ √ (e−2 α−(pi−pi−1) − e−2 α−(pi+1−pi−1)(cid:17) max √ + . max + = γ2−φ4 1 2 − γ2−φ4 1 4 + (4.3.17) (4.3.18) Then, adding I and II together and simplifying we obtain A = −α2−φ2 max 2 + e−√ (cid:16) α−(pi+1−pi) − e−√ e−√ α−(pi+1+pi−2pi−1) − e−√ (cid:16)1 α−(pi+1−pi) − 1 4 √ (pi+1 − pi)e−2 √ e−2 α−(pi+1−pi) − 1 2 √ e−2 max 4 + γ2−φ4 1 2 + α−(2pi+1−pi−pi−1) α−(pi−pi−1)(cid:17) + α−(pi−pi−1)+ √ (pi − pi−1)e−2 α−(pi−pi−1)(cid:17) . (4.3.19) 82 With similar calculations we evaluate the last inner product B. Taking B2,n = α−2− away from pulses provides φ∆iLiB2,n = −α−2− W(cid:48)(cid:48)(φi)φ∆i and B2,nLiφ∆i = α−2− φ(cid:48)(cid:48) ∆i−α−2− W(cid:48)(cid:48)(φi)φ∆i. Combining these two in the inner product B making further simplifications we obtain B = (cid:69) (cid:68) φ∆iLiB2,n + B2,nLiφ∆i,(cid:0)W(cid:48)(cid:48)(φi)(cid:1) (cid:68) (cid:68) (cid:69) + α−2− = −2α−2− (cid:68) (cid:68) ∆i,(cid:0)W(cid:48)(cid:48)(φi)(cid:1) + α−2− = −α−2− φ(cid:48)(cid:48) W(cid:48)(cid:48)(φi)φ∆i,(cid:0)W(cid:48)(cid:48)(φi)(cid:1) (W(cid:48)(cid:48)(φi))2(cid:17) (cid:16) (cid:69) ∆i,(cid:0)W(cid:48)(cid:48)(φi)(cid:1) (cid:69) φ∆i, φ(cid:48)(cid:48) z z . z z (cid:69) z (4.3.20) Here recall that φ(cid:48)(cid:48)(cid:48) ∆i = α−φ(cid:48) ∆i. Expanding W(cid:48)(cid:48)(φi) about b− and grouping common terms in these two inner products give B = α−2− = α−2− (cid:68) (cid:68) ∆i, (W(cid:48)(cid:48)(φi))2(cid:69) − α−1− φ(cid:48) φ(cid:48) (cid:68) ∆i, α2− + 2γ−α−φmaxe−√ φ(cid:48) (cid:68) −√ α−|z−pi| − α−1− φ(cid:48) α−|z−pi|(cid:69) ∆i, α− + γ e ∆i, e−√ = α−1− γ−φmax φ(cid:48) ∆i,(cid:0)W(cid:48)(cid:48)(φi)(cid:1)(cid:69) α−|z−pi| + γ2−φ2 maxe−2 (cid:69) (cid:68) + α−2− γ2−φ2 (cid:68) max √ ∆i, e−2 φ(cid:48) √ α−|z−pi|(cid:69) α−|z−pi|(cid:69) (4.3.21) . Then, we calculate the inner products in B. mi+1(cid:90) B = α−1− γ−φmax √ α−(e−√ α−(pi+1−z) − e−√ α−(z−pi−1))e−√ α−|z−pi|+ mi + α−2− γ2−φ2 max √ α−(e−√ α−(pi+1−z) − e−√ √ α−(z−pi−1))e−2 α−|z−pi|dz (4.3.22) 83 = 1 2 + 3 √ max max α−(pi+1−pi−1)) α−(pi−pi−1))+ − 1 2 max(e−√ α−(pi+1−pi) − e−√ √ α−(pi − pi−1)e−√ (cid:16)1 (e−√ α−(pi+1−pi) − e (cid:17) α−(pi−pi−1)) − 3 α−(pi−pi−1) − e + 2 max(e−√ α−(pi+1−pi−1) − e−√ α−(pi−pi−1)) √ α−(pi+1 − pi)e−√ (cid:16)1 − 1 2 α−(pi+1−pi))+ α−(3pi+1−pi−2pi−1) − e−√ max √ √ (e α−1− γ−φ2 α−1− γ−φ2 − 1 2 + α−2− γ2−φ3 − (e−√ α−1− γ−φ2 α−1− γ−φ2 + α−2− γ2−φ3 − (e − 3 2 1 2 1 2 √ + (cid:17) α−(pi+1−pi)) . max 3 α−(pi+1−pi) − e−√ α−(2pi+1+pi−3pi−1))+ (4.3.23) α−(pi−pi−1))+ Inserting (4.3.19), and (4.3.23) in (4.3.8) we obtain the evolution equations of pulse locations, (cid:16) p(cid:48) i = (cid:18) 1(cid:13)(cid:13)φ(cid:48) (cid:13)(cid:13)2 + e−√ L2 i + γ2−φ4 1 2 + + α−(2pi+1−pi−pi−1) 4 + max e−√ √ e−2 α−(pi+1−pi) − e−√ α−(pi−pi−1)(cid:17) α−(pi−pi−1)+ √ (pi − pi−1)e−2 α−(pi+1−pi) − e−√ −α2−φ2 max 2 α−(pi+1+pi−2pi−1) − e−√ (cid:16)1 √ α−(pi+1−pi) − 1 e−2 4 √ (pi+1 − pi)e−2 α−(pi+1−pi) − 1 (cid:16) (cid:16)1 δ¯λ(cid:13)(cid:13)φ(cid:48) (cid:13)(cid:13)2 2 (e−√ α−1− γ−φ2 α−((pi+1 − pi)e−√ α−(pi+1−pi) − (pi − pi−1)e−√ (cid:16)1 α−(pi+1−pi) − e−√ (e−√ α−(pi−pi−1)) − (e 1 3 + α−2− γ2−φ3 − e−√ α−(3pi+1−pi−2pi−1) − e α−(pi+1−pi) − e − 1 2 − 3 2 − 1 2 √ max max L2 √ √ √ (e + + 2 3 i + α−(pi−pi−1)(cid:17)(cid:19) α−(pi−pi−1)(cid:17) α−(pi−pi−1)) + (e−√ α−(pi−pi−1))+ + √ − 3 2 α−(pi−pi−1))+ (cid:17) α−(2pi+1+pi−3pi−1)) α−(pi+1−pi)+ . (4.3.24) Observing that(cid:13)(cid:13)φ(cid:48) the big terms e−√ (cid:13)(cid:13)2 √ φ2 max 2 α− i L2 = and neglecting exponentially small terms comparing to α−(cid:96)i where (cid:96)i represents the distance between the centers of two adjacent 84 pulses we conclude that p(cid:48) i = −α 3/2− (cid:16) e−√ α−(pi+1−pi) − e−√ α−(pi−pi−1)(cid:17) + O(δ2), (4.3.25) assuming p0 = −p1 and pn+1 = 2 ε − pn. L2 After we derived the ODE’s for pulse locations we also construct the ODE for the background parameter ¯λ evaluating the inner product in (4.3.2). ¯λ(cid:48) =(cid:10)−Π0 δI δu (Φn(p, ¯λ)), 1(cid:11). (4.3.26) Since Π0 is self-adjoint and Ker(Π0) = span{1}, we move Π0 to the right side of the inner product and obtain ¯λ(cid:48) = 0. (4.3.27) From this we conclude that there are no background dynamics. Hence, the ODE system, consisting of n + 1 evolution equations for n pulse positions and ¯λ, is ¯λ(cid:48) = 0, p(cid:48) i = −α 3/2− (cid:16) e−√ α−(pi+1−pi) − e−√ α−(pi−pi−1)(cid:17) + O(δ2). (4.3.28) It is easy to see that if the pulses are equally separated, pi+1 − pi = pi − pi−1 = (cid:96), for (cid:96) > 0 big enough for all i = 1, . . . , n then, these pulses are stationary, namely, p(cid:48) i = 0, ∀i = 1,··· , n. (4.3.29) Further, we would like to study the stability of these stationary n-pulse configurations 85 examining the Jacobian matrix of this ODE system. 2 0 γ − γ 2 γ − γ − γ 0 − γ ... 2 0 ... 0 . . . 0 . . . 0 2 0 γ − γ 2 . . . . . . . . . . . . 0 . . . 0 . . . − γ 2 0 ... γ . . . 0 − γ 2  (4.3.30)  J = 0 0 (cid:18) where γ := −α2−e−√ α−(cid:96). The matrix J has n eigenvalues λk = −α2− 1 + cos e−√ α−(cid:96) < 0, (4.3.31) (cid:18) k (cid:19)(cid:19) n + 1 for all k = 1, . . . , n. From this, we conclude that the stationary solutions are spectrally stable. Moreover, assuming that the initial data is given for two-pulse configuration U0 = φ1 +φ2 and setting (cid:96)i := pi − pi−1 we obtain the evolution equations for pulse locations as (cid:16) e−√ e−√ α−(cid:96)3 − e−√ α−(cid:96)2 − e−√ (cid:17) α−(cid:96)2 max (cid:16) p(cid:48) 1 = −α p(cid:48) 2 = −α 3/2− φ2 3/2− (cid:17) α−(cid:96)1 + O(δ2), + O(δ2). (4.3.32) If the distance between φ1 and φ2 is smaller than their distance to adjacent pulses, (cid:96)3 > (cid:96)1 > (cid:96)2 or (cid:96)1 > (cid:96)3 > (cid:96)2, then p(cid:48) 2 > 0. From this analysis, we conclude that if 1 < 0 and p(cid:48) two adjacent pulses are closer to each other than other neighbor pulses, then they repel each other. If two pulses more far apart comparing to other neighbor pulses than they attract 86 each other. 4.4 Conclusion We have shown that the pulses are attracted into and remain within an O(δ2) window of the equally spaced (periodic) distribution. Moreover, the full solution remains within an O(δ) neigborhood in H2 of the periodic n-pulse. By compactness, a subsequence of times tn with tn → ∞ as n → ∞ exists such that u(tn) converges to u∗ in the ball of radius δ of the n-periodic solution. As this is a gradient flow, u∗ must be an equilibrium and then we must have u(t) → u∗ for the whole sequence as traversing the distance between u∗ and a distinct equilibrium value infinitely many times would expend infinite energy. In particular, the flow converges to an equilibrium which is exponentially close to the periodic n-pulse. 87 APPENDIX 88 Appendix A bound for H 2 norm of the product of two H 2 functions Lemma A.0.1. Let f, g ∈ H2([R]). Then, there exists a constant C > 0 such that Proof. (cid:107)f g(cid:107)H2 = = = (1 + k2)| ˆf (k − ˜k)||ˆg(˜k)|d˜k (cid:107)f g(cid:107)H2 ≤ C (cid:107)f(cid:107)H2 (cid:107)g(cid:107)H2 . (cid:90) (1 + k2)2|(cid:99)f g(k)|2dk, (cid:90) (cid:18)(cid:90) (cid:90) (1 + k2)2 (cid:18)(cid:90) R R | ˆf (k − ˜k)ˆg(˜k)|d˜k R (A.0.1) (A.0.2) dk, dk. (cid:19)2 (cid:19)2 R R 89 Writing k = k − ˜k + ˜k and then by the inequality (a + b)2 ≤ 2a2 + 2b2, we have (cid:107)f g(cid:107)H2 ≤ (1 + (k − ˜k)2) ˆf (k − ˜k)||ˆg(˜k)|d˜k + (1 + ˜k2)| ˆf (k − ˜k)||ˆg(˜k)|d˜k (cid:17)| ˆf (k − ˜k)||ˆg(˜k)|d˜k (cid:19)2 dk, (cid:90) R R R 22(1 + (k − ˜k)2) + 22(1 + ˜k2) (cid:18)(cid:90) (cid:16) (cid:90) (cid:90) (cid:16) (cid:90) (cid:16)|(1 + k2) ˆf (k)| ∗ |ˆg(k))| + |(1 + k2)ˆg| ∗ | ˆf|(cid:17)2 (cid:90) (cid:16)|(1 + k2) ˆf (k)| ∗ |ˆg(k))|(cid:17)2 (cid:19) (cid:18)(cid:13)(cid:13)(cid:13)|(1 + k2) ˆf| ∗ |ˆg) (cid:13)(cid:13)(cid:13)|(1 + k2)ˆg| ∗ | ˆf )|(cid:13)(cid:13)(cid:13)2 (cid:13)(cid:13)(cid:13)2 dk + 26 (cid:90) + R R R R L2 . L2 = 24 = 24 ≤ 26 = 26 dk, (cid:16)|(1 + k2)ˆg| ∗ | ˆf )|(cid:17)2 dk, (cid:17)2 dk, (A.0.3) By the Young’s inequality for convolutions that is (cid:107)f ∗ g(cid:107)r ≤ (cid:107)f(cid:107)p (cid:107)g(cid:107)q for f ∈ Lp and g ∈ Lq with 1 p + 1 q = 1 r + 1, (cid:107)f g(cid:107)H2 ≤ 26 ≤ 26 (cid:13)(cid:13)(cid:13)2 (cid:13)(cid:13)(cid:13)(1 + k2)ˆg (cid:19) (cid:13)(cid:13)(cid:13) ˆf (cid:13)(cid:13)(cid:13)2 , L1 L2 (cid:19) (cid:13)(cid:13)(cid:13)2 (cid:13)(cid:13)(cid:13) ˆf L1 , (A.0.4) (cid:107)ˆg(cid:107)2 L1 + L2 L1 + (cid:107)g(cid:107)2 H2 (cid:18)(cid:13)(cid:13)(cid:13)(1 + k2) ˆf (cid:13)(cid:13)(cid:13)2 (cid:18) (cid:107)f(cid:107)H2 (cid:107)ˆg(cid:107)2 H2 (cid:107)g(cid:107)2 ≤ C (cid:107)f(cid:107)2 H2 , since (cid:107)g(cid:107)L1 ≤ c(cid:107)g(cid:107)H2 for a constant c > 0. Second variation of I The second variation of I with respect to L2 inner product is calculated taking derivative of i(τ ) twice, 90 i(cid:48)(cid:48)(τ ) = = =2 dz, ε ε ε 0 0 d dτ d2 dτ 2 (cid:16) (cid:16) z (Φn + τ v) − W(cid:48) (Φn + τ v) ∂2 (cid:17)2 z (Φn + τ v) − W(cid:48)(Φn + τ v) ∂2 (cid:17)(cid:16) (cid:90) L2 (cid:90) L2 (cid:90) L2 (cid:17) (cid:17)(cid:16) (cid:16) z v − W(cid:48)(cid:48) (Φn + τ v) v z v − W(cid:48)(cid:48) (Φn + τ v) v (cid:16) (cid:17)(cid:16)−W(cid:48)(cid:48)(cid:48) (Φn + τ v) v2(cid:17) ∂2 ∂2 z (Φn + τ v) − W(cid:48) (Φn + τ v) (cid:90) L2 (cid:17)(cid:17)2 (cid:16)(cid:16) ∂2 z v − W(cid:48)(cid:48) (Φn + τ v) v (cid:17)(cid:16)−W(cid:48)(cid:48)(cid:48) (Φn + τ v) v2(cid:17) (cid:16) ∂2 z (Φn + τ v) − W(cid:48) (Φn + τ v) ∂2 0 0 + + ε =2 dz, dz. z v − W(cid:48)(cid:48) (Φn + τ v) v ∂2 (cid:17) dz, (A.0.5) Evaluating this derivative at τ = 0, (cid:90) L2 (cid:90) L2 0 ε ε 0 (cid:16) (cid:16) z v − W(cid:48)(cid:48) (Φn) v ∂2 (cid:17)2 z − W(cid:48)(cid:48) (Φn) ∂2 (cid:17)2 + v2 + i(cid:48)(cid:48)(0) = 2 = 2 (cid:16) (cid:16) (cid:17)(cid:16)−W(cid:48)(cid:48)(cid:48) (Φn) v2(cid:17) (cid:17)(cid:16)−W(cid:48)(cid:48)(cid:48) (Φn) v2(cid:17) z Φn − W(cid:48) (Φn) ∂2 z Φn − W(cid:48) (Φn) ∂2 dz, dz. (A.0.6) When we insert the expansion for Φn in (A.0.6) and expand it about φn, we have 91 (cid:90) L2 (cid:90) L2 0 ε ε 0 i(cid:48)(cid:48)(0) =2 =2 dz, v2+ v2 + (cid:17)2 z − W(cid:48)(cid:48) (Φn) ∂2 (cid:16) (cid:17)(cid:16)−W(cid:48)(cid:48)(cid:48) (Φn) v2(cid:17) (cid:16) z Φn − W(cid:48) (Φn) ∂2 (cid:1)(cid:17)2 (cid:16) z − W(cid:48)(cid:48)(cid:0)φn + ελB2,n (cid:1)(cid:17)(cid:0)−W(cid:48)(cid:48)(cid:48)(cid:0)φn + ελB2,n (cid:16) (cid:1) − W(cid:48)(cid:0)φn + ελB2,n (cid:0)φn + ελB2,n (cid:1)(cid:17)2 (cid:16) z −(cid:0)W(cid:48)(cid:48) (φn) + ελW(cid:48)(cid:48)(cid:48) (φn)B2,n (cid:1)(cid:17)(cid:16) zB2,n −(cid:0)W(cid:48) (φn) + ελW(cid:48)(cid:48) (φn)B2,n W(cid:48)(cid:48)(cid:48) (φn) + ελW (4) (φn)B2,n (cid:18)(cid:16) (cid:17)(cid:0)W(cid:48)(cid:48)(cid:48) (φn)(cid:1) + O(ε) (cid:1)(cid:1) v2dz, (cid:17)2 −(cid:16) z φn − W(cid:48) (φn) ∂2 z − W(cid:48)(cid:48) (φn) ∂2 v2− v2dz. (cid:19) ∂2 ∂2 z ∂2 =2 + ε (cid:90) L2 (cid:16) (cid:90) L2 ε =2 0 0 ∂2 z φn + ελ∂2 Since φn solves ∂2 z u − W(cid:48) (u) = 0, at the leading order we obtain (cid:17)2 z − W(cid:48)(cid:48) (φn) ∂2 v2dz, (cid:16) (cid:90) L2 (cid:90) L2 0 ε ε 0 i(cid:48)(cid:48)(0) = 2 = 2 (Lnv)2 dz, = 2(cid:104)Ln 2v, v(cid:105). Thus, the second variation of I at the leading order is δ2I δU 2 (Φn) := L2 n. (cid:17) v2dz, (A.0.7) (A.0.8) (A.0.9) H 2 norm of the projection Π0 Lemma A.0.2. Let Π0 be the orthogonal projection of H2 onto U ⊂ H2 given in (4.2.2). Then, for any u ∈ H2 92 Proof. For any u ∈ H2, (cid:107)Π0u(cid:107)H2 = (cid:107)u(cid:107)H2 . (cid:107)Π0u(cid:107)2 H2 = (cid:104)Π0u, Π0u(cid:105)H2 0u, u(cid:105)H2 = (cid:104)Π2 ≤ (cid:107)Π0u(cid:107)H2 (cid:107)u(cid:107)H2 , (A.0.10) (A.0.11) and therefore, (cid:107)Π0u(cid:107)H2 ≤ (cid:107)u(cid:107)H2. On the other hand, since Π0 is a projection we can write u = Π0u + u(cid:48) where u(cid:48) ∈ U⊥ and obtain H2 +(cid:13)(cid:13)u(cid:48)(cid:13)(cid:13)2 H2 , (cid:107)u(cid:107)2 H2 = (cid:104)Π0u + u(cid:48), Π0u + u(cid:48)(cid:105) = (cid:107)Π0u(cid:107)2 (A.0.12) which proves that (cid:107)Π0u(cid:107)H2 ≥ (cid:107)u(cid:107)H2. 93 BIBLIOGRAPHY 94 BIBLIOGRAPHY [Cahn and Hilliard, 1958] Cahn, J. W. and Hilliard, J. E. (1958). Free energy of a nonuniform system. i. interfacial free energy. The Journal of Chemical Physics, 28(2):258– 267. [Evans, 2010] Evans, L. (2010). Partial Differential Equations: Second Edition. American Mathematical Society. [Gavish et al., 2011] Gavish, N., Hayrapetyan, G., Promislow, K., and Yang, L. H. (2011). Curvature driven flow of bilayer interfaces. [Gearhart, 1978] Gearhart, L. (1978). Spectral theory for contraction semigroups on hilbert space. Transactions of the American Mathematical Society, 236:385–394. [Gompper and Schick, 1990] Gompper, G. and Schick, M. (1990). Correlation between structural and interfacial properties of amphiphilic systems. Phys. Rev. Lett., 65:1116– 1119. [Kapitula and Promislow, 2013] Kapitula, T. and Promislow, K. (2013). Spectral and Dynamical Stability of Nonlinear Waves. Springer. [Kato, 1976] Kato, T. (1976). Perturbation Theory for Linear Operators. Springer, Berlin, Heidelberg. [Kriegl et al., 2011] Kriegl, A., Michor, W. P., and A., R. (2011). Denjoy-carleman differentiable perturbation of polynomials and unbounded operators. Integral Equations and Operator Theory, 71:407–416. [Modica, 1987] Modica, L. (1987). The gradient theory of phase transitions and the minimal interface criterion. Arch. Ration. Mech. Anal., 98:123–142. [Pego, 1989] Pego, R. L. (1989). Front migration in the nonlinear cahn-hilliard equation. Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences, 422(1863):261–278. [Promislow, 2002] Promislow, K. (2002). A renormalization method for modulational SIAM J. Math. Analysis, stability of quasi-steady patterns in dispersive systems. 33(6):1455–1482. [Promislow and Wetton, 2009] Promislow, K. and Wetton, B. (2009). Pem fuel cells: SIAM Journal on Applied Mathematics, 70(2):369–409. A mathematical overview*. 95 Copyright - Copyright Society for Industrial and Applied Mathematics 2009; Document feature - Equations; Illustrations; Diagrams; Graphs; ; Last updated - 2012-02-21. [Promislow and Zhang, 2013] Promislow, K. and Zhang, H. (2013). Critical points of functionalized lagrangians. Discrete and Continuous Dynamical Systems- Series A, 33:1231–1246. [Pr¨uss, 1984] Pr¨uss, J. (1984). On the spectrum of c0-semigroups. Transactions of the American Mathematical Society, 284(2):847–857. [R¨oger and Sch¨atzle, 2006] R¨oger, M. and Sch¨atzle, R. (2006). On a modified conjecture of de giorgi. Mathematische Zeitschrift, 254(4):675–714. [Rosen, 2004] Rosen, M., J. (2004). Surfactants and Interfacial Phenomena. John Wiley and Sons, 2004. [Rubinstein and Sternberg, 1992] Rubinstein, J. and Sternberg, P. (1992). Nonlocal IMA Journal of Applied Mathematics, reaction—diffusion equations and nucleation. 48(3):249–264. [Sandstede, 1998] Sandstede, B. (1998). Stability of multi-pulse solutions. Transactions of the American Mathematical Society, 350:429–472. [Sandstede, 2001] Sandstede, B. (2001). Stability of travelling waves. [Sternberg, 1988] Sternberg, P. (1988). The effect of a singular perturbation on nonconvex variational problems. Arch. Ration. Mech. Anal., 101:209–260. [Teubner and Strey, 1987] Teubner, M. and Strey, R. (1987). Origin of the scattering peak in microemulsions. The Journal of Chemical Physics, 87(5):3195–3200. [Wiebe et al., 2012] Wiebe, H., Spooner, J., Boon, N., Deglint, E., Edwards, E., Dance, P., and Weinberg, N. (2012). Calculation of molecular volumes and volumes of activation using molecular dynamics simulations. The Journal of Physical Chemistry C, 116(3):2240–2245. 96