STUDY OF A CLASS OF LANDAU-LIFSHITZ EQUATIONS OF FERROMAGNETISM WITHOUT EXCHANGE ENERGY By Wei Deng A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Applied Mathematics 2012 ABSTRACT STUDY OF A CLASS OF LANDAU-LIFSHITZ EQUATIONS OF FERROMAGNETISM WITHOUT EXCHANGE ENERGY By Wei Deng Landau-Lifshitz equations of ferromagnetism, which are based on several competing energy contributions, are important mathematical models for the evolution of magnetization field m of a ferromagnetic material. Many problems, such as existence, stability, regularity, asymptotic behavior, thin-film limit and numerical computation, have been well studied for the Landau-Lifshitz equations that include the so-called exchange energy. However, these problems turn out to be quite challenging for equations without the exchange energy. The main reason is that when the exchange energy is included, one automatically has the magnetization vector m ∈ L∞ ((0, ∞); H 1 (Ω)) from energy estimates, which gives some compactness and stability that are needed for using the standard methods; however, in the cases without the exchange energy, one only has m ∈ L∞ ((0, ∞); L∞ (Ω)), which is too rough to get the needed compactness and stability. In this thesis, we investigate some problems for models of reduced Landau-Lifshitz equations with no-exchange energy. In Chapter 1, we introduce the Landau-Lifshitz theory of ferromagnetism and summarize the main results of the thesis. The readers can check out the main results quickly in this chapter and then go to the corresponding chapters for details of proof, more discussions and further references. In Chapter 2, we study the quasi-stationary limit of a simple Landau-Lifshitz-Maxwell system with the permittivity parameter approaching zero and, using this quasi-stationary limit, establish the existence of global weak solutions to the reduced Landau-Lifshitz equa- tions with initial value m0 ∈ L∞ (Ω). In Chapter 3, we establish a local L2 -stability theorem for the global weak solutions in finite time. The key in the proof of stability theorem is that we split the nonlocal term Hm into two parts: one is bounded in L∞ (Ω) and the other bounded in L2 (Ω). Using this stability theorem, we also provide another proof for the existence of global weak solutions for a full expression of the no-exchange energy with applied field a(x) ∈ L∞ (Ω). In Chapter 4, we prove a higher time regularity for the regular solutions, using mainly induction method, together with several interpolation results. In this chapter, we also study the weak ω-limit sets for the so-called soft-case and study the asymptotic behaviors for the special case when Ω is ellipsoid and initial values m0 are constant. In Chapter 5, we investigate a different model called the fractional Landau-Lifshitz equations and establish the existence of global weak solutions with initial value m0 ∈ H α (Ω), where 0 < α < 1. In this new model, in contrast to the case when only the nonlocal term Hm is included, we have some compactness in H α (Ω), which enables us to apply the Galerkin method to establish the existence of global weak solution. ACKNOWLEDGMENTS I would like to gratefully and sincerely thank Dr. Baisheng Yan for his guidance, understanding, patience, and most importantly, his friendship during my graduate studies at Michigan State University. His mentorship was paramount in providing a well rounded experience which is consistent with my long-term career goals. He encouraged me to not only grow as a mathematician but also as an independent thinker. I would like to thank the Department of Mathematics at Michigan State University, especially those members of my doctoral committee for their input, valuable discussions and accessibility. In particular, I would like to thank Dr. Zhengfang Zhou, Dr. Moxun Tang, Dr. Chang Yi Wang and Dr. Ignacio Uriarte-Tuero for their assistance and guidance in getting my graduate career started on the right foot. Finally, and most importantly, I thank my parents, Kepei Wang and Jianping Deng, for their faith in me and allowing me to be as ambitious as I wanted. It was under their watchful eye that I gained so much drive and an ability to tackle challenges. Also, I thank my graduate fellow, Hongli Gao for her helpful discussions. iv TABLE OF CONTENTS Chapter 1 Introduction and Main Results . . . . . 1.1 Existence . . . . . . . . . . . . . . . . . . . . . . 1.2 Stability . . . . . . . . . . . . . . . . . . . . . . . 1.3 Existence of global solutions . . . . . . . . . . . . 1.4 Time regularity and special asymptotics . . . . . 1.5 Existence of fractional Landau-Lifshitz equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 3 6 7 8 12 Chapter 2 Existence for Reduced Landau-Lifshitz Equations 2.1 Quasi-stationary limit of Landau-Lifshitz-Maxwell systems . . 2.1.1 General assumptions and preliminaries . . . . . . . . . 2.1.2 Two Compensated Compactness Lemmas . . . . . . . . 2.1.3 Strong convergence of M . . . . . . . . . . . . . . . . 2.2 Existence for rough initial data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 18 20 22 25 36 Chapter 3 Local L2 -Stability of Solutions in Finite Time 3.1 Stability of weak solutions . . . . . . . . . . . . . . . . . . 3.1.1 Decomposition of Hm . . . . . . . . . . . . . . . . 3.1.2 Stability Theorem . . . . . . . . . . . . . . . . . . . 3.2 Existence of global weak solutions . . . . . . . . . . . . . . 3.2.1 Properties of f for smooth applied fields . . . . . . 3.2.2 Existence of global solution for smooth data . . . . 3.2.3 Existence of global weak solution for rough data . . 3.2.4 Proof of Theorem 3.2.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 38 38 40 45 45 50 51 53 Chapter 4 Higher Time Regularity and Special Asymptotics 4.1 Higher time regularity . . . . . . . . . . . . . . . . . . . . . . 4.2 Weak ω-limit sets . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.1 The energy identity . . . . . . . . . . . . . . . . . . . . 4.2.2 Weak ω-limit sets and the estimate for soft-case . . . . 4.3 Special dynamics for constant initial data on ellipsoid domains 4.3.1 Associated ODE system on R3 . . . . . . . . . . . . . 4.3.2 Lyapunov function and the special dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 59 64 64 65 68 69 71 Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 74 77 80 82 Chapter 5 Existence for Fractional Landau-Lifshitz 5.1 Notations and preliminaries . . . . . . . . . . . . . 5.2 A priori estimates . . . . . . . . . . . . . . . . . . . 5.3 Compactness . . . . . . . . . . . . . . . . . . . . . 5.4 Existence . . . . . . . . . . . . . . . . . . . . . . . v . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Bibliography . . . . . . . . . . . . . . . vi . . . . . . . . . . . . . . . . . . . 86 Chapter 1 Introduction and Main Results The well-known Landau-Lifshitz theory in ferromagnetism models the state of magnetization vector m of a ferromagnetic material occupying a domain Ω in R3 based on a formulation of the total energy E(m) consisting of several competing terms: E(m) = 1 κ a(x) · m dx + ϕ(m) dx − | m|2 dx + |Hm |2 dx. 2 Ω 2 R3 Ω Ω (1.1) We refer to [4, 34, 37, 38] for more backgrounds on such a model. The first term of E(m) is the exchange energy, penalizing the spacial change of m; this term could also be given in terms of a positive-definite quadratic form of m. The second term is the anisotropy energy due to crystallographic properties of the material. The third term is the interaction energy due to a given applied magnetic field a(x). The last term is the magnetostatic energy of the stray field Hm induced by m through a simplified Maxwell equations: curl Hm = 0, div(Hm + mχΩ ) = 0 in R3 , (1.2) where χΩ is the characteristic function of domain Ω. In this theory, the (rescaled) saturation condition: |m| = 1 is usually assumed over Ω. The dynamic Landau-Lifshitz equation governing the evolution of magnetization m = 1 m(x, t) is given by ∂t m = γm × Heff + α γ m × (m × Heff ) |m| (1.3) on Ω × [0, ∞), where γ < 0 is the electron gyromagnetic ratio, α ≥ 0 is the Landau-Lifshitz phenomenological damping parameter, and Heff is the total effective magnetic field defined by the functional derivative of E(m) as Heff = − ∂E = κ∆m − ϕ (m) + a(x) + Hm . ∂m (1.4) Note that the Landau-Lifshitz equation (1.3) can be written as a Landau-Lifshitz-Gilbert equation: ∂t m = γ(1 + α2 )m × Heff + α m × ∂t m. |m| (1.5) Most existing studies, including existence, stability, asymptotic behavior and regularity, on (1.3) or (1.5) are for models with the exchange energy, that is, when Heff = κ∆m − ϕ (m) + a(x) + Hm with κ > 0; however, few results have been established for the case when the exchange energy is excluded, because of the lack of compactness and stability in such a case. We consider the reduced Landau-Lifshitz equation (1.3) without exchange energy and study properties of the corresponding solutions. In following sections, we briefly introduce some background and our results on existence, stability, asymptotic behavior and regularity problems for reduced Landau-Lifshitz equations. 2 1.1 Existence There are many results for existence of solutions to (1.3) or (1.5) with the exchange energy, see [1, 3, 7, 8, 10, 17, 22, 27, 49]. In these cases, the initial data m0 usually need to be smooth enough (e.g. in H 1 (Ω) or H 2 (Ω)) in order to use the Galerkin method and elliptic estimates to establish the existence of solutions. Later, we will see that the most important key in the proof of existence is that we need strong convergence for some certain sequence of m. When we include exchange energy, the strong convergence condition is much easier to obtain than the case without exchange energy. Specifically, with exchange energy, most of previous research work can easily get following bound for m by applying energy estimation. m ∈ L1 ((0, ∞); H 1 (Ω))) (1.6) The bound (1.6) is also very important to many other problems, such as asymptotic behavior of solutions. Actually, most of results about Landau-Lifshitz equation with exchange energy based on (1.6). Later we will see that in our case (without exchange energy), from energy estimation, the best bound we can get for m is, m ∈ L∞ ((0, ∞); L∞ (Ω))) That is why our problem is more challenging. Existence of weak solution for rough initial data As introduced above, we will prove existence of solutions to the dynamic Landau-Lifshitz equation (1.3) for a total energy E(m) without the exchange energy or simply called reduced 3 dynamic Landau-Lifshitz equation (1.3). Therefore, we have Heff = −ϕ (m) + a(x) + Hm This Cauchy problem can be written as a quasi-stationary system:     ∂t m = F (x, m, Hm ) in Ω × (0, ∞),      3 curl Hm = 0, div(Hm + mχΩ ) = 0 in R ,        m(x, 0) = m0 (x) on Ω, (1.7) with a given initial datum m0 ∈ L∞ (Ω), where F : Ω × R3 × R3 → R3 is an appropriate function (the Landau-Lifshitz interaction function) specifically given in Chapter 2. In order to prove the existence of global solution, we wish to have tight bound like (1.6) for solutions or approximate solutions of system (1.7) and then easily extract strong convergent subsequence in L2 , which would handle the nonlinear term F (x, m, Hm ). ∂m For smooth initial data m0 ∈ H 2 (Ω) with ∂ν0 |∂Ω = 0, previous research work of Carbou and Fabrie [8] has established the existence to the similar equation without exchange energy using the singular perturbation method: include κ∆m in Heff and let κ → 0. But unfortunately, for initial data only in L∞ (Ω), their method does not directly work. We need to find other ways to obtain the compactness for approximate solutions to (1.7) with rough initial data only in L∞ . We study the system (1.7) as a quasi-stationary limit of the following Landau-LifshitzMaxwell system of electro-magnetism when the permittivity parameter is constant and approaches zero; the more general case of Maxwell equations with variable permittivity has 4 been studied in Jochmann [29]. However, the system (1.8) studied below is much simpler and the method used is quite different, but more direct.     ∂t E − curl H = 0,         ∂ (H + M χ ) + curl E = 0 in R3 × (0, ∞),  t Ω (1.8)   ∂ M = F (x, M, H) in Ω × (0, ∞),  t        (E, H)| 3  t=0 = (E0 , H0 ) on R , M |t=0 = m0 on Ω, where the initial data E0 , H0 for electric and magnetic fields E, H are any vector-fields satisfying E0 , H0 ∈ L2 (R3 ; R3 ), For div E0 = div(H0 + m0 χΩ ) = 0. (1.9) > 0, the existence of certain weak solutions to the Cauchy problem (1.8) can be obtained in the same way as studied in Joly, Metivier and Rauch [31] with = 1. Therefore, for any > 0, there exists a global weak solution M . We have following theorem for the convergence of sequence M as → 0. Theorem 1.1.1 (Chapter 2, section 2.1). We have M → m strongly in both C 0 ([0, T ]; L2 (Ω)) and L2 (ΩT ) for all 0 < T < ∞. The strong convergence result in Theorem 1.1.1 is exactly what we need to handle the nonlinear term in the Cauchy problem (1.7). We use quite different methods to get such compactness comparing to the case with exchange energy. From details in Chapter 2, Section 2.1 for proof of Theorem 1.1.1, one can notice that we never have the same bound as (1.6). To the best of our knowledge, in the reduced Landau-Lifshitz model, one should not be able to get (1.6). 5 With Theorem 1.1.1, we establish the existence of global weak solution to the Cauchy problem (1.7). Theorem 1.1.2 (Chapter 2, section 2.2). Let m0 ∈ L∞ (Ω) and let E0 , H0 be any functions satisfying (1.9). Then any function m determined by the convergence in Theorem 1.1.1 is a weak solution to the Cauchy problem (1.7). 1.2 Stability Definition 1.2.1. Given m0 ∈ L∞ (Ω), a global weak solution to the Cauchy problem (1.7) is a function 1,∞ m ∈ Wloc ([0, ∞); L2 (Ω; R3 )) ∩ L∞ ((0, ∞); L∞ (Ω; R3 )) (1.10) such that m(0) = m0 in L2 (Ω) with m = m(·, t), equation ∂t m = F (x, m, Hm ), (1.11) holds both in L∞ ((0, ∞); L2 (Ω)) and in the sense of distribution on Ω × (0, T ), for all 0 < T < ∞. In this section we will investigate the stability of global weak solutions to quasi-stationary system (1.7) including function a(x); for a similar result on the Maxwell system, see [31, Theorem 6.1] and [13, Theorem 5.1]. Theorem 1.2.1 (Chapter 3, Section 3.1). Let 0 < R, T < ∞ be given. Then there exist constants C = C(R, T ) > 0, c = c(R, T ) > 0 and ρ = ρ(R, T ) > 0 such that, for any weak solution mk to the system (1.7) with applied field ak and initial datum mk (0) = mk 0 6 satisfying ak L∞ + mk L∞ ≤ R for k = 1, 2, if µ = max{ m1 −m2 L2 , a1 −a2 L2 } ≤ c, 0 0 0 then one has, for all t ∈ [0, T ], m1 (t) − m2 (t) L2 (Ω) ≤ Cµρ . (1.12) This stability result also implies the uniqueness of weak solution to system (1.7). 1.3 Existence of global solutions Based on the previous stability theorem, we present a new method for the existence of global weak solution to (1.7) with general applied fields a and initial data m0 . First, we show the existence of global solution to (1.7) for smooth fields a and initial data m0 ∈ H 2 (Ω; R3 ). Define f (m) = Fa (x, m, Hm ). We show f : H 2 (Ω; R3 ) → H 2 (Ω; R3 ) and is locally Lipschitz ; the proof uses a critical estimate that Hm ∈ H 2 (Ω; R3 ) for all m ∈ H 2 (Ω; R3 ) (see, e.g., [8, 31]). By the abstract ODE theory in Banach spaces, problem (1.7) has a local solution belongs to H 2 (Ω; R3 ) if m0 ∈ H 2 (Ω; R3 ). Then a no-blowup result (Theorem 3.2.4) shows that the local solution is in fact global on t ∈ [0, ∞). We remark that in the special case when ϕ = 0 and a = 0 (thus Heff = Hm ), for smooth ∂m initial data m0 ∈ H 2 (Ω) with ∂ν0 |∂Ω = 0, Carbou and Fabrie [8] also established the global existence through a singular perturbation method, by including κ∆m in Heff and letting κ → 0. Once we have obtained the global existence for smooth data a and m0 , we use approximation and the stability result Theorem 1.2.1 to establish the existence for general data. Theorem 1.3.1 (Chapter 3, section 3.2). Let a ∈ L∞ (Ω; R3 ). Given any initial datum 7 m0 ∈ L∞ (Ω; R3 ), the problem (1.7) has an unique global weak solution. 1.4 Time regularity and special asymptotics Higher regularity in time In this section, we introduce the higher time-regularity result of solution to LLG equation (1.13).    m = γm × H + γαm × (m × H ) in Ω × (0, ∞),  t m m (1.13)   m(0) = m ,  0 where Ω is a bounded smooth domain in R3 and m0 ∈ H 2 (Ω; R3 ). The existence of such regular solutions has been proved in Section 3.2 by applying abstract ODE theory. Similar regularity problem has been studied by Cimrak and Keer [6] for LLG with exchange energy. In their proof, they highly take advantage of energy estimation from exchange energy and then apply induction method. Now even without exchange energy, we also get similar time regularity. Our method is actually also inspired by Cimrak and Keer, using induction method; however, because we drop off exchange energy, we do not have any a priori estimates as being used in [6]. One can refer to Section 4.1 for details of proof of Theorem 1.4.1. Theorem 1.4.1 (Chapter 4, Section 4.1). For any time T > 0 and initial unit vector m0 ∈ H 2 (Ω), the regular solution, i.e. m ∈ H 1 ([0, T ]; H 2 (Ω)), satisfies p+1 ∂t m H 2 (Ω) ≤ C 8 where C is one constant only depending on T, p, m0 H 2 (Ω) . Asymptotic behavior for constant initial data on ellipsoid domains We now consider asymptotic behavior of regular solutions to Landau-Lifshitz equation (1.13). One of difficulties is that without exchange energy m, we do not have much control for weak sequence with respect to time t. Thus, methods used in Carbou and Fabrie [7] can not be used in our case. Recently, research work about asymptotic behavior of global weak solutions to Landau-Lifshitz equation (1.13) has been done by Yan ([51],[52]). In [52], without exchange energy, Yan studied asymptotic behaviors of equation (1.13) in the weak-star convergence of L∞ (Ω, R3 ). The equilibrium set of equation (1.13) is m × Hm = 0 on Ω. Yan [51] has investigated the equilibrium set under the general framework of partial differential inclusions and the vectorial calculus of variations based on the notation quasiconvexity in Ball [2], Dacorogna [11] and Morrey [40]. He has proved the following theorem. Theorem 1.4.2 (Yan [51]). Let mj lim j→∞ Ω m weak-star in L∞ (Ω; R3 ) as j → ∞. If (|mj |2 + 2|mj × Hmj | − 1)+ dx = 0 then the weak-star limit m satisfies |m|2 + 2|m × Hm | ≤ 1 a.e. on Ω. It is very obvious that all points in equilibrium set satisfies condition in Theorem 1.4.2, thus Yan [52] described the behavior of equilibrium set in some sense. In Section 4.2, we devote to prove a similar result as Theorem 1.4.2 in a different way. We 9 first derive an energy identity for the global weak solutions to the Landau-Lifshitz equation (1.3). Theorem 1.4.3. The global weak solution m to (1.7) with bounded initial data satisfies the energy identity t E(m(t)) − E(m(s)) = γα s Ω |m × Heff |2 dxdτ ∀ 0 ≤ s ≤ t < ∞. (1.14) Furthermore, if γα < 0, then mt ∈ L2 ((0, ∞); L2 (Ω; R3 )). Therefore, the global-in-time regularity for weak solutions (even for regular solutions) is that m ∈ L∞ ((0, ∞); L∞ (Ω; R3 )) with mt ∈ L2 ((0, ∞); L2 (Ω; R3 )). But this regularity is not enough to have strong convergence as t → ∞; it would be enough if one has mt ∈ L1 ((0, ∞); L2 (Ω; R3 )) (see [32]). Therefore, it is quite challenging to study the asymptotic behaviors for even the regular solutions. The solution orbits for general initial data may not have strong ω-limit points; we thus study the weak ω-limit set: ˜ ω ∗ (m0 ) = {m | ∃ tj ↑ ∞ such that m(tj ) ˜ m weakly in L2 (Ω; R3 )}. (1.15) We give an estimate of ω ∗ (m0 ) for the so-called soft-case, where there is no anisotropy energy (ϕ = 0). Theorem 1.4.4. Let γα < 0, ϕ = 0 and a ∈ L∞ (Ω; R3 ). Then, for any m0 ∈ L∞ (Ω; R3 ) with |m0 (x)| = 1 a.e. on Ω, it follows that ˜ ˜ ˜ ω ∗ (m0 ) ⊆ {m ∈ L∞ (Ω; R3 ) | |m|2 + 2|m × (a + Hm )| ≤ 1 a.e. on Ω}. ˜ 10 (1.16) This theorem generalizes some results in [52]. For more results on further special case when a = 0, see [51, 52]. Instead of considering the asymptotic behavior of global weak solutions, in this section, we only try to determine asymptotic behavior of solutions to Landau-Lifshitz equation (1.13) under a special case when initial value m0 is constant over Ω and Ω set is ellipsoid. It mainly involves two step. The first step is that we need to write explicit form for nonlocal term Hm under our special geometry with ellipsoid domains. After some work has been done in Chapter 4, Section 4.2.1, we can write Hm = −Λm on Ω, where Λ is a positive definite matrix; the details of how we determine matrix Λ can be found in Chapter 4, Section 4.2.1. Then, we will use Lyapunov Theorem to determine global stable equilibrium point as follows, Theorem 1.4.5 (Chapter 4, Section 4.2). Let us assume that m0 is constant and Ω is ellipsoid, 3 Ω = {x ∈ R3 | i=1 x2 i < 1} ai b1 , b2 , b3 are positive numbers determined by (4.23). If bk = min{b1 , b2 , b3 }, then ±ek are asymptotically stable critical points to Landau-Lifshitz equation (1.13), where {e1 , e2 , e3 } are the standard basis vectors of R3 . 11 1.5 Existence of fractional Landau-Lifshitz equations Motivation Before we go in details about the existence result for our new model called fractional LandauLifshitz equations, let us first see motivation behind it. The micromagnetic energy is given by, E(m) = κ 1 | m|2 + ϕ(m) − a(x) · m + |Hm |2 . 2 Ω 2 R3 Ω Ω (1.17) One popular research topic for functional energy (1.17) is to explore global minimizers under non-convexity constraints |m| = 1. Another interesting problem is to investigate its limiting behavior in different asymptotic regimes. Those regimes are different with relations among following terms, t = thickness of the film l = length scale of the cross section (1.18) d = characteristic length scale of the magnetic material There are two well studied regimes: (a) The large-body limit, in which d → 0 while t is l l fixed. See [28], [18], [48] and [42]. (b) The small-aspect-ratio limit, in which t → 0 while d l l is fixed. In this case, when the external filed is constant, the asymptotic variational problem predicts a uniform magnetization; see [21]. DeSimone et al [15] considered different regimes: (a) Ω is a cylindrical domain of thickness t with cross section Ω , such as Ω = Ω × (0, t); (b) m does not depend on the thickness direction x3 ; (c) t l. Under such regime, DeSimone et al [15] derived following convergent 12 magnetostatic energy through Fourier transform: R3 |Hm |2 dx = 1 R2 | ∧− 2 ( m )|2 dx (1.19) 1 where m = (m1 , m2 ) and ∧ denotes the square root of Laplacian (− ) 2 . It is pretty interesting that the exchange energy is dissipative under this regime and magnetostatic energy reduced to (1.19). This result motivated some research work associated with energy (1.19) applied to Landau-Lifshitz equation; see [23], [24], [25], [26]. Fractional Landau-Lifshitz equation To extend the models considered in [23], [24], [25], [26], we would like to prove the existence of global weak solution to fractional Landau-Lifshitz equation (1.20) including magnetostatic term with periodic boundary condition: mt = γm × Fm + γm × (m × Fm ) in Rn , (1.20) where Fm = ∧2α m + Hm 0 < α < 1. (1.21) We focus on the existence of global weak solution(defined in Chapter 5, see definition (5.0.1) ) in special domain Ω = [0, 2π]n with periodic boundary conditions. Now, we define Hm in energy term (1.21) as,   (ξ · m) · ξ  ,  |ξ|2 Hm (ξ) =    0, ξ = 0. 13 ξ = 0. (1.22) 1 The operator ∧ denotes the square root of Laplacian (− ) 2 , so ∧β m can be understood in terms of Fourier transform: F(∧β m(x)) = |ξ|β m(ξ) (1.23) A priori estimates We mainly use Galerkin method to prove existence. Therefore, we first project our problem into finite space in L2 (Ω) and the solution can be easily established by standard ODE method. Then we are going to prove a priori estimation in order to extract strong convergence subsequence. Lemma 1.5.1 (Chapter 5, Section 5.2). Let m0 ∈ H α (Ω), then for any 0 < T < ∞, the approximate solutions mN to systems (5.8) in Chapter 5, satisfy, max 0≤t≤T mN 2 α (Ω) ≤ C1 H d where C1 only depends on initial data m0 2 α (Ω) . Moreover, for 1 ≤ r ≤ r∗ = d−α , where H d is dimension. ∂mN ≤ C2 r ∂t L (QT ) d and for 1 < r ≤ r∗ = d−α r−1 mN (t1 ) − mN (t2 ) Lr (Ω) ≤ C2 |t2 − t1 | r where C2 only depends on initial data m0 2 α (Ω) and time T , QT = (0, T ) × Ω. H It is not enough to extract strong convergence subsequence; we still need following compactness lemma, 14 Lemma 1.5.2 ([39]). Let B0 , B, B1 be three Banach space such that, B0 ⊂ B ⊂ B1 where the injections are continuous and B0 , B1 are reflexive and B0 → B is compact. Denote W = {v|v ∈ Lp0 (0, T ; B0 ), dv ∈ Lp1 (0, T ; B1 )} dt for T < ∞ and 1 < p0 , p1 < ∞. Then W equipped with the norm dv v Lp0 (0,T ;B ) + p 0 dt L 1 (0,T ;B1 ) is a Banach space and the embedding W → Lp0 (0, T ; B) is compact. Now, we can proved that there exists some m ∈ L∞ (0, T ; H α (Ω)) such that, mN m weakly in Lp (0, T ; H α (Ω)) for 1 < p < ∞ mN → m strongly in Lp (0, T ; H β (Ω)) for 1 < p < ∞, 0 ≤ β ≤ α ∂mN ∂t ∂m ∂t (1.24) weakly in Lr (QT ) for 1 < r Existence With compactness result (1.24), it is easy to establish existence. Theorem 1.5.3 (Chapter 5, section 5.4). Let 0 < α < 1 and m0 ∈ H α (Ω), then for any 0 < T < ∞, then there exists at least one global weak solution to fractional Landau-Lifshitz 15 equation (1.20) such that m ∈ L∞ (0, T ; H α (Ω)) d for 1 < r ≤ r∗ = d−α , where d is dimension. 16 r−1 C 0, r (0, T ; Lr (Ω)) Chapter 2 Existence for Reduced Landau-Lifshitz Equations In this chapter, we study the existence for reduced Landau-Lifshitz equations given by ∂t m = γm × Heff + α γ m × (m × Heff ), |m| (2.1) where Heff = −ϕ (m) + a(x) + Hm (2.2) with curl Hm = 0, div(Hm + mχΩ ) = 0 in R3 . (2.3) We establish the existence of global weak solution using two methods for different initial data. First, we study the quasi-stationary limit of a simple Landau-Lifshitz-Maxwell system with the permittivity parameter approaching zero and, using this quasi-stationary limit, establish the existence of global weak solutions to the reduced Landau-Lifshitz equations with initial value m0 ∈ L∞ (Ω). We then give a different method for the existence when the initial datum is more regular (e.g., in H 2 (Ω)) by considering the problem as an abstract ordinary differential equation in an appropriate Banach space. 17 2.1 Quasi-stationary limit of Landau-Lifshitz-Maxwell systems We study the Cauchy problem for equations (2.1)-(2.3) with a given bounded initial datum m(x, 0) = m0 (x). This Cauchy problem can be written as a quasi-stationary system:    ∂ m = F (x, m, H ) in Ω × (0, ∞),  t  m     3 curl Hm = 0, div(Hm + mχΩ ) = 0 in R ,       m(x, 0) = m0 (x) on Ω,  (2.4) with a given initial datum m0 ∈ L∞ (Ω), where F : Ω × R3 × R3 → R3 is an appropriate function (the Landau-Lifshitz interaction function), which can be written in a general form: F (x, M, H) = F(M )H + a(x, M ) (2.5) with certain specific functions F(M ) and a(x, M ); our results are valid for more general interaction functions F (x, M, H) of this form satisfying some conditions to be specified later (see (2.13)-(2.14) below). Definition 2.1.1. Given m0 ∈ L∞ (Ω), a weak solution to the Cauchy problem (2.4) is a function m ∈ W 1,∞ ((0, ∞); L2 (Ω)) ∩ L∞ ((0, ∞); L∞ (Ω)) (2.6) such that m(0) = m0 in L2 (Ω) and, with Hm defined by (2.3) with m = m(·, t), equation ∂t m = F (x, m, Hm ), 18 (2.7) holds both in L∞ ((0, ∞); L2 (Ω)) and in the sense of distribution on Ω × [0, ∞). Remark 1. We remark that, for the Landau-Lifshitz equation (2.1) with Heff given by (2.2), since F (x, m, Hm ) · m = 0, the length |m(x, t)| is preserved for weak solutions; therefore, if the initial datum m0 satisfies the saturation condition |m0 (x)| = 1 then any weak solution m(x, t) to the Cauchy problem (2.4) will also satisfy the saturation condition: |m(x, t)| = 1 for all t > 0. In general, our assumptions on interaction function F will guarantee the L∞ -norm of m(·, t) be non-increasing for t > 0 (see Section 2.1.1). ∂m In the special case of Heff = Hm , for the regular initial data m0 ∈ H 2 (Ω) with ∂ν0 |∂Ω = 0, certain weak solution of (2.4) has been obtained in [8] as limit of the regular solution mκ to the Landau-Lifshitz equation (2.1) with Heff = κ∆m + Hm when κ → 0+ . However, this singular perturbation method does not work for rough initial data m0 in L∞ (Ω). We study the Cauchy problem (2.4) as a quasi-stationary limit of the following LandauLifshitz-Maxwell system when the permittivity parameter approaches zero:     ∂t E − curl H = 0,         ∂ (H + M χ ) + curl E = 0 in R3 × (0, ∞),  t Ω (2.8)   ∂ M = F (x, M, H) in Ω × (0, ∞),  t        (E, H)| 3  t=0 = (E0 , H0 ) on R , M |t=0 = m0 on Ω, where the initial data E0 , H0 for electric and magnetic fields E, H are any vector-fields satisfying E0 , H0 ∈ L2 (R3 ; R3 ), For div E0 = div(H0 + m0 χΩ ) = 0. (2.9) > 0, the existence of certain weak solutions to the Cauchy problem (2.8) can be 19 obtained in the same way as studied in [31] for = 1. More general Landau-Lifshitz-Maxwell systems with variable dielectric permittivity (x) and magnetic permeability µ(x) have been studied in [29], together with the quasi-stationary limit as the variable (x) → 0. Long-time asymptotic problems for such systems have been addressed in [7, 30, 32]. One of the main purposes is to present a more direct way to study the asymptotic behavior of weak solution (E , H , M ) to the Cauchy problem (2.8) as → 0+ . We prove that, for all 0 < T < ∞, a subsequence M converges strongly to a function m in C 0 ([0, T ]; L2 (Ω)) and L2 (Ω × (0, T )) (see Theorem 2.1.3 below) and, furthermore, the limit m = m(x, t) is a weak solution to Cauchy problem (2.4) (see Theorem 2.2.1 below). Our direct proof is motivated by the methods of [31] with, however, a completely new parameter asymptotics, and is much different in nature from the studies based on semi-group theory in [29]. The rest of section is organized as follows. We will first give the assumptions on the interaction functions F (x, M, H) appearing in the general Cauchy problem (2.4) above and present some preliminaries, including two useful compensated compactness results. Then, we prove the strong convergence of M as → 0+ using a direct approach motivated by the work [31], which is much different from the abstract semi-group techniques used in [29]. Finally, we prove the existence of weak solution to (2.4) directly from the strong convergence theorem. 2.1.1 General assumptions and preliminaries We assume Ω is a bounded domain in R3 ; the boundedness of Ω will simplify many technical assumptions otherwise needed as in [29, 31]. Let F (x, M, H) = F(M )H + a(x, M ) be defined as in (2.5), where we assume F : R3 → 20 R3×3 and a : Ω × R3 → R3 satisfy the following condition (see also [29]): (a) F(M )T M = 0, a(x, M ) · M ≤ 0 ∀ M ∈ R3 , x ∈ Ω; (b) a(x, 0) ∈ L2 (Ω); (2.10) |a(x, M ) − a(x, M )| + |F(M ) − F(M )| ≤ C1 (R) |M − M | (c) ∀ R > 0, x ∈ Ω, |M |, |M | ≤ R, where C1 (R) is constant. In particular, we have F (x, M, H) · M ≤ 0 ∀ x ∈ Ω, M, H ∈ R3 (2.11) |F (x, M, H)| ≤ C2 (R)(|H| + 1) + |a(x, 0)| ∀ |M | ≤ R (2.12) for all x ∈ Ω, H ∈ R3 , R ≥ 0, where C2 (R) is a constant depending on R. Consider the Landau-Lifshitz function defined by L(0, H) = 0 and L(M, H) = γM × H + α γ M × (M × H), |M | M = 0, (2.13) where α, γ are constants. Then the function F (x, M, H) appearing in (2.4) is given by F (x, M, H) = F(M )H + a(x, M ) with F(M )H = L(M, H), a(x, M ) = L(M, a(x)) − L(M, ϕ (M )). (2.14) It is easily seen that this particular function F satisfies all assumptions in condition (2.10); in fact, one has F (x, M, H)·M = 0 for all x ∈ Ω and M, H ∈ R3 . This property in particular yields the equality in the length estimate (2.18) below. 21 The orthogonal decomposition L2 (R3 ; R3 ) = L2 (R3 ; R3 ) ⊕ L2 (R3 ; R3 ) is standard, where ⊥ L2 (R3 ; R3 ), L2 (R3 ; R3 ) are the subspaces of curl-free or divergence-free functions in the ⊥ sense of distributions, respectively. This decomposition can be explicitly given in terms of ˆ the Fourier transform f of f ∈ L2 (R3 ; R3 ): f = f + f⊥ , where ˆ ˆ f = (ξ · f )ξ/|ξ|2 , ˆ ˆ ˆ fˆ = f − (ξ · f )ξ/|ξ|2 = −ξ × (ξ × f )/|ξ|2 . ⊥ The projection operator P (f ) = f also extends to a bounded linear operator on Lp (R3 ; R3 ) for all 1 < p < ∞, with operator norm bounded by C0 p when p ≥ 2, where C0 is independent of p ≥ 2; see [45]. 2.1.2 Two Compensated Compactness Lemmas We give two compensated compactness results in a suitable form to be used later; both are special case of the more general results [20, 46, 47]. ∞ Let b ∈ Cc (R3 ) and [b, P ] = bP −P b : L2 (R3 ; R3 ) → L2 (R3 ; R3 ) be the commutator. We have the following special compactness result from the well-known div-curl lemma [46]. Lemma 2.1.1. Let f k 0 weakly in L2 (R3 ; R3 ) and Lp (R3 ; R3 ) for some p > 2. Then g k = [b, P ]f k → 0 strongly in L2 (Ω; R3 ) for all bounded sets Ω. Proof. From k g k = bf k − (bf k ) = (bf k )⊥ − bf⊥ , k we have div g k = − b·f⊥ , curl g k = b×f k in the sense of distributions on R3 , and so both {div g k } and {curl g k } are compact in H −1 (R3 ). By the div-curl lemma, |g k |2 = g k · g k → 0 22 in the sense of distributions on R3 . Since {g k } is bounded in Lp (R3 ) with p > 2, we have g k → 0 strongly in L2 (Ω) for all bounded sets Ω. Let U be any domain in R3 and, for 0 < T ≤ ∞, let UT = U × (0, T ). Let c ∈ R be any constant. Denote ˜ that ∆ = −1 cu = ∆u − cutt , where ∆ is the Laplacian with respect to x ∈ R3 . Note is the Laplacian operator with respect to (x, t) ∈ R3 × R and − 1 is the usual wave operator on R3 × R. The following lemma, with c = 0, will be used later; the result is a special case of more general analyses on micro-local defect measures or H-measures and orthogonality of sequences [20, 47]. Lemma 2.1.2. Let uk u and v k v weakly in L2 (UT ) as k → ∞. Suppose { c uk } is compact in H −2 (UT ) and {∂t v k } is bounded in L2 (UT ). Then uk v k → uv in the sense of distributions on UT . ∞ Proof. By considering uk − u and v k − v, we can assume u = v = 0. Given any ζ ∈ Cc (UT ), we need to show ζ(x, t)uk (x, t)v k (x, t) dxdt = 0. lim k→∞ (2.15) UT Let A, B be two regular domains in UT such that supp ζ ⊂⊂ A ⊂⊂ B ⊂⊂ UT . We solve the following Dirichlet problem:   ˜ k ∆w = ∆wk + wk = v k in B  tt (2.16)   k w = 0 on ∂B.  1 By the standard elliptic theory, this problem has a unique solution wk ∈ H0 (B) ∩ H 2 (B), which also satisfies the estimate wk H 2 (B) ≤ C1 v k L2 (B) . Hence {wk } is bounded in H 2 (B). By the Rellich compactness theorem, there exists a subsequence, still denoted by 23 k, such that {wk } converges strongly in H 1 (B) as k → ∞. From (2.16), it follows that k ˜ k ∆(wt ) = vt ∈ L2 (B) and hence by the (interior) L2 -estimates, k k k wt H 2 (A) ≤ C2 ( vt L2 (B) + wt L2 (B) ). k This shows that {wt } is bounded in H 2 (A) and hence for a subsequence, denoted again by k k k, {wt } converges strongly in H 1 (A). This implies the strong convergence of {wtt } in L2 (A) as k → ∞. Extending wk by zero onto UT , we write ζuk v k dxdt = UT = UT UT k (1 + c)ζuk wtt dxdt + k ζuk (∆wk + wtt ) dxdt UT where the first term Ik → 0 as k → ∞ since uk ζuk c wk dxdt := Ik + IIk , k 0 and {wtt } converges strongly in L2 (A). The second term can be written as IIk = uk ( c (ζwk ) − wk c ζ − 2 ζ · UT uk c (ζwk ) dxdt − = UT k wk + 2cζt wt ) dxdt uk (wk c ζ + 2 ζ · UT k wk − 2cζt wt ) dxdt. The second integral on the right-hand side approaches zero as k → ∞ because uk 0 and {wk } converges strongly in H 1 (B). Since {ζwk } is bounded in H 2 (UT ), we can estimate the first integral as follows: uk c (ζw k ) dxdt UT 24 = k k c u , ζw ≤ cu k H −2 (UT ) ζwk H 2 (U ) → 0 T 2 as k → ∞, where ·, · stands for the duality pairing between H −2 (UT ) and H0 (UT ). This completes the proof of (2.15). Finally, the lemma is proved. 2.1.3 Strong convergence of M Let 0 < < 1. Let (E , H , M ) be the weak solution to the Cauchy problem (2.8) with the 1,∞ property (E , H ) ∈ C 0 ([0, ∞); L2 (R3 )) and M ∈ Wloc ([0, ∞); L2 (Ω))∩L∞ ([0, ∞); L∞ (Ω)); loc the existence of such a weak solution has been established in [29, 31]. Estimates and weak convergence of M and H Standard estimates using (2.9), (2.11) and (2.12) and Gronwall’s inequality, yield that, for all t > 0, div E = div(H + M χΩ ) = 0 in R3 , |M (x, t)| ≤ |m0 (x)| a.e. x ∈ Ω, R3 (2.17) (2.18) ( |E (t)|2 + |H (t)|2 )dx ≤ C3 + eC3 t R3 ( |E0 |2 + |H0 |2 )dx, (2.19) where C3 is a constant depending only on a(·, 0) L2 and m0 L∞ . Since {M } is bounded in L∞ (Ω × (0, ∞)) and L∞ ((0, ∞); L2 (Ω)), there exist a function m ∈ L∞ (Ω × (0, ∞)) ∩ L∞ ((0, ∞); L2 (Ω)) and a subsequence of → 0, still denoted by , such that M m weakly* in L∞ (Ω × (0, ∞)) ∩ L∞ ((0, ∞); L2 (Ω)). (2.20) Let 0 < T < ∞ and ΩT = Ω × (0, T ) and UT = R3 × (0, T ). By energy estimate (2.19), 25 third equation of (2.8) and condition (2.12), it follows that the sequences { √ E }, {H } are bounded in L∞ ((0, T ); L2 (R3 )) and L2 (UT ) and that the sequence {∂t M } is bounded ˜ ˜ ˜ in L∞ ((0, T ); L2 (Ω)) and L2 (ΩT ). Let E, H and S be the limit of any weakly convergent subsequence of { √ E }, {H } and {∂t M } along the chosen → 0 in the respective Banach ˜ ˜ spaces. Clearly S = ∂t m. Note that Et → 0 and H → H also in the sense of distributions on UT . Hence, from the first equation of system (2.8), using (2.17), we have ˜ curl H(t) = 0, ˜ div(H(t) + m(t)χΩ ) = 0 ˜ in the sense of distributions on R3 for almost every t ∈ (0, T ); therefore, H(t) = −P (m(t)χΩ ) ˜ for almost every t ∈ (0, ∞). This implies that any weak limit H is uniquely determined by m. This uniqueness also shows that the whole sequence {H } (along the chosen → 0) converges weakly in both L∞ ((0, T ); L2 (R3 )) and L2 (UT ) for all 0 < T < ∞; the limit is given by H ∞ (t) = −P (m(t)χΩ ) = Hm(t) , (2.21) where Hm(t) is defined by m(t) through (1.2) above. Moreover, since M ∈ W 1,∞ ((0, T ); L2 (Ω)) ⊂ C 0 ([0, T ]; L2 (Ω)), we have m ∈ W 1,∞ ((0, T ); L2 (Ω)) ⊂ C 0 ([0, T ]; L2 (Ω)) ∀ 0 < T < ∞; hence, we can also assume (2.21) holds for all t > 0. Strong convergence of M The main result of this section is the following. 26 (2.22) Theorem 2.1.3. We have M → m strongly in both C 0 ([0, T ]; L2 (Ω)) and L2 (ΩT ) for all 0 < T < ∞. A closer look at the proof of [29, Theorem 1.1] shows that our theorem follows from that proof since the conductivity σ is zero here. However the proof in [29] involves other complicated techniques aimed for handling the variable permittivity (x) and permeability µ(x). We present a different and direct proof of this result in our setting, which is motivated by the methods in [31]. The rest of this section is devoted to the proof of Theorem 2.1.3. Weighted energy estimates Assume T > 0 and , δ → 0 are the chosen subsequence. We want to estimate some weighted norm of M (t) − M δ (t) in L2 (Ω) for 0 ≤ t ≤ T. In what follows, we use Cj to denote various constants depending on m0 L∞ and perhaps T or other quantities to be specified as needed. We write ∂t (M − M δ ) = F (x, M , H ) − F (x, M δ , H δ ) = F(M )H − F(M δ )H δ + a(x, M ) − a(x, M δ ) (2.23) = (F(M ) − F(M δ ))H ∞ + a(x, M ) − a(x, M δ ) + F(M )(H − H ∞ ) − F(M δ )(H δ − H ∞ ). 27 Hence 1∂ |M − M δ |2 ≤ |(F(M ) − F(M δ ))H ∞ | · |M − M δ | 2 ∂t + |a(x, M ) − a(x, M δ )| · |M − M δ | + (F(M )(H − H ∞ ) − F(M δ )(H δ − H ∞ )) · (M − M δ ) (2.24) ≤C1 |M − M δ |2 (|H ∞ | + 1) + (F(M )(H − H ∞ ) − F(M δ )(H δ − H ∞ )) · (M − M δ ), where C1 = C1 (R) is the constant in (2.10-c) with R = m0 L∞ . Let a(x, t) be the function defined by a(x, t) = |x|2 + C1 t (1 + |H ∞ (x, s)|) ds (x ∈ R3 , t ≥ 0). (2.25) 0 For all t ≥ 0, since e−a(·,t) ∈ L2 (R3 ) and M (·, t) ∈ L∞ (Ω), it follows that e−a(t) M (t) e−a(t) m(t) weakly in L2 (Ω) ∀ t ≥ 0. (2.26) Furthermore, by (2.24), 1 ∂ −2a (e |M − M δ |2 ) 2 ∂t ≤ e−2a (F(M )(H (2.27) − H ∞ ) − F(M δ )(H δ 28 − H ∞ )) · (M − M δ ). Integrating (2.27) with respect to x ∈ Ω and on time-interval (0, t), we get 1 −a(t) e (M (t) − M δ (t)) 2 2 L (Ω) 2 t ≤ 0 Ω t e−2a F(M )(H − H ∞ ) · (M − M δ ) dxds (2.28) − 0 Ω e−2a F(M δ )(H δ − H ∞ ) · (M − M δ ) dxds :=f ,δ (t) − f δ,δ (t), ρ where functions f ,δ (t) (with ρ = , δ) are defined by t ρ f ,δ (t) = 0 Ω e−2a F(M ρ )(H ρ − H ∞ ) · (M − M δ )dxds. (2.29) ρ To analyze f ,δ (t), we split it into two terms: t ρ f ,δ (t) = ρ 0 Ω t ∞ e−2a F(M ρ )(H⊥ − H⊥ ) · (M − M δ )dxds ρ + 0 Ω e−2a F(M ρ )(H − H ∞ ) · (M − M δ )dxds (2.30) ρ ρ :=g ,δ (t) + h ,δ (t). ρ ˜ ˜ ˜ ˜ By (2.17) and (2.21), H − H ∞ = −P (M ρ − m), where M ρ = M ρ χΩ and m = mχΩ . So, ρ the function h ,δ (t) in (2.30) can be rewritten as t ρ h ,δ (t) = − 0 Ω t =− 0 Ω t − 0 Ω ˜ ˜ e−2a F(M ρ )(P (M ρ − m)) · (M − M δ )dxds ˜ ˜ F(M ρ )(P (e−a (M ρ − m))) · e−a (M − M δ )dxds (2.31) ˜ ˜ F(M ρ )([e−a , P ](M ρ − m)) · e−a (M − M δ )dxds := −I3 − I4 , 29 where [e−a , P ]f = e−a f − (e−a f ) denotes the commutator operator. For I3 , we have: t |I3 | = 0 t ≤ C3 Ω e−a (M ρ − m) 2 2 + e−a (M − M δ ) 2 2 ds L (Ω) L (Ω) 0 t ≤ 4C3 ˜ ˜ F(M ρ )(P (e−a (M ρ − m))) · e−a (M − M δ )dxds 0 (2.32) e−a (M − m) 2 2 + e−a (M − M δ ) 2 2 ds. L (Ω) L (Ω) For I4 , we have the following result. Lemma 2.1.4. For any η > 0, there exists ξ = ξ(η, T, R) > 0 such that |I4 | < η/4 ∀ 0 < , δ < ξ, ∀ t ∈ [0, T ]. (2.33) Proof. Note that T |I4 | ≤ C3 0 ˜ ˜ [e−a(s) , P ](M ρ − m) L2 (Ω) ds. (2.34) We claim ˜ ˜ [e−a(s) , P ](M ρ (s) − m(s)) L2 (Ω) < ∞ 0≤t≤T ˜ ˜ lim [e−a(s) , P ](M ρ (s) − m(s)) L2 (Ω) = 0 ∀ s ∈ (0, T ). sup ρ→0 (2.35) (2.36) Once (2.35) and (2.36) are proved, by Lebesgue dominated convergence theorem, we have T lim ρ→0 0 ˜ ˜ [e−a(s) , P ](M ρ − m) L2 (Ω) ds = 0 and hence (2.33) follows by (2.34). To prove (2.35), note that, by the boundedness of P on 30 L4 (R3 ), ˜ ˜ [e−a , P ](M ρ − m) L2 (Ω) ≤ C0 e−a L4 (Ω) M ρ − m L4 (Ω) . To prove the point-wise convergence (2.36), fix s ∈ [0, T ] and note that ˜ ˜ [e−a(s) , P ](M ρ − m) L2 (Ω) ≤ I5 + I6 ˜ ˜ ˜ ˜ := [e−a(s) − b, P ](M ρ − m) L2 (Ω) + [b, P ](M ρ − m) L2 (Ω) , ∞ where b ∈ Cc (Ω). Again, by the boundedness of P on L4 (R3 ), we have I5 ≤ C0 e−a(s) − b L4 (Ω) M ρ − m L4 (Ω) ≤ C4 e−a(s) − b L4 (Ω) . ∞ Choose b ∈ Cc (Ω) so that e−a(s) − b L4 (Ω) and hence I5 are arbitrarily small. Once b ˜ is chosen, since M ρ (s) ˜ m(s) weakly in Lp (R3 ) for all p ≥ 2, by Lemma 2.1.1, we have I6 → 0 as ρ → 0. Hence (2.36) is proved. Combining (2.28)–(2.33) above, we have, for all t ∈ [0, T ] and 0 < , δ < ξ, with ξ = ξ(η, T, R) determined in Lemma 2.1.4, e−a(t) (M (t) − M δ (t)) 2 2 ≤ η + 2(g ,δ (t) − g δ,δ (t)) L t +C5 e−a(s) (M 0 − M δ) 2 2 L + e−a(s) (M − m) 2 L2 (2.37) ds, ρ where g ,δ (t) is the function defined in (2.30) by t ρ g ,δ (t) = 0 ρ Ω ∞ e−2a F(M ρ )(H⊥ − H⊥ ) · (M − M δ )dxds. 31 (2.38) From (2.37), using Gronwall’s inequality, we have t t e−a(s) (M − M δ ) 2 2 ds ≤ L 0 0 + T eT C5 η + C5 2e(t−s)C5 (g ,δ (s) − g ,δ (s)) ds t 0 e−a(s) (M − m) 2 2 ds . L Plugging this inequality into the right-hand of (2.37), we have obtained the following result. Proposition 2.1.5. For each η > 0, there exists ξ = ξ(η, T, R) > 0 such that, for all 0 < , δ < ξ and t ∈ [0, T ], e−a(t) (M (t) − M δ (t)) 2 2 ≤ G ,δ (t) − Gδ,δ (t) L t + C6 η + e−a(s) (M − m) 0 (2.39) 2 ds L2 , where C6 = C6 (R, T ) is a constant depending only on R, T and, for ρ = , δ, ρ t ρ G ,δ (t) = 2g ,δ (t) + 0 ρ 2e(t−s)C5 g ,δ (s) ds. (2.40) ρ Estimates of g ,δ (t) ρ ∞ We now study the function g ,δ (t) defined by (2.38). Since curl H ∞ = 0 and so H⊥ = 0, we have thus t ρ g ,δ (t) = 0 ρ Ω e−2a(s) F(M ρ )(H⊥ ) · (M − M δ )dxds. (2.41) e−2a(s) F(M )(H⊥ ) · (M − m)dxds. (2.42) Let t g (t) = 0 Ω We have the following result. 32 Proposition 2.1.6. There exists a constant C7 = C7 (R, T ) such that |g ,δ (t)| + |g δ,δ (t)| ≤ C7 lim g δ,δ (t) = 0 δ→0 ∀ 0 < < 1, t ∈ [0, T ] lim g ,δ (t) = g (t) δ→0 ∀ 0 < , δ < 1, t ∈ [0, T ] (2.43) (2.44) ∀ 0 < < 1, t ∈ [0, T ] (2.45) lim g (t) = 0 ∀ t ∈ [0, T ]. (2.46) →0 Proof. It is easy to see that T ρ |g ,δ (t)| ≤ C8 H ρ (s) L2 (R3 ) ds ≤ C7 0 by (2.19). The convergence (2.45) follows easily from the weak convergence M δ M ∞ . The proofs of (2.44) and (2.46) are similar; so we only give the proof of (2.44). To this end, we write g δ,δ (t) = t 0 Ω t = 0 Ω t = 0 Ω δ e−2a F(M δ )(H⊥ ) · (M − M δ ) dxds δ e−2a (F(M δ )T (M − M δ )) · H⊥ dxds (2.47) δ e−2a (F(M δ )T M ) · H⊥ dxds. Let Ωt = Ω × (0, t). Since e−2a ∈ L2 (Ωt ) for all 0 < t ≤ T, the limit (2.44) will be proved if we show that δ (F(M δ )T M ) · H⊥ 0 weakly in L2 (Ωt ) as δ → 0. (2.48) δ Since the sequence {(F(M δ )T M ) · H⊥ } is bounded in L2 (Ωt ), to show its weak convergence δ to 0, we only need to show (F(M δ )T M )·H⊥ → 0 in the sense of distributions on Ωt as δ → 0. We prove this by using Lemma 2.1.2 above. It is easy to see function g(m, n) = F(m)T n 33 restricted to the set BR = {(m, n) | |m| ≤ R, |n| ≤ R}, where R = m0 L∞ (Ω) , is Lipschitz continuous with Lipschitz constant ≤ C9 (R). We can thus extend this function g(m, n) to a Lipschitz function G(m, n) on whole (m, n) ∈ R3 × R3 with Lipschitz constant ≤ C9 (R). δ δ Note that (F(M δ )T M ) · H⊥ = G(M , M δ ) · H⊥ . Hence G(M , M δ ) L2 (Ω ) ≤ C10 < ∞. t (2.49) Since G is Lipschitz and ∂s M and ∂s M δ both exist as integrable functions on Ωt , we have ∂s (G(M , M δ )) exists as an integrable function on Ωt and |∂s (G(M , M δ ))| ≤ C9 (R) (|∂s M | + |∂s M δ |) ≤ C9 (R) (|F (x, M , H )| + |F (x, M δ , H δ )|) and hence ∂s (G(M , M δ )) L2 (Ω ) ≤ C11 < ∞. t (2.50) From first two equations of the Maxwell system (2.8) for δ > 0, with s as time-variable, we deduce δ δ δ 2 ˜δ ˜δ ∆(H⊥ ) = δ∂ss (M⊥ + H⊥ ) = δ(M⊥ + H⊥ )ss (2.51) ˜ in the sense of distributions on R3 × (0, ∞), where again M δ = M δ χΩ . Since the sequences δ ˜δ {H⊥ } and {M⊥ } are bounded in L2 (Ωt ), we have δ 0 H⊥ δ Clearly, H⊥ δ = ∆H⊥ → 0 strongly in H −2 (Ωt ) as δ → 0. (2.52) ∞ H⊥ = 0 weakly in L2 (Ωt ) as δ → 0, by (2.50), (2.52) and Lemma 2.1.2 34 δ above, we have that G(M , M δ ) · H⊥ → 0 in the sense of distributions on Ωt , as δ → 0; hence, (2.48) is proved. This completes the proof of Proposition 2.1.6. Proof of Theorem 2.1.3 We now complete the proof of Theorem 2.1.3. Taking the limit as δ → 0 in (2.39), using (2.26), Fatou’s lemma and (2.43)–(2.45), we have e−a(t) (M (t) − m(t)) 2 2 L (Ω) t ≤ G (t)+C6 η + e−a(s) (M − m) 0 (2.53) 2 ds L2 (Ω) for all 0 < < ξ and all t ∈ [0, T ], where t G (t) = 2g (t) + 2e(t−s)C5 g (s) ds. 0 From (2.53), another use of Gronwall’s inequality yields T 0 T e−a(t) (M − m) 2 2 dt ≤ C12 η + e(T −t)C6 G (t) dt L (Ω) 0 (2.54) for all 0 < < ξ. Since η > 0 is arbitrary, using (2.46), (2.40), we have T lim →0 0 e−a(t) (M − m) 2 2 dt = 0. L (Ω) Hence the sequence {e−a M } converges to e−a m strongly in L2 (ΩT ) and L2 ((0, T ); L2 (Ω)). Since a(x, t) is finite almost everywhere and e−a is not zero almost everywhere, for any sequence of {M } we can extract a subsequence converging point-wise almost everywhere on 35 Ω; the limit must be m from the strong convergence of {e−a M }. By Lebesgue’s dominated convergence theorem, we have that the whole sequence {M } converges to m strongly in L2 (ΩT ). This and (2.53) imply the convergence: M (t) − m(t) L2 → 0 for all t ∈ [0, T ]. On the other hand, the third equation of system (2.8) and (2.12) imply that: M (t1 ) − M (t2 ) L2 (Ω) ≤ C13 |t1 − t2 | ∀ t1 , t2 ∈ [0, T ]. Thus {M } is equi-continuous in C 0 ([0, T ]; L2 (Ω)). This, combined with the point-wise convergence of {M (t)} to m(t) for all t ∈ [0, T ], implies that M → m in C 0 ([0, T ]; L2 (Ω)). The proof is now completed. 2.2 Existence for rough initial data We finish proving the existence of solutions to the Cauchy problem (2.4). Theorem 2.2.1. Let m0 ∈ L∞ (Ω) and let E0 , H0 be any functions satisfying (2.9). Then any function m determined by the convergence (2.20) is a weak solution to the Cauchy problem (2.4). Proof. By (2.22), m ∈ W 1,∞ ((0, ∞); L2 (Ω)). By Theorem 2.1.3, M → m in C 0 ([0, T ]; L2 (Ω)); hence m(0) = m0 because M (0) = m0 . Also, by the strong convergence of M m and weak convergence of H F (x, M , H ) → H ∞ in L2 (ΩT ), we have the weak convergence of F (x, m, H ∞ ) in L2 (ΩT ) along the chosen sequence → 0. From the e- quation ∂t M = F (x, M , H ), it follows that ∂t m = F (x, m, H ∞ ) in L∞ ((0, ∞); L2 (Ω)) and in the sense of distribution on Ω × (0, ∞). Finally, by (2.21), H ∞ = −P (mχΩ ) = Hm and hence m is a weak solution to (2.4). 36 Chapter 3 Local L2-Stability of Solutions in Finite Time From Chapter 2, we have proved the existence of solutions with very rough initial value condition. In this chapter, we will continue to prove the stability result for global weak solutions. We also consider initial value problem (2.1)-(2.3) in Chapter 2 as a quasi-stationary system:    ∂ m = F (x, m, H ) in Ω × (0, ∞),  t  a m     3 curl Hm = 0, div(Hm + mχΩ ) = 0 in R for all t ∈ [0, ∞),       m(x, 0) = m0 (x) on Ω,  (3.1) where Fa (x, m, H), specifying the dependence on applied field a, is the Landau-Lifshitz interaction function given by Fa (x, m, H) = L(m, −ϕ (m) + a(x) + H), with L(m, n) linear in n and defined by L(m, n) = γm × n + γαm × (m × n), 37 m, n ∈ R3 . (3.2) In this chapter, we aim to prove the stability of global weak solutions(see Definition in Section 1.2) to quasi-stationary system (3.1); we first introduce the important Lemma about how we split nonlocal term Hm , then show the main stability result. The last section will be one of applications of stability theorem, which proves the existence of global weak solution to reduced Landau-Lifshitz equation with a(x) ∈ L∞ (Ω) for rough initial data. 3.1 3.1.1 Stability of weak solutions Decomposition of Hm The following lemma aims to handle the nonlocal term Hm . Actually, the stability result will be trivial if we have Hm ∈ L∞ (Ω). However, such bound is valid only when m ∈ H 2 (Ω) (see Lemma 3.2.1). In our case, we only require initial value m ∈ L∞ (Ω); therefore, Hm can not be bounded in L∞ . Following lemma enables us to split Hm into two parts: one is bounded in L∞ ; the other is bounded in L2 (Ω) (also see [13, Lemma 5.2] and [31, Lemma 6.2]). The projection operator used below is the same as Helmholtz Decomposition introduced in Chapter 2, Section 2.1 and it is pretty easy to see that H = −P (mχΩ ) with this operator. ˜ Lemma 3.1.1. Let m ∈ L∞ (Ω; R3 ) and H = −P (mχΩ ) = −m . Then, for all λ ≥ e, H = H λ + (H − H λ ) on R3 , where H λ is a function such that H λ L∞ ≤ C ln λ, H − H λ L2 ≤ with constant C = C0 m L∞ for an absolute constant C0 . 38 C|Ω|1/2 , λ (3.3) Proof. Define H λ by H λ (x) =    H(x) |H(x)| ≤ C ln λ    0  otherwise ˜ ˜ Since H = −m , where m = mχΩ , we have, for all p ≥ 2, H − Hλ 2 2 = L ˜ |m |2 dx ˜ |m |>C ln λ p−2 ˜ L ˜ ≤ m 2 p |{x : |m | > C ln λ}| p 1 ˜ L ˜ p−2 ≤ m 2p m Lp (C ln λ)p−2 1 ˜ p = m Lp . (C ln λ)p−2 However, the boundedness of P on Lp yields that, for all p ≥ 2, ˜ ˜ m Lp ≤ C0 p m Lp ≤ C0 p m L∞ (Ω) |Ω|1/p , where C0 is an abstract constant independent of p for all p ≥ 2; see [45]. Hence |Ω|(C1 p)p H − Hλ 2 2 ≤ , L (C ln λ)p−2 where C1 = C0 m L∞ . We choose C = 4eC1 and p = 4 ln λ ≥ 4 to obtain |Ω|(C1 p)p (C ln λ)2 H − Hλ 2 2 ≤ = |Ω| L (C ln λ)p−2 λ4 and hence H − H λ L2 ≤ C|Ω|1/2 ln λ C|Ω|1/2 ≤ , λ λ2 39 using ln λ ≤ λ for λ ≥ e. This proves (3.3). 3.1.2 Stability Theorem We now prove our main stability result, Theorem 1.2.1, which generalizes our previous result [13, Theorem 5.1] to the case of different applied fields a(x). A similar stability result including the different anisotropy functions ϕ(m) can also be proved; for a similar result on Maxwell systems, see [31, Theorem 6.1]. Assume mk (k = 1, 2) is any weak solution to the problem (3.1) with given applied field ak and initial datum mk satisfying 0 ak L∞ + mk L∞ ≤ R for k = 1, 2, 0 (3.4) where R > 0 is a given constant. Then, Theorem 1.2.1 will be proved once we prove the following result. Theorem 3.1.2. Given any 0 < T < ∞, there exist constants C = C(R, T ) > 0, c = c(R, T ) > 0 and ρ = ρ(R, T ) > 0 such that, if µ = max{ m1 − m2 L2 , a1 − a2 L2 } ≤ c, 0 0 then one has, for all t ∈ [0, T ], m1 (t) − m2 (t) L2 (Ω) ≤ C µρ . (3.5) Proof. Step 1. Let δm = m1 (t) − m2 (t) and δF = Fa1 (x, m1 , H1 ) − Fa2 (x, m2 , H2 ), where Hk = Hmk for k = 1, 2. Then ∂t (δm) = δF and hence ∂t ( δm(t) L2 ) ≤ ∂t (δm(t)) L2 = δF (t) L2 . 40 So we have t δm(t) L2 − δm0 L2 ≤ 0 δF (s) L2 ds. (3.6) Step 2. The function L(m, n) defined by (3.2) above can be written as L(m, n) = B(m) · n, (3.7) where B(m) is a 3×3-matrix for each m ∈ R3 ; note that each element of B(m) is a quadratic function of m. Given any mk , nk ∈ R3 (k = 1, 2), letting δm = m1 − m2 , δn = n1 − n2 , by virtue of L(m1 , n1 ) − L(m2 , n2 ) = [L(m1 , n1 ) − L(m2 , n1 )] + L(m2 , n1 − n2 ), one can write L(m1 , n1 ) − L(m2 , n2 ) = A(m1 , m2 , n1 ) · δm + B(m2 ) · δn, (3.8) where A(m1 , m2 , n1 ) is a matrix function given by A(m1 , m2 , n1 ) = 1 ∂L (tm1 + (1 − t)m2 , n1 ) dt. 0 ∂t (3.9) Step 3. By Remark 1 in Section 2.1, it follows that mk (t) L∞ ≤ R (k = 1, 2) for all t ≥ 0. From Fak (x, mk , Hk ) = −L(mk , ϕ (mk )) + L(mk , ak (x)) + L(mk , Hk ), by (3.4) and (3.8), we obtain the following point-wise estimate for δF : |δF | ≤ A|δH| + B(|H1 | + 1)|δm| + D|δa|, (3.10) where δH = H1 − H2 = Hδm , δa = a1 (x) − a2 (x), and A = A(R), B = B(R), D = D(R) are constants depending only on R. We apply Lemma 3.1.1 to function H1 (t) = −P (m1 (t)χΩ ). λ λ λ For any λ ≥ e, let H1 = H1 + (H − H1 ), where H1 is given in Lemma 3.1.1 with constant 41 C = C0 m1 (t) L∞ ≤ C0 R. So, by (3.10), we have the L2 (Ω)-norm estimate: δF L2 ≤A δH L2 + B(C ln λ + 1) δm L2 1 C|Ω| 2 +B δm L∞ + D δa L2 λ C ≤(A + B ln λ) δm L2 + + D δa L2 , λ (3.11) using δH L2 ≤ Hδm L2 (R3 ) ≤ δm L2 , where constants A , B , C depend on R. Step 4. From (3.6) and (3.11), it follows that t δm(t) L2 − δm0 L2 ≤ δF (s) L2 ds 0 t ≤ 0 C (A + B ln λ) δm(s) L2 + + D a L2 λ ds t Ct δm(s) L2 ds. + δa L2 Dt + (A + B ln λ) λ 0 = From this, a Gronwall inequality yields δm(t) L2 ≤ ≤ Ct δm0 L2 + + δa L2 Dt eA t+B t ln λ λ Ct δm0 L2 + DT δa L2 + λ e A t λB t (3.12) ∀ 0 ≤ t ≤ T. Step 5. We consider two cases. Case 1. Assume both δm0 = 0 and δa = 0. Then, by (3.12), δm(t) L2 ≤ C teA t λB t−1 . Let t0 = 1 . B +1 (3.13) If 0 ≤ t ≤ t0 , then B t − 1 < 0 and hence, by (3.13) with λ → ∞, we have 42 δm(t) = 0 for all t ∈ [0, t0 ]. With mk (t0 ) as initial datum at time t0 , we obtain δm(t) = 0 on [t0 , 2t0 ]; eventually, we have δm(t) = 0 for all t ≥ 0; hence (3.5) holds. This also shows the uniqueness of the weak solution to the system. Case 2. Assume 0 < δm0 L2 + DT δa L2 ≤ 1/e < 1. In this case, setting λ = ( δm0 L2 + DT δa L2 )−1 ≥ e in (3.12), we obtain δm(t) L2 ≤ (1 + C t)eA t ( δm0 L2 + DT δa L2 )1−B t . Let t1 = 1 2(B +1) 1 and C1 = (1 + C t1 )eA t1 > 1. Then 1 − B t ≥ 2 for all 0 ≤ t ≤ t1 ; hence 1 δm(t) L2 ≤ C1 ( δm0 L2 + DT δa L2 ) 2 ∀ 0 ≤ t ≤ t1 . Adding DT δa L2 to both sides, we obtain 1 δm(t) L2 + DT δa L2 ≤ C2 ( δm0 L2 + DT δa L2 ) 2 ∀ 0 ≤ t ≤ t1 , (3.14) where C2 = C1 + 1 depends only on R. Step 6. Combining Cases 1 and 2 in Step 5 above, with the constants t1 = t1 (R) and C2 = C2 (R) > 1 determined above, we have that, if δm0 L2 + DT δa L2 ≤ 1/e, then 1 δm(t) L2 + DT δa L2 ≤ C2 ( δm0 L2 + DT δa L2 ) 2 ∀ 0 ≤ t ≤ t1 . (3.15) Assume 1 C2 ( δm0 L2 + DT δa L2 ) 2 ≤ 1/e. (3.16) Then, by (3.15), δm(t1 ) L2 + DT δa L2 ≤ 1/e. With mk (t1 ) as initial datum at time t1 , 43 we apply (3.15) again to obtain 1 δm(t1 + t) L2 + DT δa L2 ≤C2 ( δm(t1 ) L2 + DT δa L2 ) 2 1+ 1 ≤C2 2 ( 1 ∀ 0 ≤ t ≤ t1 . δm0 L2 + DT δa L2 ) 4 We have thus proved that, if (3.16) holds then 1 1 1+ δm(t) L2 + DT δa L2 ≤ C2 2 ( δm0 L2 + DT δa L2 ) 4 ∀ 0 ≤ t ≤ 2t1 . By induction, we obtain that, for k = 1, 2, · · · , if 1 1+ 2 +···+ C2 1 2k−1 1 ( δm0 L2 + DT δa L2 ) 2k ≤ 1/e, (3.17) then δm(t) L2 + DT δa L2 ≤ 1 1+ 2 +···+ 1 2k ( C2 1 2k+1 ) L2 δm0 L2 + DT δa 1 2 ≤ C2 ( δm0 L2 + DT δa L2 ) 2k+1 ∀ 0 ≤ t ≤ 2k t1 .(3.18) Step 7. In this step, we complete the proof of the theorem. Let k be the integer such that 2k−1 t1 < T ≤ 2k t1 . Define ρ = ρ(R, T ) = 1/(2k+1 ), k 2 c = c(R, T ) = (C2 e)−2 /(1 + DT ). Assume µ = max{ δm0 L2 , δa L2 } ≤ c. Then k 2 δm0 L2 + DT δa L2 ≤ (1 + DT )µ ≤ (C2 e)−2 , 44 from which it is easily seen that (3.17) holds; so, by (3.18), 2 δm(t) L2 + DT δa L2 ≤ C2 ( δm0 L2 + DT δa L2 )ρ ∀ 0 ≤ t ≤ T. Therefore, 2 δm(t) L2 ≤ C2 (1 + DT )ρ µρ ∀ t ∈ [0, T ]; 2 this proves (3.5) with constant C = C2 (1 + DT )ρ . 3.2 Existence of global weak solutions In this section, we present a proof for the existence of global weak solution to (3.1) based on the stability theorem proved above. To this end, we introduce a nonlinear function f (m) = Fa (x, m, Hm ) = −L(m, ϕ (m)) + L(m, a(x)) + L(m, Hm ) (3.19) for m ∈ L∞ (Ω; R3 ), where Hm is defined by Maxwell equation in (3.1) and L is defined by (3.2). As before, we always assume the anisotropy function ϕ : R3 → R3 is smooth. 3.2.1 Properties of f for smooth applied fields ¯ In this subsection, we assume the applied field a belongs to C ∞ (Ω; R3 ) and show that, in this case, f : H 2 (Ω; R3 ) → H 2 (Ω; R3 ) and is locally Lipschitz. We need some estimates. Lemma 3.2.1. Let Ω ⊂ R3 be a bounded domain with smooth boundary. Then the following 45 estimates hold on H 2 (Ω; R3 ) : m L∞ (Ω) + m W 1,p (Ω) ≤ C0 m H 2 (Ω) ∀ 1 ≤ p ≤ 6, Hm H 2 (Ω) ≤ C1 m H 2 (Ω) . (3.20) (3.21) Proof. We omit the proof, but only mention that (3.20) is a simple consequence of the well1 ¯ known embeddings: H 2 (Ω) ⊂ W 1,6 (Ω) ⊂ C 2 (Ω) ⊂ L∞ (Ω) for bounded smooth domain Ω ⊂ R3 , and that estimate (3.21) has been, e.g., proved in [8]. Finally, we remark that, from (3.20) and (3.21), it follows that, with constant C2 = C0 C1 , Hm L∞ (Ω) ≤ C2 m H 2 (Ω) ∀ m ∈ H 2 (Ω; R3 ). (3.22) The main result of the subsection is the following local Lipschitz property of f on H 2 (Ω; R3 ). Proposition 3.2.2. Function f maps space H 2 (Ω; R3 ) into itself and is locally Lipschitz on H 2 (Ω; R3 ). Proof. Since f (0) = 0, the self-mapping property of f will follow from the local Lipschitz property of f on H 2 (Ω; R3 ). To prove the local Lipschitz property of f , given any two functions m1 , m2 ∈ H 2 (Ω; R3 ) satisfying max{ m1 H 2 (Ω) , m2 H 2 (Ω) } ≤ R, 46 (3.23) where R < ∞ is a constant, we need to show that f (m1 ) − f (m2 ) H 2 (Ω) ≤ L m1 − m2 H 2 (Ω) (3.24) for a (local Lipschitz) constant L = L(R) < ∞ depending on R. By (3.19), we write f (m1 ) − f (m2 ) = I1 + I2 , where I1 = L(m1 , a − ϕ (m1 )) − L(m2 , a − ϕ (m2 )) and I2 = L(m1 , Hm1 ) − L(m2 , Hm2 ). Let δm = m1 − m2 . Then, by (3.8), I2 = A(m1 , m2 , Hm1 ) · δm + B(m2 ) · Hδm , (3.25) where A, B are functions defined in Step 2 of the proof of Theorem 3.1.2 above. We also write I1 as 1 I1 = d L(m2 + tδm, a − ϕ (m2 + tδm)) dt = C(m1 , m2 , a) · δm, dt 0 (3.26) where C(m1 , m2 , a) is certain smooth function of (m1 , m2 , a) ∈ R3 × R3 × R3 . Note that C is linear in a. We aim to show Ik H 2 (Ω) ≤ L(R) δm H 2 (Ω) (k = 1, 2) for some constant L(R) depending on R. By (3.23), (3.22) and Lemma 3.2.1, it follows that, 47 for k = 1, 2, mk L∞ (Ω) + Hmk L∞ (Ω) + Hmk H 2 (Ω) + mk L4 (Ω) ≤ C3 R. (3.27) We proceed with several steps. Step 1: Estimation of I1 . Clearly, by (3.26) and (3.27), I1 L2 (Ω) ≤ C(m1 , m2 , a) L∞ δm L2 ≤ L(R) δm L2 (Ω) . We estimate the H 2 -norm. Denote by ∂j the first partial derivative with respect to xj and 2 by ∂ij the second partial derivative with respect to xj and xi (i, j = 1, 2, 3). Note that ∂j (I1 ) = ∂j (C(m1 , m2 , a)) · δm + C(m1 , m2 , a) · (δm)xj and 2 2 ∂ij (I1 ) = ∂ij (C(m1 , m2 , a)) · δm + ∂j (C(m1 , m2 , a)) · (δm)xi + ∂i (C(m1 , m2 , a)) · (δm)xj + C(m1 , m2 , a) · (δm)xi xj . Since ∂j (C(m1 , m2 , a)) = (∂m1 C) · m1 + (∂m2 C) · m2 + (∂a C) · axj has L2 -norm controlled x x j j by R, we have ∂j (I1 ) L2 ≤ ∂j (C(m1 , m2 , a)) L2 δm L∞ + C(m1 , m2 , a) L∞ (δm)xj L2 ≤ L(R) δm H 2 (Ω) . 48 2 Similarly, ∂ij (C(m1 , m2 , a)) contains terms up to second derivatives of a and terms like q 2 (∂mp mq C) · mk · ml and (∂mp C) · mx x , with certain choices of p, q, k, l ∈ {1, 2} and xi xj i j 2 i , j ∈ {i, j}. Hence ∂ij (C(m1 , m2 , a)) L2 is bounded by the quantity C(R) | m1 |2 L2 + | m2 |2 L2 + which, due to | m|2 L2 = 2 m1 L2 + 2 m2 L2 + a H2 , m 2 4 ≤ C m 2 2 , is in fact bounded by another constant H L 2 C(R). From this, similar to the term ∂j (I1 ), the L2 -norm of the first or fourth term of ∂ij (I1 ) 2 is bounded by L(R) δm H 2 (Ω) . The second and third terms of ∂ij (I1 ) can be estimated as follows: ∂j (C(m1 , m2 , a)) · (δm)xi + ∂i (C(m1 , m2 , a)) · (δm)xj L2 ≤2 (C(m1 , m2 , a)) L4 · ≤ C(R) ( m1 W 1,4 + m2 W 1,4 + (δm) L4 a L4 ) · δm W 1,4 ≤ L(R) δm H 2 (Ω) . This proves I1 H 2 (Ω) ≤ L(R) δm H 2 (Ω) . Step 2: Estimation of I2 . We write I2 = I21 + I22 with I21 = A(m1 , m2 , Hm1 ) · δm, I22 = B(m2 ) · Hδm . The term I21 is more like term I1 , except the constant field a is replaced by the fields Hm1 . Since Hm1 ∈ H 2 (Ω) and Hm1 L∞ + Hm1 H 2 (Ω) ≤ C1 m1 H 2 (Ω) ≤ C1 R, estimation resulting from Hm1 in A can be handled in a much similar way as the term a in C of I1 . 49 The term I22 is simpler but slightly different than I1 in that Hδm is in place of δm. Nevertheless this term can also be estimated in a similar fashion as I1 , using the following estimate on Hδm : Hδm L∞ (Ω) + Hδm L4 (Ω) + Hδm H 2 (Ω) ≤ C5 δm H 2 (Ω) . We eventually obtain I2 H 2 (Ω) ≤ L(R) m H 2 (Ω) . This completes the proof. 3.2.2 Existence of global solution for smooth data ¯ We continue to assume a ∈ C ∞ (Ω; R3 ) in this subsection. Let X = H 2 (Ω; R3 ). With function f : X → X defined above, we formulate the problem (1.7) as an abstract ODE on X by    dm   dt = f (m), (3.28)   m(0) = m .  0 A solution m to (3.28) on [0, T ] is a function m ∈ C([0, T ]; X) that satisfies t m(t) = m0 + f (m(s)) ds ∀ 0 ≤ t ≤ T. 0 We say m is a solution to (3.28) on [0, T ) if m is a solution on [0, T ] for all 0 < T < T (in this case T could be ∞). Theorem 3.2.3. Given any m0 ∈ X, (3.28) has a unique solution m on [0, ∞). This solution is also a global weak solution to problem (3.1). Proof. Given m0 ∈ X, since f is locally Lipschitz on X, from the abstract theory, there 50 exists T > 0 such that (3.28) has a unique solution m on [0, T ]. Let T∗ = sup T > 0 (3.28) has a unique solution on [0, T ] . We claim that T∗ = ∞, which implies that (3.28) has a unique global solution m defined on [0, ∞). Clearly, this solution is also a global weak solution to the Cauchy problem (1.7) above. Suppose T∗ < ∞. Then, by the elementary ODE theory, a solution m to (3.28) would exist on [0, T∗ ) and satisfy lim − t→T∗ m(t) X = ∞. The following asserts that this finite-time blowup is impossible; this completes the proof of Theorem 3.2.3. Theorem 3.2.4. Given any T > 0, if m is a solution to (3.28) on [0, T ), then sup t∈[0,T ) < ∞. m(t) X ≤ CT, m 0 X (3.29) The proof of this theorem involves lots of technical estimates and will be delayed to the next individual section. 3.2.3 Existence of global weak solution for rough data In this subsection, we assume both applied field a and initial datum m0 are in L∞ (Ω; R3 ). 51 ¯ Let a , m0 ∈ C ∞ (Ω; R3 ) be such that a L∞ + m0 L∞ ≤ R ∀ > 0, lim ( a − a L2 + m0 − m0 L2 ) = 0, →0+ a → a, m0 → m0 pointwise in Ω. (3.30) (3.31) (3.32) Consider the Cauchy problem (3.1) with applied field a and initial datum m0 . Then, by Theorem 3.2.3, for each > 0, (3.1) has a global weak solution m . Since m · f (m ) = 0, it follows that ∂t (|m (x, t)|2 ) = 0 and hence |m (x, t)| = |m0 (x)| for a.e. x ∈ Ω and all t > 0. This implies m (t) L∞ = m0 L∞ ≤ R. (3.33) For each n ∈ {1, 2, 3, · · · }, our stability result (Theorems 1.2.1 and 3.1.2) implies that sequence {m } is Cauchy in Banach space C([0, n]; L2 (Ω; R3 )) as → 0+ . Therefore, m → m in C([0, n]; L2 (Ω; R3 )) as → 0+ for some m ∈ C([0, n]; L2 (Ω; R3 )). (Presumably, m = mn depends on n.) Hence, by (3.31), m(0) = m0 . (3.34) We also have Hm → Hm in C([0, n]; L2 (Ω; R3 )). It follows that m → m and Hm → Hm also in L2 (Ω × (0, n)) as → 0+ . Using a subsequence, we can assume m (x, t) → m(x, t), Hm (x, t) → Hm (x, t) pointwise in Ω × (0, n). Therefore, Fa (x, m , Hm ) → Fa (x, m, Hm ) pointwise in Ω × (0, n). This shows ∂t m = 52 Fa (x, m, Hm ) in the sense of distribution on Ω × (0, n). Note also that Fa (x, m , Hm ) ∈ L2 (Ω; R3 ) uniformly on and t ∈ (0, n); this implies that ∂t m = Fa (x, m, Hm ) holds in L∞ ((0, n); L2 (Ω)) and that m ∈ W 1,∞ ([0, n); L2 (Ω; R3 )). Combining with (3.34), we have proved that m = mn is a weak solution to (3.1) on Ω×(0, n). By the uniqueness of weak solutions, we have mn+1 = mn on Ω × (0, n); therefore, the sequence {mn }∞ defines a unique function m by setting m(x, t) = mn (x, t) with n = [t] + 1. 1 It is easy to see that m is a global weak solution to (3.1). Finally, we have proved the following theorem. Theorem 3.2.5. Let a ∈ L∞ (Ω; R3 ). Given any initial datum m0 ∈ L∞ (Ω; R3 ), the problem (3.1) has a unique global weak solution. 3.2.4 Proof of Theorem 3.2.4 In this separate section, we give the proof of Theorem 3.2.4. This involves the special form of function L(m, n) and several estimates. ¯ In what follows, assume a ∈ C ∞ (Ω; R3 ), 0 < T < ∞ and m is a solution to (3.28) on [0, T ) with initial datum m0 ∈ H 2 (Ω; R3 ). Assume m0 L∞ (Ω) = R > 0. Then, similar to (3.33) above, we have m(t) L∞ = m0 L∞ = R, 1 m(t) L2 = m0 L2 ≤ R|Ω| 2 53 ∀ 0 ≤ t < T. (3.35) We would like to show sup t∈[0,T ) m(t) H 2 (Ω) ≤ CT, m < ∞. 0 H2 (3.36) Let y(t) = 1 + m(t) 2 2 = 1 + m(t) 2 2 + H (Ω) L m(t) 2 2 + L 2 m(t) 2 . L2 The goal is to show y (t) ≤ Cy(t)(1 + ln y(t)) ∀ 0 < t < T, (3.37) where C = C(R) is a constant depending on R. Once (3.37) is proved, one easily obtains that ln(y(t)) ≤ (ln(y(0)) + 1) eCT < ∞ ∀ t ∈ [0, T ), from which (3.36) follows. The rest of the section is devoted to proving (3.37). Energy estimates It is convenient to use the special structure of function L to write function f (m) as follows: f (m) = B(m) · a − B(m) · ϕ (m) + B(m) · Hm , where B(m) is a 3 × 3 matrix defined in (3.7) above, whose elements are quadratic functions of m; hence B (m) = D is a constant tensor. However, this special structure of B is not used; in fact, the following arguments are valid for arbitrary smooth functions B. 54 Differentiating equation in (3.28) with respect to xi yields dmxi = B (m) · mxi · a + B(m) · axi dt − B (m) · mxi · ϕ (m) − B(m) · ϕ (m) · mxi (3.38) + B (m) · mxi · Hm + B(m) · (Hm )xi . Again, differentiating equation (3.38) with respect to xj yields dmxi xj dt = D · mxj · mxi · a + B · mxi xj · a + B · mxi · axj + B · mxj · axi + B · axi xj − D · mxj · mxi · ϕ − B · mxi xj · ϕ + B · mxi · ϕ · mxj (3.39) − B · mxj · ϕ · mxi − B · ϕ · mxj · mxi − B · ϕ · mxi xj + D · mxj · mxi · Hm + B · mxi xj · Hm + B · mxi · (Hm )xj + B · mxj · (Hm )xi + B · (Hm )xi xj . Dot-product of (3.38) with mxi and of (3.39) with mxi xj yields the following identities: 1d 2 dt 1d 2 dt mxi 2 2 L = mxi xj 2 2 L = Ω Ω dmxi dx, dt dmxi xj dx. mxi xj · dt mxi · (3.40) (3.41) The energy estimates involve estimating the right-hand sides of (3.40) and (3.41) with terms dmxi dmxi xj dt , dt given by the right-hand sides of (3.38) and (3.39). 55 More subtle inequalities To handle the terms involved in the integrals on the right-hand sides of (3.40) and (3.41), more subtle inequalities are needed. Lemma 3.2.6. Let Ω ⊂ R3 be a bounded domain with smooth boundary. Then 1 1 2 n L4 ≤ C6 n L∞ n 2 2 , H Hn L∞ (Ω) ≤ C n ∞ L ∀ n ∈ H 2 (Ω; R3 ), (1 + ln+ ( (3.42) n H 2 )), where ln+ t = max{ln t, 0} for t > 0 and C n ∞ < ∞ depends on n L∞ (Ω) . L Proof. The first inequality of (3.42) is a consequence of the well-known Gagliardo-Nirenberg inequality: jf Lq (Rn ) ≤ C f 1−θ n ) Lr (R lf θ Lp (Rn ) , where θ = j/l ∈ (0, 1) and 1/q = θ/p + (1 − θ)/r, 1 ≤ p, r ≤ ∞. Here j = 1, l = 2, p = 2, q = 4, r = ∞ and θ = 1/2. While the second inequality of (3.42) is a Judovic-type inequality proved, e.g., in [31, Lemma 7.2]. The following result is an immediate consequence of this lemma and (3.35). Proposition 3.2.7. For the solution m(t), with y(t) defined above, it follows that m(t) 4 4 ≤ C7 y(t), L (Ω) Hm (t) L∞ (Ω) ≤ C8 (1 + ln y(t)), where C7 , C8 are constants depending on R = m0 L∞ . 56 ∀ 0 ≤ t < T, (3.43) Energy estimates (continued) and proof of (3.37) First of all, the integral on right-hand side of (3.40) is bounded by | m|2 + | m| + | m|2 |Hm | + | m|| (Hm )| dx. C(R) Ω The third term is bounded by C(R) Hm L∞ (Ω) m 2 2 and hence, by (3.43b), is bounded L by C(R)y(t)(1 + ln y(t)), while all the other terms are bounded by C(R) m 2 2 and hence H by C(R)y(t). Therefore, d dt mxi 2 2 ≤ C(R)y(t)(1 + ln y(t)), L ∀ 0 < t < T. (3.44) Similarly, the integrand of the right-hand side of (3.41) is bounded by a constant C(R) times | m|2 | 2 m| + | 2 m|2 + | m|| 2 m| + | 2 m| + | 2 m|| 2 (Hm )| +| m|2 |Hm || 2 m| + |Hm || 2 m|2 + | m|| 2 m|| (Hm )|. Integrals of terms in the first group can all be bounded by Cy(t). Integrals of the first two terms in the second group can be bounded by constant times Hm L∞ (Ω) ( m 44 + L 2 m 2 ), L2 which, by (3.43a-b), is bounded by Cy(t)(1 + ln y(t)). Finally, the integral of the last term in the second group can be estimated as follows: | m|| 2 m|| (Hm )|dx ≤ | m| · | (Hm )| L2 (Ω) | 2 m| L2 (Ω) Ω 57 ≤ (Hm ) L4 (Ω) | 2 m| L2 (Ω) , m L4 (Ω) which, by using Lemma 3.2.6, is bounded by 1 1 1 2 ≤C m 22 Hm L∞ (Ω) Hm 2 2 m H 2 (Ω) H (Ω) H (Ω) 1 1 1 1 2 2 ≤C m 22 Hm L∞ (Ω) m 2 2 m H 2 (Ω) = C m 2 2 Hm L∞ (Ω) H (Ω) H (Ω) H (Ω) 1 ≤ C y(t) · (1 + ln y(t)) 2 ≤ C y(t)(1 + ln y(t)). Therefore, by (3.41), we have obtained that d dt mxi xj 2 2 ≤ C(R)y(t)(1 + ln y(t)), L ∀ 0 < t < T. (3.45) Summing up i, j = 1, 2, 3 in (3.44) and (3.45) and using (3.33), we obtain (3.37). Remark 2. By the local Lipschitz property of f (m), from (3.36), one easily obtains sup t∈[0,T ) mt H 2 (Ω) ≤ CT, m < ∞. 0 H2 In next chapter, we prove higher time-regularity for solutions. 58 (3.46) Chapter 4 Higher Time Regularity and Special Asymptotics 4.1 Higher time regularity The higher time regularity has been studied for Landau-Lifshitz equation with exchange energy by Cimrak and Keer [6]. We study a higher time-regularity of weak solutions for simple Landau-Lifshitz equation    m = γm × H + γαm × (m × H ) in Ω × (0, ∞),  t m m (4.1)   m(0) = m ,  0 where Ω is a bounded smooth domain in R3 and m0 ∈ H 2 (Ω; R3 ). Theorem 4.1.1. For any time T > 0, the solution m to (4.1) satisfies, for p = 0, 1, 2, · · · , sup t∈[0,T ] p+1 ∂t m H 2 (Ω) ≤ C < ∞, (4.2) where C is constant depending on T, p and m0 H 2 (Ω) . Proof. We use induction on p. The case for p = 0 is already mentioned in Remark 2 in Section 3.2.4. Let us assume (4.2) holds for all powers up to p − 1. We consider the case for 59 i p. Note that ∂t (Hm ) = H∂ i m and hence, by (3.22), t i H∂ i m H 2 (Ω) ≤ C ∂t m H 2 (Ω) . t Therefore, by the induction assumption, it follows that, for all t ∈ [0, T ], i i ∂t (Hm ) H 2 (Ω) ≤ C ∂t m H 2 (Ω) ≤ CT,p, m < ∞ ∀ 0 ≤ i ≤ p. 0 H2 (4.3) Taking pth -derivatives with respect to t to equation (4.1) yields p+1 ∂t j i+j=p p+1 We need to prove ∂t j i ∂t m × ∂t Hm + γα m=γ i k ∂t m × (∂t m × ∂t Hm ). (4.4) i+j+k=p m H 2 (Ω) ≤ CT,p, m < ∞. 0 H2 Estimation of ∂tp+1 m L2 (Ω) Since, ∀ 0 ≤ i ≤ p, i i ∂t (Hm ) L∞ (Ω) ≤ C ∂t (Hm ) H 2 (Ω) ≤ CT,p, m < ∞, 0 H2 the L2 -norm of each term on the right-hand side of (4.4) can be bounded by the L∞ -norms of its factors, which are in turn bounded by constant CT,p, m . Hence we have 0 H2 p+1 ∂t m L2 (Ω) ≤ CT,p, m . 0 H2 60 (4.5) Estimation of ∂tp+1 m L2 (Ω) Taking ∂l = ∂xl on equation (4.1) yields ∂l mt = γ∂l (m × Hm ) + γα∂l (m × (m × Hm )) = γmxl × Hm + γm × (Hm )xl + γαm × (mxl × Hm ) (4.6) + γαm × (m × (Hm )xl ) + γαmxl × (m × Hm ) Taking pth derivative with respect to t on Eq. (4.6) yields p+1 ∂t j j i ∂t m × ∂t (Hm )xl i ∂t mxl × ∂t Hm + γ mxl =γ i+j=p i+j=p j j i k ∂t m × (∂t mxl × ∂t Hm ) i k ∂t m × (∂t m × ∂t (Hm )xl ) + γα γα i+j+k=p i+j+k=p j i k ∂t mxl × (∂t m × ∂t Hm ) + γα i+j+k=p p+1 In order to estimate ∂t mxl L2 (Ω) , it is sufficient to estimate the following L2 -norms. j i ∂t mxl × ∂t Hm L2 (Ω) . i+j=p j i ∂t m × ∂t (Hm )xl L2 (Ω) . i+j=p j k i ∂t m × (∂t m × ∂t (Hm )xl ) L2 (Ω) . i+j+k=p j i k ∂t m × (∂t mxl × ∂t Hm ) L2 (Ω) . i+j+k=p j i k ∂t mxl × (∂t m × ∂t Hm ) L2 (Ω) . i+j+k=p 61 All these norms can be estimated in the same way: For each of the individual cross-product integrands, use the L2 -norm of a sole factor with xl -derivative and use the L∞ -norms for the other factor or factors. All these norms can be bounded by constant CT,p, m < ∞. 0 H2 Finally, summing up l = 1, 2, 3, we have proved p+1 m L2 (Ω) ≤ CT,p, m < ∞. 0 H2 ∂t Estimation of ∂tp+1 m (4.7) L2 (Ω) Differentiating (4.6) with respect to xl and summing up over l = 1, 2, 3 yields that mt = γ m × Hm + γm × mxl × (Hm )xl Hm + γ l + γα[ m × (m × Hm ) + m × ( m × Hm ) + m × (m × (4.8) Hm )] [mxl × (mxl × Hm ) + mxl × (m × (Hm )xl ) + m × (mxl × (Hm )xl )]. + γα l p+1 Differentiating equation (4.8) p times with respect to t will yield a formula for ∂t p+1 To estimate ∂t m. m L2 (Ω) , we do not need to estimate every single term because lots of them are similar; it is sufficient to estimate the following 4 L2 -norms: i ∂t m × ∂t Hm L2 (Ω) . j (4.9) j (4.10) i+j=p i ∂t mxl × ∂t (Hm )xl L2 (Ω) . i+j=p j (4.11) j (4.12) i k ∂t m × (∂t m × ∂t Hm ) L2 (Ω) . i+j+k=p i k ∂t mxl × (∂t mxl × ∂t Hm ) L2 (Ω) . i+j+k=p 62 For (4.9), we use j j i i ∂t m × ∂t Hm L2 (Ω) ≤ ∂t Hm L∞ ∂t m L2 (Ω) . For (4.10), we use j j i i ∂t mxl × ∂t (Hm )xl L2 (Ω) ≤ ∂t m L4 (Ω) ∂t Hm L4 (Ω) . For (4.11), we use j j i k k i ∂t m × (∂t m × ∂t Hm ) L2 (Ω) ≤ ∂t m L∞ ∂t Hm L∞ ∂t m L2 (Ω) . For (4.12), we use j j i k k i ∂t mxl × (∂t mxl × ∂t Hm ) L2 ≤ ∂t Hm L∞ ∂t m L4 ∂t m L4 . Finally, from these estimates, we obtain p+1 ∂t m L2 (Ω) ≤ CT,p, m < ∞. 0 H2 (4.13) Combining (4.5), (4.7) and (4.13), we have shown that p+1 ∂t m H 2 (Ω) ≤ CT,p, m < ∞. 0 H2 This completes the induction process and hence the proof. Remark 3. Theorem 4.1.1 is also valid for the general equation (3.28) with smooth functions 63 ϕ(m) and a(x); the proof should be similar. 4.2 Weak ω-limit sets We first study the energy decay for global weak solutions to the Landau-Lifshitz equation (4.1). We write the initial value problem as    m = L(m, H ) in Ω × (0, ∞),  t eff (4.14)   m(0) = m ,  0 in terms of the Landau-Lifshitz interaction function L defined by (3.2), where the effective magnetic field Heff is given by (2.2). 4.2.1 The energy identity Let E(m) be defined by (1.1). Assume a, m0 ∈ L∞ (Ω; R3 ). Theorem 4.2.1. The global weak solution m to (4.14) satisfies the energy identity t E(m(t)) − E(m(s)) = γα s Ω |m × Heff |2 dxdτ ∀ 0 ≤ s ≤ t < ∞. (4.15) Furthermore, if γα < 0, then mt ∈ L2 ((0, ∞); L2 (Ω; R3 )). Proof. By the definition of Heff , (4.15) follows from the identity L(m, n) · n = −αγ|m × n|2 for all m, n ∈ R3 and the property d (E(m(t))) = − Heff · mt dx a.e. (0, ∞). dt Ω 64 ∞ If γα < 0, by (4.15), it follows that 0 Ω |m × Heff |2 dxdt ≤ E(m0 )/|αγ| < ∞. Also from the equation (4.14), |mt |2 =|L(m, Heff )|2 =γ 2 |m × Heff |2 + (γα)2 |m|2 |m × Heff |2 ≤C |m × Heff |2 , where constant C depends on m0 L∞ . Hence mt ∈ L2 ((0, ∞); L2 (Ω; R3 )). Remark 4. The stability result and all the regularity estimates previously proved are for finite time; the only global space for the solutions (even for the regular solutions) is m ∈ L∞ ((0, ∞); L∞ (Ω; R3 )) with mt ∈ L2 ((0, ∞); L2 (Ω; R3 )). But this space is not enough to have strong convergence as t → ∞; it would be enough if one has mt ∈ L1 ((0, ∞); L2 (Ω; R3 )) (see [32]). Therefore, it is quite challenging to study the asymptotic behaviors of even regular solutions. The solution orbits for general initial data may not have strong ω-limit points. See [52] for some recent results on the estimation of weak ω-limit set. 4.2.2 Weak ω-limit sets and the estimate for soft-case Given m0 ∈ L∞ (Ω; R3 ), let m be the global weak solution to the initial value problem (4.14) and define the weak ω-limit set for m to be ˜ ω ∗ (m0 ) = {m | ∃ tj ↑ ∞ such that m(tj ) 65 ˜ m weakly in L2 (Ω; R3 )}. (4.16) We give an estimate of ω ∗ (m0 ) for the so-called soft-case, where there is no anisotropy energy (ϕ = 0). For more results on further special case when a = 0, see [51, 52]. Theorem 4.2.2. Let γα < 0, ϕ = 0 and a ∈ L∞ (Ω; R3 ). Then, for any m0 ∈ L∞ (Ω; R3 ) with |m0 (x)| = 1 a.e. on Ω, it follows that ˜ ˜ ˜ ω ∗ (m0 ) ⊆ {m ∈ L∞ (Ω; R3 ) | |m|2 + 2|m × (a + Hm )| ≤ 1 a.e. on Ω}. ˜ (4.17) Proof. Let m be the global weak solution to (4.14) with the given initial datum m0 . Then ˜ m weakly in L2 (Ω; R3 ) for a sequence |m(t)| = 1 a.e. on Ω for all t ≥ 0. Assume m(tj ) tj ↑ ∞. In the following, we show that ˜ ˜ |m|2 + 2|m × (a + Hm )| ≤ 1 a.e. on Ω. ˜ (4.18) Let e(t) = E(m(t)). Then, by (4.15), e(t) is non-increasing and bounded and hence e(t) has limit as t → ∞; this again by (4.15) implies e(tj + 1) − e(tj ) = γα tj +1 tj m(t) × (a + Hm(t) ) 2 2 dt → 0. L Hence there exists some sj ∈ [tj , tj + 1] such that m(sj ) × (a + Hm(s ) ) L2 (Ω) → 0. j (4.19) By Theorem 4.2.1, mt ∈ L2 ((0, ∞); L2 (Ω; R3 )); hence sj m(sj ) − m(tj ) L2 ≤ tj mt (t) L2 dt ≤ 66 1 (sj − tj ) 2 sj tj mt 2 2 dt L 1 2 → 0, which yields m(sj ) ˜ m weakly in L2 (Ω; R3 ). Therefore, by (4.19), (4.18) follows from the following lemma with mj = m(sj ). This completes the proof. Lemma 4.2.3. Let mj (a) ˜ m weakly in L2 (Ω; R3 ) and satisfy the following conditions: |mj | = 1 a.e. Ω; mj × (a + Hmj ) L2 (Ω) → 0. (b) ˜ Then m satisfies the condition (4.18) above. Proof. This result can be proved by a similar method as used for [51, Theorem 1.1]. However, we present a different but direct proof based on the div-curl lemma [46]. For any m ∈ L∞ (Ω; R3 ), let Gm = mχΩ + Hm . Then div Gm = 0 on R3 . Denote Gj = a + Gmj , Hj = a + Hmj ; Then Gj ˜ G, Hj ˜ ˜ G = a + Gm , H = a + Hm . ˜ ˜ ˜ H weakly in L2 (Ω; R3 ) and, by the div-curl lemma [46], Ω Gj · Hj φ dx → Ω ∞ ˜ ˜ G · Hφ dx ∀ φ ∈ C0 (Ω). (4.20) Since mj = Gj − Hj on Ω, it follows that |mj |2 + 2|mj × (a + Hmj )| =|Gj − Hj |2 + 2|Gj × Hj | =|Gj |2 + |Hj |2 (4.21) + 2|Gj × Hj | − 2Gj · Hj . Note that function f (m, n) = |m|2 + |n|2 + 2|m × n| is convex on (m, n) ∈ R3 × R3 . Hence, 67 ∞ for all φ ∈ C0 (Ω) with φ ≥ 0, one has lim inf j→∞ Ω (|Gj |2 + |Hj |2 + 2|Gj × Hj |)φ ≥ ˜ ˜ ˜ ˜ (|G|2 + |H|2 + 2|G × H|)φ. (4.22) Ω By assumptions (a), (b), from (4.20)–(4.22), it follows that ˜ ˜ ˜ ˜ ˜ ˜ (|G|2 + |H|2 + 2|G × H| − 2G · H)φ dx ≤ Ω φ dx Ω ∞ for all φ ∈ C0 (Ω) with φ ≥ 0. This implies ˜ ˜ ˜ ˜ ˜ ˜ |G|2 + |H|2 + 2|G × H| − 2G · H ≤ 1 a.e. Ω, which, exactly, is equivalent to (4.18). This completes the proof. 4.3 Special dynamics for constant initial data on ellipsoid domains We now study a special case of (4.14) where applied field a(x) = a is constant, domain Ω is an ellipsoid, and initial datum m0 is a constant unit vector. Therefore, in (4.14), the effective magnetic field Heff is now given by Heff = −ϕ (m) + a + Hm 68 as above, but with constant vector a. In what follows, we assume ellipsoid domain Ω is given by 3 Ω = {x ∈ R3 x2 /ai < 1}, i | i=1 where ai > 0 are constants. 4.3.1 Associated ODE system on R3 Let bi (i = 1, 2, 3) be defined as below: √ a1 a2 a3 dt 1 ∞ bi = . 2 0 (ai + t) (a1 + t)(a2 + t)(a3 + t) (4.23) Note that, if Ω is a ball in R3 , all bi ’s are equal to 1/3. Theorem 4.3.1 (Pedragal and Yan [44]). For each k = 1, 2, 3, the Dirichlet-Neumann problem ω = 0 in Ωc (4.24) with boundary condition ω|∂Ω = xk , ∂ω 1 |∂Ω = (1 − )νk ∂ν bk ν is the unit normal on ∂Ω pointing outward of Ω, has an unique solution ω satisfies ω ∈ 1 Hloc (Ωc ) and | ω| ∈ L2 (Ωc ). With Theorem 4.3.1, one can prove following lemma; the proof is in [44]. Lemma 4.3.2. Let ω = (ω1 , ω2 , ω3 ) be the functions determined in Theorem 4.3.1. For each 69 k = 1, 2, 3, let uk (x) =    b x  k k x ∈ Ω,   b ω (x)  k k 1 Then uk ∈ Hloc (Rn ), F k = (4.25) x ∈ Ωc uk ∈ L2 (Rn ; Rn ), and curl F k = 0, div(−F k + ek χΩ ) = 0, in R3 , (4.26) where e1 , e2 , e3 are the standard basis vectors of R3 . From this lemma, the magneto-static stray field Hm induced by any constant field m has constant value on Ω given by Hm |Ω = −Λm (∀ m ∈ R3 ), where Λ is a diagonal matrix of positive numbers. In fact, Λ = diag (b1 , b2 , b3 ). Let m0 be a constant unit vector. Then, problem (4.14) reduces to the following ODE system on m ∈ R3 :    ˙ m = Φ(m), t > 0, (4.27)   m(0) = m ,  0 where function Φ : R3 → R3 is defined by Φ(m) = L(m, −ϕ (m) + a − Λm), ∀ m ∈ R3 . (4.28) Since m · Φ(m) = 0, system (4.27) also preserves the length of m(t). Thus we have |m(t)| = 1 for all t ≥ 0. Moreover, L(m, n) = 0 if and only if m × n = 0; hence, the 70 equilibrium points of (4.27), that is, the solutions of Φ(m) = 0 on unit sphere |m| = 1, are characterized by vectors m ∈ R3 for which there is a real number λ ∈ R such that −ϕ (m) + a − Λm = λm, |m| = 1. (4.29) This condition is equivalent to m being a critical point of the function 1 P (m) = Λm · m − a · m + ϕ(m) 2 (4.30) on unit sphere |m| = 1. In most cases, there will be at least two distinct equilibrium points for system (4.27); for example, all maximum or minimum points of P on |m| = 1 (always exist) are such points. 4.3.2 Lyapunov function and the special dynamics The dynamics of system (4.27) can be studied by the classical ODE theory. For example, we have the following result. Theorem 4.3.3. Function P is a Lyapunov function for system (4.27). The ω-limit set of (4.27) is contained in the set of all critical points of P on unit sphere. Proof. Since Heff = −ϕ (m) + a − Λm = − P (m), ∀ m ∈ R3 , it follows that for solution m = m(t) of (4.27) d P (m(t)) = dt P (m) · mt = −Heff · L(m, Heff ) = γα|m × Heff |2 ≤ 0. Hence P is a Lyapunov function for system (4.27). 71 ¯ To show the second part of the theorem, assume m(tj ) → m for a sequence tj ↑ ∞. Let p(t) = P (m(t)). Then p(t) is smooth and has limit as t → ∞. Hence p(tj + 1) − p(tj ) = p (sj ) → 0 for some sj ∈ (tj , tj + 1). Since γα < 0, this implies |m(sj ) × Heff (sj )| = |m(sj ) × P (m(sj ))| → 0. (4.31) As above, since mt ∈ L2 (0, ∞), one has |m(sj ) − m(tj )| ≤ sj tj |mt |dt ≤ 1 (sj − tj ) 2 ¯ ¯ This implies m(sj ) → m; hence, by (4.31), |m × sj tj |mt |2 dt 1 2 → 0. ¯ ¯ P (m)| = 0, which proves m is a critical point of P on unit sphere. This completes the proof. Finally, we prove a special result for a = 0 and ϕ = 0. Proposition 4.3.4. Let b1 , b2 , b3 be positive numbers determined by (4.23). If bk = min{b1 , b2 , b3 }, then ±ek are asymptotically stable equilibrium points to the system (4.27), where e1 , e2 , e3 are the standard basis vectors of R3 . Proof. Without loss of generality, let us assume b1 = min{b1 , b2 , b3 }. It is trivial to see that P has a strict relative minimum at (±1, 0, 0). According to the Lyapunov stability theorem, (±1, 0, 0) are the asymptotically stable equilibrium points. 72 Chapter 5 Existence for Fractional Landau-Lifshitz Equations In this chapter, we prove the existence of global weak solution to fractional LandauLifshitz equation including magneto-static term with periodic boundary condition: mt = γm × Fm + γm × (m × Fm ) in Rn , (5.1) where Fm = ∧2α m + Hm 0 < α < 1. (5.2) Let Ω = [0, 2π]n denote the periodic box in all directions and Z denote the dual lattice associated to Ω in Rn . The two terms in energy form (5.2), for x ∈ Ω, are defined as:   (ξ · m) · ξ  ,  |ξ|2 Hm (ξ) =    0, ξ = 0. ξ = 0. (5.3) 1 The operator ∧ denotes the square root of Laplacian (− ) 2 , so ∧β m can be understood in terms of Fourier transform: F(∧β m(x)) = |ξ|β m(ξ) 73 (5.4) Definition 5.0.1. Let m0 ∈ H α , |m0 | = 1, m is a global weak solution to equation (5.1) if m ∈ L∞ ((0, T ); H α (Ω)) ∂m ·ϕ=γ (m × Fm ) · ϕ + γ m × (m × Fm ) · ϕ QT QT QT ∂t for all ϕ ∈ C ∞ (QT ). Here QT = (0, T ) × Ω. Remark 5. We remark that, as being the same as weak solution in Chapter 2, the length |m(x, t)| is preserved for weak solutions; therefore, if the initial datum m0 satisfies the saturation condition |m0 (x)| = 1 then any weak solution m(x, t) to the fractional equation (5.1) will also satisfy the saturation condition: |m(x, t)| = 1 for all t > 0. 5.1 Notations and preliminaries Let Lp denote the space of all the p-th integrable function f in Ω with associated norm given by: 1 f (x) Lp = ( |f (x)|p ) p Ω For any r ∈ R, we define the homogeneous Sobolev space H r for all tempered distribution function f (x) such that f H r is finite, where · H r is defined through Fourier transform: f H r = ∧r f (x) L2 = ( ξ∈Z 74 1 |ξ|2r |f (ξ)|2 ) 2 For 1 ≤ p ≤ ∞ and r ∈ R, the space H r,p (Ω) consists of all f (x) such that f = ∧−r g, where g ∈ Lp (Ω). The corresponding norm is given by: f H r,p = F−1 (|ξ|r f ) Lp The following commutator estimate comes from a general result of Coifman and Meyer [9]; see also [35, 36]. Lemma 5.1.1. Suppose that s > 0 and p ∈ (1, ∞). Then ∧s (f g) − f ∧s g Lp ≤ C f Lp1 g s−1,p2 + f H s,p3 g Lp4 , H where p1 , p2 , p3 , p4 ∈ (1, ∞) satisfy 1 1 1 1 1 = + = + p p1 p2 p3 p4 and f, g are such that the right hand side terms make sense. For fractional derivative, we also have Hardy-Littlewood-Sobolev theorem; the proof can be found in Stein [45]. Lemma 5.1.2. Suppose that q > 1, p ∈ [q, ∞] and 1 s 1 + = p d q Assume that ∧s f ∈ Lq , then f ∈ Lp and there is a constant C > 0 such that f Lp ≤ C ∧s f Lq 75 As a corollary, if f = ∧−s g for g ∈ Lq , then g H −s,p = ∧−s g Lp ≤ C g Lq Later we are going to apply the Galerkin method; therefore, we need to consider following eigenvalue problem,   2α   ∧ u = µu (5.5)    with periodic boundary conditions. Because ∧−2α is a compact self-adjoint operator in L2 (Ω), there exists a completed orthogonal family of L2 (Ω), say {ωj }j∈N , being eigenvectors of ∧−2α , ∧−2α ωj = µj ωj where sequence µj is decreasing and tends to 0. By setting νj = µ−1 , we obtain, j ∧2α ωj = νj ωj where 0 < ν1 ≤ ν2 ≤ ... The family {ωj } satisfies (ωj , ωk ) = δjk , Kronecker symbol 76 (5.6) 5.2 A priori estimates We let Vn = n k=1 ωk ∈ L2 (Ω), where ωk is eigenvectors in (5.6). We look for solutions to approximate fractional Landau-Lifshitz equation (5.1) in the form of N mN = ϕk,N (t)ωk (x) (5.7) k=1 In order to show such {ϕk,N (t)} exist, we take inner product of equation (5.1) with each ωk (x), k = 1, 2, ..., N , ∂mN mN × (mN × FmN ) · ωk (x) · ωk (x) = γ mN × FmN · ωk (x) + γ Ω Ω Ω ∂t (5.8) where FmN = ∧2α mN + PN (HmN ) We define PN as the orthogonal projection, ˙ PN : H r (Ω) → VN = {ω1 , ω2 , ..., ωN } The local existence of solutions {ϕk,N (t)} is justified by standard Picard’s theorem. In order to extend the local weak solution to [0, T ] and take N → ∞, we need to establish following a priori estimate, Lemma 5.2.1. Let m0 ∈ H α (Ω), then for any 0 < T < ∞, the approximate solutions mN to systems (5.8) satisfy, max 0≤t≤T mN 2 α (Ω) ≤ C1 H 77 d where C1 only depends on initial data m0 2 α (Ω) . Moreover, for 1 ≤ r ≤ r∗ = d−α , where H d is dimension, ∂mN ≤ C2 r ∂t L (QT ) d and for 1 < r ≤ r∗ = d−α r−1 mN (t1 ) − mN (t2 ) Lr (Ω) ≤ C2 |t2 − t1 | r where C2 only depends on initial data m0 2 α (Ω) and time T . H Proof. Multiply systems (5.8) with ϕk,N (t), k = 1, 2, .., N and sum up together, then we have, d |m |2 dx = 0 dt Ω N So, mN L2 (Ω) ≤ m0 L2 (Ω) Multiply systems (5.8) with ϕk,N (t), k = 1, 2, .., N , where ϕk,N (t) satisfies, N FmN = ϕk,N (t)ωk (x) k=1 and sum up together, ∂mN ∂mN · ∧2α mN + · PN (HmN ) + |mN × FmN |2 = 0 ∂t ∂t Ω Ω Ω 78 (5.9) By using periodic boundary condition, ∂mN 1∂ | ∧α mN |2 + · PN (HmN ) + |mN × FmN |2 = 0 2 ∂t Ω ∂t Ω Ω For mapping operator PN , ∂mN ∂mN · PN (HmN ) = · HmN Ω ∂t Ω ∂t Also we have, Ω mN · HmN = mN · HmN = Ω |HmN |2 So, 1∂ 1∂ |mN × FmN |2 = 0 | ∧α mN |2 + |HmN |2 + 2 ∂t Ω 2 ∂t Ω Ω (5.10) Integrating (5.10) with respect to time from [0, T ], max 0≤t≤T ∧α mN 2 2 + max HmN 2 2 + mN × FmN 2 2 ≤ C1 (5.11) L (Ω) 0≤t≤T L (Ω) L (0,T ;L2 (Ω)) where C1 only depends on initial data ∧α m0 2 2 . By Sobolev embedding, for any L (Ω) 2d 1 ≤ p ≤ p∗ = d−2α , we have, max 0≤t≤T mN (t) Lp (Ω) ≤ k1 C1 d By Holder inequality, for 1 ≤ r ≤ r∗ = d−α , ( 2r 1 2−r |mN × FmN |r dx) r ≤ FmN L2 (Ω) ( |mN | 2−r dx) 2r ≤ k2 C1 Ω Ω 79 2r d since 2−r ≤ p∗ when r < r∗ . Moreover, we have for 1 ≤ r ≤ r∗ = d−α , ( ∂mN r |mN × FmN |r + |mN × (mN × FmN )|r | )≤ ∂t QT QT QT | T ≤ k2 C1 + 0 2r 2−r mN × FmN L2 (Ω) ( |mN | 2−r dx) 2r dt Ω (5.12) ≤ k3 C2 where C2 only depends on initial data and time T . With the same r > 1 as ∧α m0 2 2 L (Ω) defined above, we can further prove the approximation solution mN is continuous in time, t2 ∂mN dt Lr (Ω) ∂t t1 t2 ∂m N ≤ dt r ∂t L (Ω) t1 mN (t1 ) − mN (t2 ) Lr (Ω) ≤ ≤ r−1 |t2 − t1 | r ( (5.13) ∂mN r 1 | | )r ∂t QT r−1 ≤ C2 |t2 − t1 | r Based on Lemma 5.2.1, we have following conclusions for approximate solution mN , Lemma 5.2.2. For arbitrage N and time T , under the condition of Lemma 5.2.1, the initial value problem for approximation systems (5.8) has at least one time continuous and global weak solution. 5.3 Compactness In order to take N → ∞ with the approximation solutions, we still need strong convergence based on Lemma 5.2.1. The compactness lemma below can be found in [39], 80 Lemma 5.3.1. Let B0 , B, B1 be three Banach space such that, B0 ⊂ B ⊂ B1 where the injections are continuous and B0 , B1 are reflexive and B0 → B is compact. Denote W = {v|v ∈ Lp0 (0, T ; B0 ), dv ∈ Lp1 (0, T ; B1 )} dt for T < ∞ and 1 < p0 , p1 < ∞. Then W equipped with the norm dv v Lp0 (0,T ;B ) + p 0 dt L 1 (0,T ;B1 ) is a Banach space and the embedding W → Lp0 (0, T ; B) is compact. From Lemma 5.2.1, we know mN is uniformly bounded in L∞ (0, T ; H α (Ω)) W 1,r (0, T ; Lr (Ω)) With compactness lemma 5.3.1, there exists some m ∈ L∞ (0, T ; H α (Ω)) such that, mN m weakly in Lp (0, T ; H α (Ω)) for 1 < p < ∞ mN → m strongly in Lp (0, T ; H β (Ω)) for 1 < p < ∞, 0 ≤ β ≤ α ∂mN ∂t ∂m ∂t weakly in Lr (QT ) for 1 < r 81 (5.14) 5.4 Existence We are going to prove the global existence of weak solution to fractional LLG (5.1) by determining weak convergence limit of approximate solution mN when N → ∞. From compactness result(5.14), it is obvious that we have, ∂m ∂mN ·ϕ→ · ϕ , ∀ϕ ∈ C ∞ (QT ) QT ∂t QT ∂t For the nonlocal nonlinear term, we first observe that, mN × FmN 2 2 ≤ C1 L (0,T ;L2 (Ω)) which has been proved in (5.11). So there exists η ∈ L2 (0, T ; L2 (Ω)), such that, mN × FmN η weakly in L2 (0, T ; L2 (Ω)) We would like to show that η = m × Fm in L2 (0, T ; L2 (Ω)). Let ϕ ∈ H s (Ω) for s > α + d/2, QT mN × FmN · ϕ = QT mN × PN (HmN ) · ϕ + QT mN × ∧2α mN · ϕ = A1 + A2 From compactness result(5.14), A1 = QT mN × PN (HmN ) · ϕ → 82 m × Hm · ϕ QT (5.15) For term A2 , A2 = QT mN × ∧2α mN · ϕ (5.16) ∧α m =− QT α N · ∧ (mN × ϕ) In order to show that η = m × Fm in L2 (0, T ; L2 (Ω)), it is enough to prove QT ∧α mN · ∧α (mN × ϕ) ∧α m · ∧α (m × ϕ) weakly in L2 (0, T ; L2 (Ω)) QT We set commutator Φϕ (m) = [∧α , ϕ]m = ∧α (m × ϕ) − ∧α m × ϕ, then it is equivalent to prove QT ∧α mN · Φϕ (mN − m) + QT ∧α (mN − m) · Φϕ (m) → 0 From the compactness result (5.14), it is obvious, QT ∧α (mN − m) · Φϕ (m) → 0 From Lemma 5.1.1 and Lemma 5.1.2, Φϕ (mN − m) L2 (Ω) ≤ C( ϕ Lp1 mN − m α−1,p2 + ϕ H α,p3 mN − m Lp4 ) H ≤ C( ϕ Lp1 mN − m L2 (Ω) + ϕ H α,p3 mN − m H β (Ω) ) ≤ C ϕ H s (Ω) mN − m H β (Ω) (5.17) where p2 , p3 ∈ (1, ∞) 1 1 1 1 1 = + = + 2 p1 p2 p3 p 4 83 1−α 1 1 + = d p2 2 β 1 1 + = d p4 2 0<β<α d −1 2 s> Therefore, | QT ∧α mN · Φϕ (mN − m)| ≤ C ϕ H s (Ω) ∧α mN L2 (Ω ) mN − m L2 (0,T ;H β (Ω)) → 0 T So, ∀ϕ ∈ C ∞ (QT ) QT (mN × FmN ) · ϕ → (m × Fm ) · ϕ QT Furthermore, since mN → m strongly in L2 (QT ), QT (mN × FmN ) × mN · ϕ → (m × Fm ) × m · ϕ QT The theorem is summarized as below Theorem 5.4.1. Let 0 < α < 1 and m0 ∈ H α (Ω), then for any 0 < T < ∞, then there exists at least one global weak solution to fractional Landau-Lifshitz equation (5.1) such that m ∈ L∞ (0, T ; H α (Ω)) d for 1 < r ≤ r∗ = d−α . 84 r−1 C 0, r (0, T ; Lr (Ω)) BIBLIOGRAPHY 85 BIBLIOGRAPHY [1] F. Alouges and A. Soyeur, On global weak solutions for Landau-Lifshitz equations: Existence and Nonuniqueness, Nonlinear Analysis, TMA, 18(11), (1992), 1071–1084. [2] J.M. Ball, Convexity conditions and existence theorems in nonlinear elasticity, Arch. Rational Mech. Anal., 63(1977), 337-403. [3] M. Bertsch, P. Podio-Guidugli and V. Valente, On the dynamics of deformable ferromagnets. I. Global weak solutions for soft ferromagnets at rest, Ann. Mat. Pura Appl., 179(4) (2001), 331–360. [4] W. F. Brown, Micromagnetics, Interscience, New York, 1963. [5] A. Capella, C. Melcher and F. Otto, Effective dynamic in ferromagnetic thin films and the motion of Neel walls, Nonlinearity, 20(11) (2007), 2519-2537. [6] I. Cimrak and R. V. Keer, Higher order regularity results in 3D for the Landau-Lifshitz equation with an exchange field, Nonlinear Analysis, 68(2008), 1316-1331. [7] G. Carbou and P. Fabrie, Time average in micromagnetics, Journal of Differential Equations, 147 (1998), 383–409. [8] G. Carbou and P. Fabrie, Regular solutions for Landau-Lifshitz equation in a bounded domain, Differential Integral Equations, 14(2) (2001), 213–229. [9] R. Coifman and Y. Meyer, Nonlinear harmonic analysis, operator theory and P.D.E. In: Beijing Lectures in Harmonic Analysis, Princeton University Press(1986), 3-45. [10] N. H. Chang and K. Uhlenbeck, Schr¨dinger maps, Commun. Pure Appl. Math. o 53(5)(2000), 590-602. [11] B. Dacorogna, ”Direct Methods in the Calculus of Variations.”, Second Edition. Springer, New York, 2008. [12] B. Dacorogna and I. Fonseca, A-B quasiconvexity and implicit partial differential equations, Calc. Var., 14 (2) (2002), 115–149. 86 [13] W. Deng and B. Yan, Quasi-stationary limit and a degenerate Landau-Lifshitz equation of ferromagnetism, AMRX, to appear. [14] A. DeSimone, Energy minimizers for large ferromagnetic bodies, Arch Rational Mech. Anal., 125 (1993), 99–143. [15] A. DeSimone, R.V. Kohn, S. M¨ller and F. Otto, A reduced theory for thin-film microu magnetics, Comm. Pure Appl. Math. 55 (2002), 1-53. [16] A. DeSimone, R. V. Kohn, S. M¨ller and F. Otto, Recent analytical developments in u micromagnetics, The Science of Hysteresis ed G Bertotti and I Mayergoyz (Elsevier Academic Press), 2(2005), 269-381. [17] S. Ding and B. Guo, Initial-boundary value problem for higher-dimensional LandauLifshitz systems, Appl. Anal., 83 (2003), 673–697. [18] I. Fonseca and G. Leoni, Relaxation results in micromagnetics. Ricerche Mat. 49(2000), 269-304. [19] C.J. Garcia-Cervera and W. E, Effective dynamics of ferromagnetic thin films, J. Appl. Phys., 90(2001), 370-374. [20] P. G´rard, Microlocal defect measures, Comm. in PDE, 16(11)(1991), 1761–1794. e [21] G. Gioia and R. D. James, Micromagnetics of very thin films, R. Soc. Lond. Proc. Ser. A, 453(1997), 213-223. [22] B. Guo and M. Hong, The Landau-Lifshitz equation of the ferromagnetic spin chain and harmonic maps, Calc. Var., 1(1993), 311 – 334. [23] B. Guo and X. Pu, The fractional Landau-Lifshitz-Gilbert equation and the heat flow of harmonic maps, Cacl. Var., 42(2011), 1-19. [24] B. Guo and X. Pu, Regular solutions for multiplicative stochastic Landau-LifshitzGilbert equation and blow-up phenomena, Sci. China Math., 53(12)(2010), 3115-3130. [25] B. Guo, X. Pu and J. Zhang, Global weak solutions to the 1-D fractional Landau-Lifshitz equation, Discrete and Continuous Dynamical Systems Series, 14(1)(2010), 199-207. [26] B. Guo. and M. Zeng, Solutions for the fractional Landau-Lifshitz equation, J. Math. Anal. Appl., 361(2010), 131-138. 87 [27] S. Gustafson, K. Nakanishi and T. P. Tsai, Asymptotic stability, concentration, and oscillation in harmonic map heat-flow. Landau-Lifshitz, and Schrodinger maps on R2 , Commun. Math. Phys., 300(2010), 205–242. [28] R. D. James and D. Kinderlehrer, Frustration in ferromagnetic materials, Cont. Mech. Thermodyn., 2 (1990), 215–239. [29] F. Jochmann, Existence of solutions and a quasi-stationary limit for a hyperbolic system describing ferromagnetism, SIAM J. Math. Anal., 34(2) (2002), 315–340. [30] F. Jochmann, Aysmptotic behavior of the electromagnetic field for a micromagnetism equation without exchange energy, SIAM J. Math. Anal., 37(1) (2005), 276–290. [31] J. L. Joly, G. Metivier and J. Rauch, Global solutions to Maxwell equations in a ferromagnetic medium, Ann. Henri Poincar´, 1 (2000), 307–340. e [32] P. Joly, A. Komech and O. Vacus, On transitions to stationary states in a MaxwellLandau-Lifshitz-Gilbert system, SIAM J. Math. Anal., 31 (1999), 346–374. [33] M. Jurzke, C. Melcher and R. Moser, Domain walls and vortices in thin ferromagnetic films, Analysis, modeling and simulation fo multiscale problems, Springer, Berlin (2006), 249–298. [34] M. Kruz´ and A. Prohl, Recent developments in the modeling, analysis, and numerics ık of ferromagnetism, SIAM Review, 48(3) (2006), 439–483. [35] T. Kato, Liapunov, Functions and Monotonicity in the Navier-Stokes equations, Lecture Notes in Mathematics, 1450(1990), DOI: 10.1007/BFb0084898, 53-63. [36] T. Kato and G. Ponce, Commutator estimates and the Euler and Navier-Stokes equations. Commun. Pure Appl. Math., 41(7)(1988), 891-907. [37] L. D. Landau and E. M. Lifshitz, On the theory of the dispersion of magnetic permeability of ferromagnetic bodies, Phys. Z. Sowj., 8 (1935), 153–169. [38] L. Landau, E. Lifshitz and L. Pitaevskii, Electrodynamics of Continuous Media, Pergamon Press, New York, 1984. [39] J. L. Lions, Quelques methodes de resolution des problemes aux limits non lineeaires. Dunod, Paris, 1969. 88 [40] C.B. Morrey, Quasiconvexity and the lower semicontinuity of multiple intergrals, Pacific J. Math, 2(1952), 25-53. [41] C. Melcher, Thin-film limits for Landau-Lifshitz-Gilbert equations, SIAM J. Math. Anal., 42(1)(2010), 519-537. [42] P. Pedregal, Relaxation in ferromagnetism: the rigid case. J. Nonlinear Sci., 4(1994), 105-125. [43] P. Pedregal and B. Yan, On two-dimensional ferromagnetism, Proc. R. Soc. Edinburgh, 139A (2009), 575–594. [44] P. Pedregal and B. Yan, A duality method for micromagnetics, SIAM J. Math. Anal., 41(6) (2010), 2431–2452. [45] E. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton University Press, 1970. [46] L. Tartar, The compensated compactness method applied to systems of conservation laws, Systems of Nonlinear Partial Differential Equations, (J. M. Ball ed.), NATO ASI Series, Vol. CIII, D. Reidel, 1983. 263–285. [47] L. Tartar, H-measures, a new approach for studying homogenization, oscillations and concentrations effects in partial differential equations, Proc. Roy. Soc. Edinburgh, 115(A) (1990), 193–230. [48] L. Tartar, Beyond Young measures. Meccanica, 30(1995), 505-526. [49] A. Visintin, On Landau-Lifshitz equation for ferromagnetism, Japan J. Appl. Math., 2(1) (1985), 69–84. [50] B. Yan, Characterization of energy minimizers in micromagnetics, J. Math. Anal. Appl., 374(1) (2011), 230–243. [51] B. Yan, On the equilibrium set of magnetostatic energy by differential inclusion, Calc. Var., (2012), DOI: 10.1007/s00526-012-0527-y. [52] B. Yan, On stability and asymptotic behaviors for a degenerate Landau-Lifshitz equation, Preprint submitted 89