13.07.2015 Views

Semi–Lagrangian Approximation Schemes for Linear and Hamilton ...

Semi–Lagrangian Approximation Schemes for Linear and Hamilton ...

Semi–Lagrangian Approximation Schemes for Linear and Hamilton ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

✐✐Semi–Lagrangian <strong>Approximation</strong> <strong>Schemes</strong><strong>for</strong> <strong>Linear</strong> <strong>and</strong> <strong>Hamilton</strong>–Jacobi EquationsMaurizio Falcone <strong>and</strong> Roberto FerrettiJune 6, 2012


✐✐iiWARNINGThis is a draft NOT intended <strong>for</strong> free circulation. It contains the first six chapters ofthe book which will be published by SIAM. Three more chapters deal with control<strong>and</strong> game problems, front propagation <strong>and</strong> fluid dynamics.If you want to use this material, please contact the authors via email:Maurizio Falcone, falcone@mat.uniroma1.itRoberto Ferretti, ferretti@mat.uniroma3.itRome, February 2, 2011


✐✐ContentsPrefaceiii1 Models <strong>and</strong> motivations 11.1 <strong>Linear</strong> advection equation . . . . . . . . . . . . . . . . . . . . . 11.2 Nonlinear evolutive problems of <strong>Hamilton</strong>–Jacobi type . . . . . 31.3 Nonlinear stationary problems . . . . . . . . . . . . . . . . . . . 61.4 Simple examples of approximation schemes . . . . . . . . . . . . 111.5 Some difficulties arising in the analysis <strong>and</strong> in the approximation 141.6 How to use this book . . . . . . . . . . . . . . . . . . . . . . . . 152 Viscosity Solutions of First Order PDEs 172.1 The definition of viscosity solution . . . . . . . . . . . . . . . . . 172.1.1 Some properties of viscosity solutions . . . . . . . . 192.1.2 An alternative definition of viscosity solution . . . . 232.1.3 Uniqueness of viscosity solutions . . . . . . . . . . . 262.2 Viscosity solution <strong>for</strong> evolutive equations . . . . . . . . . . . . . 272.2.1 Representation <strong>for</strong>mulae <strong>and</strong> Legendre trans<strong>for</strong>m . . 282.2.2 Semiconcavity <strong>and</strong> regularity of viscosity solutions . 322.3 Problems in bounded domains . . . . . . . . . . . . . . . . . . . 352.3.1 Boundary Conditions in a weak sense . . . . . . . . 362.4 Viscosity solutions <strong>and</strong> entropy solutions . . . . . . . . . . . . . 382.5 Discontinuous viscosity solutions . . . . . . . . . . . . . . . . . . 402.6 Commented references . . . . . . . . . . . . . . . . . . . . . . . 413 Elementary building blocks 433.1 A review of ODE approximation schemes . . . . . . . . . . . . . 433.1.1 One-step schemes . . . . . . . . . . . . . . . . . . . 443.1.2 Multistep schemes . . . . . . . . . . . . . . . . . . . 453.2 Reconstruction techniques in one <strong>and</strong> multiple space dimensions 473.2.1 Symmetric Lagrange interpolation in R 1 . . . . . . . 483.2.2 Essentially Non-Oscillatory interpolation in R 1 . . . 523.2.3 Weighted Essentially Non-Oscillatory interpolationin R 1 . . . . . . . . . . . . . . . . . . . . . . . . . . 553.2.4 H<strong>and</strong>ling multiple dimensions by separation of variables. . . . . . . . . . . . . . . . . . . . . . . . . . . 633.2.5 Finite element interpolation . . . . . . . . . . . . . . 643.3 Function minimization . . . . . . . . . . . . . . . . . . . . . . . 693.3.1 Direct search methods . . . . . . . . . . . . . . . . . 70v


✐✐viContents3.3.2 Descent methods . . . . . . . . . . . . . . . . . . . . 713.3.3 Powell’s method <strong>and</strong> its modifications . . . . . . . . 723.3.4 Trust-region methods based on quadratic models . . 733.4 Numerical computation of the Legendre trans<strong>for</strong>m . . . . . . . . 733.5 Commented references . . . . . . . . . . . . . . . . . . . . . . . 754 Convergence theory 774.1 The general setting . . . . . . . . . . . . . . . . . . . . . . . . . 774.2 Convergence results <strong>for</strong> linear problems: the Lax–Richtmeyertheorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 794.2.1 Consistency . . . . . . . . . . . . . . . . . . . . . . . 794.2.2 Stability . . . . . . . . . . . . . . . . . . . . . . . . . 804.2.3 Convergence . . . . . . . . . . . . . . . . . . . . . . 814.2.4 Time-dependent evolution operators . . . . . . . . . 824.2.5 The stationary case . . . . . . . . . . . . . . . . . . 834.3 More on stability . . . . . . . . . . . . . . . . . . . . . . . . . . 854.3.1 The CFL condition . . . . . . . . . . . . . . . . . . 854.3.2 Monotonicity <strong>and</strong> L ∞ stability . . . . . . . . . . . . 864.3.3 Monotonicity <strong>and</strong> Lipschitz stability . . . . . . . . . 884.3.4 Von Neumann analysis <strong>and</strong> L 2 stability . . . . . . . 894.4 Convergence results <strong>for</strong> <strong>Hamilton</strong>–Jacobi equations . . . . . . . 914.4.1 Cr<strong>and</strong>all–Lions <strong>and</strong> Barles–Souganidis theorems . . 914.4.2 Lin–Tadmor theorem <strong>and</strong> semi–concave stability . . 984.5 Numerical diffusion <strong>and</strong> dispersion . . . . . . . . . . . . . . . . . 1004.6 Commented references . . . . . . . . . . . . . . . . . . . . . . . 1025 First-order approximation schemes 1035.1 Treating the advection equation . . . . . . . . . . . . . . . . . . 1035.1.1 Upwind discretization: the First Order Upwind scheme1045.1.2 Central discretization: the Lax–Friedrichs scheme . 1085.1.3 Semi–Lagrangian discretization: the Courant–Isaacson–Rees scheme . . . . . . . . . . . . . . . . . . . . . . 1135.1.4 Multiple space dimensions . . . . . . . . . . . . . . . 1235.1.5 Boundary conditions . . . . . . . . . . . . . . . . . . 1255.1.6 Examples . . . . . . . . . . . . . . . . . . . . . . . . 1265.2 Treating the convex HJ equation . . . . . . . . . . . . . . . . . . 1285.2.1 Upwind discretization . . . . . . . . . . . . . . . . . 1295.2.2 Central discretization . . . . . . . . . . . . . . . . . 1315.2.3 Semi–Lagrangian discretization . . . . . . . . . . . . 1335.2.4 Multiple space dimensions . . . . . . . . . . . . . . . 1385.2.5 Boundary conditions . . . . . . . . . . . . . . . . . . 1395.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1405.4 Stationary problems . . . . . . . . . . . . . . . . . . . . . . . . . 1415.4.1 The linear case . . . . . . . . . . . . . . . . . . . . . 1425.4.2 The nonlinear case . . . . . . . . . . . . . . . . . . . 1455.5 Commented references . . . . . . . . . . . . . . . . . . . . . . . 1466 High-order SL approximation schemes 1496.1 Semi-Lagrangian schemes <strong>for</strong> the advection equation . . . . . . 1496.1.1 Construction of the scheme . . . . . . . . . . . . . . 149


✐✐viiiContents8.3.2 Treatment of diffusive terms . . . . . . . . . . . . . 2498.4 Numerical Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . 2498.4.1 <strong>Linear</strong> advection . . . . . . . . . . . . . . . . . . . . 2498.4.2 Examples in R . . . . . . . . . . . . . . . . . . . . . 2498.4.3 Examples in R 2 . . . . . . . . . . . . . . . . . . . . . 2498.5 Commented references . . . . . . . . . . . . . . . . . . . . . . . 250Bibliography 251Index 265


✐✐ContentsixBasic Notationsa.e. almost everywhere (with respect to the Lebesgue measure)R d euclidean d-dimensional spacex · y scalar product ∑ di=1 x iy i of two vectors x, y ∈ R d|x| euclidean norm of x ∈ R d ,|x| = (x · x) 1/2B(x 0 , r) open ball of center x 0 <strong>and</strong> radius r in R d , {x : |x − x 0 | < r}B(x 0 , r) closed ball of center x 0 <strong>and</strong> radius r in R d , {x : |x − x 0 | ≤ r}∂E boundary of the set Eint E interior of the set EE closure of the set Eco E convex hull of the set Ea ∨ b min{a, b}, <strong>for</strong> a, b ∈ Ra ∧ b max{a, b}, <strong>for</strong> a, b ∈ RDu(x) gradient of a function u : R d → R at the point xd(x, C) distance from the point x to the closed set C∆t, ∆x discretization steps,⎧in compact <strong>for</strong>m ∆ ≡ (∆x, ∆t)⎪⎨ +1 if x > 0sgn(x) sign of x: sgn(x) := 0 if x = 0⎪⎩−1 if x < 0i imaginary unit{1 if x ∈ Ω1 Ω (x) characteristic function of the set Ω: 1 Ω (x) :=0 if x ∉ Ω{1 if i = jδ ij Kronecker’s symbol: δ ij :=0 if i ≠ j⌊x⌋ lower integer part of x: ⌊x⌋ := max{m ∈ Z : m ≤ x}I k [G](x) value at the point x of the interpolate of degree k of the function gbased on the vector G = (g(x i )) of grid values of gν(x) exterior normal to a surface at the point xy(x, t; s) the position at time s of the trajectory starting at x at time ty x,t (s; α) the position at time s of the trajectory starting at x at time t<strong>and</strong> driven by the control αy x (s; α) the position at time s of the trajectory starting at x at time t = 0<strong>and</strong> driven by the control αT V (w) total variation of the univariate function w: T V (w) = ∫ D |w′ (x)| dxω(δ) a modulus of continuityBUC(Ω) space of bounded uni<strong>for</strong>mly continuous functions on the open set ΩC k (Ω) space of functions u : Ω → R with continuousk-th derivative on a domain ΩC0 k (Ω) space of functions with compact support belonging to C k (Ω)L p (Ω) space of functions u : Ω → R such that ∫ Ω up (x) dx < +∞, 1 ≤ p < +∞L ∞ (Ω) space of functions u : Ω → R such that ess sup|u(x)| < +∞L p loc (Ω) space of functions u ∈ Lp (K) <strong>for</strong> any compact set K ⊂ Ω, 1 ≤ p ≤ +∞‖u‖ p norm of u in L p (Ω), 1 ≤ p ≤ ∞W m,p (Ω) Sobolev space of functions u : Ω → R with the first m derivatives in L p (Ω)W k,ploc (Ω) Sobolev space of functions u with the first m derivatives in Lp (K)<strong>for</strong> any compact set K ∈ Ω


✐✐xContentsspace of p-th power summable vectors or sequencesspace of bounded vectors or sequencesV , v j numerical solution, as a vector <strong>and</strong> as a value at the node x jV n , vj n numerical solution, as a vector <strong>and</strong> as a value at the node (x j , t n )V (k) , v (k)j numerical solution at the iteration k, as a vector <strong>and</strong> as a valueat the node x j‖V ‖ p norm in l⎧p , renormalized if referred to a numerical solution:⎨∑ ) 1/α(∆x‖V ‖ α := 1 · · · ∆x d j |v j| α if α < ∞⎩max j |v j | if α = ∞.B t transpose of a matrix Bρ(B) spectral radius of a square matrix B ∈ R m×m‖B‖ p <strong>for</strong> a (possibly infinite) matrix B, natural operator norm in l p :‖BV ‖ p‖B‖ p := supV ‖V ‖ pl pl ∞‖B‖ ∞ ∞-norm of a matrix B: ‖B‖ ∞ = max i∑j |b ij|‖B‖ 1 1-norm of a matrix B: ‖B‖ 1 = max j∑i |b ij|‖B‖ 2 2-norm of a matrix B: ‖B‖ 2 = √ ρ(B t B)w, w lower <strong>and</strong> upper semicontinuous envelopes:{w(x) := lim inf y→x w(y)w(x) := lim sup y→x w(y)


✐✐Chapter 1Models <strong>and</strong> motivationsThis book is devoted to the study of some class of hyperbolic Partial DifferentialEquations, <strong>and</strong> to their approximation of Semi-Lagrangian type. Throughout thebook, we will try to keep a mathematically rigorous analysis as much as possible;however, be<strong>for</strong>e getting into the most technical details <strong>and</strong> results, we would liketo start by presenting some examples <strong>and</strong> applications which have motivated muchof the research in this area. In fact, the analysis <strong>and</strong> approximation of first orderPartial Differential Equations of <strong>Hamilton</strong>–Jacobi type has been driven by theirapplication to fields such as fluid dynamics, optimal control <strong>and</strong> differential games,image processing <strong>and</strong> material science just to mention the main ones.In all the a<strong>for</strong>ementioned fields, the notion of weak solutions has had a crucialimpact in giving a sound theoretical framework <strong>for</strong> both the analytical <strong>and</strong> thenumerical study. The goal of this chapter is precisely to introduce the reader tothis theory via some model problems. We will start from the linear advectionequation, then turn to nonlinear models, either the evolutive ones arising in frontpropagation <strong>and</strong> optimal control, or the stationary ones including, e.g. the eikonalequation of the classical Shape-from-Shading problem in image processing. We willalso discuss some difficulties related to the numerical approximation of such models,<strong>and</strong> conclude the chapter by giving some hints on how to use the material presentedin the book.1.1 <strong>Linear</strong> advection equationAs a first model, we start by considering the advection equation:u t + f(x, t) · Du = g(x, t) (x, t) ∈ R d × (t 0 , T ). (1.1)Here, f : R d × (t 0 , T ) → R d is a vectorfield termed as drift <strong>and</strong> g : R d × (t 0 , T ) → Ris the source term. We look <strong>for</strong> a solution u : R d × (t 0 , T ) → R of (1.1) satisfyingthe initial conditionu(x, t 0 ) = u 0 (x) x ∈ R d . (1.2)This is a classical model describing the transport of a scalar field, e.g. the distributionof a pollutant emitted by one of more sources (represented by g), <strong>and</strong>transported by a water stream or a wind (represented by f).This physical interpretation is better clarified through the method of characteristicswhich provides an alternative characterization of the solution. In the1


✐✐2 Chapter 1. Models <strong>and</strong> motivationssimplest case, <strong>for</strong> g(x, t) ≡ 0 <strong>and</strong> f(x, t) ≡ c (constant) the solution v is given bythe representation <strong>for</strong>mulau(x, t) = u 0 (x − c(t − t 0 )) (x, t) ∈ R d × [t 0 , T ). (1.3)In fact, assume that a regular solution u exists. Deriving u with respect to t alonga curve of the <strong>for</strong>m (y(t), t) we obtaindudt (y(t), t) = Du(y(t), t) · ẏ(t) + u t(y(t), t), (1.4)so that (using the fact that u solves the advection equation) the total derivativewill identically vanish along the curves which satisfyẏ(t) = c. (1.5)Such curves have the physical meaning of flow lines along which the scalar field uis transported. They are known as characteristics <strong>and</strong>, in this particular case, arestraight lines. Then, to assign a value to the solution at a point (x, t) it sufficesto follow the unique line passing through (x, t) until it crosses the x-axis at thepoint z = x − c(t − t 0 ), z being the foot of the characteristic. Since the solutionu is constant along this line, we get (1.3). Conversely, it is easy to check that afunction of x <strong>and</strong> t in the <strong>for</strong>m (1.3) satisfies (1.1) <strong>for</strong> f ≡ c <strong>and</strong> g ≡ 0, provided u 0is differentiable.In the case of variable coefficients f <strong>and</strong> g in (1.1), we have the more general<strong>for</strong>m∫ tu(x, t) = u 0 (y(x, t; t 0 )) + g(y(x, t; s), s)ds, (1.6)t 0where y(x, t; s) is the (backward) solution at time s of the Cauchy problem, calledsystem of base characteristics{ẏ(x, t; s) = f(y(x, t; s), s)(1.7)y(x, t; t) = x.In this more general <strong>for</strong>mula, which will be proved in Chapter 2, the value u(x, t)depends on the value at the foot of the characteristic curve (which might no longerbe a straight line), as well as on the integral of the source term along the characteristicitself. Again, it can be easily shown that a function defined by (1.6)–(1.7)solves (1.1)–(1.2) if u 0 is smooth.The representation <strong>for</strong>mula (1.6)–(1.7) shows that the regularity of the initial conditionis preserved along the characteristics if both drift <strong>and</strong> source terms aresmooth enough, so that u 0 ∈ C p (R d ) <strong>and</strong> f, g ∈ C p−1 (R d × (t 0 , T )) imply thatu ∈ C p (R d × (t 0 , T )). A key point in this respect is that characteristics cannotcross, a property which only holds in the linear case.A typical result of existence <strong>and</strong> uniqueness of classical solutions to the advectionequation in one space dimension is the following.Theorem 1.1. Consider equation (1.1)–(1.2) <strong>and</strong> assume that f, g ∈ C 1 (R ×(t 0 , T )) <strong>and</strong> that u 0 ∈ C 1 (R×(t 0 , T )). Then, there exists a unique classical solutionof (1.1), i.e. a function u satisfying (1.1) at every point (x, t) ∈ R × (t 0 , T ).


✐✐1.2. Nonlinear evolutive problems of <strong>Hamilton</strong>–Jacobi type 3The proof of this result is based on the method of characteristics, which willbe better examined in Chapter 2 (a more general result in dimension d can befound in [E98]). A crucial point, beside the regularity of the coefficients <strong>and</strong> of theinitial condition, is that the field of characteristics (1, f(x, t)) should be transversalto the x-axis at t 0 so that the solution can be defined by the initial condition ona neighborhood of the x-axis. In practice, this correspond to the request that fshould not vanish over an interval.At a closer look, however, well-posedness results requiring differentiability ofthe initial condition might not be satisfactory in a number of applications whichnaturally lead to a nonsmooth definition of u 0 . For example, if we try to simulatethe advection of a pollutant spread over a precise area ω p , it would be natural toconsider a discontinuous initial condition such as{1 if x ∈ ω pu 0 (x) = 1 ωp =(1.8)0 else.Although we expect the physically relevant solution to be given again by (1.6)–(1.7), the lack of regularity in the initial condition does not enable the same <strong>for</strong>malderivation of the representation <strong>for</strong>mula. Hence, it becomes necessary to introducea notion of weak solution which could allow to treat the Cauchy problem in a moregeneral setting. One way to achieve this goal is to look <strong>for</strong> weak solutions in thesense of distributions:Definition 1.2. A function u : R × (t 0 , T ) → R in L 1 (R × (t 0 , T )) is a weaksolution of the Cauchy problem (1.1)–(1.2) if <strong>and</strong> only if <strong>for</strong> any test function φ ∈C0(R 1 × (t 0 , T )) the following integral condition is satisfied:∫ Tt 0∫Ru(x, t)φ t (x, t) + [u(x, t)f(x, t)] φ x (x, t)dx dt = (1.9)=∫ Tt 0∫Rg(x, t)ϕ(x, t) dx dtIn recent years, this strategy of definition <strong>for</strong> weak solutions has also beenextended to include bounded variation solutions. Anyway, we will not pursue thisline of work here, <strong>and</strong> rather focus on the concept of viscosity solution, which hasproved to be well suited <strong>for</strong> the case of <strong>Hamilton</strong>–Jacobi equations.1.2 Nonlinear evolutive problems of <strong>Hamilton</strong>–JacobitypeAmong all extensions to nonlinear equations, this book will mostly address itself tothe following <strong>Hamilton</strong>–Jacobi equation:{u t + H(x, Du) = 0 (x, t) ∈ R d × (0, T )v(x, 0) = v 0 x ∈ R d (1.10),where H : R d → R is called the <strong>Hamilton</strong>ian of the equation <strong>and</strong> is typicallyassumed to be convex. We soon sketch a couple of applications which fall in thisclass of problems.


✐✐4 Chapter 1. Models <strong>and</strong> motivationsFigure 1.1. A representation function <strong>for</strong> Γ 0Front propagation via the level set method A special case of the <strong>Hamilton</strong>–Jacobiequation arise in modeling a front which evolves in the normal direction with a givenvelocity c : R d → R. The level set method considers the Cauchy problem{u t + c(x)|Du| = 0 (x, t) ∈ R d × (0, T )u(x, 0) = u 0 , x ∈ R d (1.11)where u 0 : R d → R must be a proper representation of the front Γ 0 , i.e. a continuousfunction u 0 which changes sign on Γ 0 := ∂Ω. In two space dimensions, this wouldlead to the definition:Definition 1.3. Assume that Γ 0 is a closed curve in R 2 <strong>and</strong> denote by Ω 0 the regionenclosed by Γ 0 . A continuous function u 0 : R 2 → R is a proper representation ofΓ 0 if <strong>and</strong> only if it satisfies the following conditions:⎧⎪⎨ u 0 (x) < 0 x ∈ Ω 0u 0 (x) = 0 x ∈ Γ 0(1.12)⎪⎩u 0 (x) > 0 x ∈ R 2 \ Ω 0In this way, the front Γ 0 at the initial time is identified with the zero-level setof u 0 . Assume now that we let Γ 0 evolve in the normal direction with velocity c(x).To fix ideas, we can assume that the normal η is directed outwards with respect toΩ <strong>and</strong> that the velocity is positive (this situation corresponds to an expansion ofthe front, whereas an opposite sign would produce a shrinking). Moreover, let usconsider the evolution of a curve in the (x, y)-plane to simplify the presention.The front at time t will be denoted by Γ t <strong>and</strong> a generic point P ∈ Γ t will be representedas P = (x 1 (s, t), x 2 (s, t)), where s refers to the parametric representation ofΓ t (e.g. as an arc length parametrization). Since the velocity V at every point hasmagnitude c(P ) <strong>and</strong> is directed in the normal direction η, we can writeV (P ) = (ẋ 1 (s, t), ẋ 2 (s, t)) = c(P )η(P ) (1.13)


✐✐1.2. Nonlinear evolutive problems of <strong>Hamilton</strong>–Jacobi type 5Figure 1.2. The propagation of a front in the normal directionAfter describing Γ 0 as the zero-level set of u 0 , our plan is to describe the front Γ tat a generic time t as the zero-level set of the solution of a Cauchy problem, havingu 0 as its initial condition. In order to derive the equation characterizing Γ t , weconsider the total derivative of u(x 1 (s, t), x 2 (s, t)). Taking into account that thisderivative vanishes on a level set of constant value, we have:0 = d dt u(x 1(s, t), x 2 (s, t), t) == u t (x 1 (s, t), x 2 (s, t)) + Du(x 1 (s, t), x 2 (s, t)) · (ẋ 1 , ẋ 2 ) == u t (P ) + c(P )Du(P ) · η(P ) == u t (P ) + c(P )|Du(P )|where the normal direction to a level set of u (computed at P ) has been writtenin terms of u as Du(P )/|Du(P )|. In conclusion, solving (1.11) <strong>and</strong> looking at thezero-level set of the solution u(x, t) we can recover the front Γ t . Note that the choiceof the zero level is arbitrary, <strong>and</strong> in fact all level sets of u are moving according tothe same law.Once <strong>for</strong>mulated the front evolution by means of (1.11), we can drop thesmoothness assumption on Γ t , <strong>and</strong> apply the same method to more general initialconfigurations. Here, the possibility of defining weak (nondifferentiable) solutions<strong>for</strong> (1.11) plays a crucial role. For example, taking Γ 0 as the union of a numberof separate curves in the plane, if the evolution is driven by a positive velocitythe various components of the front will collide in a finite time <strong>and</strong> we will have amerging of the fronts. It is clear that, at a point where two components of the frontare tangent, we cannot expect to have a normal vector <strong>and</strong> a singularity appears inthe evolution (<strong>and</strong> in the equation). However, as we will see later on, the level setmethod relies on a weak concept of solution which is robust enough to h<strong>and</strong>le boththe onset of a singularity <strong>and</strong> the change of topology.This quick explanation of the level set method also shows its main drawback.In order to follow a (d − 1)-dimensional front we are <strong>for</strong>ced to embed the problemin R d , <strong>and</strong> this implies an extra computational cost. This remark has motivated


✐✐6 Chapter 1. Models <strong>and</strong> motivationsthe development of “local” versions of the level set method which reduce the computationalcost by taking into account only the values in a small neighbourhood ofthe front, the so-called narrow b<strong>and</strong>. More details on such methods will be given inChapter ??.Optimal control <strong>and</strong> dynamic programming A second relevant application of the<strong>Hamilton</strong>–Jacobi equation is to characterize the value function of a finite horizonoptimal control problem. Let a controlled dynamical system be described by{ẏ(t) = b(y(t), t, α(t))(1.14)y(t 0 ) = x 0where y ∈ R d is the state of the system, α : (t 0 , T ) → A ⊆ R m is the controlfunction, b : R d × (t 0 , T ) × A → R d is the controlled drift (or dynamics).Let s consider the class of admissible controls A := {α : (t 0 , T ) → A, measurable}where A is a compact subset of R m since it is known that <strong>for</strong> any given measurablecontrol α(·) the solution of (1.14) is well defined (in a weak sense). Then we canintroduce the cost functional J : A → R which will be used to select the “optimal”(i.e., the minimizing) trajectory. For the finite horizon problem, the functional isusually defined asJ(x, t 0 ; α(·)) =∫ Tt 0g(y(s), s, α(s))e −λ(s−t0) ds + e −λ(T −t0) g(y(T )) λ > 0(1.15)where g : R d × (t 0 , T ) × A → R is the running cost depending on the state of thesystem, on the time <strong>and</strong> on the control, λ ∈ R + is the discount factor which allowsto compare costs at different times by rescaling them at time t 0 <strong>and</strong> g is a terminalcost. The value function associated to this problem is defined asu(x, t) = inf J(x, t; α(·)) (1.16)α(·)The value function represents the best value we can get from our cost functional <strong>and</strong>,as it will be shown in Chapter 7, enables to construct the optimal solution. In allclassical deterministic control problems, the Dynamic Programming principle allowsto characterize the value function as the unique solution, once more in a suitablyweak sense, of a partial differential equation known as the Dynamic Programming(or Bellman) equation:⎧⎨u t + sup[−b(x, t, a) · Du − g(x, t, a)] = 0 (x, t) ∈ R d × (0, T )a∈A(1.17)⎩u(x, T ) = g(x) x ∈ R d ,where a ∈ A is a parameter spanning the control set. Note that, unless <strong>for</strong> being abackward equation, (1.17) is in the <strong>Hamilton</strong>–Jacobi <strong>for</strong>m (1.10). More details onthis technique will be given in Chapter 7, whereas a more extensive presentationcan be found in [BCD97].1.3 Nonlinear stationary problemsWe turn now to some examples of stationary <strong>Hamilton</strong>–Jacobi equations which willbe analyzed in the sequel. Problems of this kind may arise when dealing withapplications related to control theory <strong>and</strong> image processing.


✐✐8 Chapter 1. Models <strong>and</strong> motivationsFigure 1.3. A case in which d(·, Ω) is not differentiablewhere α : (0, +∞) → B(0, 1) is a measurable control. Given a closed convex targetset Ω, we want to find the minimal time necessary to drive the system from itsinitial position to the target (the minimal time represents the value function ofthis particular problem). Since the system can move in every direction, the optimaltrajectory is a straight line connecting the initial position x 0 to its unique projectionon Ω. Moreover, the velocity has at most unit norm, <strong>and</strong> there<strong>for</strong>e the minimumtime of arrival from x 0 equals the distance d(x 0 , Ω) <strong>and</strong> solves the eikonal equation(1.19).More in general (as we will see in Chapter 7 devoted to control problems <strong>and</strong>games), if the system is driven by a nonlinear controlled dynamical system of the<strong>for</strong>m {ẏ(t) = b(y(t), α(t))y(0) = x 0 ∈ R d (1.25)\ Ω,we end up with the following Bellman equation <strong>for</strong> the minimum time function:{max [−b(x, a) · Du] = 1 xa∈Au(x) = 0∈ Rd \ Ωx ∈ ∂Ω.(1.26)The characterization of the value function in terms of a stationary Bellmanequation also applies to other deterministic optimal control problems. For example,we mention that an equation of the <strong>for</strong>mλu(x) + max[−b(x, a) · Du(x) − g(x, a)] = 0 (1.27)a∈Acharacterizes the value function of an infinite horizon problem <strong>for</strong> the system (1.25),i.e. the functionu(x) := inf J(x, α(·)), (1.28)α∈Awith a cost functional J(x, α(·)) given byJ(x, α) :=∫ +∞0g(y(s), α(s))e −λs ds. (1.29)


✐✐1.3. Nonlinear stationary problems 9Figure 1.4. Optimal trajectories of the minimum time problem <strong>for</strong> a squaretarget Ω.Figure 1.5. A gray-level picture (left) <strong>and</strong> the corresponding surface (right).We refer again to Chapter 7 <strong>for</strong> a more detailed presentation of Dynamic Programmingtechniques.Shape-from-Shading An equation of eikonal type also arises in the Shape-from-Shading problem, a classical inverse problem in which one tries to reconstruct, usingthe shading in<strong>for</strong>mation, the surface which has generated a certain monochromaticpicture taken by a camera.To be more precise, we introduce the basic assumptions <strong>and</strong> notations <strong>for</strong> thisproblem. We attach to the camera a three-dimensional coordinate system (Oxyz),such that Oxy coincides with the image plane <strong>and</strong> Oz coincides with the optical


✐✐10 Chapter 1. Models <strong>and</strong> motivationsaxis. Under the assumption of orthographic projection, the visible part of the sceneis, up to a scale factor, a graph z = u(x), where x = (x 1 , x 2 ) is an image point.Horn has shown in [HB89] that the Shape-from-Shading problem can be modeledby the image irradiance equation:R(ν(x)) = I(x), (1.30)where I(x) is the grey level measured in the image at point x (in fact, I(x) is theirradiance at point x, but the two quantities are proportional) <strong>and</strong> R(ν(x)) is thereflectance function, giving the value of the light re-emitted by the surface as afunction of its orientation, i.e., of the unit normal ν(x) to the surface at a point(x, u(x)). For the graph of a function u, this normal can be expressed as:ν(x) =1√ (−p(x), −q(x), 1), (1.31)1 + p(x)2 + q(x)2where p = ∂u/∂x 1 <strong>and</strong> q = ∂u/∂x 2 , so that Du(x) = (p(x), q(x)). The irradiancefunction (or image brightness) I is given as a measure of grey level (from 0 to 255in digital images) at each pixel of the image, <strong>and</strong> in order to construct a continuousmodel, we will assume that it is renormalized in the interval [0, 1]. We also assumethat there is a unique light source at infinity, whose direction (supposed to beknown) is given by the unit vector ω = (ω 1 , ω 2 , ω 3 ) ∈ R 3 .The height function u, which is the unknown of the problem, has to be reconstructedon a compact domain Ω ⊂ R 2 , called the reconstruction domain. Recallingthat, <strong>for</strong> a Lambertian surface of uni<strong>for</strong>m unit albedo, R(ν(x)) = ω ·ν(x), <strong>and</strong> using(1.31), then (1.30) can be rewritten <strong>for</strong> any x ∈ Ω asI(x) √ 1 + |Du(x)| 2 + (ω 1 , ω 2 ) · Du(x) − ω 3 = 0, (1.32)which is a stationary first-order equation of <strong>Hamilton</strong>–Jacobi type. Points x ∈ Ωsuch that I(x) is maximal correspond to the particular situation where ω <strong>and</strong> ν(x)point in the same direction: these points are usually called “singular points”. Themost widely studied case in the Shape-from-Shading literature corresponds to afrontal light source at infinity, i.e., ω = (0, 0, 1). Then, (1.32) becomes an eikonalequation of the <strong>for</strong>m:√1 − I|Du(x)| =2 (x)I 2 (1.33)(x)It is known that the eikonal equation may not have a unique solution wheneverthe right-h<strong>and</strong> side vanishes at some point. So the existence of singular pointscorresponds to an intrinsic ill-posedness of the Shape-from-Shading problem. Infact, a simple example can show that uniqueness does not hold even <strong>for</strong> classicalsolutions. Consider a one-dimensional case in which the surface is given by u 1 (x) =1 − x 2 in the interval [−1, 1]. This surface clearly satisfies the boundary conditionsu 1 (1) = u 1 (−1) = 0 <strong>and</strong>, assuming a vertical light source with ω = (0, 1) it alsosatisfies (1.33) <strong>for</strong> the corresponding irradiance, which is given byI(x) =(0, 1) · (2x, 1)√1 + 4x2(1.34)At the point x = 0, I takes its maximum value I(0) = 1, <strong>and</strong> there<strong>for</strong>e the righth<strong>and</strong>side of (1.33) vanishes. Note that the Lambertian assumption on the surface


✐✐1.4. Simple examples of approximation schemes 11Figure 1.6. One classical solution <strong>and</strong> some a.e. solutions of the Shapefrom-Shadingproblem.implies that I only depends on the angle between the normal ν(x) <strong>and</strong> ω, so thatit is clear that the mirror image of the surface u 1 with respect to the x-axis, i.e.u 2 (x) = −1 + x 2 is still a classical solution of (1.33). It is natural to observe (seeFigure 1.6) that all mirror reflections of a part of both surfaces u 1 <strong>and</strong> u 2 withrespect to a horizontal line satisfy (1.33) almost everywhere (although this does notimply that they could be acceptable weak solutions).1.4 Simple examples of approximation schemesIn this section, we try to give some very basic ideas (in a single space dimension)about the approximation schemes <strong>for</strong> the model problems introduced so far. Westart with the case of the rigid transport problem, i.e. the linear advection atconstant speed c (to fix ideas, we assume that c > 0), which is described by theequation {u t + cu x = 0 (x, t) ∈ R × (0, T )(1.35)u(x, 0) = u 0 (x) x ∈ R.The most classical method to construct an approximation of (1.35) is to build auni<strong>for</strong>m grid in space <strong>and</strong> time (a lattice) with constant steps ∆x <strong>and</strong> ∆t, coveringthe domain of the solution:{(x j , t n ) = (j∆x, n∆t), j ∈ Z, n ∈ N, n ≤ T }. (1.36)∆tThe basic idea behind all finite difference approximation is to replace every derivativeby an incremental ratio. Thus, one obtains a finite dimensional problem whoseunknown are the values of the numerical solution at all the nodes of the lattice,so that the value u n j associated to the node (x j, t n ) should be regarded as an approximationof u(x j , t n ). For the time derivative it is natural to choose the <strong>for</strong>wardincremental ratiou(x, t + ∆t) − u(x, t)u t (x, t) ≈ (1.37)∆twhich allows, starting with the solution at the initial time t 0 , to compute explicitlythe approximations <strong>for</strong> increasing times t k > t 0 . Writing at (x j , t n ) the <strong>for</strong>wardincremental ratio in time, we getu t (x j , t n ) ≈ un+1 j − u n j∆t(1.38)


✐✐12 Chapter 1. Models <strong>and</strong> motivationsFigure 1.7. <strong>Approximation</strong> of the rigid transport problem by the upwindscheme (left, ∆t = 0.025) <strong>and</strong> by the Semi-Lagrangian scheme (right, ∆t = 0.05).For the approximation of u x we have more choices, likeu x (x j , t n ) ≈ un j+1 − un j∆xu x (x j , t n ) ≈ un j − un j−1∆xu x (x j , t n ) ≈ un j+1 − un j−12∆x(right finite difference)(left finite difference)(centered finite difference),(1.39)which are all based on the values at the nodes (x k , t n ).In this case, the choice of the approximation <strong>for</strong> u x crucially affects the convergenceof the numerical solution to the exact solution. Although centered finite differencewould in principle provide a more accurate approximation of the space derivative,as we will see later on, given the time derivative approximation (1.38), the onlychoice in (1.39) which results in a convergent scheme is to use the left incrementalratio if c > 0, the right incremental ratio if c < 0, which corresponds to the so-calledupwind scheme. In conclusion, <strong>for</strong> c > 0, we obtain a scheme in the <strong>for</strong>mu n+1j− u n j∆t+ c un j − un j−1∆x= 0. (1.40)A different way to construct the approximation of (1.35) is to consider theadvection term as a directional derivative <strong>and</strong> writecu x (x j , t n ) ≈ − un (x j − cδ) − u n jδ(1.41)where δ is a “small” positive parameter, <strong>and</strong> u n denotes an extension of the numericalsolution (at time t n ) to be computed outside of the grid. Coupling the <strong>for</strong>wardfinite difference in time with this approximation we getu n+1j− u n j∆t− un (x j − cδ) − u n jδ<strong>and</strong> finally, choosing δ = ∆t, we obtain the scheme= 0,u n+1j = u n (x j − c∆t) (1.42)


✐✐14 Chapter 1. Models <strong>and</strong> motivations1.5 Some difficulties arising in the analysis <strong>and</strong> in theapproximationAfter giving models, applications <strong>and</strong> motivations <strong>for</strong> the theory which will be theobject of this book, it is worth to point out a number of mathematical peculiaritiesof the problems under consideration.1. The analysis of linear transport problems <strong>and</strong> <strong>Hamilton</strong>–Jacobi equation leadsto drop classical solutions <strong>and</strong> introduce weak solutions in order to treat relevantapplications. The analysis of weak solutions in this framework hasstarted in the 60s with the pioneering works by S.N. Kružkov <strong>and</strong> has proceededthrough the last decades of the century when M.G. Cr<strong>and</strong>all <strong>and</strong> P.L.Lions introduced the notion of viscosity solution. Typically, viscosity solutionsare Lipschitz continuous but, at least <strong>for</strong> convex <strong>Hamilton</strong>ians, the conceptcan be extended to discontinuous solutions. Although the theory is still undergoingtechnical developments, <strong>and</strong> new classes of problems are analyzedin this framework, this book will mainly refer to the basic theory <strong>for</strong> convex<strong>Hamilton</strong>ians, which is at the moment quite an established matter.2. The lack of smoothness of viscosity solutions makes it difficult to developefficient approximations. Starting from the 80s, monotone finite differencemethods have been proposed using the relationship between <strong>Hamilton</strong>–Jacobiequations <strong>and</strong> conservation laws (this relationship will be discussed in thenext chapter). On this basis, monotone finite difference methods conceived <strong>for</strong>conservation laws have been adapted to the approximation of <strong>Hamilton</strong>–Jacobiequation. However, monotone schemes are typically very diffusive, a seriousdrawback when treating nonsmooth problems. On the other h<strong>and</strong>, Semi-Lagrangian schemes may provide less diffusive, still monotone approximationsthrough the use of large time steps. Construction <strong>and</strong> convergence theory ofSemi-Lagrangian schemes, as well as their comparison with difference schemes,will be a major subject of this book.3. In the last decades, the low efficiency of monotone scheme has motivated alarge ef<strong>for</strong>t to develop less diffusive, high-order approximation schemes. Thisongoing research, which can have a great impact on applications, has alreadyproduced a number of methods of great accuracy, although very few theoreticalresults are available. In this book, we will try to synthesize the existing results<strong>for</strong> high-order Semi-Lagrangian schemes into a rigorous convergence analysis,at least on model cases.4. The theory of viscosity solutions has been tested on a variety of applications,ranging from control <strong>and</strong> games theory to image processing, from fluid dynamicsto combustion <strong>and</strong> geophysics, <strong>and</strong> has proved to be particularly effectivein characterizing the relevant solution. Some chapters of this book will bedevoted to these topics. However, these applications often require to work inhigh dimension. Along with their theoretical study, the development of fast<strong>and</strong> accurate numerical schemes <strong>for</strong> high-dimensional problems seems to be acritical, challenging task <strong>for</strong> future research.


✐✐1.6. How to use this book 151.6 How to use this bookWe conclude this introductory chapter by giving, on the basis of our personal experience,some hints about the use of the material presented in the book. Such hintscan also be useful <strong>for</strong> a reader approaching the subject <strong>for</strong> the first time.• A basic introductory course on the analysis <strong>and</strong> approximation of <strong>Hamilton</strong>–Jacobi equation can use the first three chapters if you want to emphasize themathematical aspects.• An advanced numerical course on nonlinear hyperbolic differential equationswill include, <strong>for</strong> example, Chapters 2, 4, 3, 6.• A course on applications to front propagation, level set methods <strong>and</strong> fluiddynamics can use the material in Chapters 2, 3, 4, <strong>and</strong> 8 or 9 depending onthe emphasis to be given to different topics.• A course on applications to control theory can use the material in Chapters2, 3, 4 <strong>and</strong> 7.We will be glad to receive other suggestions <strong>for</strong> the use of this book <strong>for</strong> graduate<strong>and</strong> advanced undergraduate courses.


✐✐16 Chapter 1. Models <strong>and</strong> motivations


✐✐Chapter 2Viscosity Solutions ofFirst-Order PDEsThis chapter contains a general presentation of the theory of weak solutions in theviscosity sense <strong>for</strong> first-order partial differential equations, starting from the stationarycase (which include, e.g., the eikonal equation), <strong>and</strong> moving to evolutivemodels (which arise in control theory <strong>and</strong> front propagation).The chapter reviews basic <strong>and</strong> essential results of existence <strong>and</strong> uniqueness <strong>for</strong>viscosity solutions. Moreover, it also presents other related topics such as characteristics<strong>and</strong> Hopf–Lax <strong>for</strong>mulae, regularity results (typically, based on the notion ofsemiconcavity), boundary conditions, <strong>and</strong> the relationship with entropic solutionsof conservation laws. Most of these topics will prove useful in the construction <strong>and</strong>analysis of the approximation schemes presented in the sequel. Finally, we brieflysketch the concept of discontinuous viscosity solutions.2.1 The definition of viscosity solutionTo introduce the theory of weak solutions in the viscosity sense, we start by consideringa stationary problem of the general <strong>for</strong>mH(x, u, Du) = 0, x ∈ Ω. (2.1)Here, Ω is an open domain of R d <strong>and</strong> the <strong>Hamilton</strong>ian H : R d × R × R d → R is acontinuous real valued function on Ω×R×R d . In this first discussion, we will avoidto treat boundary conditions. In fact, boundary conditions need to be understoodin a suitable, more delicate “weak sense” which will be examined later on in thischapter.Typically, the main results in the theory of viscosity solutions assume that the<strong>Hamilton</strong>ian H(x, u, p) satisfies the following basic set of assumptions:(A1) H(·, ·, ·) is uni<strong>for</strong>mly continuous on Ω × R × R d(A2) H(x, u, ·) is convex on R d(A3) H(x, ·, p) is monotone on R.The need <strong>for</strong> a notion of weak solution arises when observing that problem (2.1) isnot expected to have a classical (i.e., C 1 (Ω)) solution, as the following counterexampleshows.17


✐✐18 Chapter 2. Viscosity Solutions of First Order PDEsA counterexampleConsider the following problem in one dimension:{|u x | = 1 x ∈ Ω ≡ (−1, 1)u(x) = 0 x ∈ ∂Ω.(2.2)Obviously, u 1 (x) = x <strong>and</strong> u 2 (x) = −x are C 1 <strong>and</strong> satisfy the equation pointwise in(−1, 1) but they do not satisfy both boundary conditions.On the other h<strong>and</strong>, if the assumption of differentiability is dropped, we mightconsider functions which satisfy the equation almost everywhere <strong>and</strong> satisfy theboundary conditions.Two such solutions would be given by the functions u 3 (x) = |x| − 1 <strong>and</strong> u 4 (x) =1 − |x|, but in fact, by collecting piecewise affine functions with slopes ±1 it ispossible to construct infinitely many a.e. solutions of the equation. There<strong>for</strong>e, it isclear that the notion of “a.e. solution” gives too many solutions <strong>and</strong> is unsuitable <strong>for</strong>a uniqueness result. One possibility to recover uniqueness is to per<strong>for</strong>m an ellipticregularization, i.e., to regularize the problem by adding a second order term −εu xx .Consider the second order problem−εu xx + |u x | = 1, x ∈ (0, 1) (2.3)with the homogeneous boundary condition u(−1) = 0, u(1) = 2. For every positiveε, this problem has a regular C 2 ((−1, 1)) solution u ε , which can be explicitly writtenasu ε (x) = x + exp ( ) ( )x−1ε − exp −1ε1 − exp ( )− 1 (2.4)εSince (2.2) is the limit problem of (2.3) as ε → 0, we would like to know thelimiting behaviour of u ε . In any compact set contained in (0, 1), u ε convergesuni<strong>for</strong>mly to u(x) = x. The main trouble appears near the boundary since theboundary condition causes the occurrence of a boundary layer of width O( √ ε). Thesolution u ε must attain the value u(x) = 2 at x = 1, <strong>and</strong> in the limit this produces adiscontinuity. Assuming that the limiting process converges to a continuous functionu, we can pass to the limitlimε→0 uε = u (2.5)<strong>and</strong> define u to be the weak solution of our problem. The definition of viscosity solutionavoids the limiting procedure related to the vanishing viscosity approximation<strong>and</strong> gives a local characterization of the weak solution.The “classical” definition We first give the most popular definition of viscositysolution, <strong>for</strong> u ∈ BUC(Ω) (the space of Bounded Uni<strong>for</strong>mly Continuous functionover the open set Ω).Definition 2.1. Let u ∈ BUC(Ω). We say that u is a viscosity solution of (2.1)if <strong>and</strong> only if, <strong>for</strong> any ϕ ∈ C 1 (Ω), the following conditions hold:(i) at every point x 0 ∈ Ω, local maximum <strong>for</strong> u − ϕ,(i.e., u is a viscosity subsolution);H(x 0 , u(x 0 ), Dϕ(x 0 )) ≤ 0


✐✐2.1. The definition of viscosity solution 19(ii) at every point x 0 ∈ Ω, local minimum <strong>for</strong> u − ϕ,(i.e., u is a viscosity supersolution).H(x 0 , u(x 0 ), Dϕ(x 0 )) ≥ 0Going back to the counterexample, we explain how such a definition allows toselect a unique a.e. solution.First, we show that any a.e. solution u which has a local minimum at x 0 cannot bea viscosity supersolution. In fact, choose ϕ ≡ c (a constant). Clearly, x 0 is a localminimum point <strong>for</strong> u − ϕ, <strong>and</strong> there<strong>for</strong>e we should have|ϕ x (x 0 )| ≥ 1which is false since ϕ x ≡ 0.Note that the same argument works in a different way <strong>for</strong> subsolutions. Take ana.e. solution u which has a local maximum at x 0 <strong>and</strong> choose ϕ ≡ c. Clearly, x 0 isa local maximum point <strong>for</strong> u − ϕ so that we should have|ϕ x (x 0 )| ≤ 1which is true, since ϕ x ≡ 0. This implies that the viscosity solution can havemaximum but not minimum points. The only viscosity solution of our problem isthere<strong>for</strong>e u(x) = 1 − |x|.2.1.1 Some properties of viscosity solutionsWe review here some properties of viscosity solutions which will be useful in thesequel. In particular, we want to examine the link between classical <strong>and</strong> viscositysolutions, as well as some other properties which are relevant <strong>for</strong> the constructionof approximation schemes.Relationship with classical solutions The first result states the local character ofviscosity solutions, along with their relationship with classical solutions.Proposition 2.2. The following statements hold true:(i) If u ∈ C(Ω) is a viscosity solution of (2.1) in Ω, then u is a viscosity solutionof the same equation in any open set Ω ′ ⊂ Ω.(ii) If u is a classical C 1 (Ω) solution, then it is also a viscosity solution.(iii) Any viscosity solution that is also regular is a classical solution.Proof. (i) If x 0 is a local maximum point on Ω ′ <strong>for</strong> u − ϕ, ϕ ∈ C 1 (Ω ′ ) then x 0is a local maximum point on Ω <strong>for</strong> u − ˜ϕ, <strong>for</strong> any test function ˜ϕ ∈ C 1 (Ω) suchthat ˜ϕ ≡ ϕ on B(x 0 , r) <strong>for</strong> some positive r. Then, by the definition of viscositysubsolutions,H(x 0 , u(x 0 ), Dϕ(x 0 )) = H(x 0 , u(x 0 ), D ˜ϕ(x 0 )) ≤ 0.(ii) Assume u ∈ C 1 is a classical solution, so that H(x, u(x), Du(x)) = 0 <strong>for</strong> every


✐✐20 Chapter 2. Viscosity Solutions of First Order PDEsx ∈ Ω. Consider a test function ϕ <strong>and</strong> a local maximum point <strong>for</strong> u − ϕ. It is notrestrictive to choose ϕ(x 0 ) = u(x 0 ). Then, we haveD(u − ϕ)(x 0 ) = 0. (2.6)This implies that Dϕ(x 0 ) = Du(x 0 ) so u is viscosity subsolution. The proof that uis a supersolution is analogous.(iii) We can always choose ϕ = u so that every point is a local maximum <strong>and</strong> alocal minimum <strong>for</strong> u − ϕ. Then, <strong>for</strong> any x ∈ Ω both inequalities will be satisfied:H(x, u(x), Du(x)) ≤ 0,H(x, u(x), Du(x)) ≥ 0.This implies that the equation is satisfied pointwise.Min/Max of viscosity solutions Viscosity subsolutions (respectively, supersolutions)are stable with respect to the max (respectively, the min) operator. Introducing,<strong>for</strong> u, v ∈ C(Ω), the notations:we have the stability result:(u ∨ v)(x) = max{u(x), v(x)}(u ∧ v)(x) = min{u(x), v(x)},Proposition 2.3. The following statements hold true:(i) Let u, v ∈ C(Ω) be viscosity subsolutions of the stationary equation in (2.1).Then, u ∨ v is a viscosity subsolution.(ii) Let u, v ∈ C(Ω) be viscosity supersolutions of the stationary equation in (2.1).Then, u ∧ v is a viscosity supersolution.Proof. Let x 0 be a local maximum point <strong>for</strong> u ∨ v − ϕ where ϕ ∈ C 1 (Ω) is our testfunction. Without loss of generality, we can assume that (u ∨ v)(x 0 ) = u(x 0 ). Sincex 0 is local maximum point <strong>for</strong> u − ϕ, we haveH(x 0 , u(x 0 ), Dϕ(x 0 )) ≤ 0which proves (i). The reverse assertion (ii) can be proved is a similar way.An important property which follows from Proposition 2.3 is that the viscositysolution u can be characterized as the maximal subsolution of the equation, i.e.,u ≥ v, <strong>for</strong> any v ∈ S (2.7)where S is the space of subsolutions, i.e. S = {v ∈ C(Ω) : condition (i) is satisfied}.Proposition 2.4. Let u ∈ C(Ω) be a viscosity subsolution of the stationary equationin (2.1), such that u ≥ v <strong>for</strong> any viscosity subsolution v ∈ C(Ω). Then, u is aviscosity supersolution <strong>and</strong> there<strong>for</strong>e a viscosity solution of (2.1).


✐✐2.1. The definition of viscosity solution 21Proof. We will prove the result by contradiction. Assume thatd := H(x 0 , u(x 0 ), Dϕ(x 0 )) < 0<strong>for</strong> some ϕ ∈ C 1 (Ω) <strong>and</strong> x 0 ∈ Ω such thatu(x 0 ) − ϕ(x 0 ) ≤ u(x) − ϕ(x)∀x ∈ B(x 0 , δ 0 ) ⊂ Ω<strong>for</strong> some δ 0 > 0. Now, consider the function w ∈ C 1 (Ω) defined asw(x) := ϕ(x) − |x − x 0 | 2 + u(x 0 ) − ϕ(x 0 ) + 1 2 δ2<strong>for</strong> 0 < δ < δ 0 . It is easy to check that, by construction,(u − w)(x 0 ) < (u − w)(x) ∀x such that |x − x 0 | = δ. (2.8)We prove now that, <strong>for</strong> some sufficiently small δ,H(x, w(x), Dw(x)) ≤ 0 ∀x ∈ B(x 0 , δ). (2.9)For this purpose, a local uni<strong>for</strong>m continuity argument shows that, <strong>for</strong> 0 < δ < δ 0 ,{|ϕ(x) − ϕ(x 0 )| ≤ ω 1 (δ),|Dϕ(x) − 2(x − x 0 ) − Dϕ(x 0 )| ≤ ω 2 (δ) + 2δ(2.10)<strong>for</strong> any x ∈ B(x 0 , δ), where the ω i , i = 1, 2, are the moduli of continuity of respectivelyϕ <strong>and</strong> Dϕ. Then, we haveNow,|w(x) − u(x 0 )| ≤ ω 1 (δ) + δ 2 x ∈ B(x 0 , δ)H(x, w(x), Dw(x)) = d + H(x, w(x), Dw(x)) = (2.11)= d + H(x, w(x), Dϕ(x) − 2(x − x 0 )) − H(x 0 , w(x 0 ), Dϕ(x 0 )).Denoting by ω the modulus of continuity of H, we can writeH(x, w(x), Dw(x)) ≤ d + ω(δ, ω 1 (δ) + δ 2 , ω 2 (δ) + 2δ),<strong>for</strong> all x ∈ B(x 0 , δ). Since d is negative, the above inequality proves (2.9) <strong>for</strong> δ > 0small enough. Let us fix such a δ, <strong>and</strong> set̂v(x) :={u ∨ w on B(x 0 , δ)u on Ω \ B(x 0 , δ).(2.12)It is easy to check that, by (2.8), ̂v ∈ C(Ω), so by Propositions (2.2) <strong>and</strong> (2.3) (i) ̂vis a subsolution of (2.1). Since ̂v(x 0 ) > u(x 0 ) the statement is proved.


✐✐22 Chapter 2. Viscosity Solutions of First Order PDEsUni<strong>for</strong>m convergence of viscosity solutions Viscosity solutions are also stablewith respect to uni<strong>for</strong>m convergence in C(Ω), as the following result shows. It isworth to point out that this property does not hold <strong>for</strong> generalized a.e. solutions.Proposition 2.5. Let u n ∈ C(Ω), n ∈ N, be a viscosity solution ofH n (x, u n (x), Du n (x)) = 0 x ∈ Ω. (2.13)Assume thatu n → u locally uni<strong>for</strong>mly in ΩH n → H locally uni<strong>for</strong>mly in Ω × R × R d .Then, u is a viscosity solution of (2.1).(2.14)Proof. Let ϕ ∈ C 1 (Ω) <strong>and</strong> x 0 be a local maximum point <strong>for</strong> u − ϕ. It is notrestrictive to assume thatu(x 0 ) − ϕ(x 0 ) > u(x) − ϕ(x)<strong>for</strong> x ≠ x 0 , x ∈ B(x 0 , r) <strong>for</strong> some r > 0. By uni<strong>for</strong>m convergence, <strong>and</strong> <strong>for</strong> n largeenough, u n − ϕ attains a local maximum at a point x n close to x 0 . Then, we haveH n (x n , u n (x n ), Dϕ n (x n )) ≤ 0.Since x n tends to x 0 <strong>for</strong> n → +∞, passing to the limit we getH(x n , u(x 0 ), Dϕ(x 0 )) ≤ 0The proof that u is also a viscosity supersolution can be obtained by a similarargument.In addition, the viscosity solution u may be characterized as the uni<strong>for</strong>m limit(<strong>for</strong> ε → 0) of classical solutions u ε of the regularized problem−εu ε xx + H(x, u ε , Du ε ) = 0, (2.15)that is,lim uε = u.ε→0 +This explains the name of viscosity solutions, since the term −εu xx corresponds tothe viscosity term in fluid dynamics. In principle, <strong>for</strong> a given ε, the regularizationterm −εu ε xx allows to use the theory of elliptic equations to establish the existenceof a regular solution u ε ∈ C 2 (Ω). Nevertheless, passing to the limit <strong>for</strong> ε → 0 isquite a technical point, since the regularizing effect of the second order term is cutdown as ε vanishes. Moreover, <strong>for</strong> what we have seen in this chapter, we do notexpect to have a regular solution <strong>for</strong> the limiting (first order) equation. Withoutproving a precise result (which can be found in [Ba93]), we try give an idea of therelationship between the definition of viscosity solution <strong>and</strong> the limiting behaviourof u ε .Assume that u ε ∈ C 2 (R) converge uni<strong>for</strong>mly as ε → 0 to some u ∈ C(R). Letus now take a regular function ϕ ∈ C 2 (R) <strong>and</strong> assume that x is a local maximumpoint <strong>for</strong> u − ϕ. By uni<strong>for</strong>m convergence, u ε − ϕ attains a local maximum at somepoint x ε <strong>and</strong>lim xε = x.ε→0 +


✐✐2.1. The definition of viscosity solution 23By the maximum properties, we getD(u ε − ϕ)(x ε ) = 0, −D 2 (u ε − ϕ)(x ε ) ≥ 0where D 2 denotes the second derivative operator. Then, using the previous inequalities<strong>and</strong> the fact that u ε is a classical solution of (2.15), we obtain−εϕ xx (x ε ) + H(x ε , u(x ε ), Dϕ(x ε )) ≤ (2.16)−εu ε xx(x ε ) + H(x ε , u(x ε ), Du(x ε )) = 0Passing to the limit <strong>and</strong> using the continuity of ϕ xx , Dϕ <strong>and</strong> H we finally getH(x, u(x), Dϕ(x)) ≤ 0.This computation explains why, from the theoretical point of view, it is reasonableto obtain viscosity solutions via an elliptic regularization of the first order equation.Such a technique, however, is not as convenient <strong>for</strong> a numerical approach, sincethe introduction of a second-order term has a smoothing action. In particular, thiseffect is apparent on the singularities of solutions, which are definitely a point ofinterest in the approximation of <strong>Hamilton</strong>–Jacobi equations.Be<strong>for</strong>e concluding the section, we give a result concerning changes of variablesin (2.1). In the sequel, this result will be useful to deal with certain control problems.Proposition 2.6. Let u ∈ C(Ω) be a viscosity solution of (2.1) <strong>and</strong> Φ : R → R bea function in C 1 (R) such that Φ ′ (t) > 0. Then v := Φ(u) is a viscosity solution ofH ( x, Φ −1 ((v(x)), Φ −1 (v(x))Dv(x) ) = 0 x ∈ Ω. (2.17)We conclude this review with a warning about the fact that viscosity solutionsare not preserved by a change of sign in the equation. This behaviour, althoughunusual, is easily understood if we go back to the definition, which only takes intoaccount the local properties at minimum <strong>and</strong> maximum points <strong>for</strong> u−ϕ. Indeed, thefact that any local maximum of u−ϕ is a local minimum of −(u−ϕ) implies that uis a viscosity subsolution of (2.1) if <strong>and</strong> only if v = −u is viscosity supersolution of−H(x, −v(x), −Dv(x)) = 0 in Ω. Similarly, u is a viscosity supersolution of (2.1)if <strong>and</strong> only if v = −u is a viscosity subsolution of −H(x, −v(x), −Dv(x)) = 0 in Ω.2.1.2 An alternative definition of viscosity solutionThe definition of viscosity solutions based on test functions is probably the mostpopular <strong>and</strong> useful. However, the original definition was based the notion on sub<strong>and</strong>superdifferentials which are st<strong>and</strong>ard tools in convex analysis.To introduce this notion, we first recall the elementary definition of a differentiablefunction:Definition 2.7. v is differentiable at x 0 if <strong>and</strong> only if there exists a linear operatorA such thatv(y) − v(x 0 ) − A(y − x 0 )lim= 0y→x 0 |y − x 0 |


✐✐24 Chapter 2. Viscosity Solutions of First Order PDEsIf v is not differentiable at x 0 we can still have a generalized version of thedifferential, based on the following idea (sketched in one dimension <strong>for</strong> simplicity).In a univariate differentiable function the tangent at x 0 is uniquely defined bytaking the limit of secants passing through points of the graph of v which convergeto x 0 , the limit slope defining the derivative v ′ . If v is not differentiable at x 0 , thisprocedure does not select a unique tangent (<strong>and</strong> a unique derivative), but rathera set of “generalized tangents”. For example, v(x) = |x| has a unique tangent <strong>for</strong>every x ≠ 0, whereas <strong>for</strong> x = 0 the “generalized tangents” are all the straight linespassing in (0, 0) <strong>and</strong> not crossing the graph of v, i.e. the lines of the set{y = mx, m ∈ (−1, 1)}.The st<strong>and</strong>ard extension of derivatives used in nonsmooth analysis is based on thisidea <strong>and</strong> corresponds to the following definition of sub- <strong>and</strong> superdifferentials.Definition 2.8. Let Ω be an open set of R d <strong>and</strong> v : Ω → R.The super-differential D + v(x) of v at x ∈ Ω, is defined as the set:{}D + v(x) := p ∈ R d v(y) − v(x) − p · (y − x): lim sup≤ 0y→x y∈Ω |y − x|The sub-differential D − v(x) of v at x ∈ Ω, is defined as the set:{}D − v(x) := q ∈ R d v(y) − v(x) − q · (y − x): lim inf≥ 0 .y→x y∈Ω |y − x|We remark that the sub- <strong>and</strong> superdifferential are (possibly empty) convex<strong>and</strong> closed subsets of R d . Moreover, if at the point x we haveD + v(x) = D − v(x)then v is differentiable in the sense of Definition 2.7. Going back to the example ofv(x) = |x|, we have, <strong>for</strong> x ≠ 0:{D + v(x) = D − −1 <strong>for</strong> x < 0v(x) =1 <strong>for</strong> x > 0so they coincide <strong>and</strong> give the st<strong>and</strong>ard derivative.At x = 0, we have{D − v(0) = q ∈ R d : lim infTaking <strong>for</strong> example q = 1, we have :=y→0|y| − q y|y|}≥ 0 ={}q ∈ R d : lim inf (1 − q sgn(y)) ≥ 0 .y→0lim inf(1 − sgn(y)) = min{2, 0} = 0.y→0In the same way, we can show that every q ∈ [−1, 1] belongs to the subdifferential.Then, D − v(0) = [−1, 1].


✐✐2.1. The definition of viscosity solution 25Considering now the superdifferential at x = 0, we have:{}D + v(0) = p ∈ R d |y| − p y: lim sup ≤ 0 =y→0 |y|{}= p ∈ R d : lim sup(1 − p sgn(y)) ≤ 0y→0On the other h<strong>and</strong>, we havelim sup(1 − p sgn(y)) = lim(1 + max(p, −p)) =y→0y→0= lim(1 + |p|) ≥ 1.y→0This shows that D + v(0) = ∅.We can now give the definition of viscosity solution based on sub- <strong>and</strong> superdifferentials.Definition 2.9. Let u ∈ BUC(Ω). We say that u is a viscosity solution of (2.1)if <strong>and</strong> only if, <strong>for</strong> any ϕ ∈ C 1 (Ω), the following conditions hold:(i) H(x, u(x), p) ≤ 0, <strong>for</strong> all p ∈ D + u(x), x ∈ Ω(i.e., u is a viscosity subsolution);(ii) H(x, u(x), p) ≥ 0, <strong>for</strong> all p ∈ D − u(x), x ∈ Ω(i.e., u is a viscosity supersolution).The following result explains the link between the two definitions.Theorem 2.10. Let Ω be an open subset of R d <strong>and</strong> u ∈ C(Ω), then(i) p ∈ D + u(x), x ∈ Ω if <strong>and</strong> only if there exists a function φ ∈ C 1 (Ω) such thatDφ(x) = p <strong>and</strong> u − φ has a local maximum point at x.(ii) q ∈ D − u(x), x ∈ Ω if <strong>and</strong> only if there exists a function φ ∈ C 1 (Ω) such thatDφ(x) = q <strong>and</strong> u − φ has a local minimum point at x.Proof. A proof of this theorem can be found in [BCD97] (Lemma 1.7, p. 29).We conclude this section with a further result linking viscosity <strong>and</strong> a.e. solutions.Theorem 2.11. The following statements hold true:(i) if u ∈ C(Ω) is a viscosity solution of (2.1), thenH(x, u(x), Du(x)) = 0at any point x ∈ Ω where u is differentiable;(ii) if u is locally Lipschitz continuous <strong>and</strong> is a viscosity solution of (2.1), thenH(x, u(x), Du(x)) = 0a.e. in Ω


✐✐26 Chapter 2. Viscosity Solutions of First Order PDEsProof. Let u be a point of differentiability <strong>for</strong> u. Then, D + u(x) ∩ D − u(x) ≠ ∅since this set contains Du(x). Moreover, this intersection reduces to a singleton, sothatDu(x) = D + u(x) = D − u(x).Hence, by Definition 2.9 we have0 ≥ H(x, u(x), Du(x)) ≥ 0which proves (i). Statement (ii) follows immediately from (i) <strong>and</strong> Rademacher’stheorem (which states that Lipschitz continuous functions are a.e. differentiable,see [E98] pp. 280–281 <strong>for</strong> the proof).2.1.3 Uniqueness of viscosity solutionsThe crucial point in the theory of viscosity solution (<strong>and</strong> also the essential advantageover the concept of solutions almost everywhere) is to prove uniqueness. This isdone via a comparison principle, also termed maximum principle.Theorem 2.12. Let u, v ∈ BUC(Ω) be respectively a sub- <strong>and</strong> a supersolution <strong>for</strong>(2.1), <strong>and</strong> letu(x) ≤ v(x) <strong>for</strong> any x ∈ ∂Ω.Then,u(x) ≤ v(x) <strong>for</strong> any x ∈ Ω.This is enough to get uniqueness, since taken two viscosity solutions of theequation they are (both) sub- <strong>and</strong> supersolutions. This implies that both u(x) ≤v(x) <strong>and</strong> u(x) ≥ v(x) <strong>for</strong> any x ∈ Ω, i.e., u(x) = v(x) in Ω.The following is a classical assumption <strong>for</strong> uniqueness.(A4) Let ω(·) be a modulus of continuity. We assume that|H(x, u, p) − H(y, u, p)| ≤ ω(|x − y|(1 + |p|))Q R (x, y, u, p)<strong>for</strong> any x, y ∈ Ω, u ∈ [−R, R] <strong>and</strong> p ∈ R d , whereQ R (x, y, u, p) ≡ max (ϕ(H(x, u, p)) , ϕ (H(y, u, p)))<strong>and</strong> ϕ : R → R + is continuous.We can now give a sufficient condition <strong>for</strong> the comparison principle to hold.Theorem 2.13. Let assumptions (A1)–(A4) be satisfied. Then, the comparisonprinciple (Theorem 2.12) holds <strong>for</strong> (2.1), i.e., the viscosity solution is unique.Proof. We only sketch the proof, which is carried out via variable doubling.The goal is to prove that M = max x∈Ω(u−v) is negative. To this end, we introducea test function depending on two variables:ψ ε (x, y) ≡ u(x) − v(y) −|x − y|2ε 2 .


✐✐2.2. Viscosity solution <strong>for</strong> evolutive equations 27Due to the penalization term, we can expect that the maximum points (x ε , y ε ) <strong>for</strong>ψ ε should have x ε <strong>and</strong> y ε close enough <strong>for</strong> ε small. Moreover, <strong>for</strong> ε → 0 + we have:M ε → M,|x − y| 2ε 2 → 0,u(x ε ) − v(y ε ) → M.This allows to pass to the limit <strong>and</strong> get the comparison result.2.2 Viscosity solution <strong>for</strong> evolutive equationsOnce fixed some basic ideas, the definition of viscosity solution can be easily extendedto the evolutive case, by taking into account also the time derivative.Definition 2.14. u ∈ BUC(Ω × (0, T ) is a viscosity solution in Ω × (0, T ) of theequationu t + H(x, t, u, Du) = 0 (2.18)if <strong>and</strong> only if, <strong>for</strong> any ϕ ∈ C 1 (Ω × (0, T )), the following conditions hold:(i) at every point (x 0 , t 0 ) ∈ Ω × (0, T ), local maximum <strong>for</strong> u − ϕ,ϕ t (x 0 , t 0 ) + (H(x 0 , t 0 , u(x 0 , t 0 ), Dϕ(x 0 , t 0 )) ≤ 0(i.e., u is a viscosity subsolution).(ii) at every point (x 0 , t 0 ) ∈ Ω × (0, T ), local minimum <strong>for</strong> u − ϕ,ϕ t (x 0 , t 0 ) + (H(x 0 , t 0 , u(x 0 , t 0 ), Dϕ(x 0 , t 0 )) ≥ 0(i.e., u is a viscosity supersolution).As one can see, this definition is very close to the one used <strong>for</strong> stationaryproblems. In fact, it could also be derived from Definition 2.1 by setting the problem(2.1) in d + 1 variables (x, t) ∈ R d × (0, T ) <strong>and</strong> making the time derivative explicit.The following comparison result refers to the evolutive problem <strong>and</strong> usuallygives the uniqueness <strong>for</strong> the particular case H(x, t, u, Du) = H(t, Du).Theorem 2.15. Assume H ∈ C((0, T ) × R d ). Let u 1 , u 2 ∈ C(R d × (0, T )) be,respectively, viscosity sub- <strong>and</strong> supersolution in R d × [0, T ] of the equationu t (x, t) + H(t, Du(x, t)) = 0.Then,sup (u 1 − u 2 ) ≤ sup(u 1 (·, 0) − u 2 (·, 0)).R d ×(0,T )R dProof. The proof can be found in [BCD97], p. 56. Note that more generalcomparison results <strong>for</strong> more general <strong>Hamilton</strong>ians can be found in [Ba98].


✐✐28 Chapter 2. Viscosity Solutions of First Order PDEs2.2.1 Representation <strong>for</strong>mulae <strong>and</strong> Legendre trans<strong>for</strong>mIn some cases it is possible to derive representation <strong>for</strong>mulaa <strong>for</strong> the viscosity solution.This <strong>for</strong>mulae have a great importance from both the analytical <strong>and</strong> thenumerical point of view, <strong>and</strong> will be derived here in two major cases: linear advectionequations <strong>and</strong> convex HJ equations.<strong>Linear</strong> advection equationLet us start by the linear case in which the representation <strong>for</strong>mula can be obtainedvia the method of characteristics.Theorem 2.16. Let u : R d × (t 0 , T ) → R be a viscosity solution of the initial valueproblem{u t (x, t) + λu(x, t) + f(x, t) · Du(x, t) = g(x, t) (x, t) ∈ R d × (t 0 , T )u(x, t 0 ) = u 0 (x) x ∈ R d (2.19).Assume that λ ≥ 0, f : R d × (t 0 , T ) → R d <strong>and</strong> g : R d × (t 0 , T ) → R are continuousin (x, t) <strong>and</strong> f is globally Lipschitz continuous with respect to x. Then,∫ tu(x, t) = e λ(t0−t) u 0 (y(x, t; t 0 )) + e λ(s−t) g(y(x, t; s), s)dst 0(2.20)where y(x, t; s) is the the position at time s of the solution trajectory passing throughx at time t, i.e., solving the Cauchy problem⎧⎨ dy(x, t; s) = f(y(x, t; s), s)ds (2.21)⎩y(x, t; t) = x.Proof. We give the proof under the additional assumption that u ∈ C 1 .Let (x, t) be fixed, <strong>and</strong> denote <strong>for</strong> shortness the solution of (2.21) as y(s) ≡ y(x, t; s).Writing the equation in (2.19) at a point (y(s), s) <strong>and</strong> multiply by e λs , we havee λs u s (y(s), s) + λe λs u(y(s), s) + e λs f(y(s), s) · Du(y(s), s) = e λs g(y(s), s).Since u is differentiable, this may also be rewritten asIntegrating (2.22) over the interval [t 0 , t] we getd [e λs u(y(s), s) ] = e λs g(y(s), s). (2.22)ds∫ te λt u(y(t), t) = e λt0 u(y(t 0 ), t 0 ) + e λs g(y(s), s)ds.t 0(2.23)Recalling that y(t) = y(x, t; t) = x <strong>and</strong> u(y(t 0 ), t 0 ) = u 0 (y(t 0 )) <strong>and</strong> dividing by e λt ,we get∫ tu(x, t) = e λ(t0−t) u 0 (y(t 0 )) + e λ(s−t) g(y(s), s)dst 0


✐✐2.2. Viscosity solution <strong>for</strong> evolutive equations 29which coincides with (2.20).We recall that a solution of (2.21) is called characteristic curve. Note that, inusing (2.20), (2.21) is integrated backwards. Note also that, if λ is strictly positive,f <strong>and</strong> g do not depend on t <strong>and</strong> the source g is bounded, then u(x, t) has a limit<strong>for</strong> t − t 0 → ∞. Setting conventionally t = 0 <strong>and</strong> letting t 0 → −∞, we obtain infact the limitu(x) ==∫ 0−∞∫ ∞0e λs g(y(x, 0; s))ds =e −λs g(y(x, 0; −s))ds (2.24)which is a regime solution <strong>for</strong> problem (2.19), or, in other terms, solves the stationaryequationλu(x) + f(x) · Du(x) = g(x)<strong>for</strong> x ∈ R d .<strong>Hamilton</strong>–Jacobi equationsConcerning HJ equations, the representation <strong>for</strong>mula is known as Hopf–Lax <strong>for</strong>mula,<strong>and</strong> is typically related to the problem:{u t + H(Du) = 0 (x, t) ∈ R d × (0, T ),u(x, 0) = u 0 (x) x ∈ R d (2.25),where H : R d → R is convex <strong>and</strong> satisfies the coercivity condition:H(p)lim = +∞. (2.26)|p|→+∞ |p|Assumption (2.26) allows to give the following definition:Definition 2.17. Let (2.26) be satisfied. We define the Legendre–Fenchel conjugate(or Legendre–Fenchel trans<strong>for</strong>m) of H <strong>for</strong> q ∈ R d as:H ∗ (q) = supp∈R d {p · q − H(p)}. (2.27)Note that the convexity assumption on H implies that H is continuous, <strong>and</strong>being also coercive in the sense of (2.26), the sup in (2.27) is in fact a maximum.In general, Legendre–Fenchel trans<strong>for</strong>m may not allow <strong>for</strong> an explicit computation– we will show in Chapter 3 a numerical procedure to approximate it. Afew examples, anyway, can be computed analytically, among which the quadratic<strong>Hamilton</strong>ian:H 2 (p) = |p|22 ,<strong>for</strong> which an easy computation givesH ∗ 2 (q) = |q|22 .


✐✐30 Chapter 2. Viscosity Solutions of First Order PDEsIt is interesting to note that a similar construction may also work <strong>for</strong> the noncoercivecase, but in this case the Legendre–Fenchel conjugate will not in general be bounded<strong>and</strong> defined everywhere. For example, takingH 1 (p) = |p|<strong>and</strong> using the definition (2.27), it is easy to check thatH ∗ 1 (q) ={0 <strong>for</strong> |q| ≤ 1+∞ elsewhere.The main result of interest here concerns two important properties of the Legendretrans<strong>for</strong>m:Theorem 2.18. Let (2.26) be satisfied. Then, the function H ∗ has the followingproperties:(i) H ∗ : R d → R is convex <strong>and</strong>(ii) H(p) = H ∗∗ (p), <strong>for</strong> any p ∈ R dH(p)lim = +∞;|p|→+∞ |p|Proof. The proof can be found in [E98], p. 122.The above theorem says that the by applying twice the Legendre trans<strong>for</strong>mto H we obtain back H itself. The definition of Legendre trans<strong>for</strong>m is very usefulto characterize the unique solution of (2.25) by means of the so-called Hopf–Laxrepresentation <strong>for</strong>mula. The following theorems provide the basic results concerningthis characterization.Theorem 2.19. The function u defined by the following Hopf–Lax <strong>for</strong>mula:[( )] x − yu(x, t) = inf v 0 (y) + tH ∗y∈R d t(2.28)is Lipschitz continuous, differentiable almost everywhere in R d ×(0, +∞) <strong>and</strong> solvesa.e. the initial value problem (2.25).Proof. The proof relies on the functional identity{ ( ) }x − yu(x, t) = min (t − s)H ∗ + u(y, s)y∈R d t − s(2.29)(which is proved in [E98] p. 126). This identity implies that u is Lipschitz continuous<strong>and</strong>, by Rademacher’s theorem, differentiable almost everywhere. Now, consider apoint of differentiability (x, t).


✐✐2.2. Viscosity solution <strong>for</strong> evolutive equations 31Step 1. u is a subsolution at (x, t).Fix p ∈ R d <strong>and</strong> <strong>and</strong> a positive parameter h. By (2.29), we have{ ( ) }x + hp − yu(x + hp, t + h) = min hH ∗ + u(y, t) ≤ hH ∗ (p) + u(x, t)y∈R d hwhich impliesNow, <strong>for</strong> h → 0, we getu(x + hp, t + h) − u(x, t)h≤ H ∗ (p)p · Du(x, t) + u t (x, t) ≤ H ∗ (p).This inequality holds <strong>for</strong> any p ∈ R d , so that we haveu t (x, t) + H(Du(x, t)) + u t (x, t) + maxp∈R d {p · Du(x, t) − H ∗ (p)} ≤ 0where we have used the fact that H = H ∗∗ .Step 2. u is a supersolution at (x, t).Let us choose z such thatu(x, t) = t H ∗ ( x − zt)+ u 0 (z).Fix a positive parameter h <strong>and</strong> set s = t − h, y = x s/t + (1 − s/t)z. This implies<strong>and</strong> there<strong>for</strong>eu(x, t) − u(y, s) ≥ tH ∗ ( x − ztAs a consequence,<strong>and</strong>, <strong>for</strong> h → 0, we obtain:Then, we finally getx − st)= (t − s)H ∗ ( x − zt= y − z ,s[+ u 0 (z) −).sH ∗ ( y − zsu(x, t) − u (( )1 − h t x +h( )tz, t − h) x − z≥ H ∗ ,htx − zt( ) x − z· Du(x, t) + u t (x, t) ≥ H ∗ .t) ]+ u 0 (z) =u t (x, t) + H((Du(x, t)) = u t (x, t) + max {p · Du(x, t) − H ∗ (p)} ≥p∈R d≥ u t (x, t) + x − z( ) x − z· Du(x, t) − H ∗ ≥ 0tt


✐✐32 Chapter 2. Viscosity Solutions of First Order PDEswhich concludes the proof.Theorem 2.20. The unique viscosity solution of (2.25) is given by the Hopf-Laxrepresentation <strong>for</strong>mula.Proof. The proof of this result can be found in [E98] p. 561.We finally note that an equivalent way of rewriting Hopf-Lax <strong>for</strong>mula is to seta := x − ytso that y = x + at. Since the minimization with respect to y corresponds to aminimization with respect to a, we also obtain:u(x, t) = infa∈R d [u 0 (x − at) + tH ∗ (a)] . (2.30)This <strong>for</strong>mula shows more clearly the link between the value of the solution at (x, t)<strong>and</strong> the values of the solution at points (x−at, 0) <strong>and</strong> it will be useful in the analysisof semi-Lagrangian schemes.2.2.2 Semiconcavity <strong>and</strong> regularity of viscosity solutionsAs we have seen earlier in this chapter, we expect viscosity solutions to be in generalonly uni<strong>for</strong>mly continuous. However, some further regularity results can be given.We will examine here the most typical of these results. A regularity property whichplays an important role in HJ equations is semiconcavity, defined as follows.Definition 2.21. A function u ∈ C(Ω) is semiconcave in the open convex set Ω ifthere exists a constant C > 0 such that, <strong>for</strong> any x, z ∈ Ω <strong>and</strong> µ ∈ [0, 1]:µu(x) + (1 − µ)u(y) ≤ u(µx + (1 − µ)y) + 1 2 Cµ(1 − µ)|x − y|2 (2.31)This definition corresponds to the request that u(x) − 1 2 C|x|2 is concave, as itcan be immediately verified. Note that an equivalent definition, when u is continuous,is to require thatu(x + h) − 2u(x) + u(x − h) ≤ C|h| 2 (2.32)<strong>for</strong> all x ∈ Ω <strong>and</strong> h ∈ R d , <strong>for</strong> |h| sufficiently small.Clearly, concave functions are also semiconcave. Moreover, C 1 functions havinga Lipschitz continuous gradient are semiconcave. Another important exampleof nondifferentiable semiconcave functions is given by marginal functions, which arefunctions defined asu(x) = inf F (x, a) (2.33)a∈Aprovided F (·, a) satisfies (2.31) uni<strong>for</strong>mly with respect to a. Marginal functionsare particularly important in control problems, since value functions belong to thisclass. We will see that semiconcavity is one of the properties that the value functioninherits from the data in many control problems.


✐✐2.2. Viscosity solution <strong>for</strong> evolutive equations 33Example: distance function from a setas usual the distance d(x, C) byd(x, C) := inf |x − y|y∈CGiven a domain C ⊂ R d , C ≠ ∅, defineThen d 2 is semiconcave in R d since the application x → |x − y| 2 is C ∞ <strong>and</strong> hasconstant second derivatives. Moreover, d itself is semiconcave on any compact sethaving strictly positive distance from C, because the mapping x → |x − y| hasalso bounded second derivatives in such a set, <strong>and</strong> that bound is uni<strong>for</strong>m <strong>for</strong> y ina bounded set. It is interesting to note that our counterexample 2.2 correspondsto the characterization of the distance function from the set R \ (−1, 1) <strong>and</strong> onlyviscosity solution u(x) = 1 − |x| is semiconcave (<strong>and</strong> is also the only semiconcavea.e. solution).Properties of semiconcave functions We review here some properties of semiconcavefunctions which are of particular interest in relationship with viscosity solutionsof HJ equations.Theorem 2.22. Let u be semiconcave in Ω. Then, u is locally Lipschitz continuousin Ω.Proof. For any fixed x ∈ Ω <strong>and</strong> any h such that x + h ∈ Ω, we haveu(x + h) − u(x) = ψ(x + h) − ψ(x) + Cx · h + C 2 |h|2where ψ(x) = u(x) − 1 2 C|x|2 is concave <strong>and</strong>, there<strong>for</strong>e, Lipschitz continuous in Ω.Since the sum of the last two terms is quadratic in h, this proves the assertion.Theorem 2.23. Let us assume that the <strong>Hamilton</strong>ian is convex <strong>and</strong> that the initialcondition u 0 is semiconcave. Then, the viscosity solution of (2.25) si semiconcave.Proof. Assume that u 0 satisfies (2.32), <strong>and</strong> let x <strong>and</strong> h be fixed. By the Hopf–Lax<strong>for</strong>mula, the solution u(x, t) of (2.25) may be written at x <strong>and</strong> x ± h as:u(x, t) = inf [u 0 (x − at) + tH ∗ (a)] = u 0 (x − at) + tH ∗ (a)a∈R du(x + h, t) = inf [u 0 (x + h − at) + tH ∗ (a)] = u 0 (x + h − a + t) + tH ∗ (a + )a∈R du(x − h, t) = inf [u 0 (x − h − at) + tH ∗ (a)] = u 0 (x − h − a − t) + tH ∗ (a − ),a∈R dwhere we have denoted by a, a + <strong>and</strong> a − the minimizers <strong>for</strong> respectively x, x−h <strong>and</strong>x + h in the Hopf–Lax <strong>for</strong>mula. On the other h<strong>and</strong>, replacing a ± by a, we obtainthe inequalitiesso that at last we get:u(x + h, t) ≤ u 0 (x + h − at) + tH ∗ (a)u(x − h, t) ≤ u 0 (x − h − at) + tH ∗ (a),u(x + h) − 2u(x) + u(x − h) ≤ u 0 (x + h − at) − 2u 0 (x − at) + u 0 (x − h − at) ≤≤ C|h| 2


✐✐34 Chapter 2. Viscosity Solutions of First Order PDEs<strong>and</strong>, by semiconcavity of u 0 , we obtain semiconcavity of uLet now u ∈ W 1,∞ (Ω), <strong>and</strong> define the setlocD ∗ u(x) := {p ∈ R d : p = limn→∞ Du(x n), x n → x}.Then, D ∗ is nonempty <strong>and</strong> closed <strong>for</strong> any x ∈ Ω. Denote by co D ∗ u(x) its convexhull. A classical result in nonsmooth analysis (see [Cl83] <strong>for</strong> the proof) states thatco D ∗ u(x) = ∂u(x),∀x ∈ Ωwhere ∂u(x) is the generalized gradient of u at x, defined by∂u(x) := {p ∈ R d : u(x, q) ≥ p · q, ∀q ∈ R d } = (2.34)= {p ∈ R d : u(x, q) ≤ p · q, ∀q ∈ R d }<strong>and</strong> u, u are respectively, the generalized directional derivatives defined byu(y + tq) − u(y)u(x; q) := lim supy→x,t→0 t+u(x; q) :=u(y + tq) − u(y)lim infy→x,t→0 + t(2.35)An important property is that, under the assumption of semiconcavity, some generalized<strong>and</strong> classical derivatives coincide.Proposition 2.24. Let u be semiconcave in Ω. Then, <strong>for</strong> all x ∈ Ω:(i) D + (u(x) = ∂u(x) = co D ∗ u(x);(ii) Either D − u(x) = ∅ or u is differentiable at x;(iii) If D + u(x) is a singleton, then u is differentiable at x;(iv) ∂u∂q (x) = min p∈D + u(x) p · q <strong>for</strong> all unit vectors q.Proof. The proof can be found in [BCD97] p. 66.Remark 2.25. Claim (ii) has an important consequence. Assume that the Legendretrans<strong>for</strong>m H ∗ is coercive <strong>and</strong> has bounded second derivatives. Then, writing theHopf–Lax representation <strong>for</strong>mula <strong>for</strong> u,u(x, t) = infa∈R d [u 0 (x − at) + tH ∗ (a)] ,we note that the infimum is actually a minimum, <strong>and</strong> that the function in squarebrackets to be minimized is semiconcave. At a minimum point a, its subdifferentialcannot be empty (in fact, it must contain at least the origin), <strong>and</strong> by the regularityof H ∗ , claim (ii) implies that u must be differentiable at the point x − at.Moreover, semiconcavity allows to clarify the link between a.e. solutions <strong>and</strong>viscosity supersolutions.


✐✐2.3. Problems in bounded domains 35Proposition 2.26. Let u be semiconcave <strong>and</strong> such thatH(x, u(x), Du(x)) ≥ 0 a.e. in Ω,where H is continuous. Then, u is a viscosity supersolution ofH(x, u(x), Du(x)) ≥ 0 in Ω,Proof. The proof can be found in [BCD97].Note that this proposition gives a sharper interpretation of counterexample(2.2). In fact, all a.e. solutions are subsolutions in the viscosity sense, but only thesemiconcave one u(x) = 1 − |x| is also a supersolution.Under suitable assumptions, it is possible to prove the semiconcavity of viscositysolutions <strong>for</strong> HJ equations of more general structure, as the following theoremshows.Theorem 2.27. Let λ > 0 <strong>and</strong> u ∈ W 1,∞ be a viscosity solution ofλu(x) + H(x, Du(x)) = 0with Lipschitz constant L u . Assume that H satisfiesx ∈ R d|H(x, p) − H(x, q)| ≤ ω(|p − q|), ∀x, p, q ∈ R d (2.36)<strong>and</strong> that, <strong>for</strong> some C > 0 <strong>and</strong> ̂L > 2L u ,H(x + h, p + Ch) − 2H(x, p) + H(x − h, p − Ch) ≥ −C|h| 2 (2.37)holds <strong>for</strong> all x, hR d , p ∈ B(0, ̂L). Then, u is semiconcave on R d .Proof. The proof can be found in [BCD97] p. 69.A final remark is that combining convexity of the <strong>Hamilton</strong>ian with semiconcaveregularity one can obtain an unexpected differentiability result <strong>for</strong> the viscositysolution.Proposition 2.28. Assume that u ∈ C(Ω) is a viscosity solution in Ω ofλu(x) + H(x, Du(x)) = 0with λ ≥ 0. Assume also that H(x, ·) is strictly convex <strong>for</strong> any fixed x ∈ Ω <strong>and</strong> that−u is semiconcave. Then u ∈ C 1 (Ω).Proof. The proof can be found in [BCD97].2.3 Problems in bounded domainsAnother peculiar point of the theory of viscosity solutions is the way boundaryconditions are satisfied. It is not surprising that boundary conditions cannot bearbitrarily assigned in first-order equations – even <strong>for</strong> advection equations we knowthat Dirichlet data can only be imposed at points of the boundary where the drift isdirected inwards. The technical point <strong>for</strong> nonlinear equations is that the equationplays a role up to the boundary.


✐✐36 Chapter 2. Viscosity Solutions of First Order PDEs2.3.1 Boundary Conditions in a weak senseIn the theory of viscosity solutions, the typical compact <strong>for</strong>m <strong>for</strong> boundary conditionsismin(H(x, u(x), Du(x)), B(x, u(x), Du(x))) ≤ 0max(H(x, u(x), Du(x)), B(x, u(x), Du(x))) ≥ 0where x ∈ ∂Ω, <strong>and</strong> B represents a suitably defined boundary operator. In order toexplain this setting, we will consider in detail the main situations of interest. Forexample, the Dirichlet condition would be <strong>for</strong>mulated in a classical <strong>for</strong>m as:u(x) = b(x)(x ∈ ∂Ω).In its weak <strong>for</strong>mulation, this corresponds to a boundary operator defined byB(x, u(x), Du(x))) ≡ u − b.We still remark that not all the boundary conditions are compatible with the equation.In this section, we briefly analyze the effect of Dirichlet, Neumann <strong>and</strong> “stateconstraints” boundary conditions, posed on parts of the boundary. First, note thatboundary conditions should be imposed in a weak sense. The condition whichdefines u as a viscosity subsolution <strong>for</strong> (2.1) requires that <strong>for</strong> any test functionϕ ∈ C 1 (Ω) <strong>and</strong> x ∈ ∂Ω local maximum point <strong>for</strong> u − ϕ,min{H(x, u(x), Dϕ(x)), B(x, u, Dϕ(x))} ≤ 0 (2.38)<strong>for</strong> a given boundary operator B. Similarly, the condition <strong>for</strong> supersolutions requiresthat <strong>for</strong> any test function ϕ ∈ C 1 (Ω) <strong>and</strong> x ∈ ∂Ω local minimum point <strong>for</strong> u − ϕ,max{H(x, u(x), Dϕ(x)), B(x, u, Dϕ(x))} ≥ 0. (2.39)The effect of the Dirichlet condition is to impose a value on u according to theabove conditions, in particular the value u(x) = b(x) is set at every point whereH(x, u(x), Dϕ(x)) ≥ 0 (<strong>for</strong> subsolutions) <strong>and</strong> H(x, u(x), Dϕ(x)) ≤ 0 (<strong>for</strong> supersolutions).The Neumann conditon, which is classically stated as∂u(x) = m(x)∂ν x∈ ∂Ωuses in its weak <strong>for</strong>mulation the boundary operatorB(x, u(x), Du(x))) ≡ ∂u∂ν − m.where ν(·) represents the outward normal to the domain Ω. A typical use of thiscondition occurs when we know (or presume) that the level curves of the surface areorthogonal to the boundary ∂Ω or to a part of it, in which case we simply choosem(x) = 0.On the other h<strong>and</strong>, in the “state constraints” boundary condition neither avalue <strong>for</strong> u nor a value <strong>for</strong> its normal derivative ∂u/∂ν(x) is imposed. In this respect,it has been sometimes interpreted as a “no boundary condition” choice althoughthis interpretation is quite sloppy. In fact, a bounded <strong>and</strong> uni<strong>for</strong>mly continuous


✐✐2.3. Problems in bounded domains 37function u is said to be a state constrained viscosity solution if <strong>and</strong> only if it is asubsolution in Ω <strong>and</strong> a supersolution in Ω (i.e., up to the boundary). Whenever abound on the L ∞ norm of the solution is known, this condition can be also stated asa Dirichlet boundary condition by simply setting b(x) ≡ C, provided the constantC satisfiesC > maxx∈Ω u(x) .By this choice (2.38) is trivially satisfied, whereas (2.39) requires (strictly)Representation <strong>for</strong>mulae in bounded domainsin a bounded domain Ω. Define byH(x, u(x), Dϕ(x)) ≥ 0. (2.40)We treat separately the linear caseΓ in := {x ∈ ∂Ω : f(x, t) · ν(x) < 0} (2.41)the subset of boundary points where the vectorfield f is pointing inward Ω. Weconsider again the linear evolutive problem (1.1), which is rewritten here:{u t + f(x, t) · Du = g(x, t) (x, t) ∈ Ω × (t 0 , T )u(x, 0) = u 0 (x)x ∈ Ω<strong>and</strong> complemented with the Dirichlet boundary conditionu(x, t) = b(x, t) x ∈ Γ in . (2.42)Note that, in terms of characteristics, the existence of an inflow boundary Γ inmeans that the characteristic y(x, t; s) may have a finite exit time from Ω <strong>for</strong> sgoing backwards. More precisely, we will set:θ(x, t) := sup { s ≤ t : y(x, t; s) ∈ R N \ Ω } . (2.43)By its definition, θ(x, t) represents the time at which the characteristic passing at(x, t) starts from the boundary. The representation <strong>for</strong>mula (written at (x, t)) maybe changed accordingly, obtainingu(x, t) =∫ tt 0∨θ(x,t)g(y(x, t; s), s)ds + u(y(x, t; t 0 ∨ θ(x, t)), t 0 ∨ θ(x, t)), (2.44)where a ∨ b = max(a, b), finite or −∞. Despite its <strong>for</strong>mal complexity, (2.44) hasa natural interpretation. If θ(x, t) < t 0 , that is, if the characteristic has run inthe interior of Ω during the time interval (t 0 , t), the solution u(x, t) has the usualrepresentation, i.e.u(x, t) =∫ tt 0g(y(x, t; s), s)ds + u(y(x, t; t 0 ), t 0 ).On the other h<strong>and</strong>, if θ(x, t) ≥ t 0 , it happens that the characteristic passing at (x, t)has started from the boundary, so the value which is propagated is the boundarycondition, computed at the intersection between the characteristic <strong>and</strong> the boundaryitself:u(x, t) =∫ tθ(x,t)g(y(x, t; s), s)ds + b(y(x, t; θ(x, t)), θ(x, t)).This generalized version of the representation <strong>for</strong>mula can be useful in assigning aDirichlet boundary condition, <strong>and</strong> in fact will be used in the numerical implementationof SL schemes.


✐✐38 Chapter 2. Viscosity Solutions of First Order PDEs2.4 Viscosity solutions <strong>and</strong> entropy solutionsIt is interesting to remark that there exists a link between viscosity solutions <strong>and</strong> entropysolutions which are the usual analytical tool in the framework of conservationlaws. Although this link is valid only in one space dimension <strong>and</strong> <strong>for</strong> a particularclass of <strong>Hamilton</strong>ians, it is the basis <strong>for</strong> some interesting remarks <strong>and</strong> also hasan important role in the construction of numerical schemes <strong>for</strong> <strong>Hamilton</strong>–Jacobiequations.Consider the two problems, an evolutive <strong>Hamilton</strong>–Jacobi equation{v t + H(v x ) = 0 (x, t) ∈ R × (0, T ),(2.45)v(x, 0) = v 0 (x) x ∈ R(where the <strong>Hamilton</strong>ian H is assumed to be convex) <strong>and</strong> the associated conservationlaw {u t + H(u) x = 0 (x, t) ∈ R × (0, T ),(2.46)u(x, 0) = u 0 (x) x ∈ R.Assume now thatv 0 (x) =∫ x−∞u 0 (ξ)dξ.Then, it turns out that this relationship between u <strong>and</strong> v is preserved also <strong>for</strong> t > 0,that is, if u is the entropy solution of (2.46), thenv(x, t) =∫ x−∞u(ξ, t)dξis the unique viscosity solution of (2.45). Viceversa, if v is the viscosity solution of(2.45), then u = v x is the unique entropy solution <strong>for</strong> (2.46).Note that v is a.e. differentiable, <strong>and</strong> the singular points <strong>for</strong> its derivativev x correspond to shocks <strong>for</strong> u. Note also that semiconcavity of v correspond to aone-sided (positive) bound on the derivative of u. This link will be also useful <strong>for</strong>numerical purposes.We briefly review the main results.Proposition 2.29. Let H ∈ C(R) be convex, <strong>and</strong> assume that v ∈ W 1,∞ (R×(0, T )is a solution of (2.45). Then, u := v x is a weak solution of (2.46).Proof. The proof can be found in [Li82], p.268.Theorem 2.30. Let H ∈ C 1 (R) be convex, v 0 ∈ W 1,∞ (R). If v ∈ W 1,∞ (R×(0, T ))is the unique viscosity solution of (2.45), then u := v x is the unique entropy solutionof (2.46).Proof. As we know, the viscosity solution v is the limit in L ∞ (R × (0, T )), asε → 0 + , of regular solutions v ε of the following problem{vt ε (x, t) + H(vx) ε = εvxx ε (x, t) ∈ R × (0, T )v ε (x, 0) = v 0 (x) x ∈ R,


✐✐2.4. Viscosity solutions <strong>and</strong> entropy solutions 39Hence we have, <strong>for</strong> any ϕ ∈ C ∞ (R × (0, T )),∫ T ∫∫ T ∫lim vx(x, ε t)ϕ(x, t) dx dt = − lim v ε (x, t)ϕ x (x, t) dx dt =ε→00 Rε→00 R∫ T ∫∫ T ∫= − lim v(x, t)ϕ(x, t) dx dt = v x (x, t)ϕ(x, t) dx dt.ε→00 R0 RObviously, the function u ε := vx ε solves the derived problem{u ε t (x, t) + H(u ε ) x = εu ε xx (x, t) ∈ R × (0, T )u ε (x, 0) = v 0x (x) x ∈ R.Since the sequence u ε converges in L 1 loc (R × (0, T )) <strong>for</strong> ε → 0+ , it must converge tothe entropy solution u of (2.46). This implies that, <strong>for</strong> any ϕ ∈ C ∞ (R × (0, T )),∫ T ∫∫ T ∫u ε (x, t)ϕ(x, t) dx dt = u(x, t)ϕ(x, t) dx dt.limε→0As a consequence,∫∫ T0<strong>and</strong> v x = u a.e. in R × (0, T )).0RRv x (x, t)ϕ(x, t) dx dt =∫ T00∫RRu(x, t)ϕ(x, t) dx dt,A converse of this result is also true. The first step is an intermediate result:Proposition 2.31. Let H ∈ C(R) be convex, <strong>and</strong> assume that u ∈ L ∞ loc(R × (0, T ))is a weak solution of (2.46). Definev(x, t) :=∫ xαu(ξ, t) dξ (2.47)<strong>for</strong> a fixed α ∈ R. Then, v ∈ W 1,∞loc(R × (0, T )) <strong>and</strong> v is a solution of (2.45) almosteverywhere.Proof. Since u ∈ L ∞ loc(R × (0, T )), there exists a set A ⊆ (0, T ) of zero Lebesguemeasure, such that <strong>for</strong> any t ∈ (0, T )\A, u is defined a.e. on R <strong>and</strong> u(·, t) ∈ L ∞ loc (R).Then, <strong>for</strong> such values of t, v(·, t) ∈ L ∞ loc(R). Morevover, <strong>for</strong> any t ∈ (0, T ) \ A <strong>and</strong>any ϕ ∈ C0 ∞ (R × (0, T )),∫v(x, t)ϕ x (x, t) dx =∫ [∫ xRR∫α= − u(x, t)ϕ(x, t) dt.R]u(ξ, t) dξ ϕ x (x, t) dξ =Thus, integrating on (0, T ), one has u = v x in the sense of distributions <strong>and</strong> a.e.Since u is a weak solution of (2.46), we have, <strong>for</strong> any ϕ ∈ C0 ∞ (R × (0, T )),∫ T ∫∫ T ∫H(u) ϕ x (x, t) dx dt = − u(x, t)ϕ t (x, t) dx dt =0= −R∫ T0∫Rv x (x, t)ϕ t (x, t) dx dt =0 R∫ T0∫Rv(x, t)ϕ tx (x, t) dx dt


✐✐40 Chapter 2. Viscosity Solutions of First Order PDEsSo there exists v t in the sense of distributions, <strong>and</strong> v t = −H(u) = −H(v x ). There<strong>for</strong>e,v ∈ W 1,∞loc(R × (0, T )), <strong>and</strong> v is a solution almost everywhere of (2.45).Now, <strong>for</strong> any u ∈ C([0, T ]; L 1 (R)) we definev(x, t) :=∫ x−∞u(ξ, t) dξ . (2.48)Clearly, <strong>for</strong> any t ∈ (0, T ), the function v(·, t) is absolutely continuous <strong>and</strong> v x = ua.e.. We can now prove the main result.Theorem 2.32. Let H ∈ C 1 (R) <strong>and</strong> u 0 ∈ L ∞ (R) ∩ L 1 (R). Assume that u ∈L ∞ (R × (0, T )) ∩ C([0, T ]; L 1 (R)) is the unique entropy solution of (2.46). Then,the function v given by (2.48) is the unique viscosity solution of (2.45) <strong>for</strong> the initialconditionv 0 (x) :=∫ x−∞u 0 (ξ) dξ.Proof. Since u ∈ C([0, T ]; L 1 (R)), we have that v ∈ L ∞ (R × [0, T ]). As in theprevious proposition, it is easy to show that v ∈ W 1,∞ (R × (0, T )) <strong>and</strong> that v is asolution a.e. of (2.45). Moreover,∫ xlim |v(x, t) − v 0(x)| ≤ lim |u(ξ, t) − u 0 (ξ)| dξ = 0.t→0 t→0−∞Now, suppose that v is not the viscosity solution of (2.45), <strong>and</strong> denote by v theunique viscosity solution. Then, by Theorem 2.30, v x is the unique entropy solutionof (2.46) <strong>and</strong> there<strong>for</strong>e, <strong>for</strong> any ϕ ∈ C ∞ 0 (R × (0, T )),∫ ∫(v − v)ϕ x dx dt = 0.Since ϕ can be arbitrarily chosen, the conclusion follows.2.5 Discontinuous viscosity solutionsWe conclude this chapter giving some ideas about the extension of the “classical”theory of viscosity solutions to the discontinuous case. This extension has beenstrongly motivated by a number of applications, e.g. to image processing <strong>and</strong> togames, where it is natural to accept discontinuous solutions.Let w : R d → R be a bounded function. We recall two definitions which willplay a key role in defining a bounded, but possibly discontinuous viscosity solution.The lower semi-continuous envelope of w is defined asw(x) := lim inf w(y),y→xwhereas the upper semi-continuous envelope asw(x) := lim sup w(y).y→x


✐✐2.6. Commented references 41With this additional tool, we can extend the notion of viscosity solutions as follows.Definition 2.33. A function u ∈ L ∞ (Ω) is a viscosity solution in Ω ofH(x, u, Du) = 0if <strong>and</strong> only if, <strong>for</strong> any ϕ ∈ C 1 (Ω) the following conditions hold:(i) at every point x 0 ∈ Ω, local maximum <strong>for</strong> u − ϕ,H(x 0 , u(x 0 ), Dϕ(x 0 )) ≤ 0(i.e., u is a viscosity subsolution);(ii) at every point x 0 ∈ Ω, local minimum <strong>for</strong> u − ϕ,i.e., u is a viscosity supersolution.H(x 0 , u(x 0 ), Dϕ(x 0 )) ≥ 0It is important to note that the extension of the comparison principle to thisnew settings is not trivial. However, in many relevant problem one can get a uniquenessresult <strong>for</strong> lower semicontinuous viscosity solution. A comprehensive introductionto these results goes beyond the scopes of this book <strong>and</strong> can be found in [Ba93].2.6 Commented referencesThe theory of viscosity solutions has started with the papers by Cr<strong>and</strong>all <strong>and</strong> Lions[CL84] on the solution of first order <strong>Hamilton</strong>–Jacobi equations. However, a veryimportant source of inspiration <strong>for</strong> this theory had come from some earlier papers byKružkov [Kr66], in which the solution <strong>for</strong> first order problems was obtained via thevanishing viscosity methods (i.e., by elliptic/parabolic regularization). These paperswere strongly related to the analysis of conservation laws, where Kružkov gaveimportant contributions (see the monographs by Leveque [L92] <strong>and</strong> Serre [Se99a,Se99b] <strong>for</strong> more in<strong>for</strong>mations on this topic). The main contribution by Cr<strong>and</strong>all<strong>and</strong> Lions was to find an intrinsic definition of the weak solution which would notrequire to pass to the limit in the regularized problem <strong>and</strong> could allow to obtainuniqueness results <strong>for</strong> a large class of stationary <strong>and</strong> evolutive nonlinear partialdifferential equations.At the very beginning the definition was based on sub- <strong>and</strong> superdifferentials<strong>and</strong> only at a later time the definition based on test functions was introduced([CEL84], see also [Is89]). Starting from these seminal papers the theory had anincreasing success in the following years, as witnessed by the books by Lions [Li82]<strong>and</strong> Barles [Ba98]. In particular, the latter contains a chapter on discontinuousviscosity solutions. More details on discontinuous viscosity solutions can be foundin the original papers by Barron <strong>and</strong> Jensen [BJ91], Frankowska [Fr93], Barles[Ba93], <strong>and</strong> Soravia [Sor93b].The theory has naturally been extended to second order fully nonlinear problems,but the presentation of these results goes beyond the scopes of this book. Forthese extensions of the theory, we refer the reader to the survey paper by Cr<strong>and</strong>all,Ishii <strong>and</strong> Lions [CIL92] <strong>and</strong> to the book by Fleming <strong>and</strong> Soner [FS93].


✐✐42 Chapter 2. Viscosity Solutions of First Order PDEsFinally, it is worth to say that the development of this theory has been largelydriven by the applications to control problems, image processing <strong>and</strong> fluid dynamics.For the applications to control problems we refer to the monographs by Bardi<strong>and</strong> Capuzzo Dolcetta [BCD97] <strong>for</strong> deterministic problems <strong>and</strong> Fleming <strong>and</strong> Soner[FS93] <strong>for</strong> stochastic problems. Moreover, in the books by Sethian [Se96] <strong>and</strong> Osher<strong>and</strong> Fedkiw [OF03] it is possible to find find a number of applications of <strong>Hamilton</strong>–Jacobi equations, in particular in the framework of level set methods.


✐✐Chapter 3Elementary buildingblocksSemi-Lagrangian (SL) schemes require an assembly of different (<strong>and</strong> partly independent)conceptual blocks, in particular a strategy to move along characteristics <strong>and</strong>a technique to reconstruct the numerical solution at the foot of a characteristic. Inthe case of HJ equations, a numerical minimization technique is also needed. Here,different choices <strong>for</strong> all such blocks are considered <strong>and</strong> explained, <strong>and</strong> their basictheoretical results are reviewed, with a greater emphasis on less established topics.3.1 A review of ODE approximation schemesThe first typical operations per<strong>for</strong>med by Semi-Lagrangian schemes is to move alongcharacteristics to locate the value to be propagated. This operation amounts toapproximate over a single time step a system of Ordinary Differential Equationslike (1.7), which will be rewritten as⎧⎨ ddt y(x 0, t 0 ; t) = f(y(x 0 , t 0 ; t), t),⎩y(x 0 , t 0 ; t 0 ) = x 0 ,t ∈ R(3.1)with an initial condition x 0 ∈ R d given at the time t 0 , <strong>and</strong> with a vector fieldf : R d × R → R d globally Lipschitz continuous with respect to its first argument.Such a task can be accomplished by different techniques, usually borrowed fromthe ODE literature. This section will review the basic ideas, which will be laterapplied to the specific situations of interest, whereas <strong>for</strong> a more extensive treatmentof the related techniques <strong>and</strong> theoretical results, we refer to specialized textbooks.Throughout the section, we will assume that a set of nodes t k = t 0 + k∆t (withk ∈ Z) is given on the time axis, with y k denoting the corresponding approximationsof y(x 0 , t 0 ; t k ).The usual concepts of consistency <strong>and</strong> stability are involved in the convergencetheory of methods <strong>for</strong> ODEs, as the basic convergence theorem states. Conceptually,consistency means that the scheme is a “small perturbation” of the exact equation,whereas stability means that the scheme satisfies (uni<strong>for</strong>mly in ∆t) a principle ofcontinuous dependence upon the initial data. Note that, among the various conceptof stability which have been developed <strong>for</strong> this problem, we will refer here to zerostability,which is the one required <strong>for</strong> convergence.43


✐✐44 Chapter 3. Elementary building blocksRather than outlining a general theory, we will sketch how this theory appliesto the two main classes of schemes to approximate (3.1), which are referred to asone-step <strong>and</strong> multistep methods. We will briefly review their general philosophy,as well as the main results of consistency, stability <strong>and</strong> convergence, <strong>and</strong> give somepractical examples, including cases of use in SL schemes.Remark 3.1. While practical schemes <strong>for</strong> (3.1) are necessarily stable, we will seethat, in the convergence theory of SL schemes, stability is not technically required.At the level of convergence, the choice of a particular scheme rather results in adifferent consistency error, although it may also affect the qualitative behaviour ofthe solution.3.1.1 One-step schemesIn one-step schemes the system (3.1) is discretized in the <strong>for</strong>m{y k+1 = y k + ∆tΦ(∆t; y k , t k , y k+1 )y 0 (x 0 , t 0 ) = x 0 .(3.2)In (3.2), the approximation y k+1 is constructed using only the in<strong>for</strong>mations availableat the time step t k . If the function Φ has a genuine dependence on y k+1 , then thescheme is termed as implicit <strong>and</strong> the computation of y k+1 itself requires to solve(3.2) as a system of nonlinear equations (or a scalar equation if the solution isscalar).ExamplesWe show in Table 3.1 some of the simplest examples of explicit <strong>and</strong> implicit onestepschemes of first <strong>and</strong> second order, namely Forward Euler (FE), Backward Euler(BE), Heun (H) <strong>and</strong> Crank–Nicolson (CN). More in general, this class of schemesincludes all Runge–Kutta type methods. Note that, although their use is clear whenthe dynamics f(·, ·) has an explicit expression, care should be taken when applyingthem to step back along characteristics, especially in nonlinear cases, since thein<strong>for</strong>mation might neither be available at any time step, nor at any point. Thisissue will be discussed in Chapter 5.Theoretical resultsConsistency The scheme (3.2) is said to be consistent with order p ≥ 1 if, oncereplaced y k , y k+1 with the exact values y(t k ) = y(x 0 , t 0 ; t k ), y(t k+1 ) = y(x 0 , t 0 ; t k+1 )of a smooth solution, the scheme satisfiesy(t k+1 ) = y(t k ) + ∆tΦ(∆t; y(t k ), t k , y(t k+1 )) + O(∆t p+1 ). (3.3)In general, we expect that (3.3) would result in obtaining an approximation errorof order O(∆t p ), <strong>and</strong> this is also true when using (3.2) <strong>for</strong> approximating characteristicsin SL schemes (in this case, this consistency error affects the term relatedto time discretization alone).


✐✐3.1. A review of ODE approximation schemes 45scheme <strong>for</strong>m orderFE Φ(∆t; y k , t k , y k+1 ) = f(y k , t k ) p = 1BE Φ(∆t; y k , t k , y k+1 ) = f(y k+1 , t k + ∆t) p = 1H Φ(∆t; y k , t k , y k+1 ) = 1 2 [f(y k, t k ) + f(y k + ∆tf(y k , t k ), t k + ∆t)] p = 2CN Φ(∆t; y k , t k , y k+1 ) = 1 2 [f(y k, t k ) + f(y k+1 , t k + ∆t)] p = 2Table 3.1. First- <strong>and</strong> second-order one-step schemes <strong>for</strong> the system (3.1)Stability The definition of zero-stability in the case of one-step schemes requiresthat, if (3.2) is used to approximate (3.1) starting from two different initial conditionsx 0 <strong>and</strong> x ′ 0 (with the corresponding numerical solutions denoted by y k <strong>and</strong> yk ′ ),then‖y k − y k‖ ′ ≤ C‖x 0 − x ′ 0‖<strong>for</strong> any k such that t k ∈ [t 0 , t 0 + T ], <strong>and</strong> with a constant C independent of ∆t.A general result states that if the function Φ is Lipschitz continuous withrespect to its second <strong>and</strong> fourth argument, the one-step scheme is zero-stable. Notethat all the examples of Table 3.1 satisfy this condition.ConvergenceFinally, we give the convergence theorem <strong>for</strong> one-step schemes.Theorem 3.2. If the scheme (3.2) is consistent <strong>and</strong> zero-stable, then, <strong>for</strong> any ksuch that t k ∈ [t 0 , t 0 + T ],‖y k − y(x 0 , t 0 ; t k )‖ → 0as ∆t → 0. Moreover, if (3.2) is consistent with order p <strong>and</strong> the solution y issmooth enough, then‖y k − y(x 0 , t 0 ; t k )‖ ≤ C∆t p .3.1.2 Multistep schemes<strong>Linear</strong> multistep schemes discretize the system (3.1) in the <strong>for</strong>m⎧⎪⎨ y k+1 = ∑ n s−1j=0 α jy k−j + ∆t ∑ n s−1j=−1 β jf(y k−j , t k−j )y 0 = x 0⎪⎩y 1 , . . . , y ns−1 given(3.4)In (3.4), the approximation y k+1 is constructed using the in<strong>for</strong>mations available atthe n s time steps t k−ns+1 to t k+1 . The startup values y 1 , . . . , y ns−1 are chosen so


✐✐46 Chapter 3. Elementary building blocksscheme α 0 α 1 β 0 β 1 orderMP 0 1 2 0 p = 2AB2 1 032− 1 2p = 2Table 3.2. Examples of multistep schemes <strong>for</strong> the system (3.1)as to be an approximation (with suitable accuracy) of the corresponding values ofthe exact solution.As <strong>for</strong> the case of one-step schemes, if β −1 ≠ 0 so that the right-h<strong>and</strong> side of (3.4)depends on y k+1 , then the computation of y k+1 requires to solve (3.4) as a systemof nonlinear equations.ExamplesIn table 3.2 we report two specific choices of the coefficients, both of which resultin a method of order p = 2, <strong>and</strong> corresponds to the midpoint scheme (MP) <strong>and</strong>to the second-order Adams–Bash<strong>for</strong>th scheme (AB2). Such cases turn out to be ofcommon use in SL schemes <strong>for</strong> environmental fluid dynamics.Theoretical resultsConsistency Consistency is checked in multistep schemes according to the sameprinciple used in one-step schemes, namely to replace the values y k with the exactvalues y(t k ) = y(x 0 , t 0 ; t k ) of a smooth solution. There<strong>for</strong>e, the multistep scheme(3.4) is said to be consistent with order p ≥ 1 ify(t k+1 ) =n∑s−1j=0n∑s−1α j y(t k−j ) + ∆tj=−1β j f(y(t k−j ), t k−j ) + O(∆t p+1 ). (3.5)It is possible to derive algebraic conditions on the coefficients α j , β j which ensurethat (3.5) is satisfied. However, it is also possible to avoid these technical aspects,<strong>and</strong> rather work directly on the <strong>for</strong>mulation (3.5).Stability Although similar to the definition adopted <strong>for</strong> one-step schemes, zerostabilityof multistep schemes is defined in a way which takes into account not onlythe two initial points y 0 = x 0 <strong>and</strong> y 0 ′ = x ′ 0, but all the startup values y 1 , . . . , y ns−1<strong>and</strong> y 1, ′ . . . , y n ′ s−1. We require there<strong>for</strong>e that, <strong>for</strong> any k such that t k ∈ [t 0 , t 0 + T ],‖y k − y k‖ ′ ≤ C max (‖y i − y ′0≤i≤n s−1i‖)with a constant C independent of ∆t.The endpoint of zero-stability analysis <strong>for</strong> linear multistep schemes is the socalledroot condition, which gives a necessary <strong>and</strong> sufficient condition in algebraic<strong>for</strong>m.


✐✐3.2. Reconstruction techniques in one <strong>and</strong> multiple space dimensions 47Theorem 3.3. A multistep method in the <strong>for</strong>m (3.4) is zero–stable if <strong>and</strong> only if,once denoted by ζ i (i = 1, . . . , n s ) the roots of the polynomialn∑s−1r(ζ) = ζ ns − α j ζ ns−j−1 ,j=0one has |ζ i | ≤ 1 <strong>for</strong> any i, <strong>and</strong> in addition, all roots such that |ζ i | = 1 are simple.Convergence The convergence theorem <strong>for</strong> multistep schemes, besides the st<strong>and</strong>ardassumptions of consistency <strong>and</strong> stability, requires convergence of the startupvalues to the corresponding values <strong>for</strong> the solution.Theorem 3.4. If the scheme (3.4) is consistent <strong>and</strong> zero-stable, <strong>and</strong> if, <strong>for</strong> 1 ≤k ≤ n s − 1,‖y k − y(x 0 , t 0 ; t k )‖ → 0,then, <strong>for</strong> any k such that t k ∈ [t 0 , t 0 + T ],‖y k − y(x 0 , t 0 ; t k )‖ → 0as ∆t → 0. Moreover, if (3.4) is consistent with order p, ‖y k − y(x 0 , t 0 ; t k )‖ =O(∆t p ) <strong>for</strong> 1 ≤ k ≤ n s − 1 <strong>and</strong> the solution y is smooth enough, then‖y k − y(x 0 , t 0 ; t k )‖ ≤ C∆t p .3.2 Reconstruction techniques in one <strong>and</strong> multiplespace dimensionsThe role of the reconstruction step in SL schemes is to recover the value of thenumerical solution at the feet of characteristics, which are not in general grid pointsthemselves. In their basic version, SL schemes do not use cell averages, <strong>and</strong> the reconstructionis rather based on the interpolation of pointwise values of the solution.We assume there<strong>for</strong>e that a grid (not necessarily structured) with nodes x j isset, <strong>and</strong> that a function v(x) is represented by the vector V of its correspondingnodal values v j . The space discretization parameter ∆x, in the basic case of a singlespace dimension with sequentially numbered nodes, is defined as∆x = sup (x k+1 − x k ), (3.6)kwhereas in multiple dimensions <strong>and</strong> unstructured cases, it requires more complexdefinitions which will be recalled when necessary.In the simplest cases (symmetric Lagrange interpolation, Lagrange finite elements),the reconstruction is set in the <strong>for</strong>m of a linear combination of suitablebasis functions:I[V ](x) = ∑ v k ψ k (x). (3.7)kWe will mostly consider functions ψ k in piecewise polynomial <strong>for</strong>m of degree r, <strong>and</strong> ifnecessary, the notation I r [V ] will be used to state more explicitly this degree. Also,the ψ k are usually chosen to be cardinal functions, that is, to satisfy ψ k (x i ) = δ ik .


✐✐48 Chapter 3. Elementary building blocksWhen non-oscillatory reconstructions are considered, the linearity with respect tothe values v k is lost <strong>and</strong> the reconstruction takes a more complex <strong>for</strong>m.Whenever useful, we will apply in the sequel the idea of reference basis functions,meaning a basis of cardinal functions which is defined on a reference grid (e.g.,a grid with unity distance between nodes), <strong>and</strong> generates the actual basis of thereconstruction by an affine trans<strong>for</strong>mation depending on the geometry of the specificgrid used.In the sequel of the section, we will review three reconstruction strategies(symmetric Lagrange interpolation, ENO, WENO) which are usually implementedwith constant step, <strong>and</strong> extended to multiple dimension by separation of variables.Moreover, a genuinely multidimensional <strong>and</strong> unstructured strategy (finite elements)will also be considered.3.2.1 Symmetric Lagrange interpolation in R 1This kind of interpolation is usually applied with evenly spaced nodes (by (3.6), ∆xwill denote the constant step), <strong>and</strong> the samples of V used <strong>for</strong> the reconstructionare taken from a stencil surrounding the interval which contains the point x, thisresulting in a piecewise polynomial interpolation. Note that, in principle, Lagrangeinterpolation can be implemented on nonuni<strong>for</strong>m grids as well as in globally polynomial<strong>for</strong>m. In the SL framework, however, the <strong>for</strong>mer choice would lead in generalto an unstable scheme, the latter to an undue numerical dispersion in the <strong>for</strong>m ofGibb’s oscillations. Lagrange interpolation is typically implemented with an odddegree, thus using an equal number of nodes on both sides of x. In fact, with sucha choice, the interpolation error is related to an even derivative of the function, thisresulting in a symmetric behaviour of the SL scheme around singularities, as it willturn out from the numerical dispersion analysis.ConstructionTaking into account that the number of nodes must be one unity larger then thedegree r of the interpolation, we obtain, <strong>for</strong> x ∈ [x l , x l+1 ], that the value I[V ](x) iscomputed by a Lagrange polynomial constructed on a set of nodes which includesr+12nodes on each side of the interval. This stencil of nodes will be denoted byS ={x k : l − r − 1 ≤ k ≤ l + r + 122<strong>and</strong> is depicted in Figure 3.1 <strong>for</strong> the lowest odd orders of interpolation. As aconsequence, the Lagrange polynomial reconstruction takes the <strong>for</strong>mI r [V ](x) = ∑ ∏ x − x iv k. (3.8)x k − x ix k ∈Sx i∈S\{x k }Since the Lagrange basis used to interpolate changes as x moves from oneinterval to the other, at a first glance is could seem that (3.7) would not be satisfiedwith a unique basis of cardinal functions. Nevertheless, the reconstruction beinglinear with respect to the values to be interpolated, such a unique basis can still besuitably defined. In fact, the interpolation operator I[·] is a linear map from thespace l ∞ of bounded sequences into the space of continuous functions on R:},I[·] : l ∞ → C 0 (R) (3.9)


✐✐3.2. Reconstruction techniques in one <strong>and</strong> multiple space dimensions 49Figure 3.1. Reconstruction stencil <strong>for</strong> linear, cubic, quintic Lagrange interpolationso that, by elementary linear algebra arguments, any basis function ψ k can bedefined to be nothing but the image of the base element e k (that is, the interpolationof a sequence such that v k = 1, v i = 0 <strong>for</strong> i ≠ k).On the other h<strong>and</strong>, since the grid is uni<strong>for</strong>m <strong>and</strong> the reconstruction is self-similar<strong>and</strong> invariant by translation, any such basis function may be written as( ) x − xkψ k (x) = ψ . (3.10)∆xHere, the function ψ plays the role of reference basis function <strong>and</strong> may be obtainedby applying the interpolation operator to the sequence e 0 , with ∆x = 1. In thesequel, y will represent the variable in the reference space, <strong>and</strong> we will possiblyuse the notation ψ [r] to denote the reference function <strong>for</strong> a Lagrange interpolationof order r. Although the interpolation procedure, as per<strong>for</strong>med by (3.8), does notrequire to know the explicit expression of ψ, this expression will be useful later <strong>for</strong>the theoretical analysis of the schemes.ExamplesWe sketch below the construction of the Lagrange reconstruction, as well as of thecorresponding reference basis function corresponding to the cases (r = 1, 3, 5) shownabove.<strong>Linear</strong> (P 1 ) interpolation We first show the simple situation of linear interpolation,which also corresponds to the P 1 case in the finite element setting. Theexpression of the interpolating polynomial <strong>for</strong> the first order (r = 1) isI 1 [V ](x) = x − x l+1v l +x − x lv l+1 =x l − x l+1 x l+1 − x l= x l+1 − x∆xv l + x − x l∆x v l+1, (3.11)


✐✐50 Chapter 3. Elementary building blockswhereas the interpolation of the sequence e 0 in the reference space results in areference function of the well-known structure⎧⎪⎨ 1 + y if − 1 ≤ y ≤ 0ψ [1] (y) = 1 − y if 0 ≤ y ≤ 1.(3.12)⎪⎩0 elsewhereNote that the value at a certain node affects the reconstruction only in the adjacentintervals.Lagrange interpolation of higher order Take now as a second example cubicreconstruction, which per<strong>for</strong>ms an interpolation using the four nearest nodes (twoon the left <strong>and</strong> two on the right of the interval containing x). The interpolatingpolynomial is computed asI 3 [V ](x) =∑l+2k=l−1v k∏i≠kx − x ix k − x i== − (x − x l)(x − x l+1 )(x − x l+2 )6∆x 3 v l−1 ++ (x − x l−1)(x − x l+1 )(x − x l+2 )2∆x 3v l −− (x − x l−1)(x − x l )(x − x l+2 )2∆x 3 v l+1 ++ (x − x l−1)(x − x l )(x − x l+1 )6∆x 3 v l+2(note that the value at a node x j affects the reconstruction in the interval (x j−2 , x j+2 )).Then, referring again to the case in which e 0 is interpolated <strong>and</strong> ∆x = 1, the explicit<strong>for</strong>m of the reference base function ψ [3] is⎧12(y + 1)(y − 1)(y − 2) if 0 ≤ y ≤ 1⎪⎨ψ [3] − 1(y) =6(y − 1)(y − 2)(y − 3) if 1 ≤ y ≤ 2.0 if y > 2⎪⎩ψ [3] (−y) if y < 0.(3.13)In turn, a similar computation <strong>for</strong> the quintic interpolation would yeld <strong>for</strong> theinterpolant the expression:I 5 [V ](x) =∑l+3k=l−2v k∏i≠kx − x ix k − x i== − (x − x l−1)(x − x l )(x − x l+1 )(x − x l+2 )(x − x l+3 )120∆x 5 v l−2 ++ (x − x l−2)(x − x l )(x − x l+1 )(x − x l+2 )(x − x l+3 )24∆x 5v l−1 −− (x − x l−2)(x − x l−1 )(x − x l+1 )(x − x l+2 )(x − x l+3 )12∆x 5 v l + · · · ++ (x − x l−2)(x − x l−1 )(x − x l )(x − x l+1 )(x − x l+2 )120∆x 5 v l+3 .


✐✐3.2. Reconstruction techniques in one <strong>and</strong> multiple space dimensions 51Figure 3.2. The basis functions <strong>for</strong> linear, cubic, <strong>and</strong> quintic interpolationAccordingly, the reference basis function would be in the <strong>for</strong>m:⎧− 1 12(y + 2)(y + 1)(y − 1)(y − 2)(y − 3) if 0 ≤ y ≤ 11⎪⎨ψ [5] 24(y + 1)(y − 1)(y − 2)(y − 3)(y − 4) if 1 ≤ y ≤ 2(y) = − 1120(y − 1)(y − 2)(y − 3)(y − 4)(y − 5) if 2 ≤ y ≤ 30 if y > 3⎪⎩ψ [5] (−y) if y < 0.(3.14)More in general, <strong>for</strong> an arbitrary odd interpolation degree r written in the <strong>for</strong>m(3.8), the reference basis function ψ [r] is⎧⎪⎨ψ [r] (y) =⎪⎩[r/2]+1∏k≠0,k=−[r/2].r∏ y − k−kk=1y − k−kif 0 ≤ y ≤ 1if [r/2] ≤ y ≤ [r/2] + 10 if y > [r/2] + 1ψ [r] (−y) if y < 0.(3.15)Note that this construction can also be per<strong>for</strong>med <strong>for</strong> an even degree of interpolation,but the result would not be a symmetric function.We show in figure 3.2 the reference basis functions <strong>for</strong> interpolation, ψ [r] (y),<strong>for</strong> r = 1, 3, 5.


✐✐52 Chapter 3. Elementary building blocksTheoretical resultsWe recall here the approximation result concerning the interpolation error <strong>for</strong> Lagrangepolynomial approximations. The proof is usually contained in any textbookin basic Numerical Analysis.Theorem 3.5. Let the interpolation I r [V ] be defined by (3.7), with ψ j defined by(3.10) <strong>and</strong> ψ defined by (3.15). Let v(x) be a uni<strong>for</strong>mly continuous function on R,<strong>and</strong> V be the vector of its nodal values. Then, <strong>for</strong> ∆x → 0,If moreover v ∈ W s,∞ (R), then‖v − I r [V ]‖ ∞ → 0. (3.16)‖v − I r [V ]‖ ∞ ≤ C∆x min(s,r+1) (3.17)<strong>for</strong> some positive constant C depending on the degree of regularity s.Remark 3.6. As we mentioned, globally polynomial interpolations show Gibb’soscillations when treating solutions with singularities, <strong>and</strong> there<strong>for</strong>e they are usuallyavoided in nonsmooth contexts (like viscosity or entropy solutions), whereaspiecewise polynomial approximations (e.g., Lagrange <strong>and</strong> finite element) have a betterper<strong>for</strong>mance. However, although confined to a neighbourhood of the singularity,high-order Lagrange reconstructions can still exhibit an oscillatory behaviour, due tothe changes of sign in the basis functions (see Figure 3.2). This motivates the search<strong>for</strong> non-oscillatory (ENO/WENO) reconstruction strategies which could reduce thisdrawback.3.2.2 Essentially Non-Oscillatory interpolation in R 1Essentially Non-Oscillatory (ENO) interpolation is a first strategy of reconstructionwhich has been developed in order to reduce Gibb’s oscillations. ENO interpolationcuts essentially such oscillations (that is, reduces over- or undershoots of thereconstruction to the order of some power of ∆x), still retaining a high order ofaccuracy in smooth regions of the domain. The basic tool <strong>for</strong> this operation is astrategy which selects, among all c<strong>and</strong>idate stencils, the stencil where the functionis smoother. In this procedure, different functions may require different stencils, sothat linearity of the reconstruction (expressed by (3.7)) is lost.ConstructionWe define the procedure to construct a reconstruction of ENO type based on theone-dimensional mesh {x j }, in order to approximate the value v(x) <strong>for</strong> x ∈ [x l , x l+1 ].Most of what follows needs not a uni<strong>for</strong>m mesh, although this is the typical situation.In any case, ∆x will be defined by (3.6).We have to construct an interpolating polynomial of degree r, with a stencilof r + 1 points, which must include x l <strong>and</strong> x l+1 <strong>and</strong> be chosen in order to avoidsingularities of v. This will be per<strong>for</strong>med using Newton’s <strong>for</strong>m of the interpolatingpolynomial,I r [V ](x) = V [x l0 ] +r∑k−1∏V [x l0 , . . . , x lk ] (x − x lm ) (3.18)k=1m=0


✐✐3.2. Reconstruction techniques in one <strong>and</strong> multiple space dimensions 53with the divided differences V [·] defined asV [x l0 ] = v(x l0 ),V [x l0 , . . . , x lk ] = V [x l 1, . . . , x lk ] − V [x l0 , . . . , x lk−1 ]x lk − x l0.Note that, in order to reconstruct v in the interval [x l , x l+1 ], we set x l0 = x l ,x l1 = x l+1 <strong>and</strong> proceed to extend the stencil by adding further adjacent points.Now, from the the basic results concerning divided differences, it is knownthat, if v ∈ C s (I) (I being the smallest interval containing x l0 , . . . , x ls ), then thereexists a point ξ ∈ I such thatV [x l0 , . . . , x ls ] = v(s) (ξ). (3.19)s!On the other h<strong>and</strong>, if v has a discontinuity in I, then( ) 1V [x l0 , . . . , x ls ] = O∆x s ,<strong>and</strong> more in general, if the discontinuity occurs at the level of the k–th derivativev (k) , with k < s, then( ) 1V [x l0 , . . . , x ls ] = O∆x s−k . (3.20)Hence, as long as it affects the construction of the interpolating polynomial, themagnitude of a divided difference is a measure of the regularity of the function onthe related stencil.Let there<strong>for</strong>e the fixed initial stencil be denoted byS 1 = {x l , x l+1 } = {x l0 , x l1 }. (3.21)At the k–th step of the procedure, we pass from a stencil S k−1 to the stencilS k = {x l0 , . . . , x lk } (3.22)by adding the point x lk . Then, if we plan to extend the stencil towards pointswhere v(x) is smoother, x lk will be chosen (between the two possible choices) as theone which minimizes the magnitude of the new divided difference. More <strong>for</strong>mally,settingl − k = min(l 0, . . . , l k−1 ) − 1, l + k = max(l 0, . . . , l k−1 ) + 1,then:• If |V [x l0 , . . . , x lk−1 , x l−]| < |V [x l0 , . . . , x lk−1 , xkl+ ]|, then l k = l − k kextended one point to the left)(the stencil is• If |V [x l0 , . . . , x lk−1 , x l−]| ≥ |V [x l0 , . . . , x lk−1 , xkl+ ]|, then l k = l + k kextended one point to the right)(the stencil isAfter the selection, the stencil is updated <strong>and</strong> the algorithm proceeds to thenext step, until k = r. The interpolation may be computed by adding each term ofthe Newton polynomial during the process, or by per<strong>for</strong>ming a Lagrange interpolationon the final stencil S = S r .


✐✐54 Chapter 3. Elementary building blocksTheoretical resultsWe review some basic results about ENO interpolation.Accuracy in smooth regions First, is is obvious that if the function to be interpolatedis smooth, then the accuracy coincides with the accuracy of Lagrangeinterpolation:Theorem 3.7. Let the interpolation I r [V ] be defined by (3.18), with the ENOprocedure of selection <strong>for</strong> the V [x l0 , . . . , x lk ]. Let v ∈ W s,∞ (R). Then,<strong>for</strong> some positive constant C.‖v − I r [V ]‖ ∞ ≤ C∆x min(s,r+1) (3.23)Bounds on the Total Variation The second property is related to the total variationof the reconstruction. We start with a preliminary result, which does notrequire the ENO procedure <strong>for</strong> computing I r , but is suitable <strong>for</strong> any piecewise polynomialinterpolation.Proposition 3.8. Let v be discontinuous in the interval [x l , x l+1 ]. Then, I r [V ] hasno extrema in the same interval.Proof. We only give a sketch of the proof, based on the case of v(x) defined as astep function:{0 x ≤ 0v(x) =1 x > 0.Let p ∈ P r be the polynomial interpolating v on the stencil S = S r , <strong>and</strong> assume thediscontinuity point x = 0 satisfiesx l < 0 < x l+1with x l , x l+1 ∈ S r . On any interval [x j , x j+1 ] not containing the discontinuity pointwe havep(x j ) = u(x j ) = u(x j+1 ) = p(x j+1 ),<strong>and</strong> there<strong>for</strong>e there exists a point ξ j ∈ (x j , x j+1 ) such thatp ′ (ξ j ) = 0.Then, we can find r − 1 different roots <strong>for</strong> p ′ (x), that is, one root <strong>for</strong> any interval(x j , x j+1 ) not containing the point x = 0. Since p ′ ∈ P r−1 , it follows that p ′ (x)cannot have further roots in the shock interval [x l , x l+1 ], that is, p(x) is monotonein this interval.Last, we give the result on the total variation of ENO interpolation. Basically,this results states that ENO interpolation is TVB, that is, it has bounded totalvariation. In addition, the increase in variation introduced is of the order of theinterpolation error in smooth regions.


✐✐3.2. Reconstruction techniques in one <strong>and</strong> multiple space dimensions 55Theorem 3.9. Let v(x) be a piecewise continuous function with derivative v (s)bounded in the continuity set. Then, there exists a function z(x) such that<strong>and</strong> satisfyingz(x) = I r [V ](x) + O(∆x min(s,r+1) ),T V (z) ≤ T V (v).Proof. The statement directly follows from the two last results, taking z(x) = v(x)in intervals where v is smooth, <strong>and</strong> z(x) = p(x) in interval containing discontinuities.3.2.3 Weighted Essentially Non-Oscillatory interpolation in R 1WENO interpolation is another strategy <strong>for</strong> non-oscillatory reconstruction. It isderived from the ENO strategy, but unlike ENO it uses all the computed in<strong>for</strong>mationas much as possible.In practice, while the ENO strategy computes twice the required number ofdivided differences to build a Newton polynomial, but at each step one differenceamong two is discarded, WENO strategy computes all the c<strong>and</strong>idate interpolatingpolynomials. Then, if the function is smooth enough on the given stencil, c<strong>and</strong>idatepolynomials are combined in order to have a further increase of the accuracy. If thefunction is nonsmooth, then the reconstruction selects only the smoothest c<strong>and</strong>idateapproximation.Although the construction of a WENO interpolation is a well-established matter,we will review here the details of the procedure, along with the main theoreticalresults, since it is usually treated <strong>and</strong> used in a different framework. In this part,we do not necessarily assume a uni<strong>for</strong>m mesh spacing.ConstructionGeneral structure To construct a WENO interpolation of degree r = 2n − 1 onthe interval [x l , x l+1 ], we start from the Lagrange polynomial built on the stencilS = {x l−n+1 , . . . , x l+n }, <strong>and</strong> written in the <strong>for</strong>mn∑Q(x) = C k (x)P k (x) (3.24)k=1where the “linear weights” C k are polynomials of degree n − 1 <strong>and</strong> the P k arepolynomials of degree n interpolating V on the stencil S k = {x l−n+k , . . . , x l+k },k = 1, . . . , n (note that all the stencils S k overlap on the interval [x l , x l+1 ], <strong>and</strong> thatthe dependence on l has been dropped in this general expression). The nonlinearweights are then constructed so as to obtain an approximation of the highest degreeif suitable smoothness indicators β k give the same result on all stencils. This leadsto defineα k (x) =C k(x)(β k + ε) 2 (3.25)(with ε a properly small parameter, usually of the order of 10 −6 ), <strong>and</strong> then thenonlinear weights asw k (x) =α k(x)∑h α h(x) . (3.26)


✐✐56 Chapter 3. Elementary building blocksThe final <strong>for</strong>m of WENO interpolation is thenI r [V ](x) =n∑w k (x)P k (x), (3.27)k=1where w k is defined by (3.25)–(3.26).The general concept of (3.27) is the following. If all the smoothness indicatorsβ k would have the same value, then w k (x) = C k (x) <strong>and</strong> the reconstruction is givenby the Lagrange polynomial (3.24), which has the highest possible degree. If someof the smoothness indicators has a “large” value with respect to the others (thismeaning that the partial stencil contains a singularity), then w k (x) ≪ C k (x) <strong>and</strong>there<strong>for</strong>e the corresponding term is discarded via the weighting process.Smoothness indicators Some degrees of freedom are left in the choice of thesmoothness indicators. The general idea is that in smooth regions all indicatorsβ k should basically estimate the same quantity, <strong>and</strong> vanish at a certain rate when∆x → 0, that isβ k = D j ∆x 2p (1 + O(∆x q )) (3.28)where D j is some constant depending on the function <strong>and</strong> on the interval index j. Ifcondition (3.28) is satisfied, then in smooth conditions we would have w k (x) ≈ C k (x)(in a <strong>for</strong>m depending on the exponent q, to be made more precise later), <strong>and</strong> thisenables that scheme to increase the accuracy by merging the in<strong>for</strong>mation fromdifferent stencils. On the other h<strong>and</strong>, if β k is associated to a stencil containing adiscontinuity of the function, then we require thatβ k = O(1) (3.29)so that asymptotically such a stencil would be discarded by the weighting procedure.Following [Sh98], a typical way of defining the smoothness indicators is basedon successive derivatives P (h)kof the polynomials P k , <strong>and</strong> more preciselyβ k =n∑h=1∫ xl+1<strong>for</strong> which (3.28), (3.29) hold with p = 1, q = 2.x l∆x 2h−1 P (h)k(x) 2 dx, (3.30)<strong>Linear</strong> weights Let us consider now the linear weights C k , which have not yet beengiven a precise expression. Due to (3.24) <strong>and</strong> to the definition of the polynomialsP k , the C k are characterized by the fact that Q(x) should be the interpolatingpolynomial of V (x) on the stencil S. Since P k interpolates V on the stencil S k , itwould be natural to require that C k should vanish at the nodes outside S k , <strong>and</strong> thatin the nodes of S the nonzero weights should have unit sum. In this way, if P k isthe polynomial that interpolates the function on S k , then (3.24) is the polynomialthat interpolates V on S.Taking into account that the number of nodes belonging to S but not to S kis precisely n, we thus infer that the linear weights must necessarily have the <strong>for</strong>m∏C k (x) = γ k (x − x h ) = γ k ˜Ck (x) (3.31)x h ∈S\S k


✐✐3.2. Reconstruction techniques in one <strong>and</strong> multiple space dimensions 57with γ k to be determined in order to have unit sum. Note that the conditionsn∑C k (x i ) = 1 (3.32)k=1<strong>for</strong> each x i ∈ S apparently constitute a set of 2n equations in n unknowns. Actually,the correct perspective to look at these conditions is the following. The left-h<strong>and</strong>side of Equation (3.32) is a polynomial of degree n−1 (computed at x i ), <strong>and</strong> on theright-h<strong>and</strong> side we have the polynomial p(x) ≡ 1. There<strong>for</strong>e, imposing that the twopolynomials coincide on more than n − 1 points is equivalent to impose that theyare identical. Now, it is easy to show that the monic polynomials ˜C k , k = 1, . . . , n,are independent <strong>and</strong> <strong>for</strong>m a basis of the space P n−1 of polynomials of degree n − 1.Conditionn∑C k (x) = 1 ∀x ∈ R (3.33)k=1uniquely determines the constants γ k as a polynomial identity.Condition (3.33) can be conveniently satisfied by imposing it on a suitable setof points. Consider first the node x l−n+1 . Then, using (3.31) into (3.24), we notethat, among all the linear weights computed at x l−n+1 , the only nonzero weight isC 1 , so thatQ(x l−n+1 ) = C 1 (x l−n+1 )P 1 (x l−n+1 ) = C 1 (x l−n+1 )v l−n+1 (3.34)<strong>and</strong> this implies that C 1 (x l−n+1 ) = 1 <strong>and</strong> allows to compute γ 1 . On the other h<strong>and</strong>,in the following node x l−n+2 , the nonzero weights are C 1 <strong>and</strong> C 2 , <strong>and</strong> with a similarargument we obtain the conditionQ(x l−n+2 ) = C 1 (x l−n+2 )P 1 (x l−n+2 ) + C 2 (x l−n+2 )P 2 (x l−n+2 ) == [C 1 (x l−n+2 ) + C 2 (x l−n+2 )]v l−n+2 , (3.35)which implies that C 1 (x l−n+2 ) + C 2 (x l−n+2 ) = 1, whence γ 2 can be computed.Following this guideline, we finally obtain the set of conditionsk∑C i (x l−n+k ) = 1 (k = 1, . . . , n), (3.36)i=1that is, more explicitly,k∑i=1γ i∏x h ∈S\S i(x l−n+k − x h ) = 1 (k = 1, . . . , n), (3.37)which is a linear triangular system in the unknowns γ i .ExamplesWe give two examples of this construction (on a uni<strong>for</strong>m mesh), including theexpressions <strong>for</strong> the smoothness indicators (3.30). A general procedure to computelinear weights will be given in the next subsection.


✐✐58 Chapter 3. Elementary building blocksSecond-third order WENO interpolation To construct a third order interpolationwe start from two polynomials of second degree, so thatI 3 [V ](x) = w L P L (x) + w R P R (x), (3.38)where P L (x) <strong>and</strong> P R (x) are second-order polynomials constructed respectively onthe nodes x l−1 , x l , x l+1 <strong>and</strong> on the nodes x l , x l+1 , x l+2 . The two linear weightsC L <strong>and</strong> C R are first degree polynomials in x, <strong>and</strong> according to the general theoryoutlined so far, they readC L = x l+2 − x3∆x , C R = x − x l−13∆x , (3.39)<strong>and</strong> the expressions of α L , α R , w L <strong>and</strong> w R may be easily recovered from the general<strong>for</strong>m.According to (3.30), the smoothness indicators have the explicit expressionsβ L = 1312 v2 l−1 + 16 3 v2 l + 2512 v2 l+1 − 13 3 v l−1v l + 7 6 v l−1v l+1 − 19 3 v lv l+1 , (3.40)β R = 1312 v2 l+2 + 163 v2 l+1 + 2512 v2 l − 13 3 v l+2v l+1 + 7 6 v l+2v l − 19 3 v lv l+1 . (3.41)Third-fifth order WENO interpolation To construct a fifth order interpolationwe start from three polynomials of third degree:I 5 [V ](x) = w L P L (x) + w C P C (x) + w R P R (x), (3.42)where the third-order polynomials P L (x), P C (x) <strong>and</strong> P R (x) are constructed respectivelyon x l−2 , x l−1 , x l , x l+1 , on x l−1 , x l , x l+1 , x l+2 <strong>and</strong> on x l , x l+1 , x l+2 , x l+3 . Theweights C L , C C <strong>and</strong> C R are second degree polynomials in x, <strong>and</strong> have the <strong>for</strong>mC L = (x − x l+2)(x − x l+3 )20∆x 2 ,C C = (x − x l−2)(x − x l+3 )10∆x 2 ,C R = (x − x l−2)(x − x l−1 )20∆x 2 ,while the smoothness indicators β C <strong>and</strong> β R have the expressionsβ C = 6145 v2 l−1 + 33130 v2 l + 33130 v2 l+1 + 6145 v2 l+2 − 14120 v l−1v l + 17930 v l−1v l+1 ++ 293180 v l−1v l+2 − 125960 v lv l+1 + 17930 v lv l+2 − 14120 v l+1v l+2 , (3.43)β R = 40790 v2 l + 72130 v2 l+1 + 24815 v2 l+2 + 6145 v2 l+3 − 119360 v lv l+1 + 43930 v lv l+2 −− 683180 v lv l+3 − 230960 v l+1v l+2 + 30930 v l+1v l+3 − 55360 v l+2v l+3 , (3.44)<strong>and</strong> β L can be obtained using the same set of coefficients of β R in a symmetric way(that is, replacing the indices l − 2, . . . , l + 3 with l + 3, . . . , l − 2).


✐✐3.2. Reconstruction techniques in one <strong>and</strong> multiple space dimensions 59Theoretical resultsIt is easy to check that, due to their structure, the linear weights <strong>for</strong> the twoexamples above are always positive in the interval [x l , x l+1 ]. We prove now thatthis result holds in general. In addition, we give an explicit expression <strong>for</strong> thelinear weights in the case of evenly spaced nodes, <strong>and</strong> prove an interpolation errorestimate.Positivity of the weights We first give a general proof of positivity <strong>for</strong> the linearweights C k (x). We point out that, due to the general structure of nonlinear weights,positivity of the <strong>for</strong>mer set of weights implies positivity of the latter.Theorem 3.10. Let {x i } be a family of consecutively numbered points of R. Then,once the Lagrange polynomial Q(x) of a given function v(x) built on the stencilS = {x l−n+1 , . . . , x l+n } is written in the <strong>for</strong>m (3.24), the linear weights C k (x)(k = 1, . . . , n) are nonnegative <strong>for</strong> any x ∈ (x l , x l+1 ). Moreover, ∑ k C k(x) ≡ 1.Proof. Let us denote by Q (h,m) (x) (with j −n+1 ≤ h ≤ j < j +1 ≤ m ≤ j +n, <strong>and</strong>m − h ≥ n) the interpolating polynomial constructed on the stencil {x h , . . . , x m },so thatQ(x) = Q (j−n+1,j+n) (x), (3.45)<strong>and</strong> moreoverWe will also write a generic Q (h,m) (x) asP k (x) = Q (j−n+k,j+k) (x). (3.46)Q (h,m) (x) = ∑ kC (h,m)k(x)P k (x), (3.47)where the summation may be extended to all k = 1, . . . , n by setting C (h,m)k= 0whenever the stencil of P k is not included in {x h , . . . , x m }. There<strong>for</strong>e, the finallinear weights will beC k (x) = C (j−n+1,j+n)k(x). (3.48)We proceed by induction on the degree of Q (h,m) , given by i = deg Q (h,m) =m − h, starting with i = n up to i = 2n − 1. First we prove that the claim ofthe Theorem extends to all polynomials Q (h,m) (x) defined in (3.47). The claimis obviously true <strong>for</strong> Q (j−n+k,j+k) (x) by (3.46). In this case the coefficients readC (h,m)k= 1, C s(h,m) ≡ 0, s ≠ k. Then, we assume the claim is true <strong>for</strong> the set oflinear weights C (h,m)k(x) of a generic Q (h,m) (x) such that m − h = i <strong>and</strong> add a nodeto the right (adding a node to the left leads to an analogous computation). Byinductive assumption we have, <strong>for</strong> any h <strong>and</strong> m such that h ≤ j < j + 1 ≤ m <strong>and</strong>m − h = i, that (3.47) holds with C (h,m)k(x) ≥ 0. On the other h<strong>and</strong>, by elementaryinterpolation theory arguments (Neville’s recursive <strong>for</strong>m of the interpolatingpolynomial),Q (h,m+1) (x) = x m+1 − xQ (h,m) (x) +x − x hQ (h+1,m+1) (x). (3.49)x m+1 − x h x m+1 − x hNote that deg Q (h,m) = deg Q (h+1,m+1) = i, <strong>and</strong> that both fractions which multiplyQ (h,m) (x) <strong>and</strong> Q (h+1,m+1) (x) are positive as long as x ∈ (x l , x l+1 ). There<strong>for</strong>e, using


✐✐60 Chapter 3. Elementary building blocks(3.47) into (3.49), we getQ (h,m+1) (x) = ∑ [xm+1 − xC (h,m)k(x) + x − x ]hC (h+1,m+1)k(x) P k (x).x m+1 − x h x m+1 − x hk(3.50)The term in square brackets is C (h,m+1)k(x) <strong>and</strong> it is immediate to check that it isnonnegative. Iterating this argument up to Q (j−n+1,j+n) (x) completes the proof ofnonnegativity <strong>for</strong> all the linear weights C k (x).Lastly, setting v(x) ≡ 1, we also have Q(x) ≡ 1, <strong>and</strong>, <strong>for</strong> any k, P k (x) ≡ 1.Plugging this identities into (3.24) we obtain ∑ k C k(x) ≡ 1.Explicit <strong>for</strong>m of the linear weights We turn now to the problem of giving anexplicit expression to the linear weights in the situation of evenly spaced nodes.The expression of the weights <strong>for</strong> the linear scheme is∏C k = γ kx h ∈S\S k(x − x h ),so that the problem is to provide the expression of the constants γ k by applying tothe case of a uni<strong>for</strong>m grid the general procedure already outlined. The endpoint ofthis analysis is given by the following theorem.Theorem 3.11. Assume the space grid is evenly spaced with step ∆x. Then, thelinear weights (3.31) have the explicit <strong>for</strong>mC k (x) = (−1)n+k∆x n−1 n!(n − 1)!(n − k)!(k − 1)!(2n − 1)!∏x h ∈S\S k(x − x h ) (3.51)<strong>and</strong> satisfy the positivity condition C k (x) > 0 <strong>for</strong> any k ∈ {1, . . . , n}, x ∈ [x l , x l+1 ].Proof. First observe that the polynomial C k (x) does not change sign in the interval[x l , x l+1 ]. More precisely, the sign of it, in this interval, is given by sgn(γ k )(−1) n+k .There<strong>for</strong>eC k (x) > 0 ∀x ∈ [x l , x l+1 ], ∀k, n ⇔ sgn(γ k ) = (−1) n+k . (3.52)For simplicity, set l = n in the definition of the stencil S, so that S = {x 1 , . . . , x 2n }.Let us introduce the reference stencils˜S = S/∆x, ˜Sk = S k /∆x.Then the polynomials C k can be written as∏C k (∆xη) = ˜γ k (η − h),h∈ ˜S\ ˜S kwhere η is the variable in the reference space, ˜γ k = γ k ∆x n−1 . The dimensionlessconstants ˜γ k satisfy the conditionk∑˜γ ii=12n∏h=1h∉{i,...,i+n}(i − h) = 1.


✐✐3.2. Reconstruction techniques in one <strong>and</strong> multiple space dimensions 61Introducing now the matrixa ki =∏ 2nh=1∏ i+nh=ithe system can be written as(k − h)=(k − h)(k − 1)!(2n − k)!(n + i − k)!(k − i)! (−1)n+i , (3.53)k∑a ki˜γ i = 1.i=1Let µ i ≡ (−1) n+i˜γ i . Then, such constants satisfy the triangular systemk∑|a ki |µ i = 1 (3.54)i=1<strong>and</strong> according to (3.52) C i (x) > 0, x ∈ [x n , x n+1 ] if <strong>and</strong> only if µ i > 0.The solution to system (3.54) can be explicitly given. In fact, we will provenow thatµ i =n!(n − 1)!(n − i)!(i − 1)!(2n − 1)!(i = 1, . . . , n). (3.55)Actually, plugging (3.55) <strong>and</strong> (3.53) into (3.54) we obtain the set of conditionsthat we want to check:k∑ (k − 1)!(2n − k)!n!(n − 1)!(n + i − k)!(k − i)!(n − i)!(i − 1)!(2n − 1)! = 1.i=1These relations can be rearranged in the <strong>for</strong>m(k − 1)!(2n − k)!(2n − 1)!j=0k∑i=1( ) ( )n n − 1k − i i − 1<strong>and</strong> hence, rewriting the first term <strong>and</strong> shifting the summation index as j = i − 1,k−1∑( ) ( ) ( )n n − 1 2n − 1= ,k − 1 − j j k − 1which is in turn a special case of the identity (see [AS64])N∑j=0with N = k − 1, <strong>and</strong> m = n − 1.( ( ) n m=N − j)j( ) n + m,N= 1,Remark 3.12. We can further prove here that the µ i are inverse of integers. Infact, note first that by (3.55), µ i = µ n−i+1 , so that it suffices to prove the claim <strong>for</strong>i ≤ (n + 1)/2. Then, we rewrite the µ i as⎧n!if i = 1⎪⎨ (2n − 1)!µ i =⎪⎩(n − i + 1) · · · (n − 1)n(n − 1)!(i − 1)!(2n − 1)!if i > 1.(3.56)


✐✐62 Chapter 3. Elementary building blocksThe claim is obvious if i = 1, whereas <strong>for</strong> i > 1 we derive from (3.56)µ i =(2n − 2i + 2) · · · (2n − 2)2 i−1 (i − 1)!(n + 1) · · · (2n − 1) .Now, if i ≤ (n+1)/2, then 2n−2i+2 ≥ n+1 <strong>and</strong> there<strong>for</strong>e any term of the product(2n−2i+2) · · · (2n−2) can be simplified with a term of the product (n+1) · · · (2n−1).This completes the proof of the claim.We report in the following table the value of 1/µ i , <strong>for</strong> values of n up to n = 8.n 1/µ i1 12 3, 33 20, 10, 204 210, 70, 70, 2105 3024, 756, 504, 756, 30246 55440, 11088, 5544, 5544, 11088, 554407 1235520, 205920, 82368, 61776, 82368, 205920, 12355208 32432400, 4633200, 1544400, 926640, 926640, 1544400, 4633200, 32432400Table 3.3. Values of 1/µ i <strong>for</strong> the first 8 values of nInterpolation error estimates We finally prove here the convergence result <strong>for</strong>WENO interpolations, restricting to the smooth case.Theorem 3.13. Let v(x) be a C ∞ function, <strong>and</strong> the interpolation I r [V ] be definedby the WENO <strong>for</strong>m (3.27) with the smoothness indicators β k satisfying (3.28).Then,‖v − I r [V ]‖ ∞ ≤ C∆x min(2n,n+q+1) . (3.57)Proof. First, note that if (3.28) is satisfied, thenw k (x) = C k (x) + ω k (x, ∆x q ) (3.58)where <strong>for</strong> any k, ω k (x, ∆x q ) = O(∆x q ), <strong>and</strong>, since both the w k <strong>and</strong> the C k haveunity sum, we also have ∑ k ω k(x, ∆x q ) = 0. Then,|v(x) − I r [V ](x)| ≤∣ v(x) − ∑ k∣ ∣∣∣∣ C k (x)P k (x)∣ + ∑C k (x)P k (x) − ∑kkw k (x)P k (x)∣ .For the first term it is clear that∣ v(x) − ∑ kC k (x)P k (x)∣ ≤ C∆x2n , (3.59)


✐✐3.2. Reconstruction techniques in one <strong>and</strong> multiple space dimensions 63whereas <strong>for</strong> the second, using (3.58) <strong>and</strong> the fact that the P k are themselves interpolatingpolynomials of degree n,∑C k (x)P k (x) − ∑ ∣ ∣∣∣∣ w k (x)P k (x)∣∣ = ∑[C k (x) − w k (x)]P k (x)∣ =kkk∑=ω k (x, ∆x q ) [ v(x) + O(∆x n+1 ) ]∣ ∣ ∣∣∣=∣k=∣ v(x) ∑ ω k (x, ∆x q ) + O(∆x n+q+1 )∣ =k= O(∆x n+q+1 ) (3.60)where in the last display we have used the fact that the ω k (x, ∆x q ) have zero sum.Finally, (3.57) follows from (3.59), (3.60).3.2.4 H<strong>and</strong>ling multiple dimensions by separation of variablesLagrange interpolation When passing to multiple space dimensions, the extensionof Lagrange interpolation (that is, in practice, the definition of suitable cardinalbasis functions) can be per<strong>for</strong>med in different <strong>for</strong>ms. In the simplest case, cardinalbasis functions are defined by product of one-dimensional functions, in the <strong>for</strong>mthat follows. Let the space grid of points x j be uni<strong>for</strong>m <strong>and</strong> orthogonal, <strong>and</strong> j =(j 1 , . . . , j d ) be the multi-index associated to a given node. Now, once ensured thatthe ψ j are cardinal functions, the <strong>for</strong>m (3.7) still defines an interpolant of v. On theother h<strong>and</strong>, on a uni<strong>for</strong>m orthogonal grid, a basis of Lagrange cardinal functionscan be defined byd∏ψ j (ξ) = ψ jk (ξ k ). (3.61)k=1where, in order to avoid ambiguities, we have denoted the space variable by ξ =(ξ 1 , . . . , ξ d ). In (3.61), it should be understood that ψ j (which is indexed by amulti-index <strong>and</strong> has an argument in R d ) refers to the d–dimensional grid, whereasψ jk (which has a simple index, <strong>and</strong> has a scalar argument) refers to the grid on thek–th variable. It is clear that (3.61) defines a cardinal function, <strong>and</strong> in fact{1 if i 1 = j 1 , . . . , i d = j dψ j (x i ) =0 else.Figure 3.3 shows the bidimensional version of the cubic interpolation reference basisfunction of Figure 3.2.A different interpretation of this procedure can be given by splitting the operationimplied by (3.7). Working <strong>for</strong> simplicity in two dimensions, <strong>and</strong> plugging(3.61) into (3.7), we haveI[V ](ξ) = ∑ ∑v(x j1,j 2)ψ j1 (ξ 1 )ψ j2 (ξ 2 ) =j 1 j 2⎡⎤= ∑ ⎣ ∑ v(x j1,j 2)ψ j1 (ξ 1 ) ⎦ ψ j2 (ξ 2 ). (3.62)j 2 j 1


✐✐64 Chapter 3. Elementary building blocks2.01.51.0Z0.50.0!0.52.52.01.51.0Y0.0!1.0!2.0!2.0!1.00.0X1.01.52.02.5Figure 3.3. Reference basis function <strong>for</strong> cubic interpolation in R 2In practice, the process appears to be the result of two (in general, d) nested onedimensionalinterpolations. The upper diagram of Figure 3.4 shows the pointsinvolved in this process <strong>for</strong> a linear Lagrange interpolation. First, the point ξ is locatedin a rectangle of the grid, say [l 1 ∆x 1 , (l 1 +1)∆x 1 ]×[l 2 ∆x 2 , (l 2 +1)∆x 2 ]. Then,in the first phase of the interpolation, two one-dimensional linear reconstructionsare per<strong>for</strong>med to recover the interpolated values at (ξ 1 , l 2 ∆x 2 ) <strong>and</strong> (ξ 1 , (l 2 +1)∆x 2 ).In the second phase, a linear interpolation is per<strong>for</strong>med between these two valuesto recover the interpolated value at (ξ 1 , ξ 2 ).Once arranged the same points in a tree (as shown in the lower diagram),the interpolation is computed by advancing from the leaves towards the root (interms of data structures, the interpolated value is computed as the tree is visitedin postorder).In general, this procedure can be applied to any dimension d, as well as anyorder of interpolation. The dimension appears as the depth of the tree, whereasthe number of nodes required <strong>for</strong> one-dimensional reconstruction coincides with thenumber of sons of a given node of the tree.ENO/WENO interpolation The nonlinear nature of non-oscillatory reconstructionsprevents them to be extended to R d in the <strong>for</strong>m (3.7), (3.61). On the otherh<strong>and</strong>, the mechanism of successive one-dimensional interpolations may also be appliedto ENO or WENO (or in general, to nonlinear) reconstructions, although inthis case, due to the existence of various c<strong>and</strong>idate stencils, the total number ofpoints involved is considerably higher. It is also possible, however, to define nonoscillatoryinterpolants based on genuinely multidimensional stencils, possibly onunstructured grids.3.2.5 Finite element interpolationFinite element interpolation is a reconstruction strategy which, still interpolating agiven function in a piecewise polynomial <strong>for</strong>m, does not need any particular structureof the space grid, <strong>and</strong> is intrinsically multidimensional, although it also includesthe one-dimensional case.


✐✐// a circle with ceborder C0(t=0,2*pi) { x = 5 * cos(t); y = 5 * sin(t)border C1(t=0,2*pi) { x = 2+0.3 * cos(t); y = 3*sin(border C2(t=0,2*pi) { x = -2+0.3 * cos(t); y = 3*sin3.2. Reconstruction techniques in one <strong>and</strong> multiple space dimensions 65mesh Th = buildmesh(C0(60)+C1(-50)+C2(-50));plot(Th,ps="electroMesh");fespace Vh(Th,P1);Vh uh,vh; //problem Electro(uh,vh) = //int2d(Th)( dx(uh)*dx(vh) + dy(uh)*dy(vh) )+ on(C0,uh=0) //+ on(C1,uh=1)+ on(C2,uh=-1) ;Electro; // solve the problem, seeplot(uh,ps="electro.eps",wait=true);Figure 3.4. Dimensional splitting <strong>for</strong> linear interpolation in R 2Figure 9.3: Disk with two elliptical holesFigure 3.5. Tessellation of a complex geometry into triangles, includingmesh adaptationFigure 9.4: Ecated in right hA definite advantage of finite element interpolations is the ability to treatcomplex geometries, which in general do not allow <strong>for</strong> a structured grid. We showin Figure 3.5 an example, which also illustrates the possibility of per<strong>for</strong>ming localrefinements <strong>for</strong> accuracy purposes.


✐✐66 Chapter 3. Elementary building blocksConstructionBy definition, a finite element is a triple (K, Σ, V r ), where• K is the reference domain of the element, <strong>and</strong> also in practice the kind ofelementary figure into which the computational domain is split. In one spacedimension K is necessarily an interval; in R 2 typical choices are triangles, <strong>and</strong>less frequently rectangles, <strong>and</strong> so on;• V r is the space of polynomial in which the interpolation is constructed. Twotypical choices we will consider here are P r (the space of polynomials in dvariables of degree no larger than r) <strong>and</strong> Q r (the space of polynomials ofdegree no larger than r with respect to each variable);• Σ is a set of functionals γ i : P r → R (usually referred to as degrees of freedom)which completely determine a polynomial of P r . In the case of Lagrangefinite elements, which is the one of interest here, the degrees of freedom arethe values of the polynomial on a unisolvent set of nodes in K (with someabuse of notation, they are identified with the nodes), <strong>and</strong> the polynomial isexpressed in the basis of Lagrange functions associated to such set of nodes.We will not explain in detail the construction of a Lagrange finite elementinterpolation in its greatest generality (this topic is treated in the specialized literature),but rather sketch the basic ideas by working on some one- <strong>and</strong> twodimensionalexamples. As a starting point, the computational domain should be(approximately) tessellated with non-overlapping elements obtained by affine trans<strong>for</strong>mationsof the reference element. Interpolation nodes are placed on each elementby mapping the nodes of the reference element. As outlined <strong>for</strong> symmetric Lagrangeinterpolation, the basis function ψ k used in (3.7) can be conceptually obtained byinterpolating the sequence e k .ExamplesTo make the main ideas clear, we present here the construction of the P 1 , P 2 , Q 1<strong>and</strong> Q 2 Lagrange finite element spaces in one <strong>and</strong> two space dimensions. For ageneral treatment of the topic, we refer the reader to classical monographs on finiteelement schemes.P 1 <strong>and</strong> P 2 finite elements in R In one dimension, the only possibility is to definethe reference element as an interval. In order to interpolate with a P 1 or P 2 basis,respectively two <strong>and</strong> three nodes must be placed in the reference interval. To obtaina continuous interpolant, two of them must be placed at the interface betweenelements, that is, in the extreme points of the reference element. There<strong>for</strong>e, if thereference interval keeps unity distance between nodes, in the P 1 case it could bedefined as the interval [0, 1] with the nodes η 0 = 0, η 1 = 1 (here <strong>and</strong> in the sequel,we use η as a variable in the reference space). As a result, the Lagrange referencebasis functions (shown in the left plot of Figure 3.6) will be given byl 0 (η) = 1 − η, l 1 (η) = η,while a generic point x ∈ [x l , x l+1 ] will be carried in the reference interval [0, 1] bythe trans<strong>for</strong>mationη =x − x l.x l+1 − x l


✐✐3.2. Reconstruction techniques in one <strong>and</strong> multiple space dimensions 67The cardinal basis function ψ k associated to a node x k of a one-dimensional gridreads then:⎧x − x k−1if x ∈ (x k−1 , x k ]x k − x k−1⎪⎨ψ k (x) = 1 − x − x kif x ∈ (x k , x k+1 )x k+1 − x k⎪⎩0 elsewhereNote that, unless <strong>for</strong> allowing a nonuni<strong>for</strong>m grid, P 1 finite element interpolationcoincides with first-order Lagrange interpolation. This correspondence, however, islost <strong>for</strong> higher interpolation orders.In the case of P 2 interpolation, we can choose the reference interval as [0, 2] <strong>and</strong>place the three nodes at η = 0, 1, 2. The three Lagrange reference basis functions(shown in the right plot of Figure 3.6) are now given byl 0 (η) = 1 2 (η − 1)(η − 2), l 1(η) = 1 − (η − 1) 2 , l 2 (η) = 1 η(η − 1),2<strong>and</strong>, if x ∈ [x l , x l+2 ], the trans<strong>for</strong>mation from the variable x to the reference variableη isη = 2x − x l.x l+2 − x lThe cardinal basis function ψ k is constructed in a way which parallels the P 1 case,in particular⎧ ( ) 2 x − xk−1⎪⎨ 1 −− 1 if x ∈ (x k−1 , x k+1 )x k − x k−1ψ k (x) =⎪⎩0 elsewhereif x k is internal to an element, <strong>and</strong>⎧12⎪⎨ψ k (x) =12( x − xk−2x k − x k−1) ( x − xk−2x k − x k−1− 1( ) ( )x − xkx − xk− 1− 2x k+1 − x k x k+1 − x k)if x ∈ (x k−2 , x k ]if x ∈ (x k , x k+2 )⎪⎩0 elsewhereif x k is at the interface between two elements.Note that in the P 2 case the situations in which x k is internal or extreme ofan element leads to different <strong>for</strong>ms, even on an uni<strong>for</strong>mly spaced mesh. More ingeneral, in the P r case (with r > 1), the <strong>for</strong>m of ψ k depends on the position of x kwithin the element. Thus, in any of these cases the interpolation fails in general tobe translation invariant.We show in Figure 3.6 two examples of reference intervals <strong>and</strong> basis functions<strong>for</strong> P 1 <strong>and</strong> P 2 finite element interpolation in one space dimension.


✐✐68 Chapter 3. Elementary building blocks1.21.2110.80.80.60.60.40.40.20.200-0.2-0.2-0.40 0.2 0.4 0.6 0.8 1-0.40 0.5 1 1.5 2Figure 3.6. P 1 <strong>and</strong> P 2 reference elements <strong>and</strong> basis functions in RP 1 , P 2 , Q 1 <strong>and</strong> Q 2 finite elements in R 2 At the increase of dimensions, thegeometry of the reference element allows <strong>for</strong> an increasing number of options. InR 2 , two main situations of interest can be recognized: the situations in which thereference element is respectively a triangle (leading to P r finite element spaces) <strong>and</strong>a rectangle (leading to Q r f.e. spaces).On triangular elements, it is necessary to construct the space of polynomialsof degree r, with r = 1, 2 in our example. We will not give the explicit expressionof the reference basis functions, but rather show the choice of the nodes. In the P 1case, the polynomial space has dimension three, <strong>and</strong> the nodes are placed in thevertices of the element. In the P 2 case, the space has dimension six, <strong>and</strong> the nodesare typically placed at vertices <strong>and</strong> midpoints. Upper line of Figure 3.7 shows theposition of nodes in the two situations. Clearly, the k–th Lagrange reference basisfunction is a polynomial of degree r taking the value l k (η m ) = δ km at a genericnode η m . Note that the restriction of a Lagrange basis function to one side of thetriangle gives a polynomial of degree respectively r = 1 <strong>and</strong> r = 2, <strong>and</strong> the numberof nodes on the side is enough to determine univocally this polynomial. Thus, theinterpolation remains continuous passing from one element to the other.On quadrilateral elements, the Lagrange polynomial are constructed as tensorproducts of one-dimensional Lagrange bases, as it has been discussed concerningmultidimensional Lagrange reconstruction. There<strong>for</strong>e, the Q 1 finite element hasfour nodes at the vertices of the quadrilateral, whereas the Q 2 finite element hasnine nodes (vertices, midpoints <strong>and</strong> center of the quadrilateral element). The lowerline of Figure 3.7 reports the position of nodes in this case. Again, the number ofnodes on each side of the quadrilateral implies that the interpolation is continuousat the interface of two elements.Theoretical resultsThe approximation result concerning finite element interpolation follows the generalresult <strong>for</strong> polynomial approximations. However, when dealing with nonuni<strong>for</strong>mspace grids, care should be taken to avoid that triangular elements could degeneratewhen refining the grid, typically by flattening towards segments. While we leave a<strong>for</strong>mal theory to specialized literature, we simply remark here that grid refinementin a finite element setting requires that the largest size ∆x of the elements vanishes,<strong>and</strong> that the grid is nondegenerate. This means that, once denoted by R K theradius of the element K <strong>and</strong> by r K the radius of the inscribed circle, their ratio


✐✐3.3. Function minimization 69Figure 3.7. P 1 –P 2 (upper) <strong>and</strong> Q 1 –Q 2 (lower) reference elements in R 2remains bounded <strong>for</strong> all K:R K≤ C. (3.63)r KLast, we give the approximation result <strong>for</strong> finite element interpolations.Theorem 3.14. Let the interpolation I r [V ] be defined by a P r finite element spaceon a space grid satisfying (3.63). Let v(x) be a uni<strong>for</strong>mly continuous function onR, <strong>and</strong> V be the vector of its nodal values. Then, <strong>for</strong> ∆x → 0,If moreover v ∈ W s,∞ (R), then‖v − I r [V ]‖ ∞ → 0. (3.64)‖v − I r [V ]‖ ∞ ≤ C∆x min(s,r+1) (3.65)<strong>for</strong> some positive constant C depending on the degree of regularity s.3.3 Function minimizationAs a building block of Semi-Lagrangian schemes <strong>for</strong> nonlinear problems, we will beconcerned with the problem of minimization of a function of N variables:f(x ∗ ) = minx∈R N f(x),<strong>and</strong> with the related numerical techniques. Since in our case the function to beminimized has no explicit <strong>for</strong>m, our interest here is mainly focused on derivativefreemethods, i.e. methods which minimize a function without making use of the


✐✐70 Chapter 3. Elementary building blocksFigure 3.8. Updating of a simplex in the basic version of simplex methodanalytic expressions of gradient or hessian. Without aiming at completeness, a listof the main approaches used to per<strong>for</strong>m this operation includes:• Direct search methods• Descent methods (adapted to work without derivatives)• Powell’s method <strong>and</strong> its modifications• Trust-region methods based on quadratic models.We will briefly review the general philosophy of the various classes of methods.The huge amount of related studies cannot be condensed in such a brief survey;rather than giving implementation details, we will sketch the general ideas <strong>and</strong>refer the reader to suitable, specialized literature <strong>and</strong> software packages.3.3.1 Direct search methodsThe name of direct search methods is generally used to collect methods of heterogeneousnature, which share the feature of minimizing a function without using, eitherin exact or approximate <strong>for</strong>m, its gradient. Direct search methods rather proceedby computing the function at a discrete, countable set of points.The most common schemes of this class are based on the simplex technique(not to be confused with the algorithm used in <strong>Linear</strong> Programming), whose basicidea dates back to the ’60s <strong>and</strong> is shown in Figure 3.8.Let a simplex of vertices V 1 , . . . , V m be constructed in R N . Once computedthe corresponding values y 1 , . . . , y m of f at the vertices, the highest value y h isselected, <strong>and</strong> the corresponding vertex V h is reflected with respect to the centroidC of the simplex made by the remaining vertices. Then, the new simplex obtainedby replacing the node V h with V r is accepted if the value of the function at V r is notthe highest value in the new simplex. Otherwise, the second highest value of theoriginal simplex is selected, <strong>and</strong> the procedure is repeated. If all possible updateof the simplex lead to find the highest value on the reflected node, the algorithmstops. The size of the simplex should then be considered as a tolerance in locatingthe minimum, <strong>and</strong> to refine the search the algorithm should be restarted with asimplex of smaller size.


✐✐3.3. Function minimization 71In a later (<strong>and</strong> more widely used) version, the algorithm has been restatedreplacing the reflection of V h with a discrete line search. In this version the geometryof simplices may change, <strong>and</strong> while this allows the scheme to adapt to thegeometry of the function, it can also cause the simplices to degenerate, thus stallingthe algorithm. Specific tools have been devised to h<strong>and</strong>le this situation, the mostobvious one being the restart.3.3.2 Descent methodsIn descent methods, an initial estimate of the minimum point x 0 is given, <strong>and</strong> theupdating of the approximation x k is per<strong>for</strong>med via the <strong>for</strong>mulax k+1 = x k + β k d k , (3.66)that is, the scheme moves from x k to x k+1 along the direction d k with step β k . Asa rule (which also gives the name to this class of schemes), the direction d k shouldbe a descent direction, i.e. (d k , ∇f(x k )) < 0, β k should be positive, <strong>and</strong> f(x k+1 )


✐✐72 Chapter 3. Elementary building blocksFigure 3.9. Construction of conjugate directions in Powell’s methodthey are based on search directions d k which are A–conjugate, that is,(Ad i , d i ) > 0, (Ad i , d j ) = 0 (i ≠ j).In the case of quadratic functions, it can be proved that such a scheme convergesat most in n steps. Its adaptation to nonquadratic function (the socalledconjugate gradient, which requires to compute ∇f at each iteration)has superlinear convergence, while the cost per iteration is O(N 2 ).To turn back to the problem of implementing schemes of this class withoutcomputing derivatives, we remark that in principle it is possible to replace gradient<strong>and</strong>, possibly, hessian of f by their finite difference approximations, although thecomplexity of this operation is N + 1 computations of f <strong>for</strong> the gradient <strong>and</strong> evenhigher <strong>for</strong> the hessian. Whenever the computation of f has a critical complexity(<strong>and</strong> this is definitely the case when using minimization algorithms within a Semi-Lagrangian scheme), the minimization is better accomplished by cheaper techniques.3.3.3 Powell’s method <strong>and</strong> its modificationsThis method is genuinely derivative-free, but search directions are rather computedexploiting a geometric property of quadratic functions, which allows to constructconjugate directions without computing the gradient, nor the hessian. The basicidea is sketched in Figure 3.9 in the case of two space dimensions.Let a quadratic function f (shown through its elliptical level curves) <strong>and</strong> asearch direction d 1 be given. Consider the minima of f along two parallel lineswith direction d 1 (we recall that at a line-constrained minimum the search directionis tangent to the level curve). Then, the direction d 2 through the two minima isconjugate with d 1 . In Figure 3.9, this is recognized by the fact that d 2 passes throughthe center of the ellipses – this agrees with the fact that the second conjugatedirection brings to the global minimum, as it happens <strong>for</strong> a quadratic function intwo dimensions.


✐✐3.4. Numerical computation of the Legendre trans<strong>for</strong>m 73The technique can be extended to generate conjugate directions <strong>for</strong> a genericdimension N, <strong>and</strong> produces a scheme in the <strong>for</strong>m (3.66) of a descent method. Being aconjugate direction method, Powell’s method converges in N iterations <strong>for</strong> quadraticfunctions of N variables, <strong>and</strong> is expected to be superlinear <strong>for</strong> smooth nonquadraticfunctions.Clearly, a certain number of technical details (which will not be reviewed here)should be set up to make the idea work in practice, but this method has inspired alarge number of efficient algorithms. Among the most successful related software,we mention the routine PRAXIS, which is freely available as an open-source Fortrancode.3.3.4 Trust-region methods based on quadratic modelsIn this class of schemes, the iterative minimization of the function f is carriedout by working at each step on an approximate (typically, quadratic) model of thefunction, which is associated to a so-called trust region, which is a set in which themodel is considered as a god approximation of the function.Let x k denote the current iterate. We write the quadratic model <strong>for</strong> thefunction f at x k as<strong>and</strong> the trust region as the ballm k (x k + s) = f(x k ) + (g k , s) + 1 2 (H ks, s), (3.69)B k = {x ∈ R N : ‖x − x k ‖ ≤ ∆ k }.Although the vector g k <strong>and</strong> the (symmetric) matrix H k play the role of respectivelythe gradient <strong>and</strong> the hessian of f, their construction should not use any in<strong>for</strong>mationon derivatives. In practice, they are typically constructed by interpolating a certainset of previous iterates where the function f has already been computed.Once the model (3.69) has been constructed at the iterate x k , a new c<strong>and</strong>idatepoint x + kis computed as the minimum point <strong>for</strong> m k in the trust region. If the valuef(x + k ) represent a sufficient improvement of f(x k), then the iterate is updated asx k+1 = x + k, <strong>and</strong> typically a <strong>for</strong>mer point is removed from the set of point used inthe interpolation. If not, the trust region is reduced by decreasing ∆ k , <strong>and</strong>/or newpoints are added to improve the quality of the approximation.The robustness of the algorithm is crucially related to the geometric propertiesof the set of points used to construct the model (3.69), so various recipes have beenproposed to generate this set, to iteratively update it, <strong>and</strong> to improve it wheneverthe algorithm is unable to proceed. Once more, we will not present the technicaldetails, but rather point out that one such algorithm, NEWUOA, has available opensource implementations.3.4 Numerical computation of the Legendretrans<strong>for</strong>mIn the construction of SL schemes <strong>for</strong> <strong>Hamilton</strong>–Jacobi equations, a crucial roleis played by the Lax–Hopf <strong>for</strong>mula, <strong>and</strong> there<strong>for</strong>e by the Legendre trans<strong>for</strong>m H ∗of the <strong>Hamilton</strong>ian function H. The possibility of an explicit computation of H ∗


✐✐74 Chapter 3. Elementary building blocksonly holds <strong>for</strong> a small class of functions, so we will address here the problem of itsnumerical approximation, focusing in particular on “fast” algorithms.We recall that the Legendre trans<strong>for</strong>m of the function H(p) is defined byH ∗ (q) = supp∈R d {p · q − H(p)}, (3.70)it takes finite values if the coercivity condition (2.26) is satisfied, <strong>and</strong> is a strictlyconvex function of q whenever H is a strictly convex function of p. Moreover, inthe specific framework of interest here, the sup is in fact a maximum, <strong>and</strong> can besearched <strong>for</strong> on a bounded set. In multiple dimensions, a “factorization <strong>for</strong>mula”holds, so thatH ∗ (q) = supp 1∈R{{p 1 q 1 + supp 2∈Rp 2 q 2 + · · · + sup {p d q d − H(p)} · · ·p d ∈R}}.On the basis of this decomposition, it is possible to construct multidimensionalalgorithms <strong>for</strong> the Legendre trans<strong>for</strong>m by successive increases in the dimension.For this reason, we will sketch the basic algorithm in a single space dimension, thatis, using (3.70) with d = 1.Since the set in which we look <strong>for</strong> the max is bounded, it can be discretized bysetting up a grid of points p 1 , . . . , p N , with corresponding values H(p 1 ), . . . , H(p N )(note that in this case, the subscript refers to the index of a point, the problem beingone-dimensional). A parallel grid q 1 , . . . , q M (typically, with M ∼ N) is set up inthe variable q. The Discrete Legendre Trans<strong>for</strong>m algorithm consists in computing(3.70) at the discrete points q k , with the maximum obtained over the discrete set{p 1 , . . . , p N }, that isH ∗ (q k ) =max {p iq k − H(p i )} (k = 1, . . . , M). (3.71)i∈{1,...,N}In this <strong>for</strong>m, the algorithm is convergent under natural assumptions, but has thedrawback of a quadratic complexity. This has prompted the search <strong>for</strong> faster implementations,<strong>and</strong> in particular with complexity O(N log N) or even O(N) as theversion outlined here.The ordered set of vertices (p i , H(p i )) defines a piecewise linear approximationof H(p) as shown in Figure 3.10, where the approximation between the nodes p i<strong>and</strong> p i+1 has the slopes i = H(p i+1) − H(p i ). (3.72)p i+1 − p iNote that, H being convex, this sequence is nondecreasing.Given q, the problem of computing H ∗ (q) is solved once known the argmax ofpq − H(p). On the other h<strong>and</strong>, a certain value p is in this argmax, if <strong>and</strong> only if q isincluded in the subdifferential of H at p. In the discretized version, depending onthe fact that p is a node or is internal to an interval of the grid, this subdifferentialmay be written as{D − [s i−1 , s i ] if p = p iH(p) =s i if p ∈ (p i , p i+1 )(this is shown in Figure 3.10) <strong>and</strong> consequently the argmax (as a function of q) hasthe <strong>for</strong>m{p i if q ∈ [s i−1 , s i ]argmax {pq − H(p)} =(3.73)(p i , p i+1 ) if q = s i .


✐✐3.5. Commented references 75Figure 3.10. Computation of the Discrete Legendre Trans<strong>for</strong>m with piecewiselinear approximationIt suffices there<strong>for</strong>e to create the table of slopes (3.72) <strong>and</strong> to compare a given valueof q with the values of s i to find the argmax <strong>and</strong> hence H ∗ (q). Since a grid isalso set up in the q-domain, the comparison of the q k with the s i in (3.73) simplyrequires a merging of the two vectors. The final algorithm per<strong>for</strong>ms only operationsof linear complexity.3.5 Commented referencesMost of the numerical techniques reviewed in this chapter are analysed in detail indedicated monographs. Classical textbook on the approximation of Ordinary DifferentialEquations are [HNW93], [But03], but many others are available. The basictheory of polynomial interpolation is contained in any textbook in basic NumericalAnalysis, <strong>and</strong> any specific reference is useless. Finite element approximations are asomewhat more specialized (although well-established) issue, <strong>and</strong> among the manymonographs dedicated to this topic, we can quote [Cia02] <strong>and</strong> [BS08].On the contrary, to our knowledge no monograph has been devoted yet tonon-oscillatory schemes. A classical review on the basic ideas of ENO <strong>and</strong> WENOinterpolation is [Sh98] (from which we have reported most results concerning ENOinterpolation). WENO interpolation has been first studied in detail in [CFR05],proving in particular the results concerning positivity of the weights. Moreover, inaddition to what has been presented in this chapter, other reconstruction techniqueshave been applied to Semi-Lagrangian schemes. In particular, we mention here theso-called “shape preserving interpolations” (see [RW90] <strong>and</strong> the references therein),<strong>and</strong>, more recently, the Radial Basis Functions interpolations (see [Buh03] <strong>for</strong> ageneral review on RBF interpolation, <strong>and</strong>, <strong>for</strong> example, [Is03] <strong>for</strong> its application toSL schemes). The algorithm <strong>for</strong> tensorized multidimensional interpolations basedon a tree structure has been proposed in [CFF04], whereas a different approachto polynomial interpolation in high dimensions is provided by the so-called sparsegrids, <strong>for</strong> which an extensive review is given in [BG04].A general, up-to-date reference on optimization is [NW06]. Concerning directsearch methods, a more detailed review can be found in [KLT03], while a parallel(although less recent) review devoted to trust region methods is presented in


✐✐76 Chapter 3. Elementary building blocks[CST97]. Documentation <strong>and</strong> theory about two public domain optimization codes,PRAXIS <strong>and</strong> NEWUOA, can be found in respectively [Bre73] <strong>and</strong> [Pow08].Finally, the algorithm of Fast Legendre Trans<strong>for</strong>m (of complexity O(N log N))has been first proposed in [Br89], <strong>and</strong> further developed in [Co96]. More recently,the possibility of a linear-time (i.e., O(N)) algorithm has been shown in [Lu97].This algorithm corresponds to what has been presented in this chapter.


✐✐Chapter 4Convergence theoryThis chapter presents the main results of convergence <strong>for</strong> numerical schemes. In thecase of the Lax–Richtmeyer equivalence theorem <strong>for</strong> linear problems, the notions ofconsistency <strong>and</strong> stability are reviewed, with a special emphasis on monotone <strong>and</strong> L 2stability which play a crucial role <strong>for</strong> respectively the low-order <strong>and</strong> the high-orderschemes. In the case of convex HJ equations, we present the convergence results ofCr<strong>and</strong>all–Lions, Barles–Souganidis <strong>and</strong> Lin–Tadmor, which require ad hoc stabilityconcepts (monotonicity in the first two cases, uni<strong>for</strong>m semiconcavity <strong>for</strong> the third).4.1 The general settingWe will carry out the discretization in the usual framework of difference schemes.Time is discretized with a (fixed) time step ∆t, so that t k = k∆t, whereas discretizationwith respect to space variables will require to set up a space grid in thecomputational domain (we make the st<strong>and</strong>ing assumption that this grid is nondegenerate,in a sense to be made explicit). We write a generic node as x j , j ∈ I, <strong>for</strong>a given set I of indices. Possible choices include:• unbounded structured uni<strong>for</strong>m meshes in one space dimension, <strong>for</strong> which j ∈I = Z <strong>and</strong> x j = j∆x;• unbounded structured uni<strong>for</strong>m meshes in multiple space dimensions, <strong>for</strong> whichthe node index j = (j 1 , . . . , j d ) is a multiindex, j ∈ I = Z d <strong>and</strong> x j =(j 1 ∆x 1 , . . . , j d ∆x d ) <strong>for</strong> a set of space discretization parameters ∆x 1 , . . . , ∆x d ≤∆x (these parameters are usually required to be linearly related one anotherto avoid degeneracy of the grid);• bounded structured uni<strong>for</strong>m meshes in one or multiple space dimensions, <strong>for</strong>which I is the product of finite sets of indices, but nodes remain evenly spaced;• bounded unstructured meshes in one or more space dimensions (typicallybased on triangulations), <strong>for</strong> which I = {1, . . . , n n }, <strong>and</strong> each node is locatedwith its own coordinates. Suitable geometric conditions on trianglesensure in this case that the grid does not degenerate (see Chapter 3).Note that, in the unbounded case, it is usual to assume that the analyticalsolution be compactly supported. This makes it easier to develop a convergence77


✐✐78 Chapter 4. Convergence theorytheory in Hölder norms different from ‖ · ‖ ∞ , <strong>and</strong> will be implicitly done in thesequel.We denote by vjn be the desired approximation of u(x j , t n ), by V n <strong>and</strong> U(respectively, U(t)) the sets of nodal values <strong>for</strong> the numerical solution at time t n ,<strong>and</strong> <strong>for</strong> the exact solution u(x) (respectively, u(x, t)). W also denote by W <strong>and</strong> Φ(respectively, W (t) <strong>and</strong> Φ(t)) the sets of nodal values of generic functions w(x) <strong>and</strong>φ(x) (respectively, w(x, t) <strong>and</strong> φ(x, t)). In general, we will refer to the set of nodalvalues as to a (possibly infinite) vector.Convergence of numerical schemes will be analysed in a specific norm, <strong>and</strong>throughout the book we will always use normalized Hölder norms, defined in thestructured uni<strong>for</strong>m case by⎧⎨∑ ) 1/α(∆x‖W ‖ α := 1 · · · ∆x d j |w j| α if α < ∞(4.1)⎩max j |w j | if α = ∞.The natural matrix norm associated to the vector norm ‖ · ‖ α is defined as usual bywhich leads in particular, <strong>for</strong> α = 1, 2, ∞, to the <strong>for</strong>ms∑‖B‖ 1 = sup |b ij |j∈Ii∈I(‖B‖ 2 = sup ρ m (B t B)‖BW ‖ α‖B‖ α := sup , (4.2)W ≠0 ‖W ‖ α‖B‖ ∞ = supi∈Im∈I∑|b ij |j∈I) 1/2(where ρ m (·) is a generic eigenvalue of the matrix inside the brackets). Note thatnatural matrix norms are insensitive to the scaling factor (∆x 1 · · · ∆x d ) 1/α appearingin (4.1).In the unstructured case, once written the numerical solution as a linear combinationof basis function in the <strong>for</strong>m∑w j ψ j (x), (4.3)ja proper way of defining a Hölder norm is⎧⎛∣ ∫ ∣∣∣∣∣α ⎞1/α⎪⎨ ∑⎝ w‖W ‖ α :=j ψ j (x)dx⎠if α < ∞R d j ∣⎪⎩max j |w j | if α = ∞.(4.4)Once fixed the initial condition V 0 by setting v 0 j = u 0(x j ), the numerical schemewill be defined <strong>for</strong> n ≥ 0 by the iteration{V n+1 = S(∆; t n , V n )V 0 = U 0(4.5)


✐✐4.2. Convergence results <strong>for</strong> linear problems: the Lax–Richtmeyer theorem 79where ∆ = (∆x, ∆t) denotes the discretization parameters, with possible constraintsposed by stability <strong>and</strong>/or consistency requirements. If the dependence on the discretizationparameters needs not to be explicitly taken into account, we will simplywrite S(t n , V n ) instead of S(∆; t n , V n ); furthermore, if the evolution operator doesnot depend on t, we will write the scheme as S(∆; V n ), or in the simplified <strong>for</strong>mS(V n ). Whenever useful, we will adopt the notation S ∗ (·) in which the asterisk, ifspecified, st<strong>and</strong>s <strong>for</strong> a particular choice of the scheme. We will also denote by S j (·)the j–th component of S, that is, the scheme computed at x j .When dealing with linear problems, we will typically consider (<strong>and</strong> this isdefinitely the case when applying the equivalence theorem) schemes which are linearthemselves, that is, schemes in which the operator S(∆; t n , ·) is affine. There<strong>for</strong>e,in the linear theory we examine schemes of the <strong>for</strong>m{V n+1 = S(∆; t n , V n ) = B(∆; t n )V n + G n (∆)V 0 (4.6)= U 0 .In the same spirit as <strong>for</strong> the general case, the dependence on ∆ <strong>and</strong>/or t n willpossibly be omitted if redundant or unnecessary at all.4.2 Convergence results <strong>for</strong> linear problems: theLax–Richtmeyer theoremWe start the presentation of convergence theory <strong>for</strong> numerical methods with themore classical linear case, <strong>and</strong> in particular with time-invariant evolution operators.We assume there<strong>for</strong>e that the model to be approximated has the <strong>for</strong>m{u t (x, t) + Au(x, t) = g(x, t), (x, t) ∈ Ω × [0, T ](4.7)u(x, 0) = u 0 (x)where A is a differential operator, <strong>and</strong> with suitable boundary conditions on ∂Ω, orwith Ω ≡ R d . Equation (4.7) will be assumed to be well-posed in some space H, tobe specified.Accordingly, the scheme will be assumed to be in the <strong>for</strong>m{V n+1 = S(∆; V n ) = B(∆)V n + G n (∆)V 0 (4.8)= U 0 .It should be recalled from the very start that at the present state of the theory,the linear case is the only situation which allows <strong>for</strong> a general convergence theoryof numerical schemes. This theory is basically summarized in the Lax–Richtmeyerequivalence theorem.4.2.1 ConsistencyThe first requirement on a practical numerical scheme is that it should be a smallperturbation of the original equation. This idea, which is common to both the linear<strong>and</strong> the nonlinear case, is <strong>for</strong>malized in the concept of consistency.We define the local truncation error, or consistency error, asL(∆; t, U(t)) = 1 [U(t + ∆t) − S(∆; t, U(t))] (4.9)∆t


✐✐80 Chapter 4. Convergence theoryDefinition 4.1. Let U(t) be the set of nodal values of a solution u(x, t), <strong>and</strong> T > 0.The scheme S is said to be consistent if, <strong>for</strong> any t ∈ [0, T ],‖L(∆; t, U(t))‖ → 0 (4.10)as ∆ → 0, <strong>for</strong> any u 0 in a dense set of initial conditions.Remark 4.2. In the usual practice, at least in the linear case, the dense set ofsolutions on which (4.10) is tested coincides with C ∞ , which is dense in essentiallyall spaces of interest in the numerical approximation of PDEs. Moreover, oncechosen a smooth solution, the local truncation error typically satisfies a uni<strong>for</strong>mbound on [0, T ], so that <strong>for</strong> a consistent schemeThis point will be implicitly assumed in the sequel.‖L(∆; t, U(t))‖ ≤ τ(∆) → 0. (4.11)Remark 4.3. The idea of consistency remains basically unchanged in the nonlinearcase, but its <strong>for</strong>mal statement may vary somewhat in the various <strong>for</strong>mulations, asit will be shown in section 4.4.4.2.2 StabilityThe second key assumption is that the small perturbations introduced by the discretizationshould not be amplified by the scheme to an uncontrolled magnitude.This leads to the definition of stability.Definition 4.4.n∆t ∈ [0, T ],The scheme (4.6) is said to be stable if, <strong>for</strong> any n such that<strong>for</strong> some constant M s > 0 independent of ∆.‖B(∆) n ‖ ≤ M s (4.12)Remark 4.5. The norm used in (4.12) must be an operator (matrix) norm compatiblewith the norm used in (4.10).Remark 4.6. While consistency is a st<strong>and</strong>ing assumption which also applies tononlinear situations, stability as stated by (4.12) may only be applied to linearschemes. In fact, some <strong>for</strong>m of stablity is also necessary in nonlinear situations,but no such general concept exists, <strong>and</strong> each case is treated in a specific way.Assumption (4.12) is not easy to check in its greatest generality. Easier sufficientconditions <strong>for</strong> (4.12) to hold may be derived using the inequality (which holds<strong>for</strong> any matrix norm)‖B n ‖ ≤ ‖B‖ n .Then, a first trivial condition which implies stability is‖B‖ ≤ 1so that we would also have ‖B‖ n ≤ 1 <strong>for</strong> any n. This condition may be weakenedas‖B‖ ≤ 1 + C∆t (4.13)


✐✐4.2. Convergence results <strong>for</strong> linear problems: the Lax–Richtmeyer theorem 81<strong>for</strong> some C independent of the discretization parameters. In fact, taking into accountthat n ≤ T/∆t, we have‖B‖ n ≤ (1 + C∆t) n ≤≤ (1 + C∆t) T/∆t ≤≤ e CT .More in general, if ‖B‖ ∼ 1 + C∆t α , then the scheme is stable if α ≥ 1, unstable ifα < 1.4.2.3 ConvergenceThe definition of convergent scheme is related to the possibility of approximatingnot only a particular set of solutions (e.g., smooth solutions), but any solution inthe space H. So we will define a convergent scheme as follows.Definition 4.7.n∆t ∈ [0, T ],The scheme S is said to be convergent if, <strong>for</strong> any n such that‖V n − U(t n )‖ → 0 (4.14)<strong>for</strong> any initial condition u 0 ∈ H, as ∆ → 0.We have now all elements to state the main result, known as the Lax–Richtmeyerequivalence theorem:Theorem 4.8. Let the scheme (4.8) be consistent. Then, it is convergent if <strong>and</strong>only if it is stable. Moreover, if the solution is smooth enough to satisfy (4.11),then, <strong>for</strong> any n such that n∆t ∈ [0, T ],<strong>for</strong> some positive constant C.‖V n − U(t n )‖ ≤ Cτ(∆) (4.15)Proof. We give a sketch of the proof, <strong>for</strong> the part related to sufficiency (stabilityimplies convergence), <strong>and</strong> in particular to (4.15). First, the scheme <strong>and</strong> theconsistency condition are rewritten at a generic step k ∈ [0, n − 1] as respectivelyV k+1 = B(∆)V k + G k (∆),U(t k+1 ) = B(∆)U(t k ) + G k (∆) + ∆tL(∆; t k , U(t k )).Subtracting now both sides, <strong>and</strong> defining the error at the step k ase k = U(t k ) − V k ,we havee k+1 = B(∆)e k + ∆tL(∆; t k , U(t k )).Taking now into account that e 0 = 0 by the choice of the initial condition, we obtain


✐✐82 Chapter 4. Convergence theoryat successive time steps:e 1 = ∆tL(∆; t 0 , U(t 0 ))e 2 = B(∆)e 1 + ∆tL(∆; t 1 , U(t 1 )) == ∆t (L(∆; t 1 , U(t 1 )) + B(∆)L(∆; t 0 , U(t 0 ))).e n = B(∆)e n−1 + ∆tL(∆; t n−1 , U(t n−1 )) == ∆t ( L(∆; t n−1 , U(t n−1 )) + · · · + B(∆) n−1 L(∆; t 0 , U(t 0 )) ) =n−1∑= ∆t B(∆) i L(∆; t n−i−1 , U(t n−i−1 ))i=0so that, due to the stability assumption, we can bound the error e n as:‖e n ‖ ≤ n∆tM s τ(∆) ≤ T M s τ(∆),<strong>and</strong> using also the assumption of consistency, we obtain that e n → 0, <strong>and</strong> in particular(4.15), once defined C = T M s .Remark 4.9. Note that the computation assumes that the consistency error wouldvanish, so that in principle it only holds <strong>for</strong> smooth solutions. On the other h<strong>and</strong>,smooth solutions are a dense set, so it is possible to prove convergence to all solutionsin H by a density argument.Remark 4.10. The reverse implication (convergence implies stability) requires theuse of the principle of uni<strong>for</strong>m boundedness (Banach–Steinhaus theorem). We donot show this part of the proof, but remark that it allows to interpret the Courant–Friedrichs–Lewy condition, which will be described later, as a necessary conditionof stability.4.2.4 Time-dependent evolution operatorsWe turn now to the case of time-dependent evolution operators, that is{u t (x, t) + A(t)u(x, t) = g(x, t), (x, t) ∈ Ω × [0, T ]u(x, 0) = u 0 (x),(4.16)complemented again with suitable boundary conditions. Note that, in some respect,(4.16) might be considered as a model problem <strong>for</strong> a number of nonlinear models,in which linearization would lead to local evolution operators depending on thesolution, <strong>and</strong> there<strong>for</strong>e on time.A scheme <strong>for</strong> (4.16) should be assumed to be in the more general <strong>for</strong>m (4.6),<strong>and</strong> adapting the Lax–Richtmeyer equivalence theorem to this case is quite straight<strong>for</strong>ward.First, the general definition of consistency (4.10) directly applies to ascheme in the <strong>for</strong>m (4.6). Second, retracing the proof of the equivalence theorem,the expression of the error at the n-th time step is replaced bye n = B(∆; t n−1 )e n−1 + ∆tL(∆; t n−1 , U(t n−1 )) == ∆t ( L(∆; t n−1 , U(t n−1 )) + B(∆; t n−1 )L(∆; t n−2 , U(t n−2 ) + · · · ++B(∆; t n−1 )B(∆; t n−2 ) · · · B(∆; t 1 )L(∆; t 0 , U(t 0 )) )


✐✐4.2. Convergence results <strong>for</strong> linear problems: the Lax–Richtmeyer theorem 83<strong>and</strong> it is immediate to see that the <strong>for</strong>mulation (4.12) of stability should be replacedwith the requirement that, <strong>for</strong> any n such that n∆t ∈ [0, T ],‖B(∆; t n )B(∆; t n−1 ) · · · B(∆; t 1 )‖ ≤ M s<strong>for</strong> some constant M s > 0 independent of ∆. In practice, in what follows we willnever apply the general stability condition (4.12), but rather the sufficient condition(4.13), which obviously works also in the time-dependent case once the constant Cis independent of the time step considered.4.2.5 The stationary caseTo conclude this review on linear convergence theory, we sketch the basic ideas totreat stationary problems. In particular, we consider here problems of the linear<strong>for</strong>mAu(x) = g(x) (4.17)<strong>and</strong>, <strong>for</strong> our specific purposes, schemes in the fixed point <strong>for</strong>mV = S(∆; V ) (4.18)in which a component v j of the vector V is intended to approximate u(x j ). In orderto ensure that the discrete problem has a unique solution, we typically assume thatS is a contraction, that is‖S(∆; V ) − S(∆; W )‖ ≤ L S < 1<strong>for</strong> V , W in some set U, such that S(∆; U) ⊆ U (note that in general the Lipschitzconstant L S should be assumed to depend on ∆). In this framework, the numericalsolution may be computed as a limit of the fixed point iteration{V (k+1) = S ( ∆; V (k))V (0) (4.19)∈ U.The similarity between (4.19) <strong>and</strong> (4.5) is more than <strong>for</strong>mal. An iteration like(4.19) is often the result of applying a scheme in the time-dependent <strong>for</strong>m (4.5) tocompute the stationary state (4.17) of the evolutive equation (4.7), provided such astationary state exists. In this strategy, which produces the so-called time-marchingschemes, the discrete solution is computed as a regime state of the scheme, just asthe exact solution is a regime state of the equation. Also, <strong>for</strong> all these reasons,when considering a time-marching scheme we might find it useful to introduce atime discretization parameter even if the equation to be approximated is stationary.As it has been remarked <strong>for</strong> the evolutive case, in the linear theory we considerschemes with S(·) affine, that is, schemes of the <strong>for</strong>mV = S(∆; V ) = B(∆)V + G. (4.20)In this situation, the condition <strong>for</strong> S to be a contraction readsL S = ‖B(∆)‖ < 1 (4.21)<strong>and</strong> the invariant set U coincides with the whole domain of S. Note that, in theproof of convergence, (4.21) will appear as a stability condition. Note also that, ifthe fixed point system (4.20) is recast as a linear system in the <strong>for</strong>m(I − B(∆))V = G,


✐✐84 Chapter 4. Convergence theory<strong>and</strong> if condition (4.21) is en<strong>for</strong>ced in the ∞-norm (which typically happens in monotoneschemes), then by (4.21) the matrix I − B(∆) is diagonally dominant, <strong>and</strong>there<strong>for</strong>e nonsingular. Hence, the practical implementation of the scheme needsnot to be carried out in iterative <strong>for</strong>m.ConsistencyWhile the general idea of plugging the exact solution in the scheme also applies toschemes in the <strong>for</strong>m (4.18) or (4.20), a correct scaling should be adopted to definethe consistency error. We define the local truncation error <strong>for</strong> the scheme (4.18), asL(∆; U) =<strong>and</strong> the scheme is said to be consistent if11 − L S[U − S(∆; U)] , (4.22)‖L(∆; U)‖ ≤ τ(∆) → 0 (4.23)as ∆ → 0, <strong>for</strong> any u(x) in a dense set of solutions. When applying this definitionto the particular case of (4.20), the local truncation error is given byL(∆; U) =1[U − B(∆)U − G] . (4.24)1 − ‖B(∆)‖Remark 4.11. Typically, in time-marching schemes, the contraction coefficientsatisfies L S ∼ 1 − C∆t, <strong>and</strong> there<strong>for</strong>e the definition of local truncation error (4.22)matches the definition (4.9), when applied to a stationary solution.ConvergenceWe define convergence <strong>for</strong> the stationary scheme (4.18) in the obvious way, that is,requiring that‖V − U‖ → 0,as ∆ → 0. We can now give a convergence result which adapts the Lax–Richtmeyertheorem to the framework of linear stationary schemes in the <strong>for</strong>m (4.20).Theorem 4.12. Let the scheme (4.20) be consistent <strong>and</strong> satisfy (4.21). Then, itis convergent. Moreover, if the solution is smooth enough to satisfy (4.23), then‖V − U‖ ≤ τ(∆). (4.25)Proof. Following the same guidelines of the equivalence theorem, the scheme <strong>and</strong>the consistency condition are rewritten asV (k+1) = B(∆)V (k) + G,U = B(∆)U + G + (1 − ‖B(∆)‖)L(∆; U).Subtracting both sides, <strong>and</strong> defining the error at the iteration k ase (k) = V (k) − U,


✐✐4.3. More on stability 85we obtaine (k+1) = B(∆)e (k) + (1 − ‖B(∆)‖)L(∆; U).We are interested in giving a bound on the limitlimk→∞ e(k) = lim[U − V (k)]k→∞which corresponds to the error on the fixed point solution V of (4.20). Since thissolution is unique, it does not depend on V (0) <strong>and</strong> we can assume that V (0) = U,so that e (0) = 0, <strong>and</strong> we obtain <strong>for</strong> successive iterations:e (1) = (1 − ‖B(∆)‖)L(∆; U)e (2) = B(∆)e (1) + (1 − ‖B(∆)‖)L(∆; U) == (1 − ‖B(∆)‖)(I + B(∆))L(∆; U).e (k) = B(∆)e (k−1) + (1 − ‖B(∆)‖)L(∆; U) == (1 − ‖B(∆)‖) [ I + B(∆) + · · · + B(∆) k−1] L(∆; U).Now, giving an upper bound on the sum in square brackets as a geometric series,we can estimate the error e (k) as:‖e (k) 1‖ ≤ (1 − ‖B(∆)‖)‖L(∆; U)‖ ≤ τ(∆),1 − ‖B(∆)‖<strong>and</strong>, by consistency, we obtain that e (k) → 0, <strong>and</strong> in particular (4.25) if the solutionis smooth enough.4.3 More on stabilityWe turn back here to examine in greater detail the concept of stability, <strong>and</strong> inparticular some special <strong>for</strong>ms of this general assumption. Most of what will be saidhere can be applied to linear as well as nonlinear problems <strong>and</strong> schemes.4.3.1 The CFL conditionIn 1928, long be<strong>for</strong>e the theory of convergence of numerical schemes <strong>for</strong> PDEs hadbecome an established matter, Courant, Friedrichs <strong>and</strong> Lewy published a paperin which the concept of domain of dependence was singled out as a key point <strong>for</strong>the convergence of numerical schemes. To sketch the general idea, we first definethe analytical domain of dependence D d (x, t) of the solution u at a point (x, t), asthe smallest set such that, if the initial condition u 0 of (4.7) is changed on the setR N \ D d (x, t), u(x, t) remains unchanged. The discrete counterpart of this set isthe numerical domain of dependence D ∆ d (x, t) which, <strong>for</strong> a point (x, t) = (x i, t n ) inthe space-time grid, is the set of all nodes x j such that u 0 (x j ) affects the value v n i .Figure 4.1 illustrates these two concepts.The Courant, Friedrichs <strong>and</strong> Lewy (CFL) necessary condition states that, <strong>for</strong>any point (x, t), the analytical domain of dependence must be included in the limsupof numerical domains of dependence, that isD d (x, t) ⊆ Dd(x, 0 t) = lim sup Dd ∆ (x, t). (4.26)∆→0


✐✐86 Chapter 4. Convergence theoryFigure 4.1. Analytical <strong>and</strong> numerical domain of dependenceMore explicitly, the set Dd 0 (x, t) at the right-h<strong>and</strong> side of (4.26) is the set of all points<strong>for</strong> which any open neighbourhood contains points of Dd ∆ (x, t) <strong>for</strong> some value of ∆,as ∆ → 0.The proof of the CFL condition is very simple: if (4.26) is not satisfied, thenthere exist subsets of D d (x, t) which have no intersection with Dd ∆ (x, t), at least as∆ → 0. There<strong>for</strong>e, since changing the value of v 0 in these subsets changes the valueof v(x, t) but not the value of the numerical approximation, the scheme cannot beconvergent.On the other h<strong>and</strong>, by the equivalence theorem (which was stated about thirtyyears later) we can interpret this situation in more general terms. Since a consistent,convergent scheme is necessarily stable, if a consistent scheme violates the CFLcondition, then this scheme is unstable.4.3.2 Monotonicity <strong>and</strong> L ∞ stabilityFor a number of differential problems (<strong>and</strong> in particular <strong>for</strong> the model problemswhich have been considered in the previous chapter), two important qualitativeproperties are satisfied:• If u 0 is changed by adding a constant c, the corresponding solution is u(x, t)+c.In particular, in lack of source terms, constant solutions are preserved, thatis, if u 0 ≡ c, then u(x, t) ≡ c <strong>for</strong> any x <strong>and</strong> t > 0.• starting from two different initial conditions u 0 <strong>and</strong> w 0 such that u 0 (x) ≥w 0 (x), this ordering is preserved also <strong>for</strong> t > 0, that is, u(x, t) ≥ w(x, t);This makes it natural to require similar properties from the numerical scheme.In fact, we will see that this is a common, although very restrictive, notion ofstability in the nonlinear case.The first property is usually expressed in terms of invariance with respect tothe addition of constants, as follows:Definition 4.13. The scheme S is said to be invariant with respect to the additionof constants ifS(V + c) = S(V ) + c (4.27)<strong>for</strong> any vector V , where the sum of a vector with the constant c is to be intendedcomponent by component.


✐✐4.3. More on stability 87It is clear that, in order to preserve constants, (4.27) should be complementedwith the assumption that S(0) = 0.In particular, when the scheme is linear, the matrix B = (b ij ) must satisfyS i (V + c) = BV + Bc + G= S(V ) + c ∑ jb ij<strong>and</strong> there<strong>for</strong>e (4.27) is satisfied if <strong>and</strong> only if ∑ j b ij = 1 <strong>for</strong> any i (note that, here<strong>and</strong> in the sequel, the dependence of the vector G on the time step is irrelevant).The concept of monotonicity requires some more words. We first define amonotone scheme:Definition 4.14. The scheme S is said to be monotone if, <strong>for</strong> any ∆,S(∆; V ) − S(∆; W ) ≥ 0 (4.28)<strong>for</strong> any couple of vectors V <strong>and</strong> W such that V − W ≥ 0, this inequality to beintended component by component.Usually, checking monotonicity of a scheme is relatively easy.scheme is linear, we may write <strong>for</strong> the i–th component:First, if theS i (V ) − S i (W ) = ∑ jb ij v j + g i − ∑ jb ij w j − g i == ∑ jb ij (v j − w j )so that monotonicity is satisfied if <strong>and</strong> only if all entries b ij of the matrix B arepositive. If a linear scheme is invariant <strong>for</strong> the addition of constants <strong>and</strong> monotone,then it is stable in l ∞ . In fact, under these assumptions the entries on a row of thematrix B must be positive <strong>and</strong> have unity sum, <strong>and</strong> hence the matrix has unity∞–norm.If the scheme is nonlinear, en<strong>for</strong>cing the conditionS i (V ) − S i (W ) ≥ 0<strong>for</strong> any V ≥ W requires that, <strong>for</strong> any i <strong>and</strong> j,∂∂v jS i (V ) ≥ 0.Note that any (linear or nonlinear) scheme which is monotone <strong>and</strong> preserves constantsis also L ∞ stable. In fact, defining two vectors W <strong>and</strong> W such that w j ≡sup x v 0 (x), w j ≡ inf x v 0 (x), we obtain thatW ≤ V 0 ≤ W , (4.29)<strong>and</strong> due to the two properties mentioned above, successive steps satisfyW = S n (W ) ≤ V n ≤ S n (W ) = W (4.30)


✐✐88 Chapter 4. Convergence theory(where S n denotes the n-iterated composition S ◦ S ◦ · · · ◦ S) so that the numericalsolution is uni<strong>for</strong>mly bounded in L ∞ .More in general, it can be proved that if a scheme is invariant with respect tothe addition of constants <strong>and</strong> is monotone, then it is nonexpansive in L ∞ , that is‖S(V ) − S(W )‖ ∞ ≤ ‖V − W ‖ ∞ . (4.31)In fact, set the constant c = ‖(V − W ) + ‖ ∞ . Then, we also have that V ≤ W + c,so that monotonicity of the scheme implies that<strong>and</strong> there<strong>for</strong>eS(V ) ≤ S(W + c) = S(W ) + cS(V ) − S(W ) ≤ ‖(V − W ) + ‖ ∞which implies (4.31), once the roles of V <strong>and</strong> W are interchanged. Note that, inthe linear case, this is directly implied by the fact that B has unity ∞–norm.4.3.3 Monotonicity <strong>and</strong> Lipschitz stabilityA further property of monotone scheme, which is crucial in treating nonlinear problems,is nonexpansivity in the Lipschitz norm. This characteristic is not a trivialconsequence of monotonicity <strong>and</strong> conservation of constants, since it not only requiresthat L ∞ be conserved, but that Lipschitz constant be conserved too. Thisgap requires an additional assumption of invariance by translation which will be<strong>for</strong>mally stated as follows.Definition 4.15. The scheme S is said to be invariant by translation if, definedthe translation operator Θ i such thatwe have, <strong>for</strong> any i = 1, . . . , d,(Θ i V ) j = V j+ei ,S(∆; Θ i V ) = Θ i S(∆; V ) (4.32)Clearly, this definition requires a uni<strong>for</strong>m infinite space grid. Also, we do notexpect that it could be satisfied when treating variable coefficient equations. Now,define the i–th partial right incremental ratio at the point x j asD i,j [V ] = v j+e i− v j∆x i(i = 1, . . . , d). (4.33)Making use of the operator Θ iright–h<strong>and</strong> side of (4.33) aswe can write the maximum value over j of theIf this computation is per<strong>for</strong>med on S(V ), we have‖D i [V ]‖ ∞ = ‖Θ iV − V ‖ ∞∆x i, (4.34)‖D i [S(V )]‖ ∞ = ‖Θ iS(V ) − S(V )‖ ∞∆x i= ‖S(Θ iV ) − S(V )‖ ∞∆x i, (4.35)


✐✐4.3. More on stability 89so that, applying the property of L ∞ –nonexpansivity of the scheme, we finallyobtain‖D i [S(V )]‖ ∞ ≤ ‖Θ iV − V ‖ ∞∆x i= ‖D i [V ]‖ ∞ . (4.36)There<strong>for</strong>e, since none of the partial incremental ratios is exp<strong>and</strong>ed, this also holds<strong>for</strong> the Lipschitz constant.4.3.4 Von Neumann analysis <strong>and</strong> L 2 stabilityIt is clear that condition (4.12) is strictly related to the eigenvalues of B. Since anymatrix norm is bounded from below by the spectral radius, it would be natural torequire that, <strong>for</strong> a generic eigenvalue ρ m of B, the inequalityρ m (B) ≤ 1 + C∆t (4.37)should hold <strong>for</strong> a stable scheme. Condition (4.37) is usually referred to as VonNeumann stability condition <strong>and</strong> <strong>for</strong> the moment it will be considered a necessarycondition. Indeed, if (4.37) is not satisfied, then successive powers of B cannot beuni<strong>for</strong>mly bounded, <strong>and</strong> the scheme is clearly unstable. Moreover, in general (4.37)is still far from being a sufficient condition, <strong>for</strong> the reasons we will soon explain.Let B be set in Jordan canonic <strong>for</strong>m by a trans<strong>for</strong>mation T , so thatB(∆) = T (∆) −1 Λ(∆) T (∆) (4.38)(here, we have explicitly indicated that all these matrices depend on the discretizationparameters). Now, successive powers of B have the <strong>for</strong>m<strong>and</strong> there<strong>for</strong>eB(∆) n = T (∆) −1 Λ(∆) n T (∆) (4.39)‖B(∆) n ‖ ≤ ∥ ∥ T (∆)−1 ∥ ∥ · ‖Λ(∆) n ‖ · ‖T (∆)‖ . (4.40)Giving a bound on successive powers of B via (4.40) may not be easy, either. Threemain technical difficulties still hold:i) The trans<strong>for</strong>mation T itself depends on ∆, so that it is necessary to furthergive a uni<strong>for</strong>m bound on the condition number K(T ) = ‖T −1 ‖ · ‖T ‖ as afunction of ∆;ii) If Λ has multiple eigenvalues, then the fundamental solutions associated tosuch eigenvalues are of the kind n ν ρ n . Even if |ρ| < 1, it is clear that thesesolutions are asymptotically stable, but may have extrema which are not easyto locate <strong>and</strong> bound;iii) The structure of eigenvectors is unknown in general, <strong>and</strong> this makes it difficultto look <strong>for</strong> eigenvalues.Points i) <strong>and</strong> ii) can be overcome by assuming B to be a normal matrix, thatis, a matrix which could be brought to diagonal <strong>for</strong>m by means of an orthonormaltrans<strong>for</strong>mation T . Actually, if this is the case, then all eigenvalues are simple (<strong>and</strong>hence, the associated fundamental solutions are of the <strong>for</strong>m ρ n ). Moreover, since Tis orthonormal, working in the 2–norm we also have, <strong>for</strong> any ∆,K 2 (T (∆)) = ∥ ∥ T (∆)−1 ∥ ∥2‖T (∆)‖ 2= 1.


✐✐90 Chapter 4. Convergence theoryTaking also into account that the 2–norm of a diagonal matrix is given by its spectralradius, we get there<strong>for</strong>e‖B(∆) n ‖ 2≤ K 2 (T (∆)) ‖Λ(∆) n ‖ 2= sup |ρ m (B(∆)) n | , (4.41)m<strong>and</strong> condition (4.37) becomes also sufficient.Concerning point iii), if one restricts to circulating matrices, which are a subclass ofnormal matrices, then the structure of eigenvectors is also known. More precisely, ifn n is the number of nodes <strong>and</strong> B ∈ R nn×nn is a circulating matrix, any eigenvectorV has components of the <strong>for</strong>m v j = z j , with z a n n –th root of the unity. Writingmore explicitly such root of the unity in the <strong>for</strong>m z = e imθ with θ = 2π/n n <strong>and</strong>m = 0, 1, . . . , n n − 1, we getv j = e i 2πmjnn , (4.42)<strong>and</strong> using this <strong>for</strong>m in the relationshipρv j = ∑ lb jl v l , (4.43)which defines the eigenvalues, we obtainρ m e i 2πmjnn= ∑ lb jl e i 2πmlnn (4.44)that is, <strong>for</strong> any j,ρ m = ∑ lb jl e i 2πm(l−j)nn . (4.45)In the literature, the eigenvalue ρ m is also referred to as the amplification factorassociated to the harmonic component m. It depends on ∆ via the coefficients b jl .Assuming that B is a circulating matrix is really restrictive, but still correspondsto a real (in a sense, the simpler) situation of application of a given scheme.In particular, most schemes can be brought to this <strong>for</strong>m when applied to constantcoefficient equations in one space dimension, with periodic boundary conditions <strong>and</strong>with constant space discretization step. This is precisely the framework in whichVon Neumann condition (4.37) becomes sufficient <strong>for</strong> L 2 stability. Actually, thiscan be the only practical stability analysis <strong>for</strong> a number of schemes, <strong>and</strong> is widelyreputed to give at least an indication <strong>for</strong> more general cases.In general, we may not be really interested in computing the eigenvalue ρ mrelated to a specific harmonic component, but rather to prove the bound (4.37).Since (4.45) shows that there exists a mapping between the boundary of the unitydisk of C <strong>and</strong> a curve (always in C) containing the eigenvalues, it is possible toneglect the position of the specific eigenvalue on this curve, <strong>and</strong> rather focus on thefact that this curve itself be contained in the unity disc. On the other h<strong>and</strong>, if onekeeps the discretization steps constant <strong>and</strong> increases the number of nodes n n (thismeans enlarging the interval over which the equation is posed), then it becomesclear that the roots of the unity become denser on the original curve, <strong>and</strong> there<strong>for</strong>ethe eigenvalues become denser on the trans<strong>for</strong>med curve. This makes it reasonableto treat the phaseω = 2πmn n∈ [0, 2π] (4.46)


✐✐4.4. Convergence results <strong>for</strong> <strong>Hamilton</strong>–Jacobi equations 91as a continuous parameter, <strong>and</strong> to carry out the stability analysis on the basis ofthe condition∣ ∑∣∣∣∣|ρ(ω)| =b jl (∆) e iω(l−j) ≤ 1 + C∆t. (4.47)∣lIn this <strong>for</strong>m, the Von Neumann analysis is suitable to treat both the case of periodiccondition <strong>and</strong> the case posed on the whole of R (which is obtained as the limit case<strong>for</strong> n n → ∞).Remark 4.16. Clearly, the eigenvalues of the matrix B(∆) could be estimatedby means of Gershgorin’s theorem, but condition (4.37), when ρ m (B) is estimatedwith respectively row or column sums, becomes equivalent to (4.13) written withrespectively the ∞– or the 1–norm of B.4.4 Convergence results <strong>for</strong> <strong>Hamilton</strong>–JacobiequationsWe move in this section to the study of the convergence theory <strong>for</strong> the convex HJequation,{u t (x, t) + H(Du(x, t)) = 0 (x, t) ∈ R d × [0, T ],(4.48)u(x, 0) = u 0 (x).Once out of the framework of linear equations (<strong>and</strong> linear schemes), no convergenceresult of the same generality of Lax–Richtmeyer theorem exists. In practice, whileconsistency of a scheme can be defined in essentially the same way <strong>for</strong> a nonlinearscheme, stability becomes a more subtle topic, since uni<strong>for</strong>m boundedness ofnumerical solutions does not suffice in general <strong>for</strong> convergence.There<strong>for</strong>e, different nonlinear stability concepts have been developed <strong>for</strong> nonlinearequations <strong>and</strong> schemes. In the framework of <strong>Hamilton</strong>–Jacobi equations twomain concepts have been proposed: the more classical idea of monotone stability(which leads to the convergence theorems of Cr<strong>and</strong>all–Lions <strong>and</strong> Barles–Souganidis)<strong>and</strong> the idea of uni<strong>for</strong>m semiconcavity used in the Lin–Tadmor convergence theorem.We also mention that, in the specific field of Semi-Lagrangian schemes, Lipschitzstability can be an equally useful concept.4.4.1 Cr<strong>and</strong>all–Lions <strong>and</strong> Barles–Souganidis theoremsWe collect together these two results which make use of monotonicity as a stabilityassumption.The Cr<strong>and</strong>all–Lions theorem is inspired by the result of convergence of monotoneconservative schemes <strong>for</strong> conservation laws, there<strong>for</strong>e it assumes the schemeto have a structure which parallels the structure of conservative schemes. On thecontrary, Barles–Souganidis theorem does not assume any particular structure <strong>for</strong>the scheme, <strong>and</strong> is suitable <strong>for</strong> more general situation (including second-order HJequations), provided a comparison principle holds <strong>for</strong> the exact equation. It alsorequires a more technical definition of consistency.


✐✐92 Chapter 4. Convergence theory<strong>Schemes</strong> in differenced <strong>for</strong>m: the Cr<strong>and</strong>all–Lions theoremWe present this result referring to the case of two space dimensions, the extensionto an arbitrary number of dimensions being straight<strong>for</strong>ward. With a small abuse ofnotation, we rewrite equation (4.48) asu t + H(u x1 , u x2 ) = 0. (4.49)Cr<strong>and</strong>all–Lions theorem works in the framework of difference schemes, so we assumethat the space grid is orthogonal <strong>and</strong> uni<strong>for</strong>m, ∆x i being the space step along thei–th direction. We define an approximation of the partial derivative u xi at the pointx j by the right (partial) incremental ratio (4.33), that isD i,j [V ] = v j+e i− v j∆x i(i = 1, 2).In parallel with the definition of schemes in conservative <strong>for</strong>m <strong>for</strong> conservation laws,we define here the class of schemes in differenced <strong>for</strong>m.Definition 4.17. A scheme S is said to be in differenced <strong>for</strong>m if it has the <strong>for</strong>mv n+1j = v n j − ∆tH (D 1,j−p [V n ], . . . , D 1,j+q [V n ]; D 2,j−p [V n ], . . . , D 2,j+q [V n ]) ,(4.50)<strong>for</strong> two multiindices p <strong>and</strong> q with positive components, <strong>and</strong> <strong>for</strong> a Lipschitz continuousfunction H (called the numerical <strong>Hamilton</strong>ian).In practice, (4.50) defines schemes in which the dependence on V n appears onlythrough its finite differences, computed on a rectangular stencil of points aroundx j . The differenced <strong>for</strong>m of a scheme lends itself to an easier <strong>for</strong>mulation of theconsistency condition, which is given in the followingDefinition 4.18. A scheme in differenced <strong>for</strong>m is consistent if, <strong>for</strong> any a, b ∈ R,H(a, . . . , a; b, . . . , b) = H(a, b). (4.51)Remark 4.19. We point out that the previous definition matches the usual one.To show this fact, we rewrite the scheme in the <strong>for</strong>mv n+1j− v n j∆t+ H (D 1,j−p [V n ], . . . , D 1,j+q [V n ]; D 2,j−p [V n ], . . . , D 2,j+q [V n ]) = 0.For the first term, it is clear that, once replaced V n by U(t n ),u(x j , t n+1 ) − u(x j , t n )∆t→ u t (x j , t n ),whereas <strong>for</strong> the second, it is necessary to notice that, <strong>for</strong> i = 1, 2 <strong>and</strong> <strong>for</strong> all fixedmultiindex k,D i,j+k [U(t n )] = u xi (x j , t n ) + O(∆x).There<strong>for</strong>e, using this estimate in the scheme, we obtainH (D 1,j−p [U(t n )], . . . , D 1,j+q [U(t n )]; D 2,j−p [U(t n )], . . . , D 2,j+q [U(t n )]) =


✐✐4.4. Convergence results <strong>for</strong> <strong>Hamilton</strong>–Jacobi equations 93= H (u x1 (x j , t n ), . . . , u x1 (x j , t n ); u x2 (x j , t n ), . . . , u x2 (x j , t n )) + O(∆x),where all the error terms o(∆x) have been collected due to the Lipschitz continuityof H. On the other h<strong>and</strong>, applying (4.51), we obtain at the point (x j , t n ):H(u x1 , . . . , u x1 ; u x2 , . . . , u x2 ) + O(∆x) = H(u x1 , u x2 ) + O(∆x)which correspond to the usual notion of consistency.Also, in the nonlinear case we expect that monotonicity may or may not holddepending on the speed of propagation of the solution, this speed being related tothe Lipschitz constant of u 0 (in fact, in general monotonicity does depend on thespeed of propagation, but in the linear case this speed is given <strong>and</strong> unrelated tou 0 ). We will say that the scheme is monotone on [−R, R] if (4.14) is satisfied <strong>for</strong>any V <strong>and</strong> W such that |D i,j [V ]|, |D i,j [W ]| ≤ R.We have now all elements to state the Cr<strong>and</strong>all–Lions convergence theorem.Theorem 4.20. Let H : R 2 → R be continuous, u 0 in (4.48) be bounded <strong>and</strong>Lipschitz continuous (with Lipschitz constant L) on R 2 , <strong>and</strong> vj 0 = u 0(x j ). Let thescheme (4.50) be monotone on [−(L+1), L+1] <strong>and</strong> consistent, <strong>for</strong> a locally Lipschitzcontinuous numerical <strong>Hamilton</strong>ian H. Then, there exists a constant C such that,<strong>for</strong> any n ≤ T/∆t,∣ vnj − u(x j , t n ) ∣ ≤ C∆t1/2(4.52)<strong>for</strong> ∆t → 0, ∆x i = λ i ∆t, (i = 1, 2).Remark 4.21. We omit the proof, which is very technical. As a pure convergenceresult, Cr<strong>and</strong>all–Lions theory is generalized by the Barles–Souganidis theorem, althoughin the latter result no convergence estimate is obtained.Remark 4.22. In the sequel, this convergence result will be applied to first-order,three-point stencil schemes in one space dimension. In this particular setting, adifferenced scheme takes the simpler <strong>for</strong>m (in which the index i of the variable isclearly omitted)v n+1j = vj n − ∆tH (D j−1 [V n ], D j [V n ]) , (4.53)involving only the points v n j−1 , vn j<strong>and</strong> vn j+1in the computation of vn+1 j .Exploiting the comparison principle: the Barles–Souganidis theoremWhile it still requires monotonicity, Barles–Souganidis convergence theory [BaS91]gives a more abstract <strong>and</strong> general framework <strong>for</strong> convergence of schemes, includingthe possibility of treating second-order, degenerate <strong>and</strong> singular equations. Roughlyspeaking, this theory states that any monotone, stable <strong>and</strong> consistent scheme convergesto the exact solution provided there exists a comparison principle <strong>for</strong> thelimiting equation.The Cauchy problem under consideration is:{u t + H(x, u, Du, D 2 u) = 0 on R d × (0, T ),u = u 0 on R d × {t = 0}(4.54)


✐✐94 Chapter 4. Convergence theoryThe function H : R d ×R×R d ×M d → R (where M d is the space of d×d symmetricmatrices) is a continuous <strong>Hamilton</strong>ian, possibly depending on the Hessian matrixof second derivatives, <strong>and</strong> is assumed to be elliptic, i.e.H(x, w, p, A) ≤ H(x, w, p, B) (4.55)<strong>for</strong> all w ∈ R, x, p ∈ R d , A, B ∈ M d such that A ≥ B, that is, such that A − B isa positive semidefinite matrix. In particular, this <strong>for</strong>m includes the first-order HJequation (4.48).We will assume that a comparison principle holds true <strong>for</strong> (4.54), i.e. we assume thatif u <strong>and</strong> v are respectively a super- <strong>and</strong> a sub-solution of (4.54) on R d × (0, T ) → R,<strong>and</strong> u(·, 0) ≤ v(·, 0), then u ≤ v.We consider a scheme in the general <strong>for</strong>m (4.5). First, we require the property(4.27) of invariance with respect to the addition of constants. Then, a generalizedconsistency condition is assumed as follows.Definition 4.23. Let ∆ m ≡ (∆x m , ∆t m ) be a generic sequence of discretizationparameters, (x jm , t nm ) be a generic sequence of nodes in the space–time grid suchthat, <strong>for</strong> m → ∞,(∆x m , ∆t m ) → 0 <strong>and</strong> (x jm , t nm ) → (x, t). (4.56)Let φ ∈ C ∞ (R d × (0, T ]). Then, the scheme S is said to be consistent iflim infm→∞φ(x jm , t nm ) − S jm (∆ m ; Φ(t nm−1))∆t m≥ φ t (x, t) ++H(x, φ(x, t), Dφ(x, t), D 2 φ(x, t)),φ(x jm , t nm ) − S jm (∆ m ; Φ(t nm−1))lim sup≤ φ t (x, t) +m→∞∆t m+H(x, φ(x, t), Dφ(x, t), D 2 φ(x, t)).(4.57)Here, the index of the sequence is m, j m <strong>and</strong> n m denoting the correspondingindices of a node with respect to the m–th space-time grid, <strong>and</strong> we recall that byΦ or Φ(t) we denote the vector of node values <strong>for</strong> respectively φ(x) <strong>and</strong> φ(x, t).Moreover, H <strong>and</strong> H denote here lower <strong>and</strong> upper semicontinuous envelopes of H:H(x, φ(x, t), Dφ(x, t), D 2 φ(x, t)) =H(x, φ(x, t), Dφ(x, t), D 2 φ(x, t)) =lim inf H(y, φ(y, s), Dφ(y, s),(y,s)→(x,t) D2 φ(y, s)),lim sup H(y, φ(y, s), Dφ(y, s), D 2 φ(y, s)).(y,s)→(x,t)Note that, if H is continuous (<strong>and</strong> this is definitely the case <strong>for</strong> the first-order HJequation (4.48)), then in (4.57) the lim inf <strong>and</strong> the lim sup must coincide, <strong>and</strong> thedefinition reduces to the usual definition of consistency, unless <strong>for</strong> the fact that inprinciple φ needs not be a solution.The st<strong>and</strong>ard definition of monotonicity is also replaced by a generalized monotonicityassumption stated as follows.


✐✐4.4. Convergence results <strong>for</strong> <strong>Hamilton</strong>–Jacobi equations 95Definition 4.24. Let (∆x m , ∆t m ) <strong>and</strong> (x jm , t nm ) be generic sequences satisfying(4.56). Then, the scheme S is said to be monotone (in the generalized sense) if itsatisfies the following conditions:if v jm ≤ φ jm then S jm (∆ m ; V ) ≤ S jm (∆ m ; Φ) + o(∆t m ) (4.58)if φ jm ≤ v jm then S jm (∆ m ; Φ) ≤ S jm (∆ m ; V ) + o(∆t m ). (4.59)<strong>for</strong> any smooth function φ(x).Also in this case, we have that if a scheme is monotone in the usual <strong>for</strong>m(4.14), then it also satisfies (4.58)–(4.59).Now, consider a numerical solution V n (with vjn <strong>and</strong> its piecewise constant (intime) interpolation v ∆t defined as:{v ∆t I[V n ](x) if t ∈ [t n , t n+1 ) ,(x, t) =v 0 (x) if t ∈ [0, ∆t).Here, I[V n ] is assumed to be a general interpolation operatorI[V n ](x) =∑ψ l (x)vl n , (4.60)x l ∈S(x)where {ψ l } is a basis of cardinal functions in R d (which in particular satisfy theproperty ∑ l ψ l(x) ≡ 1), <strong>and</strong> S(x) is the stencil of nodes involved <strong>for</strong> interpolatingat the point x. We assume <strong>for</strong> the moment that it is contained in a ball of radiusO(∆x) around x, <strong>and</strong> delay a detailed treatment of interpolation techniques toChapter 3.The interpolation operator has also to verify a relaxed monotonicity property:if v j ≤ φ j <strong>for</strong> any j such that x j ∈ S(x), then I[V ](x) ≤ I[Φ](x) + o(∆t) (4.61)if φ j ≤ v j <strong>for</strong> any j such that x j ∈ S(x), then I[Φ](x) ≤ I[V ](x) + o(∆t) (4.62)where V <strong>and</strong> φ denote vectors of node values of respectively a generic numericalsolution, <strong>and</strong> of a smooth function φ(x). Moreover, I[·] satisfies|I[Φ](x) − Φ(x)| = o(∆t). (4.63)Note that, once ∆t <strong>and</strong> ∆x are related one another, bounds (4.61)–(4.63) (whichare usually written in terms of the space discretization parameter) may also beunderstood in terms of ∆t.We can now state the extended version of the convergence result given in [BaS91]:Theorem 4.25. Assume (4.27), (4.57) <strong>and</strong> (4.58)–(4.63). Let u(x, t) be the uniqueviscosity solution of (4.54). Then v ∆t (x, t) → u(x, t) locally uni<strong>for</strong>mly on R d ×[0, T ]as ∆ → 0.Proof. Let the bounded functions u, u be defined byu(x, t) =lim sup(y,s)→(x,t)∆t→0v ∆t (y, s), u(x, t) = lim inf(y,s)→(x,t)∆t→0v ∆t (y, s). (4.64)


✐✐96 Chapter 4. Convergence theoryWe claim that u(x, t), u(x, t) are respectively a sub- <strong>and</strong> a super-solution of (4.54).Assume <strong>for</strong> the moment that the claim is true; then by the comparison principleu(x, t) ≤ u(x, t) on R d × (0, T ]. Since the opposite inequality is obvious by thedefinition of u(x, t) <strong>and</strong> u(x, t), we haveu = u = u<strong>and</strong> u is the unique continuous viscosity solution of (4.54). This fact together with(4.64) also imply the locally uni<strong>for</strong>m convergence of v ∆t to u.Let us prove the previous claim. Let (x, t) be a local maximum of u − φ onR d × (0, T ] <strong>for</strong> some φ ∈ C ∞ (R d × (0, T ]). Without any loss of generality, we mayassume that (x, t) is a strict global maximum <strong>for</strong> u − φ <strong>and</strong> that u(x, t) = φ(x, t).Then, by a st<strong>and</strong>ard result from viscosity theory, there exist two sequences ∆t m ∈R + <strong>and</strong> (y m , τ m ) ∈ R d × [0, T ], which are global maximum points <strong>for</strong> v ∆tm − φ, <strong>and</strong>as m → ∞:∆t m → 0, (y m , τ m ) → (x, t), v ∆tm (y m , τ m ) → u(x, t).Then, <strong>for</strong> any x <strong>and</strong> t we havev ∆tm (x, t) ≤ φ(x, t) + ξ m (4.65)with ξ m = ( v ∆tm − φ ) (y m , τ m ) (note that u(x, t) = φ(x, t), <strong>and</strong> hence ξ m → 0).Since, in general, (y m , τ m ) is not a grid point, we need to reconstruct the valueattained by v ∆tm at such points. By the definition of v ∆t , there exist a t nm suchthat τ m ∈ [t nm , t nm+1) <strong>and</strong> v ∆tm (y m , τ m ) = v ∆tm (y m , t nm ). Furthermore, by thedefinition of I[·] in (4.60), there exist a set of nodes S(y m ) such thatI [V nm ] (y m ) =∑x j∈S(y m)ψ j (y m )v nmj . (4.66)Next, we apply (4.65) at t = t nm−1, x = x j ∈ S(y m ) <strong>and</strong> deduce, from (4.27) <strong>and</strong>the monotonicity property (4.58), thatS j(∆m ; V nm−1) ≤ S j (∆ m ; Φ(t nm−1)) + ξ m + o(∆t m ).Recalling that the left-h<strong>and</strong> side is nothing but v nmj , we havewhich yelds, applying (4.60), (4.61):v ∆tm (y m , τ m ) ≤v nmj ≤ S j (∆ m ; Φ(t nm−1)) + ξ m + o(∆t m ),∑x j∈S(y m)ψ j (y m )S j (∆ m ; Φ(t nm−1)) + ξ m + o(∆t m ).Now, by the definition of ξ m , we getφ(y m , τ m ) ≤∑ψ j (y m )S j (∆ m ; Φ(t nm−1)) + o(∆t m ) (4.67)x j∈S(y m)We claim now that φ(y m , τ m ) = φ(y m , t nm )+O(∆t 2 m). In fact, either τ m = t nm(<strong>and</strong> the claim obviously holds), or τ m ∈ (t nm−1, t nm ). In the latter case, since


✐✐4.4. Convergence results <strong>for</strong> <strong>Hamilton</strong>–Jacobi equations 97(v ∆tm − φ)(y m , ·) has a maximum in τ m <strong>and</strong> v ∆tm is constant in (t nm−1, t nm ), thenφ t (y m , τ m ) = 0 <strong>and</strong> we have φ(y m , τ m ) = φ(y m , t nm ) + O(∆t 2 m).Using the previous claim in (4.67), we haveφ(y m , t nm ) ≤∑ψ j (y m )S j (∆ m ; Φ(t nm−1)) + o(∆t m ) (4.68)<strong>and</strong>, by (4.63),x j∈S(y m)φ(y m , t nm ) = I[Φ(t nm )](y m ) + o(∆t m ) =Now, (4.68) <strong>and</strong> (4.69) implylim infm→∞∑x j∈S(y m)∑x j∈S(y m)ψ j (y m )φ(x j , t nm ) + o(∆t m ).ψ j (y m ) φ(x j, t nm ) − S j (∆ m ; Φ(t nm−1))∆t m+ o(1) ≤ 0.Finally, by the consistency property (4.57), we obtain the desired result:φ t (x, t) + H(x, φ(x, t), Dφ(x, t), D 2 φ(x, t)) ≤ 0.(4.69)The proof that u is a super-solution follows the same arguments, unless <strong>for</strong> replacing(4.58) with (4.59). We leave this adaptation to the reader.Note that the consistency condition (4.57) might be re<strong>for</strong>mulated so as to avoidany dependence on the variable t. In fact, adding <strong>and</strong> subtracting φ(x j , t n−1 ), weobtainφ(x j , t n ) − S j (Φ(t n−1 ))= φ(x j, t n ) − φ(x j , t n−1 )+ φ(x j, t n−1 ) − S j (Φ(t n−1 )).∆t∆t∆t(4.70)In the right–h<strong>and</strong> side of (4.70), the first term necessarily converges to φ t (x j , t n ),so that (4.57) is equivalent toH(x, φ(x), Dφ(x), D 2 φ(x)) ≤ lim infm→∞≤ lim supm→∞φ(x jm ) − S jm (∆ m ; Φ)≤ (4.71)∆t mφ(x jm ) − S jm (∆ m ; Φ)∆t m≤ H(x, φ(x), Dφ(x), D 2 φ(x)),<strong>for</strong> a function φ(x) depending on x alone, <strong>and</strong> x jm → x.Once in the <strong>for</strong>m (4.71), the definition of consistency error parallels the correspondingdefinition (4.22) <strong>for</strong> time-marching schemes. In fact, Barles–Souganidistheorem can also be recast to work <strong>for</strong> schemes in the time-marching <strong>for</strong>m (4.18)–(4.19), as follows.Theorem 4.26. Let u(x) be the unique viscosity solution in R d of the equationH(x, u, Du, D 2 u) = 0.Assume that the scheme is in the <strong>for</strong>m (4.18), with S(∆, ·) a contraction. Assumemoreover that (4.27), (4.71), (4.58)–(4.63) hold, <strong>and</strong> that L S = 1 − C∆t + o(∆t).Then, I[V ](x) → u(x) locally uni<strong>for</strong>mly on R d as ∆ → 0.


✐✐98 Chapter 4. Convergence theory4.4.2 Lin–Tadmor theorem <strong>and</strong> semi–concave stabilityIn Lin–Tadmor convergence theory, a different concept of stability is singled out.Once more, this appears to be an adaptation to HJ equations of a parallel resultobtained <strong>for</strong> conservation laws: as Cr<strong>and</strong>all–Lions theorem derives from the classicalconvergence theory of monotone conservative schemes, Lin–Tadmor theorem derivesfrom the Lip ′ –stability theory.We start by giving the main concept of semi-concave stability:Definition 4.27. A family of approximate solutions u ɛ of (4.48) is said to besemi-concave stable if there exists a function k(t) ∈ L 1 ([0, T ]) such that<strong>for</strong> t ∈ [0, T ].D 2 u ɛ (x, t) ≤ k(t)I (4.72)More explicitly, condition (4.72) means that D 2 u ɛ (x, t) − k(t)I is negativesemidefinite, that is (since we are dealing with symmetric matrices), that all eigenvaluesof D 2 u are bounded from above by k(t). The core of the theory is a firstabstract result of convergence <strong>for</strong> perturbed semi-concave solutions.Theorem 4.28. Consider problem (4.48) <strong>for</strong> a semi-concave initial condition u 0with compact support, <strong>and</strong> assume the family u ɛ is semi-concave stable. Define thetruncation error associated to u ɛ asThen, <strong>for</strong> any t ∈ [0, T ],F (x, t) = u ɛ t(x, t) + H(Du ɛ (x, t)). (4.73)‖u(t) − u ɛ (t)‖ L 1 (R d ) ≤ C 1 ‖u 0 − u ɛ (0)‖ L 1 (R d ) + C 2 ‖F ‖ L 1 (R d ×[0,T ]). (4.74)Proof. Let the error be defined bye(x, t) = u(x, t) − u ɛ (x, t).By (4.48) <strong>and</strong> (4.73), the error satisfies the advection equation{e t (x, t) + G(x, t) · De(x, t) = F (x, t),e(x, 0) = u 0 (x) − u ɛ (x, 0)(4.75)where the advecting speed G is defined as the averageG(x, t) =∫ 10∂H (ηDu(x, t) + (1 − η)Du ɛ (x, t) ) dη. (4.76)∂pIn order to estimate the norm ‖e(T )‖ of the error at a given time T , we considernow the dual equation,{ψ t (x, t) + div ( G(x, t)ψ(x, t) ) = 0,(4.77)ψ(x, T ) = ψ T (x),


✐✐4.4. Convergence results <strong>for</strong> <strong>Hamilton</strong>–Jacobi equations 99where ψ(T ) is smooth <strong>and</strong> has support in a compact set Ω(T ). From (4.75) <strong>and</strong>(4.77), using Green’s <strong>for</strong>mula, the L 2 scalar product ( e(t), ψ(t) ) satisfies∫d ( )e(t), ψ(t) + e(x, t)ψ(x, t)G(x, t) · n ds = ( F (t), ψ(t) )dt∂Ω(t)with Ω(t) a set depending on t, including the support of ψ(x, t) at T = t. Hence,the boundary integral vanishes at T , <strong>and</strong> we obtain∫( ) ( ) T( )e(T ), ψT = e(0), ψ(0) + F (t), ψ(t) dt ≤0≤ ‖e(0)‖ L 1 (R d )‖ψ(0)‖ L ∞ (R d ) + ‖F ‖ L 1 (R d ×[0,T ])which implies, via Hölder’s inequality,‖ψ(0)‖ L∞ (R‖e(T )‖ L1 (R d ) ≤ supd )‖e(0)‖ψ T‖ψ T ‖L1 (R d ) +L∞ (R d )supt∈[0,T ]‖ψ(t)‖ L ∞ (R d ),sup t∈[0,T ] ‖ψ(t)‖ L∞ (R+ supd )‖F ‖ψ T‖ψ T ‖L1 (R d ×[0,T ]), (4.78)L∞ (R d )<strong>and</strong> hence (4.74), as soon as we are able to bound on ‖ψ(t)‖ ∞ over the interval[0, T ], in terms of the final condition ψ T .To this end, we consider again (4.77), which givesddt ‖ψ(t)‖ L ∞ (R d ) + ( div G(x 0 , t) ) ‖ψ(t)‖ L∞ (R d ) = 0. (4.79)To derive (4.79), we multiply (4.77) by the sign of ψ(x, t) <strong>and</strong> denote by x 0 themaximum point of |ψ| at the time t. By Gronwall’s inequality, we obtain from(4.79), <strong>for</strong> t ∈ [0, T ]:( ∫ )T( )‖ψ(t)‖ L ∞ (R d ) ≤ ‖ψ T ‖ L ∞ (R d ) exp sup div G(x, s) ds , (4.80)xwhere by the definition (4.76), the divergence of G reads⎛⎞∫ 1div G = ⎝η ∑ ∂ 2 H ∂ 2 u+ (1 − η) ∑ ∂ 2 H ∂ 2 u ɛ⎠ dη. (4.81)0 ∂pi,j i ∂p j ∂x i ∂x j ∂pi,j i ∂p j ∂x i ∂x j(note that, in (4.81), arguments have been omitted since we are looking <strong>for</strong> a uni<strong>for</strong>mbound). The first sum in (4.81) may be estimated as∑ ∂ 2 H ∂ 2 u= trace (D 2 H D 2 u) =∂p i ∂p j ∂x i ∂x ji,jt= trace ( (D 2 H) 1/2 D 2 u (D 2 H) 1/2) ≤≤ d · k(s) · β.Here, we have used the facts that eigenvalues of D 2 u are bounded from above byk(t), <strong>and</strong> eigenvalues of D 2 H are bounded from above by β, <strong>and</strong> that the trace is


✐✐100 Chapter 4. Convergence theorythe sum of eigenvalues. Since (unless <strong>for</strong> replacing u with u ɛ ) a similar estimateholds <strong>for</strong> the second sum in (4.81), we finally obtain()‖ψ(t)‖ L ∞ (R d ) ≤ ‖ψ T ‖ L ∞ (R d ) expwhich proves (4.78), <strong>and</strong> there<strong>for</strong>e (4.74).dβ∫ T0k(s)dsIn practice, this result needs some refinement in order to be applied to a numericalscheme. In particular, the concept of semi-concave stability must be weakened<strong>and</strong> the measure of the local truncation error should be made more explicit.Discrete semi-concave stability The assumption of semi-concave stability, as statedby (4.72), is too strict <strong>for</strong> numerical solutions produced by a real scheme. In fact,a reconstruction of the solution, typically in piecewise polynomial <strong>for</strong>m, may notsatisfy the upper bound on second derivatives at the interface between two cells ofthe grid, even in a monotone scheme. To overcome this problem, a discrete counterpartof the semi-concave stability is defined <strong>for</strong> the numerical solution v ∆ by abound on the second directional incremental ratios in the <strong>for</strong>m:v ∆ (x + δ, t) − 2v ∆ (x, t) + v ∆ (x − δ, t)‖δ‖ 2 ≤ k(t). (4.82)Here, the function k(t) ∈ L 1 ([0, T ]) plays the same role as in the original definition,<strong>and</strong> δ is a vector whose norm should remain bounded away from zero, more precisely‖δ‖ ≥ C∆x. (4.83)It is possible to prove (see [LT01]) that (4.82)–(4.83) imply that each numericalsolution is close to a function of a family satisfying (4.72), so that theorem 4.28may be applied.Computation of the truncation error The second ingredient of the Lin–Tadmorconvergence theorem, the L 1 estimation of the truncation error, may also be difficultin general. In particular, (4.73) shows a somewhat reversed approach in measuringthe truncation error, that is, by plugging the numerical solution into the exactequation.An easier expression can be derived <strong>for</strong> Godunov-type schemes (as Semi-Lagrangian schemes are), taking into account that in this case numerical errorsare generated only by the projection step, <strong>and</strong> not by the evolution operator, whichis in principle exact. This expression will be introduced later, in the analysis of SLschemes <strong>for</strong> HJ equations.4.5 Numerical diffusion <strong>and</strong> dispersionConvergence analysis is concerned with the asymptotic behaviour of a scheme. Inparticular, its endpoint is the possibility of generating numerical solutions as closeas necessary to the exact one, although, in practice, we might be more interestedin examining the qualitative response of the scheme <strong>for</strong> a finite, practical value ofthe discretization parameters. In a different <strong>and</strong> complementary analysis, we regard


✐✐4.5. Numerical diffusion <strong>and</strong> dispersion 101Figure 4.2. Advection by a diffusive (left) or dispersive (right) scheme.the values vjn of the numerical solutions as samples of a function v(x, t) defined onR × [0, T ], so that vjn = v(x j, t n ). The function v is characterized in turn as thesolution of a Partial Differential Equation obtained by a perturbation of the exactPDE. This perturbed equation is known as modified equation.Typically, such an analysis is per<strong>for</strong>med on the constant coefficient advectionequation, <strong>and</strong> as a result the modified equation usually takes the <strong>for</strong>mv t + cv x = ν ∂kv + o(ν). (4.84)∂xk The coefficient ν = ν(∆) depends on the discretization parameters <strong>and</strong>, <strong>for</strong> convergencereasons, one expects that ν(∆) → 0 with the same order of the consistencyerror. Note that, in (4.84), we have made explicit the relevant perturbation, whereasthe o(ν(∆)) accounts <strong>for</strong> all the higher-order terms. In fact, we do not expect ingeneral that a finite number of perturbing terms could suffice to completely characterizethe function v.For k = 2 the modified equation appears as an advection–diffusion equation,<strong>and</strong> in this case we term the behaviour of the scheme as diffusive, <strong>and</strong> the coefficientν as the numerical viscosity (a positive number in stable schemes). If k > 2,the scheme is termed as dispersive. We show in Fig. 4.2 the typical response ofrespectively a diffusive <strong>and</strong> a dispersive scheme on the advection of a characteristicfunction. In the first case, since the fundamental solution is represented by theadvection of a gaussian kernel (which is positive), the modified equation satisfies amaximum principle – this behaviour is in fact characteristic of monotone schemes.In the second case, the fundamental solution is no longer positive <strong>and</strong> this resultsin the occurrence of under- <strong>and</strong> overshoots.Rather then outlining a general procedure <strong>for</strong> obtaining the modified equation,we will delay its practical construction to future examples. In difference schemes,the usual technique is to express the numerical samples vjn as Taylor expansionson v, but this technique is unsuitable <strong>for</strong> Semi-Lagrangian schemes. To per<strong>for</strong>ma diffusion/dispersion analysis on SL schemes, we derive there<strong>for</strong>e an analogous ofthe representation <strong>for</strong>mula (1.3) <strong>for</strong> equation (4.84). Retracing the proof of (1.3),we can writed∂k[v(ξ + ct, t)] = ν v(ξ + ct, t) + o(ν)dt ∂xk


✐✐102 Chapter 4. Convergence theoryso that integrating between t n <strong>and</strong> t n+1 we obtain∫ tn+1∂ kv(x j , t n+1 ) = v(x j − c∆t, t n ) + νt n∂x k v(x j − c(t n+1 − t), t)dt + o(ν∆t),<strong>and</strong> since, by a zeroth order approximation,∫ tn+1∂ kνt n∂x k v(x j − c(t n+1 − t), t)dt = ν∆t ∂k∂x k v(x j − c∆t, t n ) + O(ν∆t 2 ),we finally getv(x j , t n+1 ) = v(x j − c∆t, t n ) + ν∆t ∂k∂x k v(x j − c∆t, t n ) + o(ν∆t). (4.85)In practice, in the operation of a SL scheme, the term associated to the k–th spacederivative of v is generated by the interpolation error, <strong>and</strong> this makes it possible torecover all terms of the modified equation, as it will be seen later on.4.6 Commented referencesThe Lax–Richtmeyer equivalence theorem has been first published in [LR56], thuscompleting a theoretical study which had started from the work of Courant, Friedrichs<strong>and</strong> Lewy (written in 1928, reprinted <strong>and</strong> translated into English in 1967 [CFL67]).By that time, the concept of stability had already been extensively studied <strong>and</strong>singled out as a crucial requirement <strong>for</strong> convergence of schemes, along with variousstability criteria, including Von Neumann condition [VR50]. Good reviews onthe classical theory of convergence <strong>for</strong> difference schemes (including the study ofmodified equations) have been given in [RM67], <strong>and</strong>, more recently, in [Str89].In the field of HJ equations, convergence results <strong>for</strong> monotone schemes havebeen <strong>for</strong>mulated at the very start of the theory of viscosity solutions. In additionto the convergence theorem, Cr<strong>and</strong>all–Lions paper [CL84] also adapts to differenceschemes the results of L ∞ -nonexpansivity <strong>and</strong> of Lipschitz stability <strong>for</strong> monotoneschemes which had been first proved in a simple <strong>and</strong> abstact way in [CT80].Barles–Souganidis theorem appears in its first version in [Sou85a], but has beenrefined in [BaS91] to include second-order equations, like the equation of MeanCurvature Motion. We have chosen to present a version of the theorem more suitable<strong>for</strong> real schemes [CFF10], since the original has a more abstract <strong>for</strong>mulation. Finally,Lin–Tadmor convergence theory has been proposed in [LT01], <strong>and</strong> is inspired by asimilar result obtained <strong>for</strong> conservation laws [T91].


✐✐Chapter 5First-order approximationschemesThis chapter is devoted to a comparison of the basic SL scheme (proposed byCourant, Isaacson <strong>and</strong> Rees in 1952) with upwind <strong>and</strong> centered (Lax–Friedrichs)schemes. For each of the schemes, the presentation includes• Construction of the scheme• Consistency• Stability (CFL condition, monotonicity, Von Neumann analysis, uni<strong>for</strong>m semiconcavity)• Convergence estimates• Numerical viscosityA comparison among the schemes is made, <strong>and</strong> simple numerical examples inone space dimension are presented to support our theoretical study. Note that, <strong>for</strong>the sake of clarity, most of the presentation is made on the model problems <strong>and</strong> inone space dimension, but the way to treat more general cases (including multipledimensions) is sketched at the end of the sections, <strong>and</strong> further extensions of interest<strong>for</strong> applications will be considered in the following chapters. A subsection is alsodevoted to the treatment of boundary conditions.5.1 Treating the advection equationIn this section, we introduce the schemes on the simpler case of the advectionequation. We will start with the basic case of the constant coefficient case,{u t (x, t) + cu x (x, t) = 0 (x, t) ∈ R × [0, T ], c > 0(5.1)u(x, 0) = u 0 (x)<strong>and</strong> generalize the ideas to the case with variable coefficients,{u t (x, t) + f(x, t)u x (x, t) = g(x, t) (x, t) ∈ R × [0, T ]u(x, 0) = u 0 (x).(5.2)Note that the simplified model (5.1) is also useful to per<strong>for</strong>m L 2 (Von Neumann)stability analysis, as well as the study of numerical viscosity.103


✐✐104 Chapter 5. First-order approximation schemes5.1.1 Upwind discretization: the First Order Upwind schemeAs we have seen, in the physical <strong>and</strong> mathematical modeling of advection the advectionspeed (c or f(x, t)) is recognized as the vectorfield which makes the solutionpropagate. A first, natural strategy is there<strong>for</strong>e to let the numerical scheme mimicthis behaviour. More explicitly, upwind schemes are geometrically constructed sothat the numerical domain of dependence follows the direction of propagation ofthe solution.Construction of the schemeAlthough the upwind scheme could be derived in a different framework, like the fullydiscrete structure of conservative schemes, we will sketch its construction here byusing the method of lines, that is, introducing an intermediate semi-discrete version,which in our case will be understood as a <strong>for</strong>mulation in which space derivatives(but not time derivatives) have been replaced by finite differences. In particular,we first focus on the constant coefficient case (5.1), then turn to the more generalcase (5.2).The constant coefficient equation Starting with the constant-coefficient equation(5.1), <strong>and</strong> approximating the space derivative u x (x j , t) with the incremental ratiobetween x j−1 <strong>and</strong> x j , the semi-discrete scheme computed at x j reads˙v j (t) + c v j(t) − v j−1 (t)∆x= 0. (5.3)A further discretization in time of (5.3) by means of the <strong>for</strong>ward Euler method givesv n+1j− v n j∆t+ c vn j − vn j−1∆x= 0, (5.4)that is,v n+1j = vj n − c∆t [vn∆x j − vj−1n ]. (5.5)Setting λ = c∆t/∆x, the scheme also takes the more explicit <strong>for</strong>mv n+1j = λv n j−1 + (1 − λ)v n j . (5.6)A key point in the first-order upwind scheme (which also causes the presence ofthe term upwind) is that the incremental ratio used to approximate u x should beper<strong>for</strong>med on the correct side of x j . While the importance of this point will bediscussed in relationship with stability, we note that in the previous discussion, thechoice of a left incremental ratio is related to the positive sign of c. A negative signwould rather require to use the right incremental ratio.The variable coefficient equation When applying the upwind strategy to equation(5.2), we must there<strong>for</strong>e discriminate on the basis of the sign of f(x, t): wheneverf(x j , t n ) > 0, the space derivative is replaced by a left incremental ratio, <strong>and</strong>vice versa if f(x j , t n ) < 0 (in case f(x j , t n ) = 0, the equation simply says thatu t (x j , t n ) = g(x j , t n )). It is convenient to consider the upwind scheme in the <strong>for</strong>m(5.4), which immediately gives, <strong>for</strong> f(x j , t n ) > 0,v n+1j− v n j∆t+ f(x j , t n ) vn j − vn j−1∆x= g(x j , t n ), (5.7)


✐✐5.1. Treating the advection equation 105<strong>and</strong> <strong>for</strong> f(x j , t n ) < 0,v n+1j− v n j∆t+ f(x j , t n ) vn j+1 − vn j∆x= g(x j , t n ). (5.8)To sum up, the <strong>for</strong>m of the scheme <strong>for</strong> f(x, t) changing sign is⎧vj n − f(x j , t n ) ∆t [vn∆x j − v n ]j−1 + ∆tg(xj , t n )⎪⎨if f(x j , t n ) > 0v n+1j = vj n + ∆tg(x j, t n ) if f(x j , t n ) = 0Consistency⎪⎩ vj n − f(x j , t n ) ∆t∆x[vnj+1 − vjn ]+ ∆tg(xj , t n ) if f(x j , t n ) < 0.(5.9)In order to analyse consistency, we apply the upwind scheme in the <strong>for</strong>m (5.9) toa solution. For example, if f(x j , t n ) > 0, we estimate a single component of theconsistency error (4.9) asL Upj (∆; t, U(t)) = 1 [u(x j , t + ∆t) − u(x j , t) +∆t+f(x j , t) ∆t]∆x (u(x j, t) − u(x j−1 , t)) − ∆tg(x j , t) == u(x j, t + ∆t) − u(x j , t)∆t−g(x j , t) =+ f(x j , t) u(x j, t) − u(x j−1 , t)∆x−= u t (x j , t) + O(∆t) + f(x j , t)u x (x j , t) + O(∆x) − g(x j , t) == O(∆t) + O(∆x)where the two last displays use the fact that the error in discretizing the space<strong>and</strong> time derivatives is of first order if the solution is smooth (or at least, if ithas bounded second derivatives in space <strong>and</strong> time), <strong>and</strong> that (5.2) is satisfied by asolution. Since an analogous computation can be carried out with f(x j , t n ) > 0, wecan conclude that ∣ ∣∣L Upj (∆; t, U(t)) ∣ ≤ C(∆t + ∆x)<strong>and</strong> there<strong>for</strong>e, passing to a generic normalized Hölder norm (in particular, the norms‖ · ‖ 2 <strong>and</strong> ‖ · ‖ ∞ ), we can state the consistency of the scheme by∥ L Up (∆; t, U(t)) ∥ ∥ ≤ C(∆t + ∆x). (5.10)Remark 5.1. Consistency of the upwind scheme would not require any particularrelationship between the discretization steps, although a balance of the two termssuggests using steps of the same order of magnitude. In practice, this choice is alsorequired <strong>for</strong> stability, as we will soon show.


✐✐106 Chapter 5. First-order approximation schemesFigure 5.1. Numerical domain of dependence <strong>for</strong> the upwind scheme.StabilityIn the upwind scheme, all stability criteria lead to the same condition on the discretizationsteps, at least in the linear case. As we will see, this is no longer true innonlinear situations.CFL condition For simplicity, let us refer to the constant coefficient case with c >0, that is, to the scheme in the version (5.6). The numerical solution vj n at (x j, t n )depends on v n−1j−1 <strong>and</strong> vn−1 j , <strong>and</strong> in practice, stepping back in time, we construct anumerical domain of dependence in the <strong>for</strong>m of a triangle, which is shown in Figure5.1 along with the characteristic passing through (x j , t n ). Due to the choice ofapproximating space derivatives with left finite differences, the triangle st<strong>and</strong>s onthe left of x j – this agrees with the fact that, if c > 0, the characteristic passingthrough (x j , t n ) is located precisely on this side. To fulfill the CFL condition, wemust further impose that the foot of the characteristic x j − cn∆t (which representsthe analytical domain of dependence D d (x j , t n )) should be included in the numericaldomain of dependence [x j − n∆x, x j ]. This happens if cn∆t < n∆x, or moreexplicitlyλ = c∆t ≤ 1. (5.11)∆xClearly, the constant slope 1/c of characteristics in the x–t plane should be replaced,in the case of variable coefficients, by their minimum (positive or negative) slope±1/‖f‖ ∞ , so that in this case the CFL condition takes the <strong>for</strong>m‖f‖ ∞ ∆t∆x(note that ‖ · ‖ ∞ st<strong>and</strong>s here <strong>for</strong> the norm in L ∞ (R × [0, T ])).≤ 1 (5.12)Monotonicity We recall that the general condition <strong>for</strong> the monotonicity of a linearscheme in the <strong>for</strong>m V n+1 = B(∆; t n )V n +G n (∆) (see Chapter 4) is that the entriesof the matrix B(∆; t n ) should be positive <strong>for</strong> any n. In the case of the upwind


✐✐5.1. Treating the advection equation 107scheme, <strong>and</strong> assuming that f(x j , t n ) > 0, the only nonzero entries on the j–th rowareb Upj,j−1 (∆; t n) = f(x j , t n ) ∆t∆xb Upj,j (∆; t n) = 1 − f(x j , t n ) ∆t∆x(note that they have unity sum so that the scheme is also invariant with respect tothe addition of constants). Since the first one is necessarily positive, the scheme ismonotone provided 1 − f(x j , t n ) ∆t∆x> 0 <strong>for</strong> any j <strong>and</strong> n, this being satisfied if‖f‖ ∞ ∆t∆x< 1. (5.13)Repeating the computations at points where f(x j , t n ) < 0, we obtain that monotonicityis ensured under the same condition.Von Neumann analysis We refer here to the scheme in the <strong>for</strong>m (5.6). Given aneigenvector V of the scheme, in the <strong>for</strong>m v j = e ijω , the condition which characterizesρ as an eigenvalue readsρv j = λv j−1 + (1 − λ)v j ,that is,ρ = (1 − λ) + λ cos ω + iλ sin ωSumming up, eigenvalues are located on a circle of the complex plane centered at1 − λ <strong>and</strong> with radius λ (note that this is precisely the boundary of the Gershgorindiscs of the matrix B Up ). There<strong>for</strong>e, it results that <strong>for</strong> λ ∈ [0, 1] the eigenvalues arecontained in the unity disc, <strong>and</strong> the Von Neumann condition is satisfied. Fig. 5.2shows the eigenvalues of B Up <strong>for</strong> different values of λ ∈ [0, 1].Convergence estimatesFinally, we summarize the convergence issues in a theorem. Taking into accountthe consistency estimate (5.10), as well as the monotonicity condition (5.13), whichalso ensures l ∞ stability, we have at last the followingTheorem 5.2. Let f, g ∈ W 1,∞ , u be the solution of (5.2) <strong>and</strong> vjn be defined by(5.9). Then, <strong>for</strong> any j ∈ Z <strong>and</strong> n ∈ [1, T/∆t],∣∣v j n − u(x j , t n ) ∣ → 0 (5.14)as ∆t, ∆x → 0, with the CFL condition (5.13).Moreover, if u has bounded second derivatives in space <strong>and</strong> time, then‖V n − U(t n )‖ ∞≤ C∆x. (5.15)Numerical viscosityIn order to derive the modified equation, we start from the Taylor expansionsv(x j , t n+1 ) = v + ∆tv t + ∆t22 v tt + O(∆t 3 )v(x j−1 , t n ) = v − ∆xv x + ∆x22 v xx + O(∆x 3 ),


✐✐108 Chapter 5. First-order approximation schemes10.80.60.40.20−0.2−0.4−0.6−0.8−1−1 −0.5 0 0.5 1Figure 5.2. Eigenvalues of the upwind scheme <strong>for</strong> λ = 0.2, 0.6, 1 <strong>and</strong> 50 nodes.where we have kept the convention of omitting the arguments <strong>for</strong> functions computedat (x j , t n ). Plugging the expansions in the <strong>for</strong>m (5.6) of the scheme, weobtainv+∆tv t + ∆t22 v tt+O(∆t 3 ) = c∆t] ([v − ∆xv x + ∆x2∆x2 v xx + O(∆x 3 ) + 1 − c∆t )v∆x<strong>and</strong> hence, simplifying the terms in v, dividing by ∆t <strong>and</strong> taking into account thatdiscretization steps vanish at the same rate,v t + cv x = 1 2 (c∆xv xx − ∆tv tt ) + O(∆x 2 ). (5.16)On the other h<strong>and</strong>, deriving this equation first with respect to t, then with respectto x, <strong>and</strong> eliminating the mixed term v xt , we obtainwhich, plugged into (5.16), gives at lastv tt = c 2 v xx + O(∆x), (5.17)v t + cv x = c∆x2 (1 − λ)v xx + O(∆x 2 ). (5.18)5.1.2 Central discretization: the Lax–Friedrichs schemeThe consistency rate of the upwind scheme is mostly limited by the fact that, usingonly two points in discretizing the space derivative, the approximation cannot gobeyond the first order. An attempt to increase the consistency rate would there<strong>for</strong>eimply the use of a larger stencil of points. In the class of central schemes, thisis typically accomplished by enlarging the stencil symmetrically. As we will see,usually this enlargement also causes a more dispersive behaviour of the method.


✐✐5.1. Treating the advection equation 109Construction of the schemeThe simplest situation <strong>for</strong> a symmetric stencil is the use of x j−1 , x j e x j+1 . In thiscase, the space derivative u x (x j , t) can be approximated asu x (x j , t) = u(x j+1, t) − u(x j−1 , t)2∆x+ O(∆x 2 ). (5.19)In principle, this larger stencil allows to increase accuracy, although this result isnot achieved in the Lax–Friedrichs scheme.The constant coefficient equation A first, naive way of using (5.19) into (5.1)would lead to the semi-discrete scheme˙v j (t) + c v j+1(t) − v j−1 (t)2∆x= 0.However, an explicit one-step time discretization of this scheme, <strong>and</strong> in particularthe one obtained by applying the <strong>for</strong>ward Euler scheme,v n+1j= vj n − c∆t [vn2∆x j+1 − vj−1n ],results in an unstable scheme. The classical way to stabilize the scheme is either toturn to a midpoint two-step time approximation (this leads to the so-called leapfrogscheme), or to modify it in the Lax–Friedrichs <strong>for</strong>mv n+1j = vn j−1 + vn j+1− c∆t [vn2 2∆x j+1 − vj−1n ], (5.20)which clearly cannot be derived via the method of lines, although it also falls inthe class of conservative schemes. In terms of the parameter λ = c∆t/∆x, Lax–Friedrichs scheme may also be rewritten asv n+1j= 1 + λ2vj−1 n + 1 − λ v n2j+1. (5.21)Note that, in this case, no assumption on the sign of c is necessary.The variable coefficient equation Passing to the variable coefficient equation iseasier in Lax–Friedrichs scheme, since it is no longer required to take into accountthe sign of the advection term. By analogy with the upwind scheme, with a straight<strong>for</strong>wardmodification of the constant coefficient case we finally obtain the <strong>for</strong>mv n+1j = vn j−1 + vn j+1− f(x j , t n ) ∆t [vn22∆x j+1 − vj−1n ]+ ∆tg(xj , t n ). (5.22)ConsistencyBe<strong>for</strong>e applying the definition of consistency, let us remark that, <strong>for</strong> a smoothsolution u(x, t), by elementary interpolation arguments we haveu(x j−1 , t n ) + u(x j+1 , t n )2= u(x j , t n ) + O(∆x 2 ). (5.23)


✐✐110 Chapter 5. First-order approximation schemesThen, applying the Lax–Friedrichs scheme (5.22) to a solution <strong>and</strong> using (5.23), weestimate a component of the consistency error (4.9) asL LFj (∆; t, U(t)) = 1 [u(x j , t + ∆t) − u(x j , t) + O(∆x 2 ) +∆t+f(x j , t) ∆t]2∆x (u(x j+1, t) − u(x j−1 , t)) − ∆tg(x j , t) == u(x ( )j, t + ∆t) − u(x j , t) ∆x2+ O +∆t∆t+f(x j , t) u(x j+1, t) − u(x j−1 , t)− g(x j , t) =2∆x( ) ∆x2= u t (x j , t) + O(∆t) + O +∆t+f(x j , t)u x (x j , t) + O(∆x 2 ) − g(x j , t) =( ) ∆x2= O(∆t) + O∆twhere only the relevant terms have been considered. The consistency estimate inHölder norm <strong>for</strong> the Lax–Friedrichs scheme reads there<strong>for</strong>e∥ L LF (∆; t, U(t)) ∥ ( )≤ C ∆t + ∆x2 . (5.24)∆tRemark 5.3. Note that, as in the case of the upwind scheme, the balance of thetwo terms requires steps of the same order of magnitude, <strong>and</strong> this choice is also fine<strong>for</strong> stability reasons. However, in the case of the LF scheme, setting ∆t of the orderof ∆x 2 or smaller would result in a nonconsistent scheme.StabilityAnalysis of the Lax–Friedrichs scheme also leads to stability conditions which coincideamong the various stability frameworks.CFL condition Referring again to the constant coefficient case, <strong>and</strong> to the schemein the version (5.21), the numerical solution v n+1j at (x j , t n+1 ) depends on vj−1 n <strong>and</strong>vj+1 n . This causes a numerical domain of dependence in the <strong>for</strong>m shown in Fig. 5.3.Note that the triangle is symmetrical (so that no special care should be taken aboutthe sign of the velocity c) <strong>and</strong> enlarges itself at the same rate as <strong>for</strong> the upwindscheme, so that we satisfy the CFL condition under the same relationship|λ| = |c|∆t∆x≤ 1. (5.25)As in the case of the upwind scheme, the condition may be extended to variablecoefficient equations by replacing the constant advection speed c with the norm‖f‖ ∞ , so that the requirement becomes‖f‖ ∞ ∆t∆x≤ 1. (5.26)


✐✐5.1. Treating the advection equation 111Figure 5.3. Numerical domain of dependence <strong>for</strong> the Lax–Friedrichs scheme.Monotonicity We apply again the general condition <strong>for</strong> the monotonicity of alinear scheme. In the the Lax–Friedrichs scheme, the only nonzero entries on thej–th row areb LFj,j−1(∆; t n ) = 1 (1 + f(x j , t n ) ∆t )2∆xb LFj,j+1(∆; t n ) = 1 (1 − f(x j , t n ) ∆t ).2∆xOnce again, they have unity sum, <strong>and</strong> one of the two is necessarily positive. There<strong>for</strong>e,the scheme is invariant <strong>for</strong> the addition of constants, <strong>and</strong> is monotone providedthe other term is positive <strong>for</strong> any j. This leads to the same condition obtained <strong>for</strong>the upwind scheme, that is‖f‖ ∞ ∆t≤ 1. (5.27)∆xVon Neumann analysis We refer here to the scheme in the <strong>for</strong>m (5.21). Given aneigenvector V of the scheme, an eigenvalue ρ satisfies the conditionρv j = 1 + λ2that is, using the <strong>for</strong>m v j = e ijω ,v j−1 + 1 − λ v j+12ρe ijω = 1 + λ e i(j−1)ω + 1 − λ e i(j+1)ω22which gives the eigenvalue ρ asρ = 1 + λ e −iω + 1 − λ e iω =22= cos ω − iλ sin ω.In practice, eigenvalues are located on an ellipse centered at the origin, with unityhorizontal semi-axis <strong>and</strong> vertical semi-axis λ. The ellipse itself <strong>and</strong> the eigenvaluesare there<strong>for</strong>e contained in the unity disc if λ ∈ [0, 1]. Fig. 5.4 shows the eigenvaluesof B LF <strong>for</strong> different values of λ.


✐✐112 Chapter 5. First-order approximation schemes10.80.60.40.20−0.2−0.4−0.6−0.8−1−1 −0.5 0 0.5 1Figure 5.4. Eigenvalues of the Lax–Friedrichs scheme <strong>for</strong> λ = 0.2, 0.6, 1<strong>and</strong> 50 nodes.Convergence estimatesAgain, we summarize the convergence study in a theorem. Consistency followsfrom (5.24), once assumed that ∆x = o(∆t 1/2 ), <strong>and</strong> stability from the monotonicitycondition (5.27). As <strong>for</strong> the upwid scheme, the smoothness required on the solutionto achieve first-order convergence is to have bounded second derivatives in space<strong>and</strong> time, but also that ∆x ∼ ∆t. We can there<strong>for</strong>e state the convergence result as:Theorem 5.4. Let f, g ∈ W 1,∞ , u be the solution of (5.2) <strong>and</strong> vjn be defined by(5.22). Then, <strong>for</strong> any j ∈ Z <strong>and</strong> n ∈ [1, T/∆t],∣ vnj − u(x j , t n ) ∣ → 0 (5.28)as ∆t, ∆x → 0, with ∆x = o(∆t 1/2 ) <strong>and</strong> the CFL condition (5.27).Moreover, if u has bounded second derivatives in space <strong>and</strong> time <strong>and</strong> ∆x = c∆t <strong>for</strong>some constant c, then‖V n − U(t n )‖ ∞≤ C∆x. (5.29)Numerical viscosityFirst, we compute the Taylor expansionsv(x j , t n+1 ) = v + ∆tv t + ∆t22 v tt + O(∆t 3 )v(x j−1 , t n ) = v − ∆xv x + ∆x22 v xx + O(∆x 3 )v(x j+1 , t n ) = v + ∆xv x + ∆x22 v xx + O(∆x 3 )


✐✐5.1. Treating the advection equation 113(again, we have omitted the arguments <strong>for</strong> functions computed at (x j , t n )). Usingsuch expansions in (5.20), we havev + ∆tv t + ∆t22 v tt + O(∆t 3 ) = v + ∆x22 v xx + O(∆x 3 ) − c∆t [2∆xvx + O(∆x 3 ) ]2∆xso that by simplifying, dividing by ∆t <strong>and</strong> using again the fact that discretizationsteps vanish at the same rate, we getv t + cv x = ∆x22∆t v xx − ∆t2 v tt + O(∆x 2 ), (5.30)<strong>and</strong> using again (5.17) (which can be derived in the same way <strong>for</strong> (5.30)), we obtainat lastv t + cv x = 1 ( )∆x22 ∆t − c2 ∆t v xx + O(∆x 2 ). (5.31)Comparing the expression of the numerical viscosity with the corresponding expression<strong>for</strong> the upwind scheme, it is apparent that the Lax–Friedrichs scheme has amore viscous behaviour. In fact, taking into account that |λ| ≤ 1,( )1 ∆x22 ∆t − c2 ∆t = |c|∆x ( ) 12 |λ| − |λ|≥ |c|∆x (1 − |λ|)2so that the numerical viscosity is greater in the LF than in the upwind scheme. Forexample, if |λ| = 1/2, then the viscosity of the upwind scheme is |c|∆x/4, whereas<strong>for</strong> the LF scheme we get a viscosity of 3|c|∆x/4 (this is not surprising, whenconsidering that the numerical domain of dependence is twice as large). Note alsothat in the LF scheme the numerical viscosity is unbounded <strong>for</strong> vanishing Courantnumbers. This is related to the loss of consistency pointed out in remark 5.3.5.1.3 Semi–Lagrangian discretization: theCourant–Isaacson–Rees schemeIn many respects, Semi-Lagrangian schemes are upwind schemes. They have beenfirst proposed in the field of hyperbolic systems by Courant, Isaacson <strong>and</strong> Rees in[CIR52] in a <strong>for</strong>m which precisely gives the first order upwind scheme when appliedto the advection equation. The possibility of making them work at large Courantnumbers has been recognized later, as soon as they have been applied to problemswith a relevant computational complexity.Construction of the schemeThe construction of the Courant–Isaacson–Rees scheme is per<strong>for</strong>med by a discretizationon the representation <strong>for</strong>mula (1.6), (1.7), rather then on the equation itself.Following this strategy, an intermediate semi-discrete stage would be continuous inspace, discrete in time. As in the previous examples, we start from the constantcoefficient case, then turn to the more general case.The constant coefficient equation In this case, the representation <strong>for</strong>mula <strong>for</strong>the solution takes the simpler <strong>for</strong>m (1.3). Note that we are interested in making


✐✐114 Chapter 5. First-order approximation schemesthe scheme evolve on a single time step, from the time t n to t n+1 , <strong>and</strong> to assign avalue to a given node x j . There<strong>for</strong>e, (1.3) should be rewritten asu(x j , t n+1 ) = u(x j − c∆t, t n ). (5.32)We point out that, unless <strong>for</strong> computing the solution only at discrete points of thespace-time grid, no numerical approximation has been introduced yet. However,since the point x j − c∆t is not in general a grid point, the unknown value u(x j −c∆t, t n ) has to be replaced by a numerical reconstruction obtained by interpolationbetween node values. In the original, first-order version of the CIR scheme, this isper<strong>for</strong>med by a P 1 interpolation.We obtain there<strong>for</strong>e the numerical scheme:v n+1j = I 1 [V n ](x j − c∆t) (5.33)complemented with the initial condition v 0 j = u 0(x j ).In order to write the right-h<strong>and</strong> side more explicitly, assume first that c∆t < ∆x,so thatx j − c∆t ∈ (x j−1 , x j ].Setting now λ = c∆t/∆x, the P 1 interpolation of the sequence V n at this pointmay be written asI 1 [V n ](x j − c∆t) = λv n j−1 + (1 − λ)v n j ,so that the explicit <strong>for</strong>m of (5.33) becomes:v n+1j = λv n j−1 + (1 − λ)v n j . (5.34)Note that, despite being obtained through a different procedure, the result preciselycoincides with the upwind discretization. But, although Courant, Isaacson <strong>and</strong> Reesoriginally proposed their scheme in this version, since the representation <strong>for</strong>mula(5.32) holds <strong>for</strong> any ∆t, it is also possible to use Courant numbers beyond theunity, this meaning that the foot of the characteristic starting at x j is no longer inan adjacent cell. Denoting by ⌊·⌋ the integer part, any λ > 1 is split asλ = λ ′ + ⌊λ⌋<strong>and</strong> the cell in which the reconstruction is per<strong>for</strong>med is ( x j−⌊λ⌋−1 , x j−⌊λ⌋]. Hence,<strong>for</strong> Courant numbers exceeding the unity, (5.33) takes the <strong>for</strong>mv n+1j = λ ′ v n j−⌊λ⌋−1 + (1 − λ′ )v n j−⌊λ⌋ . (5.35)The variable coefficient equation In the more general situation of (5.2), we startfrom the representation <strong>for</strong>mula in the <strong>for</strong>m (1.6), (1.7), written at the node x j <strong>and</strong>on a single time step:u(x j , t n+1 ) =∫ tn+1t ng(y(x j , t n+1 ; s), s)ds + u(y(x j , t n+1 ; t n ), t n ) (5.36)where y(x, t; s) solves (1.7).In (5.36), in addition to the reconstruction of the value of u(y(x j , t n+1 ; t n ), t n ), twomore approximations are required. First, the point y(x j , t n+1 ; t n ) itself should be


✐✐5.1. Treating the advection equation 115approximated, since the exact solution of (1.7) is not known. Second, the integralshould be evaluated by some quadrature <strong>for</strong>mula. In the simplest situation, thepoint y(x j , t n+1 ; t n ) is replaced by its Euler approximation,x j − ∆tf(x j , t n+1 ) ≈ y(x j , t n+1 ; t n ), (5.37)<strong>and</strong> the integral in (5.36) by the rectangle quadrature on [t n , t n+1 ],∆tg(x j , t n+1 ) ≈∫ tn+1t ng(y(x j , t n+1 ; s), s)ds. (5.38)Plugging now (5.37), (5.38) <strong>and</strong> the P 1 reconstruction of u into (5.36), <strong>and</strong> incorporatingthe initial condition, we obtain a scheme in the <strong>for</strong>m{v n+1j = ∆tg(x j , t n+1 ) + I 1 [V n ](x j − ∆tf(x j , t n+1 ))vj 0 = u (5.39)0(x j ).ConsistencyIn analysing consistency of the CIR scheme, instead of using the equation, we willrather compare the scheme with the representation <strong>for</strong>mula (1.6), (1.7). LetS CIRj (∆; t, V ) = ∆tg(x j , t + ∆t) − I 1 [V ](x j − ∆tf(x j , t + ∆t)). (5.40)First, note that under smooth data, the elementary approximations introduced inthe scheme satisfy the error estimates:|x j − ∆tf(x j , t + ∆t) − y(x j , t + ∆t; t)| = O(∆t 2 ), (5.41)∫ t+∆t∣ ∆tg(x j, t + ∆t) − g(y(x j , t + ∆t; s), s)ds∣ = O(∆t2 ), (5.42)t‖u − I 1 [U]‖ ∞= O(∆x 2 ). (5.43)Bounds (5.41)–(5.43) are classical <strong>for</strong> respectively the Euler scheme, the rectanglequadrature <strong>and</strong> the P 1 interpolation. Writing now u(x j , t + ∆t) by means of (5.36),we may now estimate the left-h<strong>and</strong> side of (5.44) as∣∣L CIRj(∆; t, U(t)) ∣ = 1 ∣∣u(x j , t + ∆t) − SjCIR (∆; t, U(t)) ∣ ≤∆t≤ 1 ∫ t+∆tg(y(x∆t∣j , t + ∆t; s), s)ds + u(y(x j , t + ∆t; t), t) −t−∆tg(x j , t + ∆t) − I 1 [U(t)](x j − ∆tf(x j , t + ∆t))∣ ≤[∣≤ 1 ∣∣∣∣ ∫ t+∆tg(y(x j , t + ∆t; s), s)ds − ∆tg(x j , t + ∆t)∆t∣ +t+ |u(y(x j , t + ∆t; t), t) − I 1 [U(t)](y(x j , t + ∆t; t))| ++ |I 1 [U(t)](y(x j , t + ∆t; t)) − I 1 [U(t)](x j − ∆tf(x j , t + ∆t))|].


✐✐116 Chapter 5. First-order approximation schemesThe first two terms in the last display can be directly estimated by (5.42) <strong>and</strong>(5.43), whereas <strong>for</strong> the third we note that the Lipschitz constant of the reconstructedsolution I 1 [U] is no larger than the Lipschitz constant L u of u, so that applyingagain (5.41) we get∣ LCIRj (∆; t, U(t)) ∣ 1 [≤ O(∆t 2 ) + O(∆x 2 ) + L u O(∆t 2 ) ]∆twhich in turn implies the estimate∥ L CIR (∆; t, U(t)) ∥ ( )≤ C ∆t + ∆x2 . (5.44)∆tThis consistency estimate is similar to the corresponding estimate <strong>for</strong> the Lax–Friedrichs scheme (consistency requires that ∆x 2 = o(∆t) <strong>and</strong> its order is maximizedwhen ∆t ∼ ∆x). However, when taking into account that the CIR scheme turnsout to be unconditionally stable, <strong>and</strong> hence not restricted to operate with stepsof the same order of magnitude, the interplay between the two discretization stepsbecomes less trivial. This effect will be analysed in detail later in this section, <strong>and</strong>again in chapter 6.StabilityThe main issue in the stability analysis of the CIR scheme (<strong>and</strong> of SL schemes ingeneral) is its unconditional stability, which allows <strong>for</strong> large Courant numbers. Weanalyse this feature of the scheme from different viewpoints.CFL condition The fulfillment of the CFL condition is obtained in the CIR schemeby shifting the reconstruction stencil along characteristics, in the neighbourhood ofthe point x j −c∆t. This causes some <strong>for</strong>m of self-adaptation of the numerical domainof dependence, although at the additional cost of locating the foot of characteristics.Another advantage is the lower number of grid points involved, this generallyimplying a lower numerical viscosity. Figure 5.5 shows this point at a comparisonwith the corresponding situations <strong>for</strong> upwind <strong>and</strong> Lax–Friedrichs schemes (Figures5.1 <strong>and</strong> 5.3). In particular, an enlargement of the numerical domain of dependenceof O(∆x) at each step is multiplied by a number of steps of O(1/∆t), turning intoa numerical domain of dependence of radius O(∆x/∆t). If the refinement is madewith diverging Courant numbers (<strong>and</strong> this can really be the case in Semi-Lagrangianschemes), then this domain collapses towards the foot of the characteristic.Monotonicity Rewriting the interpolation operator I 1 [V ] in terms of the basisfunctions ψ [1]i , the CIR scheme takes the <strong>for</strong>mSjCIR (∆; t n , V ) = ∆tg(x j , t n+1 ) − ∑ iv i ψ [1]i (x j − ∆tf(x j , t n+1 )) (5.45)so thatb CIRji (∆; t n ) = ψ [1]i (x j − ∆tf(x j , t n+1 )). (5.46)Now, since ψ [1]i (x) ≥ 0 <strong>and</strong> ∑ i ψ[1] i (x) ≡ 1 (in fact, the latter property is true <strong>for</strong>Lagrange interpolation of any order), the CIR scheme is invariant <strong>for</strong> the additionof constants <strong>and</strong> monotone <strong>for</strong> any ∆x <strong>and</strong> ∆t.


✐✐5.1. Treating the advection equation 117Figure 5.5. Numerical domain of dependence <strong>for</strong> the CIR scheme.Von Neumann analysis As usual, we refer here to the scheme in the <strong>for</strong>m (5.33).Given an eigenvector V of the scheme, in the <strong>for</strong>m v j = e ijω , the condition whichcharacterizes ρ as an eigenvalue readsthat is,ρv j = λ ′ v j−⌊λ⌋−1 + (1 − λ ′ )v j−⌊λ⌋ ,ρe ijω = λ ′ e i(j−⌊λ⌋−1)ω + (1 − λ ′ )e i(j−⌊λ⌋)ωwhich allows to express the eigenvalue ρ asρ = e i⌊λ⌋ω [ λ ′ e −iω + (1 − λ ′ ) ] . (5.47)In the right-h<strong>and</strong> side of (5.47), the first is a pure phase term, (depending onlyon ⌊λ⌋) <strong>and</strong> is there<strong>for</strong>e irrelevant <strong>for</strong> stability, whereas the second, relevant termdepends only on λ ′ . Restricting there<strong>for</strong>e to the second term, <strong>and</strong> rearranging(5.47), we obtain (unless <strong>for</strong> replacing λ with λ ′ ) the same conclusions as <strong>for</strong> theupwind scheme, that is|ρ| = |(1 − λ ′ ) + λ ′ cos ω + iλ ′ sin ω| (5.48)There<strong>for</strong>e, if λ ∈ [0, 1], eigenvalues are located on a circle centered at 1 − λ <strong>and</strong>with radius λ (here, the CIR scheme coincides with the upwind scheme). If λ > 1,a further phase shift is superimposed on this curve. In both cases, λ ′ ∈ [0, 1] <strong>and</strong>the Von Neumann condition is satisfied. Figure 5.6 shows the eigenvalues of B CIR<strong>for</strong> two different values of λ having the same fractional part.Convergence estimatesFinally, we summarize this convergence study. We point out that (5.41), (5.42)require the functions f <strong>and</strong> g to be in W 1,∞ , whereas (5.43) requires that thesolution u is of W 2,∞ . Moreover, if u ∈ W 1,∞ , then (5.43) might be replaced by‖u − I[U]‖ ∞≤ C u ∆x.


✐✐118 Chapter 5. First-order approximation schemes10.80.60.40.20−0.2−0.4−0.6−0.8−1−1 −0.5 0 0.5 1Figure 5.6. Eigenvalues of the CIR scheme <strong>for</strong> λ = 0.25, 2.25 <strong>and</strong> 50 nodes.Taking into account the conditional consistency of the scheme, we have at last thefollowingTheorem 5.5. Let f, g ∈ W 1,∞ , u be the solution of (5.2) <strong>and</strong> vjn be defined by(5.39). Then, <strong>for</strong> any j ∈ Z <strong>and</strong> n ∈ [1, T/∆t],∣ vnj − u(x j , t n ) ∣ → 0 (5.49)as ∆t → 0, ∆x = o ( ∆t 1/2) .Moreover, if u ∈ L ∞ ([0, T ], W s,∞ (R)) (s = 1, 2), then‖V n − U(t n )‖ ∞≤ C)(∆t + ∆xs . (5.50)∆tNumerical viscosityTo compute the numerical viscosity of the CIR scheme, we first plug a regularextension of the numerical solution into the scheme <strong>and</strong> use the estimate <strong>for</strong> P 1Lagrange interpolation error to express the value I 1 [V n ](x j − c∆t) asI 1 [V n ](x j − c∆t) = v(x j − c∆t, t n ) − λ′ (λ ′ − 1)∆x 2v xx (ξ j ),2with ξ j an unknown point located in the same interval of x j −c∆t. Next, taking intoaccount that the computation of the second derivative in the interpolation error isper<strong>for</strong>med at a point which differs from x j − c∆t by an O(∆x), we get from(5.33):v(x j , t n+1 ) = v(x j − c∆t, t n ) + λ′ (1 − λ ′ )∆x 2[v xx (x j − c∆t) + O(∆x)]2


✐✐5.1. Treating the advection equation 119which is in the <strong>for</strong>m (4.85), with k = 2, once set the numerical viscosity asν = λ′ (1 − λ ′ )∆x 2.2∆tThe viscosity coefficient is positive, depends on λ ′ , <strong>and</strong> vanishes if λ ′ = 0 (this isconsistent with the fact that if the feet of characteristics coincide with grid nodes,the solution is advected without errors). In order to carry out a worst-case analysis,we note that its largest value is attained <strong>for</strong> λ ′ = 1/2. We can then deduce that,in the worst case, the numerical solution corresponding to a local interpolation offirst order solves the modified equationv t + cv x = ∆x28∆t v xx + o( ∆x2∆t). (5.51)Note that the viscous term originates from the accumulation of interpolation errors,<strong>and</strong> can be reduced by using large Courant numbers.High–order characteristics tracking in the CIR schemeThe first-order discretization in (5.32) is suitable <strong>for</strong> the homogeneous, constant coefficientcase, where it provides the exact upwinding along characteristics. However,in nonhomogeneous problems, as well as in the variable coefficient case, it could leadto an undesired error in time discretization. In practice, in order to balance the twoerror terms, the scheme could be <strong>for</strong>ced again to work at small Courant numbers(hence, with an unnecessary computational complexity), in spite of its unconditionalstability <strong>for</strong> large time steps. Another good reason to work at large time steps isthe reduction of numerical viscosity, as it has just been noticed. There<strong>for</strong>e, a moreaccurate approximation of characteristics <strong>and</strong> integral is desirable, <strong>and</strong> in fact canbe incorporated in the scheme without changing its stability properties.Looking at the proof of consistency, it is clear that the consistency rates in (5.41)<strong>and</strong> (5.42) should be increased in parallel, the lower of the two being the relevantrate in the first term of the global estimate (5.44). Following [FF98], a general wayto do this is to introduce the augmented system:(ẏ(x, ) ( )t; s) f(y(x, t; s), s)=(5.52)˙γ(x, t; s) −g(y(x, t; s), s)with the initial conditions y(x, t; t) = x, γ(x, t; t) = 0. In what follows, we denoteby respectively X ∆ (x, t; s) <strong>and</strong> G ∆ (x, t; s) the approximations of y(x, t; s) <strong>and</strong>γ(x, t; s), where typically x = x j , t = t n+1 <strong>and</strong> s = t n . For example, in the classicalCIR scheme (5.39) the Euler/rectangle rule discretization used corresponds to theapplication of the Euler scheme to (5.52), so that( )X ∆ (x j , t n+1 ; t n )G ∆ =(x j , t n+1 ; t n )( ) ( )xj f(xj , t− ∆tn+1 )0 −g(x j , t n+1 )With this modifications, the scheme (5.39) can be rewritten more generally as{v n+1j = G ∆ (x j , t n+1 ; t n ) + I 1 [V n ] ( X ∆ (x j , t n+1 ; t n ) )v 0 j = u 0(x j ).(5.53)(5.54)


✐✐120 Chapter 5. First-order approximation schemesAssuming now that (5.41), (5.42) are replaced by the respective estimates of orderp, ∣ ∣ X ∆ (x j , t n+1 ; t n ) − y(x j , t n+1 ; t n ) ∣ ∣ = O(∆t p+1 ), (5.55)∣ G∆ (x j , t n+1 ; t n ) −∫ tn+1t ng(y(x j , t n+1 ; s), s)ds∣ = O(∆tp+1 ), (5.56)<strong>and</strong> retracing the proof of (5.44), we get a consistency error bounded by∥ L CIR (∆; t, U(t)) ∥ ( )≤ C ∆t p + ∆x2 . (5.57)∆tUsually, <strong>for</strong> (5.55), (5.56) to hold it is necessary to assume at least that f, g ∈ W p,∞ .Since it is also immediate to check that monotonicity still holds <strong>for</strong> the scheme inthe version (5.54), we can re<strong>for</strong>mulate Theorem 5.5 in the more general <strong>for</strong>mTheorem 5.6. Let f, g ∈ W p,∞ , u be the solution of (5.2) <strong>and</strong> v n j be definedby (5.54). Assume moreover that (5.55), (5.56) hold. Then, <strong>for</strong> any j ∈ Z <strong>and</strong>n ∈ [1, T/∆t], ∣ ∣ v n j − u(x j , t n ) ∣ ∣ → 0 (5.58)as ∆t → 0, ∆x = o ( ∆t 1/2) .Moreover, if u ∈ L ∞ ([0, T ], W s,∞ (R)) (s = 1, 2), then)‖V n − U(t n )‖ ∞≤ C(∆t p + ∆xs . (5.59)∆tNote that, in comparison with the fully first-order CIR scheme, in this latterversion the terms of the consistency error are balanced under increasing Courantnumbers if p > 1. Introducing the relationship ∆t = ∆x α , the consistency error isO ( ∆x αp + ∆x 2−α) <strong>and</strong> its order is maximal with the choice α = 2/(p + 1) whichbalances the two terms. For example, using a second-order scheme in (5.60), wewould have an optimal relationship <strong>for</strong> α = 2/3.Examples: one-step schemes Applying a one-step scheme to (5.52), <strong>and</strong> restrictingto a single backward time step, we have:( ) ( ( (X ∆ (x, t; t − ∆t) x Φf −∆t; x, t, XG ∆ = − ∆t(x, t; t − ∆t) 0)∆ (x, t; t − ∆t) ) )(Φ −g −∆t; x, t, X ∆ (x, t; t − ∆t) ) (5.60)where the function Φ is split according to (5.52), <strong>and</strong> X ∆ is understood as thesolution of a system in implicit schemes, in which the right-h<strong>and</strong> side of (5.60)depends genuinely on X ∆ itself. Table 5.1 puts in the <strong>for</strong>m (5.60) the same examplesof table 3.1.In these examples, we have taken <strong>for</strong> granted that the functions f <strong>and</strong> g can becomputed <strong>for</strong> any generic argument, that is, that they have an explicit expression.In a number of situation, like the nonlinear advection in fluid dynamics models, itcould rather happen that• both the advection <strong>and</strong> the source term result from previous computations<strong>and</strong>/or physical measures, <strong>and</strong> are only known at grid nodes;


✐✐5.1. Treating the advection equation 121scheme <strong>for</strong>m orderFE(Φ ) f −∆t; x, t, X∆= f(x, t) p = 1(Φ ) −g −∆t; x, t, X∆= −g(x, t)BE Φ f(−∆t; x, t, X∆ ) = f ( X ∆ , t − ∆t ) p = 1Φ −g(−∆t; x, t, X∆ ) = −g ( X ∆ , t − ∆t )(H Φ ) f −∆t; x, t, X∆= 1 2[f(x, t) + f(x − ∆tf(x, t), t − ∆t)] p = 2(Φ ) −g −∆t; x, t, X∆= − 1 2[g(x, t) + g(x − ∆tf(x, t), t − ∆t)]CN(Φ ) [ (f −∆t; x, t, X∆= 1 2 f(x, t) + f X ∆ , t − ∆t )] p = 2(Φ ) [ (−g −∆t; x, t, X∆= − 1 2 g(x, t) + g X ∆ , t − ∆t )]Table 5.1. First- <strong>and</strong> second-order one-step schemes <strong>for</strong> the augmentedsystem (5.52)• the advection term is known at previous time steps, but not yet at the stept n+1 .The first situation requires that f <strong>and</strong> g be interpolated to be computedat a generic point. Assuming that f is interpolated with degree q <strong>and</strong> there<strong>for</strong>ereconstructed with error O(∆x q+1 ), the accuracy assumption (5.55) changes into∣ X ∆ (x, t; t − ∆t) − y(x, t; t − ∆t) ∣ ∣ ≤ O(∆tp+1 ) + O ( ∆t∆x q+1) ,<strong>and</strong> the consistency error into∥ L CIR (∆; t, U(t)) ∥ ()≤ C ∆t p + ∆x q+1 + ∆x2∆t(clearly, the same ideas also apply to the source term g). For example, if a secondorderscheme is used to move along characteristics, the optimal relationship withoutinterpolation <strong>for</strong> f would be, as we said above, ∆x = ∆t 3/2 . Introducing theinterpolation of f, we have in turn∥∥L CIR (∆; t, U(t)) ∥ (≤ C ∆t 2 + ∆t 3 (q+1)) 2 .It suffices there<strong>for</strong>e to interpolate f with degree q = 1 to preserve the consistencyrate obtained without interpolation. Note also that in case a constant Courantnumber is kept (∆x ∼ ∆t), the consistency rate is preserved if q + 1 ≥ p, so a linearinterpolation is still enough <strong>for</strong> a second-order time discretization.


✐✐122 Chapter 5. First-order approximation schemesThe second situation, that is, f being unknown at the (n + 1)–th time step, isusually unsuitable to be treated with the one-step schemes of table 5.1, unless witha first order approximation. Since under minimal smoothness assumptions we havef(x, t) = f(x, t − ∆t) + O(∆t),g(x, t) = g(x, t − ∆t) + O(∆t),then, replacing the values of f <strong>and</strong> g computed at t in the scheme (5.39) with thesame values computed at t − ∆t, we would obtain <strong>for</strong> X ∆X ∆ (x, t; t − ∆t) = x − ∆tf(x, t − ∆t) + O ( ∆t 2)(a similar error holds <strong>for</strong> G ∆ ) so that X ∆ , G ∆ are computed using the values atthe previous time step <strong>and</strong> without any loss in the consistency rate. An alternativechoice is given by the backward Euler scheme, which only requires the knowledge off at t − ∆t. In this case, the foot of characteristic is located by solving the systemX ∆ (x, t; t − ∆t) = x − ∆tf ( X ∆ (x, t; t − ∆t), t − ∆t ) ,whereas the source term is given by G ∆ (x, t; t − ∆t) = g(X ∆ (x, t; t − ∆t), t − ∆t).The system is already in fixed-point <strong>for</strong>m, <strong>and</strong> could be solved <strong>for</strong> X ∆ by theiterationXk+1 ∆ = x − ∆tf ( Xk ∆ , t − ∆t )provided the right-h<strong>and</strong> side is a contraction, that is, provided∆t < 1‖J f ‖ ,J f denoting the Jacobian matrix of the vectorfield f. Note that this introduces abound on ∆t, but this bound depends on the derivatives of the advection speedrather than on its magnitude. Note also that the method is first-order, so thatit gives no improvement in accuracy over the explicit Euler method used in theclassical CIR scheme.Examples: multistep schemes The most classical application of multistep schemes<strong>for</strong> the approximation of the foot of characteristics involves the midpoint method(see table 3.1). In this case, the equation of characteristics is integrated 2∆t backin time, <strong>and</strong> more precisely X ∆ (x j , 2∆t) solves the fixed-point system( x + XX ∆ ∆ )(x, t; t − 2∆t)(x, t; t − 2∆t) = x − 2∆tf, t − ∆t . (5.61)2Note that, according to the definition of the midpoint method, the vector fieldf should be computed at the point X ∆ (x, t; t − ∆t). Along a smooth trajectory,however, we haveX ∆ (x, t; t − ∆t) = x + X∆ (x, t; t − 2∆t)2+ O(∆t 2 ),so that by the same arguments used <strong>for</strong> the explicit Euler method, the approximatemidpoint method (5.61) still results in a second-order scheme. Accordingly, thesource term is computed as( x + XG ∆ ∆ )(x, t; t − 2∆t)(x, t; t − 2∆t) = 2∆tg, t − ∆t .2


✐✐5.1. Treating the advection equation 123Thus, the scheme would compute the approximate solution at time t n+1 by advectingthe approximate solution at t n−1 along the vectorfield computed at t n (such astructure is often referred to as a three-time-level scheme). In general, the latterfixed-point system can also be solved iteratively under the same time-step limitationas <strong>for</strong> the backward Euler scheme, <strong>and</strong> we expect that in general the value of f atthe right-h<strong>and</strong> side should be obtained by interpolation. Since the resulting schemeis second-order in time, as remarked above, linear interpolation on f is enough topreserve the consistency rate.5.1.4 Multiple space dimensionsWe briefly sketch here how the techniques introduced in the previous sections canbe adapted to treat more general cases, in particular problems in higher dimension<strong>and</strong> posed on bounded domains.To treat the case of multiple dimensions, it is convenient to rewrite the advectionequation,u t + f(x, t) · Du = g(x, t) R d × [0, T ],in the equivalent <strong>for</strong>mu t +d∑f i (x, t)u xi = g(x, t) R d × [0, T ]. (5.62)i=1For simplicity, we review the various approaches referring to the two-dimensionalcase. Once given the general <strong>for</strong>m, we briefly discuss consistency <strong>and</strong> monotonicity.Upwind scheme The upwind scheme may be better explained referring to thehomogeneous, constant coefficient equationu t + c 1 u x1 + c 2 u x2 = 0.Given that the consistency of the scheme is insensitive to the choice of taking rightor left incremental ratios, the crucial point is to obtain a stable (in particular,monotone) scheme. Keeping the strategy of approximating space derivatives by leftincremental ratios, we would obtainv n+1j 1,j 2− vj n 1,j 2vj n + c 1,j 2− vj n 1−1,j 2vj n1 + c 1,j 2− vj n 1,j 2−12 = 0,∆t∆x 1∆x 2which is clearly consistent. Solving this relationship <strong>for</strong> v n+1j 1,j 2, we obtainv n+1j 1,j 2= vj n 1,j 2− c 1∆t [vn∆x j1,j 2− v n ] c 2 ∆t [j 1−1,j 2 − vn1 ∆x j1,j 2− v n ]j 1,j 2−1 =(2= 1 − c 1∆t− c )2∆tvj n ∆x 1 ∆x 1,j 2+ c 1∆tvj n2 ∆x 1−1,j 2+ c 2∆tvj n1 ∆x 1,j22−1,<strong>and</strong> it is immediate to see that all coefficients are nonnegative (<strong>for</strong> partial Courantnumbers c i ∆t/∆x i small enough) if <strong>and</strong> only of c 1 , c 2 ≥ 0. This suggests thateach of the terms f i (x, t)v xi should be discretized in an upwind <strong>for</strong>m: with a leftincremental ratio if f i (x j , t n ) > 0, with a right incremental ratio if f i (x j , t n ) < 0.We give up the boring, but conceptually simple, description of the details.


✐✐124 Chapter 5. First-order approximation schemesLax–Friedrichs scheme In extending the Lax–Friedrichs scheme to (5.62), it shouldbe noted that the average which replaces vj n 1,j 2should include all points involvedin the finite difference approximations of space derivatives, <strong>and</strong> there<strong>for</strong>e in thiscase (d = 2) the four points vj n 1−1,j 2, vj n 1+1,j 2, vj n <strong>and</strong> 1,j 2−1 vn j 1,j 2+1 . This gives theschemev n+1 (j 1,j 2− 1 4 vnj1−1,j 2+ vj n 1+1,j 2+ vj n + 1,j 2−1 vn j 1,j 2+1)+∆t+f 1 (x j , t n ) vn j 1+1,j 2− v n j 1−1,j 2∆x 1+ f 2 (x j , t n ) vn j 1,j 2+1 − vn j 1,j 2−1∆x 2= g(x j , t n )(again, this discretization may be shown to be consistent using the same argumentsof the one-dimensional case), <strong>and</strong> solving <strong>for</strong> v n+1j 1,j 2we get:v n+1j 1,j 2= 1 (vn4 j1−1,j 2+ vj n 1+1,j 2+ vj n 1,j 2−1 + vj n )1,j 2+1 −−f 1 (x j , t n ) ∆t [vn∆x j1+1,j 2− v n ]j 1−1,j 2 −1−f 2 (x j , t n ) ∆t∆x 2[vnj1,j 2+1 − v n j 1,j 2−1]+ ∆tg(xj , t n ).In this way, the nonzero coefficients of the scheme are of the <strong>for</strong>mb LFjk (∆; t n ) = 1 4 ± f i(x j , t n ) ∆t∆x i(<strong>for</strong> suitable values of the multiindex k <strong>and</strong> of the index i), <strong>and</strong> again can be madenonnegative by keeping the partial Courant numbers small enough.Courant–Isaacson–Rees scheme In more than one space dimension, the CIRscheme can still be written in a <strong>for</strong>mally unchanged way, that is,v n+1j = G ∆ (x j , t n+1 ; t n ) + I 1 [V n ] ( X ∆ (x j , t n+1 ; t n ) ) .In this case, however, it should be understood that:• The augmented system (5.52) is now in dimension d + 1: the approximationof the source term remains scalar, whereas characteristics are curves in R d .Thus, X ∆ is now a point of R d itself.• The reconstruction I 1 [V ](·) is a multidimensional P 1 or Q 1 interpolation.The monotonicity of the scheme is proved following the same argument as ina single space dimension: we have in fact that the coefficients of the scheme aregiven byb CIRji (∆; t n ) = ψ [1] (i X ∆ (x j , t n+1 ; t n ) ) ,<strong>and</strong> are nonnegative if the basis functions <strong>for</strong> interpolation are nonnegative (inmultiple dimension, this is clearly true <strong>for</strong> tensorized linear bases <strong>and</strong> P 1 finiteelements).


✐✐5.1. Treating the advection equation 1255.1.5 Boundary conditionsThe basic theory has been outlined so far without any reference to boundary conditions,which naturally arise when posing a problem on a bounded domain. Asituation which allows to keep essentially the same framework of unbounded gridsis the case of periodic conditions given on a d-dimensional torus [0, P 1 ]×· · ·×[0, P d ].Here, the periodicity is imposed in classical difference schemes by identifying thecorresponding nodes (<strong>and</strong> values) on opposite sides of the torus, whereas in SLschemes a point outside the torus is identified with an internal point via the relationshipu(x 1 ± P 1 , . . . , x d ± P d ) = u(x 1 , . . . , x d ).The case of Dirichlet conditions is more interesting. Assume now that Ω ≢ R d<strong>and</strong> that Dirichlet data are imposed on the inflow part of the boundary ∂Ω:u(x, t) = b(x, t) (x, t) ∈ Γ in × [0, T ]. (5.63)We review the main points in the implementation of this boundary condition in theupwind, LF <strong>and</strong> CIR schemes, avoiding the most technical issues.Upwind scheme To fix ideas, assume that the computational domain is a rectangle.Due to the definition of the scheme, nodes on each side the boundary areassigned a value based on internal nodes if the vector field f points outwards, thatis, on the outflow boundary Γ out , <strong>and</strong> this agrees with the fact that Dirichlet conditionscannot be en<strong>for</strong>ced on this portion of the boundary. Nodes on Γ in areassigned a value by (5.63), while nodes placed at the angles may be at the interfacebetween Γ in <strong>and</strong> Γ out . In this situation, not all the partial incremental ratios maybe computed using internal nodes, so on such a node we <strong>for</strong>ce the boundary condition(5.63). Clearly, more complex geometries require to treat a larger number ofpossible situations.Lax–Friedrichs scheme In the Lax–Friedrichs scheme, assigning a value to a nodealways requires that other nodes exist around it. So, while this strategy allows toassign a boundary value to a node in Γ in (since it needs not to be computed by thescheme), it does not allow to compute the values at nodes of Γ out , since it wouldrequire a further layer of nodes. On the other h<strong>and</strong>, if the boundary condition bis imposed on the whole of ∂Ω, what typically happens is the development of aboundary layer on Γ out . To get rid of this effect, a simple response would be totreat in upwind <strong>for</strong>m the nodes on this part of the boundary.Courant–Isaacson–Rees scheme The CIR scheme requires a more complex treatmentof the Dirichlet condition. First, we rewrite the extended version (2.44) of therepresentation <strong>for</strong>mula between t n <strong>and</strong> t n+1 :u(x j , t n+1 ) =∫ tn+1t n∨θ(x j,t n+1)g(y(x j , t n+1 ; s), s)ds ++u(y(x j , t n+1 ; t n ∨ θ(x j , t n+1 )), t n ∨ θ(x j , t n+1 )). (5.64)As we have discussed in Chapter 2, this amounts to single out a case in which thesolution is propagated from an internal point, <strong>and</strong> a case in which it propagates fromthe boundary. The numerical counterpart of this procedure requires to determine


✐✐126 Chapter 5. First-order approximation schemesthe point, if any exists, at which the discrete trajectory X ∆ crosses the boundaryof Ω. If we assume <strong>for</strong> simplicity that the boundary ∂Ω is defined as the zero levelset of a continuous function ϕ : R N → R, so that x ∈ Ω if <strong>and</strong> only if ϕ(x) < 0,then we can define a variable time step δ(x j , t n+1 ) as the minimum between ∆t <strong>and</strong>the solution of the equationϕ ( X ∆ (x j , t n+1 ; t n+1 − δ(x j , t n+1 )) ) = 0.Note that we expect at most one solution in [t n , t n+1 ], <strong>for</strong> ∆t small enough <strong>and</strong> “nonpathological”sets Ω. Note also that, <strong>for</strong> consistency reasons, δ(x j , t n+1 ) should alsoreplace ∆t in the computation of X ∆ (e.g., in the construction of a one-step scheme).Once computed the variable step, we split the set of indices as I = I i ∪ I b ,so that δ(x j , t n+1 ) = ∆t <strong>for</strong> j ∈ I i (internal nodes), δ(x j , t n+1 ) ≤ ∆t <strong>for</strong> j ∈ I b(near-boundary nodes). Then, the CIR scheme is modified asv n+1j =⎧G ∆ (x j , t n+1 ; t n ) + I 1 [V n ] ( X ∆ (x j , t n+1 ; t n ) )⎪⎨if j ∈ I iG⎪⎩∆ (x j , t n+1 ; t n+1 − δ(x j , t n+1 ))++b ( X ∆ (x j , t n+1 ; t n+1 − δ(x j , t n+1 )), t n+1 − δ(x j , t n+1 ) ) if j ∈ I b .(5.65)5.1.6 ExamplesIn order to evaluate the schemes which have been presented in this section, we usethe simple one-dimensional example{u t (x, t) − (x − ¯x)u x (x, t) = 0 (x, t) ∈ (0, 1) × (0, 1)(5.66)u(x, 0) = u 0 (x)with three initial conditions u 0 of different regularity <strong>and</strong> with bounded support.In particular, the first is a bounded discontinuous function:whereas the second is Lipschitz continuous:<strong>and</strong> the third has bounded second derivative:u 0 (x) = 1 [0.25,0.5] (x), (5.67)u 0 (x) = max(1 − 16(x − 0.25) 2 , 0), (5.68)u 0 (x) = max(1 − 16(x − 0.25) 2 , 0) 2 . (5.69)In (5.66), characteristics are curved lines <strong>and</strong> this makes time discretization significantin SL schemes. Also, characteristics converge towards the point ¯x = 1.1, so thatthe initial condition is advected rightwards <strong>and</strong> the support shrinks <strong>for</strong> increasingtime.The test has been carried out with Upwind, Lax–Friedrichs <strong>and</strong> Courant–Isaacson–Rees schemes, the last being implemented both in its classical version(5.39) <strong>and</strong> in the version (5.54) with a second-order time discretization (denotedas CIR2). In the first three cases, grid refinement has been per<strong>for</strong>med by keepinga constant ∆t/∆x relationship, <strong>and</strong> more precisely keeping the Courant number


✐✐5.1. Treating the advection equation 127L ∞ solutionn n Upwind LF CIR CIR225 1.93 · 10 −1 2.45 · 10 −1 1.06 · 10 −1 1.1 · 10 −150 1.66 · 10 −1 2.3 · 10 −1 1.16 · 10 −1 7.35 · 10 −2100 1.35 · 10 −1 2.2 · 10 −1 7.59 · 10 −2 6.31 · 10 −2200 1.1 · 10 −1 1.9 · 10 −1 6.64 · 10 −2 4.71 · 10 −2400 9.2 · 10 −2 1.56 · 10 −1 5.38 · 10 −2 4.01 · 10 −2rate 0.27 0.16 0.24 0.36W 1,∞ solutionn n Upwind LF CIR CIR225 1.55 · 10 −1 2.39 · 10 −1 1.29 · 10 −1 3.36 · 10 −250 1.09 · 10 −1 2.09 · 10 −1 7.62 · 10 −2 1.78 · 10 −2100 7.05 · 10 −2 1.72 · 10 −1 4.1 · 10 −2 8.43 · 10 −3200 4.39 · 10 −2 1.28 · 10 −1 2.21 · 10 −2 4.42 · 10 −3400 2.66 · 10 −2 8.68 · 10 −2 1.16 · 10 −2 1.99 · 10 −3rate 0.64 0.37 0.87 1.02W 2,∞ solutionn n Upwind LF CIR CIR225 1.52 · 10 −1 2.17 · 10 −1 1.32 · 10 −1 2.97 · 10 −250 1.14 · 10 −1 1.96 · 10 −1 7.52 · 10 −2 1.44 · 10 −2100 7.61 · 10 −2 1.68 · 10 −1 4.01 · 10 −2 6.09 · 10 −3200 4.63 · 10 −2 1.32 · 10 −1 2.08 · 10 −2 2.6 · 10 −3400 2.63 · 10 −2 9.27 · 10 −2 1.06 · 10 −2 1.05 · 10 −3rate 0.63 0.31 0.91 1.21Table 5.2. Errors in the 2-norm, Upwind, Lax–Friedrichs, Courant–Isaacson–Rees (with first- <strong>and</strong> second-order time discretization) schemes.less than 0.6 in Upwind <strong>and</strong> LF schemes, <strong>and</strong> less than 6 in CIR scheme. Inthe case of the second-order time discretization <strong>for</strong> the CIR scheme, we have usedthe relationship ∆t ∼ ∆x 2/3 (more precisely, ∆t ≈ 2∆x 2/3 ) which optimizes theconsistency rate.We list in Table 5.2 the numerical errors in the 2-norm <strong>for</strong> the various schemes,along with overall convergence rates computed on the extreme steps of the refinement.Figure 5.7 shows the behaviour of the schemes on the advection of thecharacteristic function (5.67) (the exact solution is shown in solid line).The apparent point in the comparison of the various schemes is numerical diffusion.The accuracy of the LF scheme is reduced by its highly diffusive behaviour,but ranging from the Upwind to the CIR2 scheme the convergence rate has no dramaticchange, although it is clear how the possibility of working at large Courantnumbers reduces the numerical diffusion. In terms of absolute accuracy, the firstorderCIR scheme is about twice as accurate as the Upwind scheme, <strong>and</strong> the increasein accuracy is even more apparent <strong>for</strong> the CIR2 scheme, especially as the solution


✐✐128 Chapter 5. First-order approximation schemesFigure 5.7. Numerical results <strong>for</strong> the advection of a characteristic functionobtained via Upwind (upper left), Lax–Friedrichs (upper right), Courant–Isaacson–Rees (lower left) <strong>and</strong> Courant–Isaacson–Rees with second-order time discretization(lower right) schemes, 200 nodes.gets smoother (here, the theoretical convergence rate would be 4/3 ≈ 1.33).5.2 Treating the convex HJ equationIn this section, we show how the schemes introduced can be adapted to the onedimensional,convex <strong>Hamilton</strong>–Jacobi equation{u t (x, t) + H(u x (x, t)) = 0 (x, t) ∈ R × [0, T ],(5.70)u(x, 0) = u 0 (x).We will make the st<strong>and</strong>ing assumption that H is convex, <strong>and</strong> that there existsα 0 ∈ R such that{H ′ (α) ≤ 0 if α ≤ α 0 ,H ′ (5.71)(α) ≥ 0 if α ≥ α 0 .We also define:M H ′(L) = max[−L,L] |H′ | (5.72)which corresponds to the maximum speed of propagation of a solution with Lipschitzconstant L.


✐✐5.2. Treating the convex HJ equation 1295.2.1 Upwind discretizationIn adapting the upwind scheme to the nonlinear case, it should be taken into considerationthat the speed of propagation of the solution is H ′ (v x ). While it is perfectlyclear how to construct an upwind scheme <strong>for</strong> a speed of constant sign, care shouldbe taken at points where the speed changes sign, in order to obtain a monotonescheme.Construction of the schemeThe construction outlined will follow the guidelines of [CL84], in which monotoneschemes <strong>for</strong> HJ equations are derived from monotone schemes <strong>for</strong> conservation laws,<strong>and</strong> theory is carried out accordingly. The differenced <strong>for</strong>m of the upwind schemeis (4.53),v n+1j = v n j − ∆tH Up (D j−1 [V n ], D j [V n ]) , (5.73)where the numerical <strong>Hamilton</strong>ian H Up is defined by⎧H(α) if α, β ≥ α 0 ,⎪⎨H Up H(β) + H(α) − H(α 0 ) if α ≥ α 0 , β ≤ α 0 ,(α, β) =H(α 0 ) if α ≤ α 0 , β ≥ α 0 ,⎪⎩H(β) if α, β ≤ α 0 .(5.74)Note that the situation in which speed changes sign is subject to a different h<strong>and</strong>ling,depending on the fact that characteristics converge or diverge.ConsistencySince the scheme is in differenced <strong>for</strong>m, it actually suffices to apply proposition 4.18.If α = β = a, then the numerical hamiltonian (5.74) satisfiesH Up (a, a) = H(a), (5.75)<strong>and</strong> the consistency condition (4.51) is satisfied. Note that, in (5.74), the second<strong>and</strong> third case only occur if a = α 0 .StabilityBy construction, schemes in differenced <strong>for</strong>m (in particular, upwind <strong>and</strong> Lax–Friedrichs in what follows) are necessarily invariant <strong>for</strong> the addition of constants.There<strong>for</strong>e, the main stability issues in this context will be CFL condition <strong>and</strong> monotonicity.CFL condition Since the maximum speed of propagation of the solution is M H ′(L),the condition which keeps characteristics within the numerical domain of dependenceisM H ′(L)∆t≤ 1. (5.76)∆xIn this case, this restriction is really only necessary – as we will see, monotonicityrequires a stronger condition.


✐✐130 Chapter 5. First-order approximation schemesMonotonicityscheme as∂S Upj∂v iFirst, we write the partial derivative of the j–th component of the[ ]∂HUp∂D j−1 [V ](∆; V ) = δ ij − ∆t+ ∂HUp ∂D j [V ]∂α ∂v i ∂β ∂v i(5.77)where α <strong>and</strong> β are the dummy variables used in the definition (5.74), <strong>and</strong> δ ij is theKronecker symbol. It is clear that⎧1∂D j−1 [V ]⎪⎨ ∆xif i = j= − 1∂v∆xif i = j − 1i ⎪⎩0 otherwise⎧1∂D j [V ]⎪⎨ ∆xif i = j + 1= − 1∂v∆xif i = ji ⎪⎩0 otherwise,so that, substituting into (5.77), we obtain the more explicit <strong>for</strong>m∂∂v iS Upj (∆; V ) =⎧[ ⎪⎨1 − ∆t−∆t ∂HUp ∂D j−1[V ]∂α∂H Up ∂D j−1[V ]∂α ∂v j−∆t ∂HUpBy the definition of H Up we have∂v j−1if i = j − 1+ ∂HUp∂β]∂D j[V ]∂v jif i = j∂D j[V ]∂β ∂v ⎪⎩j+1if i = j + 10 otherwise.∂H Up∂α (α, β) = {H ′ (α) ≥ 0 if α ≥ α 0 ,0 otherwise.∂H Up∂β (α, β) = {H ′ (β) ≤ 0 if β ≤ α 0 ,0 otherwise.(5.78)(5.79)(5.80)Looking at the signs of the various terms, it is apparent that∂∂v iS Upj (∆; V ) ≥ 0 (i ≠ j), (5.81)whereas, <strong>for</strong> i = j, we have∂H Up ∂D j−1 [V ]∣ ∂α ∂v j+ ∂HUp∂β∣∂D j [V ] ∣∣∣≤ 2M H ′(L V )∂v j ∆x(5.82)where L V denotes the Lipschitz constant of the sequence V . There<strong>for</strong>e, using (5.82)in (5.78), we obtain that the scheme is monotone if∆t∆x ≤ 12M H ′(L V ) . (5.83)Note that, in contrast to the linear case, this condition is more stringent then theCFL condition.


✐✐5.2. Treating the convex HJ equation 131Convergence estimatesFinally, we give the convergence result, which follows from consistency (5.75), monotonicity(5.83), <strong>and</strong> theorem 4.20.Theorem 5.7. Let H satisfy the basic assumptions, u 0 ∈ W 1,∞ , u be the solution of(5.70) with L as its Lipschitz constant <strong>and</strong> v n j be defined by (5.73) with v0 j = u 0(x j ).Then, <strong>for</strong> any j ∈ Z <strong>and</strong> n ∈ [1, T/∆t],∣∣v n j − u(x j , t n ) ∣ ∣ ≤ C∆t 1/2 (5.84)as ∆t → 0, with 2M H ′(L + 1)∆t ≤ ∆x.5.2.2 Central discretizationIn treating the Lax–Friedrichs scheme, we will follow again the guidelines of [CL84].Rather than using more general <strong>for</strong>ms of the scheme, we will restrict here to theparticular <strong>for</strong>m that directly generalizes the linear case.Construction of the schemeThe simplest way to recast Lax–Friedrichs scheme <strong>for</strong> the HJ equation is to defineit in the <strong>for</strong>mv n+1j = vn j−1 + vn j+1− ∆tH ( D c2j[V n ] ) , (5.85)where Dj c[V n ] is the centered difference at x j defined byD c j[V n ] = vn j+1 − vn j−12∆x= D j−1[V n ] + D j [V n ]. (5.86)2This definition of the LF scheme completely parallels the linear case, <strong>and</strong> is alsosuitable to be treated in the framework of The Cr<strong>and</strong>all–Lions theorem. In fact,once recalled thatv n j−1 + vn j+12(5.85) can be written in the differenced <strong>for</strong>mby setting= v n j + ∆x2 (D j[V n ] − D j−1 [V n ]) ,v n+1j = v n j − ∆tH LF (D j−1 [V n ], D j [V n ])( ) α + βH LF (α, β) = H − ∆x (β − α).2 2∆tNote that, as <strong>for</strong> the advection equation, no special care is necessary to determinethe direction of propagation <strong>for</strong> the solution (that is, to compare α <strong>and</strong> β with α 0 ),since the stencil is symmetric.ConsistencyThe Lax–Friedrichs scheme (5.85) satisfies condition (4.51), <strong>and</strong> in fact( ) a + aH LF (a, a) = H = H(a). (5.87)2Consistency is there<strong>for</strong>e satisfied.


✐✐132 Chapter 5. First-order approximation schemesStabilityWe examine again the issues of CFL condition <strong>and</strong> monotonicity, which in this casegive the same restriction on the discretization steps.CFL condition Taking into account that the maximum speed of propagation isM H ′(L), the CFL condition readsM H ′(L V )∆t∆x≤ 1 (5.88)as <strong>for</strong> the upwind scheme. In this case, this condition is necessary <strong>and</strong> sufficient,since it also ensures monotonicity (as we will soon show).Monotonicity In examining monotonicity, it is convenient to refer to the LFscheme in the <strong>for</strong>m (5.85). Clearly, the j–th component SjLF (∆; V ) depends onlyon the values v j±1 , so thatOn the other h<strong>and</strong>, if i = j ± 1, we have∂∂v iS LFj (∆; V ) = 0 (i ≠ j ± 1)∂∂v j±1S LFj (∆; V ) = 1 2 − ∆tH′ ( D c j[V ] ) ∂Dc j [V ]∂v j±1=where we have used the fact that= 1 2 ∓ ∆t2∆x H′ ( D c j[V ] ) ,∂D c j [V ]∂v j±1= ± 12∆x .There<strong>for</strong>e, if L V is the Lipschitz constant of the sequence V , the scheme is monotoneprovided∆t∆x ≤ 1M H ′(L V ) . (5.89)Convergence estimatesAgain, the convergence result is obtained from consistency (5.87), monotonicity(5.89), applying theorem 4.20.Theorem 5.8. Let H satisfy the basic assumptions, u 0 ∈ W 1,∞ , u be the solution of(5.70) with L as its Lipschitz constant <strong>and</strong> v n j be defined by (5.85) with v0 j = u 0(x j ).Then, <strong>for</strong> any j ∈ Z <strong>and</strong> n ∈ [1, T/∆t],as ∆t → 0, with M H ′(L + 1)∆t ≤ ∆x.∣ vnj − u(x j , t n ) ∣ ∣ ≤ C∆t1/2(5.90)


✐✐5.2. Treating the convex HJ equation 1335.2.3 Semi–Lagrangian discretizationUnless <strong>for</strong> replacing the <strong>for</strong>mula of characteristics (1.6)–(1.7) with a suitable generalization,the extension of the Courant–Isaacson–Rees approach to HJ equations isrelatively straight<strong>for</strong>ward. In this chapter we analyse the monotone version of theSL scheme, that is, the version obtained with P 1 interpolation.Construction of the schemeSemi–Lagrangian discretization of the HJ equation follows the same steps seen <strong>for</strong>the advection equation: namely, what is really discretized is the representation<strong>for</strong>mula <strong>for</strong> the solution. In the case of convex HJ equations, the <strong>for</strong>mula underconsideration is the Hopf–Lax <strong>for</strong>mula (2.30). Once rewritten in a single spacedimension, <strong>and</strong> at a point (x j , t n+1 ) of the space–time grid, it reads:u(x j , t + ∆t) = mina∈R [∆tH∗ (a) + u(x j − a∆t, t)] == ∆tH ∗ (ā j ) + u(x j − ā j ∆t, t) (5.91)(note that, now <strong>and</strong> then, we might find it useful to introduce the explicit notationā j to denote the minimizer at the node x j in (5.91)). In the special case of (5.70),characteristics are straight lines, so that no special care should be taken about theaccuracy of time discretization (more complex situations will be considered at theend of this section). It is still necessary, however, to replace the value u(x j − a∆t, t)by a space reconstruction. At this stage, <strong>and</strong> in parallel with the CIR scheme, thiswill be chosen to be the P 1 interpolation I 1 . Once set t = t n , the resulting schemeis there<strong>for</strong>e:{vn+1j= minα∈R [∆tH∗ (α) + I 1 [V n ](x j − α∆t)]v 0 j = u 0(x j ).(5.92)Since the SL scheme is not in differenced <strong>for</strong>m, the convergence analysis will becarried out in the framework of respectively Barles–Souganidis <strong>and</strong> Lin–Tadmortheorems.ConsistencyIn order to develop a twofold convergence theory, consistency will be proved separatelyin two different frameworks. In the first, we use the usual notion (which issuitable <strong>for</strong> the use of Barles–Souganidis theorem), which treats consistency as apointwise concept. In the second, we adopt the L 1 notion of consistency used inthe Lin–Tadmor theory, in the particular <strong>for</strong>m suitable <strong>for</strong> Godunov-type schemes.Pointwise theory As it has been done <strong>for</strong> the CIR scheme, the scheme will becompared with the representation <strong>for</strong>mula <strong>for</strong> the solution, (2.30) in this case. WedefineSjSL (∆; V ) = minα∈R [∆tH∗ (α) + I 1 [V ](x j − α∆t)] == ∆tH ∗ (ᾱ j ) + I 1 [V ](x j − ᾱ j ∆t). (5.93)where ᾱ j denotes the minimizer at the node x j in (5.92).


✐✐134 Chapter 5. First-order approximation schemesLet now u be a smooth solution of (5.70). First, we recall that <strong>for</strong> any smoothu the error estimate (5.43) is satisfied:‖u − I 1 [U]‖ ∞≤ C∆x 2 .Writing now u(x j , t + ∆t) by means of (5.91), we give a first unilateral estimate onthe left-h<strong>and</strong> term of (5.94) asL SLj (∆; t, U(t)) = 1 [u(xj , t + ∆t) − Sj SL (∆; t, U(t)) ] =∆t= 1 ∆t [∆tH∗ (ā j ) − u(x j − ā j ∆t, t)−− ∆tH ∗ (ᾱ j ) + I 1 [U(t)](x j − ᾱ j ∆t)] ≤≤ 1 ∆t [∆tH∗ (ᾱ j ) − u(x j − ᾱ j ∆t, t)−− ∆tH ∗ (ᾱ j ) + I 1 [U(t)](x j − ᾱ j ∆t)] ≤≤ 1 ∆t ‖u(t) − I 1[U(t)]‖ ∞≤ C ∆x2∆tNote that the inequality follows from using ᾱ j instead of the exact minimizer ā j in(5.91). By reversing the roles of ᾱ j <strong>and</strong> ā j we get the opposite inequality,−L SLj (∆; t, U(t)) = 1 ∆t[u(xj , t + ∆t) − SjSL (∆; t, U(t)) ] ≤ C ∆x2∆t ,<strong>and</strong> there<strong>for</strong>e,∣∣L SLj (∆; t, U(t)) ∣ ≤ C ∆x2∆t . (5.94)Note that this estimate would suggest that the scheme achieves its best result whengoing to the final time in a single time step. In practice, this advantage is cutdown by some side effects of very large time steps. First, in situation in whichcharacteristics are not straight lines, errors in characteristics tracking should betaken into consideration (see the consistency estimate <strong>for</strong> the linear case). Second,it might be useful to approximate the solution at intermediate times as well. Third,large time steps enlarge the numerical domain of dependence, thus requiring a morecareful (<strong>and</strong> computationally more complex) procedure <strong>for</strong> the global minimumsearch. In respect of this latter argument, it should be noted that in order to obtainany computational advantage from the use of large time steps, it is necessary thatthe complexity of the minimum search would have a weak dependence on the sizeof the set in which this search is per<strong>for</strong>med.L 1 theory Be<strong>for</strong>e computing the consistency error in the sense (4.73) of Lin <strong>and</strong>Tadmor, we recall from [LT01] a result which simplifies the computation wheneverthe scheme is in Godunov <strong>for</strong>m – which definitely occurs <strong>for</strong> SL schemes. For ourpurposes, we will consider a scheme to be in Godunov <strong>for</strong>m when it is based on therepeated application of two operators:• An exact evolution operator E(·) which make the numerical solution evolve(exactly) on a single time step, so that starting from v ∆ (t n ), we define the


✐✐5.2. Treating the convex HJ equation 135numerical solution on (t n , t n+1 ) asv ∆ (t) = E(t − t n )v ∆ (t n ), t n < t < t n+1 ; (5.95)• An operator of projection P ∆x which defines v ∆ (t n+1 ) by projecting the approximationdefined above (computed at the end of the time step) on thespecific space discretization:v ∆ (t n+1 ) = P ∆x v ∆ (t − n+1 ). (5.96)In our case, this operator can be defined as the P 1 interpolation of v ∆ :P ∆x v ∆ (t − n+1 ) = I [1 V ∆ (t − n+1 )] .In the class of Godunov schemes, the local truncation error (4.73) can beexpressed by means of the projection error as follows.Theorem 5.9. [LT01] Let the approximation v ∆ be defined by (5.95)–(5.96). Then,the norm of the consistency error (4.73) may be bounded as:‖F ‖ L 1 (R d ×[0,T ]) ≤ T ∆tmax ∥ v ∆ (t −0


✐✐136 Chapter 5. First-order approximation schemes∫ xj+1I 1 [V ](x) = v(x j ) + x − x jv ′ (ξ)dξ =∆x x j∫ [xj+1x − x j= v(x j ) +v ′ (x j ) +x j∆x∫ ξx jv ′′ (σ)dσ(we denote by 1 [xj,x] the characteristic function of the interval [x j , x] <strong>and</strong> recall thatthe derivative of the P 1 –interpolate is the integral mean value of v ′ on [x j , x j+1 ]).We can there<strong>for</strong>e express the interpolation error as|v(x) − I 1 [V ](x)| =∣ v′ (x j )+∫ xj+1(1 [xj,x](ξ) − x − x jx j∆x) ∫ ξ∫ xj+1(1 [xj,x](ξ) − x − x jx j∆x])dξ +dξv ′′ (σ)dσdξx j∣ . (5.99)Note that, in (5.99), the first integral is identically zero, <strong>and</strong> hence, giving a boundon the second integral by Hölder’s inequality,∫ xj+1∣ ∣∣∣|v(x) − I 1 [V ](x)| ≤ 1 [xj,x](ξ) − x − x ∫jξ∆x ∣ dξ · sup |v ′′ (σ)|dσ ≤ξ x j≤ ∆xx j∫ xj+1x j|v ′′ (σ)|dσ, (5.100)where we have taken into account that the first term is the integral of a function takingvalues in [0, 1], <strong>and</strong> that the second is maximized setting ξ = x j+1 . Integrating(5.100) on the interval [x j , x j+1 ], we have∫ xj+1x j∫ xj+1|v(x) − I 1 [V ](x)|dx ≤ ∆x 2 |v ′′ (σ)|dσ,<strong>and</strong> last, summing over all elementary intervals, we obtainx j‖v − I 1 [V ]‖ L1 (R) ≤ ∆x 2 ‖v ′′ ‖ L1 (R). (5.101)Finally, in order to apply Theorem 5.9, we must recall that neither the evolutionoperator, nor the P 1 interpolation exp<strong>and</strong> the L 1 norm of the second derivative, asit can be easily seen by interpreting the solution as the integral of the solution ofthe associated conservation law (2.46). We have there<strong>for</strong>eStability‖F ‖ L1 (R×[0,T ]) ≤ T ‖u ′′ ∆x 20‖ L1 (R)∆t . (5.102)Along with some remarks on the CFL condition, we will address here the twoconcepts of nonlinear stability used in the two different convergence theories, namelymonotonicity <strong>and</strong> uni<strong>for</strong>m semiconcavity.CFL condition Looking at the basic definition (5.92) of the scheme, one could inferthat, since α is allowed to vary over the whole of R, then the numerical domainof dependence also coincides with R <strong>and</strong> the CFL condition is always satisfied. Inpractice, whenever the Lipschitz constant of u 0 (hence, of u(t) <strong>for</strong> positive t) isknown, it is possible to restrict the search <strong>for</strong> a minimum to a smaller subset of R,still without violating the CFL condition.


✐✐5.2. Treating the convex HJ equation 137Monotonicity First, note that the SL scheme is invariant <strong>for</strong> the addition of constantssince, <strong>for</strong> a Lagrange interpolation of any order, I[V + c](x) ≡ I[V ](x) + c.To check that the SL scheme is monotone, consider two sequences V <strong>and</strong> Wsuch that V − W ≥ 0 componentwise. We have:SjSL (∆; V ) = minα∈R [∆tH∗ (α) + I 1 [V ](x j − α∆t)] == ∆tH ∗ (ᾱ j ) + I 1 [V ](x j − ᾱ j ∆t),SjSL (∆; W ) = minα∈R [∆tH∗ (α) + I 1 [W ](x j − α∆t)] == ∆tH ∗ (˜α j ) + I 1 [W ](x j − ˜α j ∆t) ≤≤ ∆tH ∗ (ᾱ j ) + I 1 [W ](x j − ᾱ j ∆t)(as be<strong>for</strong>e, the last inequality is obtained by replacing the exact minimizer <strong>for</strong> W ,˜α j , with ᾱ j ). There<strong>for</strong>e,S SLj(∆; V ) − S SL (∆; W ) ≥ I 1 [V ](x j − ᾱ j ∆t) − I 1 [W ](x j − ᾱ j ∆t),j<strong>and</strong> since P 1 interpolation is monotone itself (that is, I 1 [V ] − I 1 [W ] ≥ 0), we getSjSL (∆; V ) − Sj SL (∆; W ) ≥ 0. (5.103)Uni<strong>for</strong>m discrete semiconcavity Assume that at the n-th step the discrete semiconcavityassumptionvr−h n − 2vr n + vr+h n ≤ C∆x 2 (5.104)holds <strong>for</strong> the discrete solution at any node x r with an increment ±h∆x (h being aninteger). Then, using the notation of (5.93), at the (n + 1)–th step we havev n+1j−h − 2vn+1 j + v n+1j+h = ∆tH∗ (ᾱ j−h ) + I 1 [V n ](x j−h + ᾱ j−h ∆t) −−2∆tH ∗ (ᾱ j ) − 2I 1 [V n ](x j + ᾱ j ∆t)] ++∆tH ∗ (ᾱ j+h ) + I 1 [V n ](x j+h + ᾱ j+h ∆t) ≤≤ ∆tH ∗ (ᾱ j ) + I 1 [V n ](x j−h + ᾱ j ∆t) −−2∆tH ∗ (ᾱ j ) − 2I 1 [V n ](x j + ᾱ j ∆t) ++∆tH ∗ (ᾱ j ) + I 1 [V n ](x j+h + ᾱ j ∆t)where we have bounded the sum from above by replacing the minimizers <strong>for</strong> thenodes x j±h , i.e., ᾱ j±h , with the minimizer <strong>for</strong> the node x j . Now, <strong>for</strong> I 1 [V n ](x j +ᾱ j ∆t) = ∑ i vn i ψ[1] i (x j + ᾱ j ∆t), by periodicity of the grid we can setψ [1]i (x j + ᾱ j ∆t) = ψ [1]i±h (x j±h + ᾱ j ∆t) = β iso that I 1 [V n ](x j±h + ᾱ j ∆t) = ∑ i β iv n i±h . Recalling that the P 1 reconstructionconsists of a convex combination of the values in neighbouring nodes, we have thatβ i ≥ 0 <strong>and</strong> ∑ i β i ≡ 1. By we obtain there<strong>for</strong>ev n+1j−h − 2vn+1 j + v n+1j+h ≤ ∑ iβ i v n i−h − 2 ∑ iβ i v n i + ∑ iβ i v n i+h == ∑ i≤ ∑ iβ i (v n i−h − 2v n i + v n i+h) ≤β i C∆x 2 = C∆x 2 , (5.105)


✐✐138 Chapter 5. First-order approximation schemes<strong>and</strong> the solution satisfies the same semiconcavity estimate at the (n + 1)–th step.ConvergenceIn view of the consistency estimate (5.94) <strong>and</strong> the monotonicity result (5.103), convergenceof the first-order SL scheme follows from the Barles–Souganidis theorem.Theorem 5.10. Let H satisfy the basic assumptions, u 0 ∈ W 1,∞ , u be the solutionof (5.70) <strong>and</strong> vjn be defined by (5.92) with v0 j = u 0(x j ). Then, <strong>for</strong> any j ∈ Z <strong>and</strong>n ∈ [1, T/∆t], ∣ vj n − u(x j , t n ) ∣ → 0 (5.106)locally uni<strong>for</strong>mly, as ∆t → 0, with ∆x = o ( ∆t 1/2) .The parallel result, obtained by means of the Lin–Tadmor convergence theorem,uses the consistency estimate (5.102) <strong>and</strong> the uni<strong>for</strong>m semiconcavity result(5.105), <strong>and</strong> may be stated as follows.Theorem 5.11. Let H satisfy the basic assumptions, u 0 ∈ W 1,∞ , u be the solutionof (5.70) <strong>and</strong> v n j be defined by (5.92) with v0 j = u 0(x j ). Assume moreover that u 0is semi-concave <strong>and</strong> compactly supported. Then, <strong>for</strong> any j ∈ Z <strong>and</strong> n ∈ [1, T/∆t],‖u(t n ) − I 1 [V n ]‖ L1 (R) ≤ C ∆x2∆t(5.107)as ∆t → 0, with ∆x = o ( ∆t 1/2) .5.2.4 Multiple space dimensionsAs it has been done <strong>for</strong> the advection equation, we briefly turn to the n–dimensionalproblem:u t + H(u x1 , . . . , u xd ) = 0 R d × [0, T ]. (5.108)Most of the work of extending one-dimensional schemes <strong>for</strong> HJ equations to themultidimensional case follows the same principles of the linear case. In particular,schemes in differenced <strong>for</strong>m have the general structure (4.50), so that <strong>for</strong> examplethe two-dimensional version of the Lax–Friedrichs scheme readsv n+1j 1,j 2= 1 (vn4 j1−1,j 2+ vj n 1+1,j 2+ vj n 1,j 2−1 + vj n )1,j 2+1 +( vnj1+1,j+∆tH2− vj n 1−1,j 2, vn j − )1,j 2+1 vn j 1,j 2−1,∆x 1∆x 2<strong>and</strong> both consistency <strong>and</strong> monotonicity may be proved by the very same argumentsused <strong>for</strong> the one-dimensional case. Also, the d-dimensional <strong>for</strong>m of the SL scheme,v n+1j= minα∈R d [∆tH ∗ (α) + I 1 [V n ](x j − α∆t)]retains the same <strong>for</strong>mal structure of the one-dimensional <strong>for</strong>m, although (as it hasbeen noticed <strong>for</strong> the linear scheme) reconstruction <strong>and</strong> minimization are now per<strong>for</strong>medin R d . Again, the proofs of consistency (which is obtained by comparisonwith the Hopf–Lax <strong>for</strong>mula) <strong>and</strong> monotonicity follow the same criteria used in asingle space dimension.


✐✐5.2. Treating the convex HJ equation 1395.2.5 Boundary conditionsAs we have seen in Chapter 2, boundary conditions <strong>for</strong> HJ equations must beinterpreted in a weak sense, with the only exception of periodic conditions which canbe numerically implemented as it has been sketched <strong>for</strong> linear problems. Concerningboundary conditions of other kinds, we will limit ourselves here to the basic ideas<strong>for</strong> their implementation in SL schemes. In this case, the typical technique has beenoutlined in the treatment of Dirichlet conditions <strong>for</strong> the advection equation, whichrequire a variable step algorithm to bring the foot of characteristics precisely on theboundary. In addition, it might happen that boundary conditions are applied atnodes belonging to a neighborhood of the boundary instead that just on the boundarynodes (as it is usual in classical finite difference schemes). This point will be usefulin treating Neumann condition.We consider three different boundary conditions: Dirichlet, Neumann <strong>and</strong>State Constraint.Dirichlet boundary conditions The boundary operator <strong>for</strong> Dirichlet conditions isB(x, u, Dϕ(x)) = u(x) − b(x), where b is the value imposed on the boundary ∂Ω.Characteristics ending outside Ω should be brought to the boundary following thesame idea used in (5.65). As a result, (5.92) is modified asv n+1j = min [∆tH ∗ (α) + I 1 [V n ](x j − ατ)] (5.109)α∈R dτ≤∆tx j−ατ∈Ωwhere the value I 1 [V n ](x j − ατ) should be understood asI 1 [V n ](x j − ατ) = b(x j − ατ)if x j − ατ ∈ ∂Ω. The minimization in (5.109) should be interpreted as follows. Theminimum is searched <strong>for</strong> with τ = ∆t. If the minimum is found at an internal point(i.e., <strong>for</strong> x j − ᾱ j ∆t ∈ Ω), then nothing else is necessary. If no minimum point isfound in the interior of Ω, then among all boundary points which can be writtenas x j − ᾱ j τ, it is necessary to select the one which achieves the minimum withrespect to α ∈ R d <strong>and</strong> τ ≤ ∆t. In the chapter on Dynamic Programming, we willre-interpret this boundary condition as a stopping cost.Neumann boundary conditionsHere, the boundary operator isB(x, u, Du) = ∂u (x) − m(x),∂νwhere m is a given function. In order to assign a value to points outside (but closeto) Ω, we can use the directional derivative on the boundary to writeu(x + δν(x)) = u(x) + δm(x), x ∈ ∂Ω. (5.110)Then, Neumann boundary conditions can be en<strong>for</strong>ced by using directly (5.110) toreplace the value I 1 [V n ](x j − α∆t) whenever x j − α∆t /∈ Ω. A different way is toextend the grid outside the domain Ω with the introduction of a neighborhood inwhich the values are assigned according to (5.110) <strong>and</strong> reconstructed as usual byinterpolation in (5.92).


✐✐140 Chapter 5. First-order approximation schemesFor example, if m(x) ≡ 0 <strong>and</strong> Ω is a rectangle of R 2 , then we just need to addan extra “frame” of nodes surrounding the boundary at a distance δ outside Ω, thecorresponding values being defined asu(x + δν(x)) = u(x),x ∈ ∂Ω.Clearly, the points x j − α∆t must be kept within this enlarged domain, either by asuitable choice of δ, or by a variable step technique.State constraints boundary conditions This boundary condition is the easiest toimplement. In fact, we recall that what is needed is just to extend the solution uoutside Ω by settingu(x) = C, x ∈ R d \ Ωwhere C is an upper bound <strong>for</strong> the L ∞ (Ω) norm of the solution. In the numericalimplementation, in order to preserve the regularity of the function to be minimizedin (5.92), it could be necessary to implement again a variable step algorithm, with Cas a Dirichlet boundary datum. Note that, in a Dynamic Programming framework,this “large” value could be interpreted as a penalization (made through the stoppingcost) of trajectories which violate the state constraints.5.3 ExamplesWe present a new set of simple numerical examples, using the one-dimensionalmodel problem{u t (x, t) + 1 2 |u x(x, t)| 2 = 0 (x, t) ∈ (0, 1) × (0, T )(5.111)u(x, 0) = u 0 (x)with T = 0.05 <strong>and</strong> two different initial conditions u 0 with bounded support. Thefirst is a Lipschitz continuous function (the same used <strong>for</strong> the tests of subsection5.1.6):u 0 (x) = max(1 − 16(x − 0.25) 2 , 0), (5.112)whereas the second, obtained by a simple change of sign, is also semiconcave:u 0 (x) = − max(1 − 16(x − 0.25) 2 , 0). (5.113)Using the initial condition (5.112), the solution eventually develops a singularitywith nonempty superdifferential. After the onset of the singularity, the exact solutionreads{(|x−12|− 1 4) 2u(x, t) = 2tif 1 4 ≤ x ≤ 3 40 elseOn the other h<strong>and</strong>, using the initial condition (5.113), the solution has the expression(∣ ∣x −1∣ 2 )2u(x, t) = min2t + 1 − 1, 0 .16Again, the test is per<strong>for</strong>med with Upwind, Lax–Friedrichs <strong>and</strong> Semi-Lagrangianschemes. In this case, the refinement has been carried out with ∆t = ∆x/40 <strong>for</strong>Upwind scheme, ∆t = ∆x/20 <strong>for</strong> Lax–Friedrichs scheme <strong>and</strong> ∆t = 0.01 (fixed)


✐✐5.4. Stationary problems 141W 1,∞ initial conditionn n Upwind LF SL25 1.13 · 10 −1 2.84 · 10 −1 8.82 · 10 −250 1.01 · 10 −1 2.51 · 10 −1 3.53 · 10 −2100 6.62 · 10 −2 1.89 · 10 −1 1.81 · 10 −2200 4.08 · 10 −2 1.27 · 10 −1 8.26 · 10 −3400 2.42 · 10 −2 8.0 · 10 −2 3.94 · 10 −3rate 0.56 0.46 1.12Semiconcave initial conditionn n Upwind LF SL25 6.42 · 10 −2 3.64 · 10 −1 2.11 · 10 −250 3.58 · 10 −2 1.97 · 10 −1 5.02 · 10 −3100 1.92 · 10 −2 9.76 · 10 −2 1.24 · 10 −3200 9.97 · 10 −3 4.83 · 10 −2 3.07 · 10 −4400 5.08 · 10 −3 2.4 · 10 −2 7.63 · 10 −5rate 0.91 0.98 2.03Table 5.3. Errors in the ∞-norm <strong>for</strong> problem (5.111), Upwind, Lax–Friedrichs <strong>and</strong> Semi-Lagrangian schemes.<strong>for</strong> Semi-Lagrangian scheme. Table 5.3 shows numerical errors in the ∞-norm <strong>and</strong>Figure 5.8 compares exact with numerical solutions <strong>for</strong> the three schemes in bothexamples.A first clear outcome of this example is that, although both initial conditionsare Lipschitz continuous, having a semiconcave initial condition (<strong>and</strong> hence,a uni<strong>for</strong>mly semiconcave solution) makes a difference. In the first test the schemesbasically respect the theory, by keeping a convergence rate of about 1/2 <strong>for</strong> Upwind<strong>and</strong> Lax–Friedrichs schemes, of about 1 <strong>for</strong> Semi-Lagrangian scheme (note that thisis the order of truncation error <strong>for</strong> a Lipschitz solution in the ∞-norm, once thetime step is kept constant). On the other h<strong>and</strong>, in the second test convergence ratesare almost doubled. A heuristic explanation of this effect is provided by Remark2.25: the <strong>Hamilton</strong>ian <strong>and</strong> its conjugate being smooth <strong>and</strong> coercive, the foot ofa characteristic must be a point of differentiability <strong>for</strong> u. Roughly speaking, thismeans that a semiconcave solution always propagates from regular points towardssingularities, <strong>and</strong> this reduces numerical errors.Note that, in spite of a similar convergence rate, LF scheme gives a poor approximationin terms of absolute accuracy, even with respect to Upwind scheme. SLscheme has the best per<strong>for</strong>mances in both convergence rate <strong>and</strong> absolute accuracy,still being monotone.5.4 Stationary problemsIn this section, we will sketch how the theory introduced so far can be adapted tostationary problems, respectively of linear <strong>and</strong> <strong>Hamilton</strong>–Jacobi type. As a generalstrategy, stationary equations will be regarded as limit states of evolutive equations,


✐✐142 Chapter 5. First-order approximation schemesUpwindLax–FriedrichsSemi-LagrangianFigure 5.8. Numerical results <strong>for</strong> problem (5.111) with Lipschitz (left) <strong>and</strong>semiconcave (right) initial condition, obtained via Upwind (upper), Lax–Friedrichs(center) <strong>and</strong> Semi-Lagrangian (lower) schemes, 200 nodes.<strong>and</strong> as a consequence it will be natural to consider schemes in the time-marching<strong>for</strong>m.5.4.1 The linear caseWhen posed on the whole of R, a general <strong>for</strong>m of the stationary linear advectionequation isu(x) + f(x)u x (x) = g(x) x ∈ R, (5.114)


✐✐5.4. Stationary problems 143whose solution u(x) can be considered as the limit <strong>for</strong> t → ∞ of a solution u(x, t)of the equationu t (x, t) + u(x, t) + f(x)u x (x, t) = g(x) x ∈ R, (5.115)regardless of the initial condition used. Note that equations having a differentcoefficient of the zeroth order term u(x) may be brought to this <strong>for</strong>m by a rescaling.We will proceed to write the various schemes in the time-marching <strong>for</strong>m (4.19)by applying them to (5.115). Note that the extension to more general situations(multiple space dimensions, boundary conditions) can be carried out in a fairlystraight<strong>for</strong>ward way by following the guidelines given <strong>for</strong> evolutive problems.Upwind discretizationFollowing the general strategy outlined in the first part of the chapter, (5.115) wouldbe discretized, <strong>for</strong> f(x j ) > 0, asv n+1j<strong>and</strong> <strong>for</strong> f(x j , t n ) < 0, as− v n j∆t+ v n j + f(x j ) vn j − vn j−1∆x= g(x j ),v (k+1)j =⎪⎨v n+1j− v n j∆t⎪⎩ (1 − ∆t)v (k)+ v n j + f(x j ) vn j+1 − vn j∆x= g(x j ).Once replaced the index of the time step by the index of iteration, the <strong>for</strong>m of thescheme <strong>for</strong> f(x) changing sign is there<strong>for</strong>e⎧(1 − ∆t)v (k)j − f(x j ) ∆t [ ]v (k)j − v (k)j−1 + ∆tg(x j ) if f(x j ) > 0∆x(1 − ∆t)v (k)j + ∆tg(x j ) if f(x j ) = 0j− f(x j ) ∆t [ ]v (k)j+1∆x− v(k) j + ∆tg(x j ) if f(x j ) < 0.(5.116)Here <strong>and</strong> in the whole section, since we look <strong>for</strong> the limit solution, the choice of theinitial vector V (0) is irrelevant.Convergence Once put the scheme in the <strong>for</strong>m (4.20), it is easy to check that,under the slightly more restrictive CFL condition‖f‖ ∞ ∆t∆x< 1 − ∆t (5.117)the scheme satisfies assumption (4.21), <strong>and</strong> more precisely ‖B Up (∆)‖ ∞ = 1 − ∆t.There also follows (see Subsection 4.2.5) that it satisfies the consistency estimate(5.10), <strong>and</strong> converges by Theorem 4.12. We have there<strong>for</strong>e:Theorem 5.12. Let f, g ∈ W 1,∞ , u be the solution of (5.114) <strong>and</strong> v j = lim k v (k)j ,with v (k)j defined by (5.116). Then, <strong>for</strong> any j ∈ Z,|v j − u(x j )| → 0 (5.118)


✐✐144 Chapter 5. First-order approximation schemesas ∆t, ∆x → 0, with the CFL condition (5.117).Moreover, if u has bounded second derivative, then‖V − U‖ ∞≤ C∆x. (5.119)Central discretizationFor simplicity, Lax–Friedrichs scheme will be adapted to equation (5.115) using oncemore the approximationv(x j ) = v(x j−1) + v(x j+1 )2+ O(∆x 2 )in treating the zeroth-order term. With this choice, the time-marching <strong>for</strong>m of theLF scheme isv (k+1)j= (1 − ∆t) v(k) j−1 + v(k) j+1− f(x j ) ∆t []v (k)j+122∆x− v(k) j−1 + ∆tg(x j ). (5.120)Convergence Comparing (5.120) with (4.20), it is easily seen that the LF schemesatisfies the stability condition ‖B LF (∆)‖ ∞ = 1 − ∆t under the CFL constraint(5.117). Moreover, as <strong>for</strong> the upwind scheme, LF scheme in this <strong>for</strong>m retains theconsistency estimate of the evolutive scheme, i.e. (5.24). There<strong>for</strong>e, applying Theorem4.12, we obtain the following convergence result:Theorem 5.13. Let f, g ∈ W 1,∞ , u be the solution of (5.114) <strong>and</strong> v j = lim k v (k)j ,with v (k)j defined by (5.120). Then, <strong>for</strong> any j ∈ Z,|v j − u(x j )| → 0 (5.121)as ∆t, ∆x → 0, with ∆x = o ( ∆t 1/2) <strong>and</strong> the CFL condition (5.117).Moreover, if u has bounded second derivative <strong>and</strong> ∆x = c∆t <strong>for</strong> some constant c,then‖V − U‖ ∞≤ C∆x. (5.122)Semi-Lagrangian discretizationTo write the time-marching version of the CIR scheme, we start from the representation<strong>for</strong>mula (5.36) which is changed into its more general versionu(x, t) =∫ tt−∆te s−t g(y(x, t; s))ds + e −∆t u(y(x, t; t − ∆t), t − ∆t). (5.123)Here, the ODE satisfied by characteristics is driven by a vector field f(x) whichdoes not depend on t. We can there<strong>for</strong>e set conventionally t = 0 <strong>and</strong> omit theinitial time in the notation of y, thus obtaining the representation <strong>for</strong>mulau(x, t) ==∫ 0−∆t∫ ∆t0e s g(y(x; s))ds + e −∆t u(y(x; −∆t), t − ∆t) =e −s g(y(x; −s))ds + e −∆t u(y(x; −∆t), t − ∆t).


✐✐5.4. Stationary problems 145In order to treat this case, the augmented system (5.52) is modified in the <strong>for</strong>m(ẏ(x; ) ( )s) f(y(x; s))=˙γ(x; s) −e −s (5.124)g(y(x; s))with the initial conditions y(x; 0) = x, γ(x; 0) = 0. Denoting again by X ∆ (x; s) <strong>and</strong>G ∆ (x; s) the approximations of y(x; s) <strong>and</strong> γ(x; s), <strong>and</strong> replacing the time index bythe iteration index, we finally obtain the fixed point <strong>for</strong>mv (k+1)j = G ∆ (x j ; −∆t) + e −∆t I 1[V (k)] (X ∆ (x j ; −∆t) ) . (5.125)Convergence Following the arguments used <strong>for</strong> the evolutive case, we can checkthat the scheme satisfies the consistency estimate (5.57), <strong>and</strong> that ‖B CIR (∆)‖ ∞ =e −∆t = 1 − ∆t + O ( ∆t 2) . It is there<strong>for</strong>e possible to apply Theorem 4.12 to provethe following convergence result:Theorem 5.14. Let f, g ∈ W p,∞ , u be the solution of (5.114) <strong>and</strong> v j = lim k v (k)with v (k)j defined by (5.125). Assume moreover that (5.55), (5.56) hold. Then, <strong>for</strong>any j ∈ Z,|v j − u(x j )| → 0 (5.126)as ∆t → 0, ∆x = o ( ∆t 1/2) .Moreover, if u ∈ W s,∞ (R) (s = 1, 2), then‖V − U‖ ∞≤ C(∆t p + ∆xs∆tj ,). (5.127)5.4.2 The nonlinear caseIn adapting the various schemes to stationary HJ equations, we refer to the stationarymodel which in some sense parallels (5.114), that isu(x) + H(u x (x)) = 0 x ∈ R. (5.128)As be<strong>for</strong>e, we consider time-marching schemes, either in differenced <strong>for</strong>m or ofSemi-Lagrangian type.Discretization in differenced <strong>for</strong>mIn differenced time-marching schemes, the schemes are applied to the evolutiveequationu t + u + H(u x ) = 0,whose solution converges to a regime state satisfying (5.128). Keeping the schemein its most general <strong>for</strong>m, ad adding the zeroth order term, we havev n+1j− v n j∆twhich is clearly a consistent scheme.iteration index, we obtainv (k+1)j= (1 − ∆t)v (k)j+ v n j + H (D j−p [V n ], . . . , D j+q [V n ]) = 0,Hence, replacing the time index with the+ ∆tH(D j−p[V (k)] [, . . . , D j+q V (k)]) . (5.129)


✐✐146 Chapter 5. First-order approximation schemesrepeating the computations per<strong>for</strong>med <strong>for</strong> the evolutive case, we can show that boththe upwind <strong>and</strong> the Lax–Friedrichs scheme satisfy the assumptions of the stationaryversion of Barles–Souganidis theorem (Theorem 4.26), with a contraction coefficientof L S = 1−∆t, provided a suitable upper bound on the Courant number is satisfied.Both schemes are there<strong>for</strong>e convergent to the viscosity solution of (5.128):Theorem 5.15. Let H satisfy the basic assumptions, u be the solution of (5.128)<strong>and</strong> v j = lim k v (k)j , with v (k)j defined by (5.129). Assume moreover that the numerical<strong>Hamilton</strong>ian H is consistent <strong>and</strong> monotone in the sense of Theorem 4.20. Then,<strong>for</strong> any j ∈ Z,|v j − u(x j )| → 0 (5.130)as ∆ → 0.Semi-Lagrangian discretizationIn the case of Semi-Lagrangian discretization, we start from a generalized <strong>for</strong>m ofthe Hopf–Lax <strong>for</strong>mula, which applies to the solution of (5.128):[(u(x) = min 1 − e−∆t ) H ∗ (a) + e −∆t u(x − a∆t) ] .a∈RA detailed derivation of this representation <strong>for</strong>mula will be shown in the chapterdevoted to Dynamic Programming, whereas <strong>for</strong> the moment we simply use it toconstruct a SL type discretization. Replacing the value u(x − a∆t) with its P 1numerical reconstruction, we have the scheme (in iterative <strong>for</strong>m):v (k+1)j= mina∈R[ (1− e−∆t ) H ∗ (a) + e −∆t I 1[V (k)] (x − a∆t)]. (5.131)Adapting the arguments used <strong>for</strong> the evolutive case, it can be shown once morethat the scheme is consistent <strong>for</strong> ∆x = o ( ∆t 1/2) <strong>and</strong> monotone <strong>for</strong> any ∆t/∆xrelationship. We can there<strong>for</strong>e apply Theorem 4.26 to state the convergence result:Theorem 5.16. Let H satisfy the basic assumptions, u be a bounded <strong>and</strong> uni<strong>for</strong>mlycontinuous solution of (5.114) <strong>and</strong> v j = lim k v (k)j , with v (k)j defined by (5.131).Then, <strong>for</strong> any j ∈ Z,|v j − u(x j )| → 0 (5.132)as ∆t → 0, ∆x = o ( ∆t 1/2) .5.5 Commented referencesThe linear theory <strong>for</strong> classical monotone difference schemes like Upwind or Lax–Friedrichs is widely treated in any textbook on difference schemes. A relativelyrecent reference on this subject is [Str89]. However, the application of such schemesto <strong>Hamilton</strong>–Jacobi equations is a much less established topic, <strong>and</strong> we are notaware of other monographs covering the subject. Unless <strong>for</strong> some more explicitcomputation, the theory reported here is basically taken from [CL84].The Courant–Isaacson–Rees has been proposed in [CIR52], <strong>and</strong> although theprocedure of construction outlined in the paper clearly implies interpolation, itsendpoint basically coincides with an upwind scheme. Moreover, the possibility ofworking at large Courant numbers was not recognized by that time. Most authors


✐✐5.5. Commented references 147believe [W59] to be the first work proposing a Semi-Lagrangian technique. In fact,[W59], along with the later paper [Ro81], has originated a considerable streamlineof literature concerning the application of SL techniques to Numerical Weather Prediction<strong>and</strong> environmental Fluid Dynamics, <strong>for</strong> which [SC91] represents a classical(although no longer up-to-date) review paper.In a completely independent way, SL schemes have been developed in thecomputational study of plasma physics. To our knowledge, the first paper presentingthis approach is [CK76], whereas further developments may be found in[GBSJFF90], [SBG99], [BM08].Once more independently, the Hopf–Lax representation <strong>for</strong>mula (or its equivalentin Optimal Control, the Dynamic Programming Principle) has been used inorder to construct numerical schemes <strong>for</strong> HJ equations. This approach has originatedin the 70s with the probabilistic techniques of the so-called Markov chainapproximations, <strong>for</strong> which a recent review can be found in the monograph [KuD01].After the emergence of the theory of viscosity solutions, a re<strong>for</strong>mulation of thisapproach has been given <strong>for</strong> deterministic problems. A review on this streamline ofresearch is given in [F97]. We also mention the papers [FF94], in which the couplingbetween a monotone reconstruction <strong>and</strong> a high-order time discretization has beenfirst proposed, <strong>and</strong> [FF98, FF02] which analyze the SL scheme <strong>for</strong> respectively theadvection <strong>and</strong> the HJ equation in a more numerical framework.


✐✐148 Chapter 5. First-order approximation schemes


✐✐Chapter 6High-order SLapproximation schemesIt is well known that, despite the increasing number of accurate <strong>and</strong> efficient highorderschemes available in the literature, the related convergence theory is usuallypoorer. This chapter presents the state-of-art of theoretical analysis <strong>for</strong> high-orderSemi-Lagrangian schemes in the fields of transport <strong>and</strong> <strong>Hamilton</strong>–Jacobi equations.The presentation is dedicated to SL schemes built by means of Lagrange, finiteelement <strong>and</strong> non-oscillatory reconstructions. In its structure, the chapter parallelsthe previous one, unless <strong>for</strong> the introduction of different <strong>and</strong> more complex toolsat the level of stability analysis. Also, a subsection of one-dimensional examples isincluded <strong>for</strong> both the linear <strong>and</strong> the nonlinear model, <strong>and</strong> a qualitative study of theinterplay between the two discretization steps is presented.6.1 Semi-Lagrangian schemes <strong>for</strong> the advectionequationIn turning our attention to high-order Semi-Lagrangian schemes, we get back to thecase of the one-dimensional linear advection equation, which is rewritten here as{u t (x, t) + f(x, t)u x (x, t) = g(x, t) (x, t) ∈ R × [t 0 , T ](6.1)u(x, 0) = u 0 (x) x ∈ R.As it has been done in the previous chapter, we will also examine in detail the casewith constant coefficients, namely,(with c > 0) whenever useful to illustrate the main ideas.6.1.1 Construction of the schemeu t + cv x = 0 (6.2)A first step to obtain a fully high-order discretization of the representation <strong>for</strong>mula(5.32), (5.36) has already been presented concerning the CIR scheme <strong>and</strong> the possibilityto track characteristics in a more accurate way. This increase in the accuracyof time discretization <strong>for</strong> the augmented system (5.60) is summarized in Theorem5.6. The remaining step is to replace the P 1 reconstruction used in the CIR schemewith an interpolation of order r, which will be denoted by I r .149


✐✐150 Chapter 6. High-order SL approximation schemesWith this further improvement, the SL scheme can be rewritten as{v n+1j = G ∆ (x j , t n+1 ; t n ) + I r [V n ] ( X ∆ (x j , t n+1 ; t n ) )vj 0 = u 0(x j ).(6.3)Note that, as it has been remarked in the first-order case, this <strong>for</strong>m also applies toa scheme in d space dimensions, as soon as the augmented system (5.52) is set indimension d + 1 <strong>and</strong> I r is a d-dimensional reconstruction.6.1.2 ConsistencyConsistency analysis <strong>for</strong> the scheme (6.3) requires the basic estimates (5.55)–(5.56),which read∣ X ∆ (x j , t n+1 ; t n ) − y(x j , t n+1 ; t n ) ∣ = O(∆t p+1 ),∫ tn+1∣ G∆ (x j , t n+1 ; t n ) − g(y(x j , t n+1 ; s), s)ds∣ = O(∆tp+1 ),as well as the corresponding estimate <strong>for</strong> interpolation:t n‖u − I r [U]‖ ∞= O(∆x r+1 ). (6.4)Besides assuming that f, g ∈ W p,∞ <strong>for</strong> (5.55), (5.56) to hold, (6.4) requires typicallythat u ∈ W r+1,∞ . Then, retracing the proof of (5.44), (5.57), <strong>and</strong> incorporatingthe interpolation error (6.4), we obtain the consistency estimate∥ L SL (∆; t, U(t)) ∥ ()≤ C ∆t p + ∆xr+1. (6.5)∆tNote that the scheme ( is)still conditionally consistent, but under a weaker constraint,that is, ∆x = o ∆t 1r+1 .6.1.3 StabilityWhile high-order characteristics tracking does not change the stability propertiesof the CIR scheme, the introduction of a space reconstruction of degree r > 1 does.In this situation, the scheme is no longer monotone, <strong>and</strong> a completely differentstability analysis must be per<strong>for</strong>med.A short digression: the Lagrange–Galerkin schemeIn this small detour, we introduce a class of schemes which will be useful in the sequelto analyse the stability of SL schemes under Lagrange reconstructions. Similar toSL schemes, Lagrange–Galerkin (LG) schemes are constructed by a discretizationof the representation <strong>for</strong>mula in the <strong>for</strong>m (5.32) or (5.36). Unlike the SL schemes,however, LG schemes per<strong>for</strong>m the space reconstruction by means of a Galerkinprojection.To give a more precise expression, assume <strong>for</strong> simplicity that g ≡ 0, so thatthe representation <strong>for</strong>mula <strong>for</strong> the solution would bev(x, t n+1 ) = v(y(x, t n+1 ; t n ), t n ). (6.6)


✐✐6.1. Semi-Lagrangian schemes <strong>for</strong> the advection equation 151The first step is to write the numerical solution at time t n , v n h , in the base {φ i}:v n h(x) = ∑ iv n i φ i (x). (6.7)Note that we have used <strong>for</strong> the numerical solution the notation vh n , usual in Galerkinschemes (here, typically h denotes the space discretization step ∆x), <strong>and</strong> that wehave discriminated the Galerkin basis function φ i from the basis functions ψ i used<strong>for</strong> interpolation. In fact, the Galerkin basis functions have a conceptually differentrole, <strong>and</strong> in particular need not to be cardinal functions. Rather, they are assumedto be in L 2 , <strong>and</strong> to be asymptotically dense as ∆x → 0. The steps to turn therepresentation <strong>for</strong>mula (6.6) into a LG scheme are:• replacing the exact solution by its approximation (6.7) <strong>and</strong> the exact displacementalong characteristics y with its approximation X ∆ ;• multiplying both sides by a test function w h = ∑ i wn i φ i(x);• integrating with respect to the space variable.This amounts to define the scheme by∫∫v n+1h(ξ)w h (ξ)dξ = vhn (X ∆ (ξ, t n+1 ; t n ) ) w h (ξ)dξ, (6.8)RRwhere the equality has to hold <strong>for</strong> any w h in the space generated by the base {φ i }.Hence, using (6.7) <strong>and</strong> using as test functions the basis functions φ j , the resultingscheme is set in the <strong>for</strong>m:∫∑∫∑v n+1i φ i (ξ)φ j (ξ)dξ = vi n (φ i X ∆ (ξ, t n+1 ; t n ) ) φ j (ξ)dξRiRiwhere j ranges over the set of admissible indices I. This condition is actuallyen<strong>for</strong>ced as∑∫v n+1i φ i (ξ)φ j (ξ)dξ = ∑ ∫v n (i φ i X ∆ (ξ, t n+1 ; t n ) ) φ j (ξ)dξ. (6.9)iR i RNote now that the LG scheme, as it has been defined, ( is stable in L 2 . Forexample, in the constant coefficient case, we have that vhn X ∆ (ξ, t n+1 ; t n ) ) =vh n (ξ − c∆t)) is a pure translation, so that ∥ ( vnhX ∆ (·, t n+1 ; t n ) )∥ ∥2= ‖vh n(·))‖ 2 .There<strong>for</strong>e, using w h = v n+1has a test function in (6.8), <strong>and</strong> applying Hölder’s inequality,we get∫∥ vn+1∥ 2 h 2 = v n (h X ∆ (ξ, t n+1 ; t n ) ) v n+1h(ξ)dξ ≤R∥ ∥≤ ‖vh‖ n ∥v n+12h<strong>and</strong> this shows that the scheme is stable in the L 2 norm. More in general, the LGscheme is stable whenever the approximate evolution operator E ∆ defined byE ∆ (t − t n )vh(x) n = vhn (X ∆ (x, t; t n ) )∥2


✐✐152 Chapter 6. High-order SL approximation schemessatisfies <strong>for</strong> t − t n → 0 + the bound:∥ E ∆ (t − t n ) ∥ ∥ ≤ 1 + C(t − tn ).in which the left-h<strong>and</strong> side is understood as the norm of an operator mapping L 2into L 2 .It is worth to point out that, in fact, the definition (6.8) of LG schemesdoes not allow in general <strong>for</strong> an exact implementation – this happens because thebasis function are de<strong>for</strong>med by the approximate advection X ∆ , <strong>and</strong> the integralsappearing at the right-h<strong>and</strong> side of (6.8) cannot be exactly evaluated, unless <strong>for</strong>constant advection speed. Thus, the real implementation of LG schemes requiressome specific techniques, like approximate integration through quadrature <strong>for</strong>mulaeor area-weighting (we will get back to this latter technique when treating variablecoefficient equations). Rather then giving details on the implementation of LGschemes, our interest here is to use the stability of their exact version (6.8) toprove stability of SL schemes, via a result of equivalence between the two classes ofschemes. This tool will be the used to treat the case of Lagrange reconstructions.The constant coefficient caseFor the constant coefficient case, the SL scheme (6.3) takes the <strong>for</strong>m{v n+1j = I[V n ](x j − c∆t) = ∑ i vn i ψ i(x j − c∆t)vj 0 = u 0(x j )(6.10)in which we have expressed the interpolation in terms of the basis functions. Dependingon the strategy of reconstruction used, the scheme would require a differenttheory. Here, we will address two cases: the first is the case of odd order Lagrangereconstruction, <strong>for</strong> which the theory is essentially complete, the second is the caseof finite element reconstructions, <strong>for</strong> which a partial Von Neumann type analysiswill be presented.Lagrange reconstructions In this situation, the reconstruction is invariant bytranslation <strong>and</strong> Von Neumann analysis (by means of Fourier methods) gives a necessary<strong>and</strong> sufficient stability condition. The Von Neumann condition (4.47) readsin this case as∣ ∑∣∣∣∣|ρ(ω)| =ψ∣ l (x j − c∆t) e iω(l−j) ≤ 1 + C∆t.lAlthough an explicit proof of this inequality has been given in [BM08], such aproof is very technical <strong>and</strong> an easier way consists in showing that the SL scheme isequivalent to a Lagrange–Galerkin scheme. We recast there<strong>for</strong>e (6.9), <strong>for</strong> the caseunder consideration, as∑∫v n+1i φ i (ξ)φ j (ξ)dξ = ∑ ∫vin φ i (ξ − c∆t)φ j (ξ)dξ. (6.11)iR i RThe equivalence between (6.10) <strong>and</strong> (6.11) is stated by the following


✐✐6.1. Semi-Lagrangian schemes <strong>for</strong> the advection equation 153Theorem 6.1. Let the functions ψ i be defined by( ) ξψ i (ξ) = ψ∆x − i , (6.12)<strong>for</strong> some reference function ψ such thatψ ∈ W 2,1 (R) (6.13)ψ(y){= ψ(−y) (6.14)ψ(i) =1 if i = 00 if i ∈ Z, i ≠ 0.(6.15)Then, there exists a basis {φ i } such that the SL scheme (6.10) is equivalent to theLG scheme (6.11) if <strong>and</strong> only the function ψ(y) has a real nonnegative Fouriertrans<strong>for</strong>m:ˆψ(ω) ≥ 0. (6.16)Proof. We look <strong>for</strong> a set of basis functions <strong>for</strong> the LG scheme in the <strong>for</strong>mφ i (ξ) = √ 1 ( ) ξφ∆x ∆x − i(6.17)<strong>for</strong> some reference function φ to be determined. Comparing (6.10) <strong>and</strong> (6.11), weobtain the set of conditions to be satisfied:∫φ i (ξ)φ j (ξ)dξ = δ ij (6.18)∫RRφ i (ξ − c∆t)φ j (ξ)dξ = ψ i (x j − c∆t). (6.19)Note that, by assumption (6.15), condition (6.18) is in fact included in (6.19), asit can be seen setting ∆t = 0. Using the definitions (6.12), (6.17) <strong>for</strong> ψ k <strong>and</strong> φ k ,(6.19) can be rewritten as1∆x∫R( ξ − c∆tφ∆x) ( ) ( )ξ− i φ∆x − j xj − c∆tdξ = ψ− i∆xthat is, after defining η = ξ/∆x − j <strong>and</strong> λ = c∆t/∆x:∫φ(η − λ + j − i)φ(η)dη = ψ(−λ + j − i).R(6.20)This ultimately amounts to find a function φ such that∫φ(η + y)φ(η)dη = ψ(y). (6.21)RThe left-h<strong>and</strong> side of (6.21) is the autocorrelation integral (see [P77]) of the unknownfunction φ. Working in the Fourier domain <strong>and</strong> trans<strong>for</strong>ming both sides of(6.21) we have:| ˆφ(ω)| 2 = ˆψ(ω). (6.22)


✐✐154 Chapter 6. High-order SL approximation schemesNow, since ψ is a real <strong>and</strong> even function of y, its Fourier trans<strong>for</strong>m ˆψ is also a real<strong>and</strong> even function of ω. Moreover, the assumption ψ ∈ W 2,1 (R) implies that ˆψ(ω)is bounded <strong>and</strong> decays like O(ω −2 ) <strong>for</strong> ω → ±∞.There<strong>for</strong>e, ˆψ being also nonnegative, its square root ˆψ(ω) 1/2 is real, even <strong>and</strong> nonnegative.In addition, ˆψ 1/2 is also bounded <strong>and</strong> decays like O(ω −1 ), <strong>and</strong> this impliesthat ˆψ 1/2 ∈ L 2 (R). Finally, looking at the inverse Fourier trans<strong>for</strong>m F −1 as an operatormapping L 2 (R) into L 2 (R), we obtain that the solution φ defined byφ(y) = F −1 { ˆψ(ω)1/2 } (6.23)is a well-defined even real function of L 2 (R) solving (6.21).In conclusion, the SL scheme (6.10) is L 2 stable, being equivalent to a stablescheme. Note that stability of LG schemes is intended in terms of the L 2 norm‖vh n‖ 2, but in our case this coincides with stability in the discrete norm ‖V n ‖ 2 . Infact, we have∫‖vh‖ n 2 2 = vh(x) n 2 dx =R∫ ( ) ⎛ ⎞∑= vi n φ i (x) ⎝ ∑ vj n φ j (x) ⎠ dx =R ij∫vi n vjn φ i (x)φ j (x)dx == ∑ i,j= ∑ iR(v n i ) 2 = 1∆x ‖V n ‖ 2 2where we have applied the orthogonality relationship (6.18). Thus, continuous <strong>and</strong>discrete norm coincide up to the scaling factor 1/∆x.Remark 6.2. Due to a theorem of Bochner (see [Sch99]), the functions havingpositive Fourier trans<strong>for</strong>m could also be characterized as positive semi-definitefunctions, defined as follows.Definition 6.3. A complex-valued function g : R d → R is said to be positivesemi-definite ifn∑ n∑a k g(x k − x j )ā j ≥ 0 (6.24)k=1 j=1<strong>for</strong> any x k ∈ R d , a k ∈ C (k = 1, . . . , n) <strong>and</strong> <strong>for</strong> all n ∈ N.Remark 6.4. We explicitly note that (6.23) needs not have (<strong>and</strong> in fact, has not ingeneral) a unique solution. Equating ˆψ <strong>and</strong> | ˆφ| 2 , we neglect any in<strong>for</strong>mation aboutthe phase diagram of φ, <strong>and</strong> this in turn makes it possible to have multiple solutions.We will get back to this point when treating the case with variable coefficients.An example: P 1 interpolation Although the case of P 1 interpolation needsnot this kind of stability analysis, we start by treating this simple case, <strong>for</strong> which


✐✐6.1. Semi-Lagrangian schemes <strong>for</strong> the advection equation 155explicit computations can be per<strong>for</strong>med. This will allow to illustrate some remarkablepoint, in particular the lack of uniqueness <strong>for</strong> the solution. We recall that inthis case the reference function has the <strong>for</strong>m⎧⎪⎨ 1 + y if − 1 ≤ y ≤ 0ψ [1] (y) = 1 − y if 0 ≤ y ≤ 1(6.25)⎪⎩0 elsewherewhose Fourier trans<strong>for</strong>m of is:ˆψ [1] (ω) = 2 − 2 cos ωω 2 =( )sinω 22( ω) 2(6.26)2Now, taking the square root of this trans<strong>for</strong>m as in (6.23), we get∣∣sinˆφ ω ∣[1] 2(ω) =∣ ω ∣<strong>and</strong> accordingly,2(6.27){∣ } ∣sin ω ∣φ [1] (y) = F −1 2∣ ω ∣. (6.28)This solution, numerically computed by FFT, is shown along with the referencefunction (6.25) in the upper row of Figure 6.1.On the other h<strong>and</strong>, a different (<strong>and</strong> possibly, more natural) solution can be pickedup by noting that ˆψ [1] (ω) is also the squared magnitude of2ˆφ [1] (ω) = sin ω 2ω2whose inverse Fourier trans<strong>for</strong>m is explicitly computable as{φ [1] 1 if − 1/2 ≤ y ≤ 1/2(y) =0 elsewhere.(6.29)(6.30)This example shows more precisely the effect of losing uniqueness by neglecting thephase in<strong>for</strong>mation – in fact, we can obtain an infinity of solutions to (6.21) whichdiffer in the phase term. This is not a major problem here, since at this stage weonly need existence of a solution. When dealing with variable coefficient equations,however, we will need a more careful choice of the solution.Lagrange interpolation of odd order As shown in Chapter 3, the general<strong>for</strong>m of the reference basis function <strong>for</strong> Lagrange interpolation of odd degree is⎧[r/2]+1∏ y − kif 0 ≤ y ≤ 1−kk≠0,k=−[r/2]⎪⎨ψ [r] (y) =.(6.31)r∏ y − kif [r/2] ≤ y ≤ [r/2] + 1−k⎪⎩ k=10 if y > [r/2] + 1


✐✐156 Chapter 6. High-order SL approximation schemes<strong>and</strong> extended by symmetry <strong>for</strong> y < 0. Since (6.31) is piecewise polynomial <strong>and</strong>compactly supported, its second derivative is the sum of a bounded compactlysupported term, plus a finite number of Dirac masses, <strong>and</strong> hence ψ ∈ W 2,1 (R). Thesymbolic computation of the Fourier trans<strong>for</strong>ms (see [Fe10], [Fe11]) shows that, <strong>for</strong>all odd orders r ≤ 13, they have the structureˆψ [r] (ω) = a 0 + a 2 ω 2 + · · · + a r−1 ω r−1 (sin ω ) r+1,(6.32)ω r+1 2where the polynomial contains only positive terms of even degree (a 2m > 0, a 2m+1 =0). All Fourier trans<strong>for</strong>ms of the <strong>for</strong>m (6.32) are there<strong>for</strong>e nonnegative. A naturalconjecture would be that this structure holds <strong>for</strong> any odd value of r, however ratherthan proving such a property, we simply note here that it holds <strong>for</strong> any order ofpractical interest.The basis functions ψ [r] (y) <strong>and</strong> the solutions φ [r] (y) = F −1 { ˆψ [r] (ω) 1/2 } <strong>for</strong> r = 3<strong>and</strong> r = 5 are shown in respectively the middle <strong>and</strong> the lower row of figure 6.1. Inthis case, direct <strong>and</strong> inverse trans<strong>for</strong>ms have been computed numerically by FFT.Finite element reconstructions In this situation, since finite element reconstructionsfail to be translation invariant (this point has already been remarked in Chapter3) the technique used <strong>for</strong> Lagrange reconstructions cannot be applied <strong>and</strong> we willrather per<strong>for</strong>m a Von Neumann analysis. Un<strong>for</strong>tunately, due to the lack of translationinvariance, B is no longer a circulating matrix <strong>and</strong> there<strong>for</strong>e such analysis onlygives a necessary condition.Assume that (6.2) is set on the interval [0, 1] with periodic conditions, thatthe interval is split into N subintervals, the interpolation degree being r on eachsubinterval. The total number of nodes is n n = Nr <strong>and</strong> ∆x = 1/(Nr) is the spacestep between nodes. We assume that the Courant number λ = c∆t/∆x is in (0, 1),whereas the case of large Courant numbers can be recovered as a byproduct of thisanalysis, as it has been shown <strong>for</strong> the CIR scheme.In the j–th subinterval, the reconstruction depends on the values of the numericalsolution at the nodes x (j−1)r , . . . , x jr . Accordingly, the matrix B has theblock circulating structure⎛⎞B 1 0 0 · · · 0 0 B 2B 2 B 1 0 · · · 0 0 0B =0 B 2 B 1 · · · 0 0 0⎜· · · · · · · · · · · · · · · · · · · · ·.⎟⎝ 0 0 0 · · · B 2 B 1 0 ⎠0 0 0 · · · 0 B 2 B 1with B 1 , B 2 ∈ R r×r , <strong>and</strong> B 2 of the <strong>for</strong>mB 2 = ( 0 | b ) ,b being a column vector of R r . More in detail, setting k = jr (that is, denoting byx k the right extremum of the j–th subinterval), we can write the j–th block of thescheme as⎧u ⎪⎨n+1k= l 0 (λ)u n k + · · · + l r(λ)u n k−r.(6.33)⎪⎩u n+1k−r+1 = l 0(λ + r − 1)u n k + · · · + l r(λ + r − 1)u n k−r


✐✐6.1. Semi-Lagrangian schemes <strong>for</strong> the advection equation 157linearcubicquinticFigure 6.1. Reference basis function <strong>for</strong> interpolation <strong>and</strong> equivalent LGreference basis functions obtained via (6.23), <strong>for</strong> linear (upper), cubic (middle) <strong>and</strong>quintic (lower) reconstructionwhere l 0 (·), . . . , l r (·) are the Lagrange basis functions expressed in a reference intervalwith unity step between nodes, <strong>and</strong>, by periodicity, u n 0 ≡ u n n nif j = 1. Notethat, <strong>for</strong>mally, the sign of the reference variable in the l i is reversed with respectto the physical variable because of the positive sign of λ, which corresponds to anegative shift along characteristics.


✐✐158 Chapter 6. High-order SL approximation schemesNow, assuming that ρ is an eigenvalue <strong>for</strong> the eigenvector v, we have from (6.33):⎧[ l0 (λ) − ρ ] v k + · · · + l r (λ)v k−r = 0⎪⎨.(6.34)l 0 (λ + r − 1)v k + · · · +⎪⎩+ [ l r−1 (λ + r − 1) − ρ ] v k−r+1 + l r (λ + r − 1)v k−r = 0.Since (6.34) has r equations <strong>and</strong> r + 1 unknowns, it is possible, <strong>for</strong> example viaGaussian elimination, to express v k−r as a function of v k in the <strong>for</strong>mv k−r = h(ρ, λ) v k (6.35)with h(·, ·) an iterated composition of rational functions. Then, repeating the procedure<strong>for</strong> N subintervals <strong>and</strong> imposing the periodicity condition we obtain:h(ρ, λ) N = 1, (6.36)that is, h(ρ, λ) must be a N–th root of the unity. As it has been shown in theclassical case of circulating matrices, we can pass to the modulus, obtaining theequation of the curve containing the eigenvalues in the implicit <strong>for</strong>m∣ h(ρ, λ)∣ ∣ = 1. (6.37)Note that this latter condition suits the case of N → ∞, in which the N-th rootsof the unity become dense on the unit circle of C. A different interpretation, whentreating an unbounded grid, is that (6.37) states that an eigenvector v should bebounded in l ∞ .Last, the Von Neumann condition is satisfied if:∣ h(ρ, λ)∣ ∣ = 1 implies that |ρ| ≤ 1. (6.38)We show in Fig. 6.2 the eigenvalues of the P 2 <strong>and</strong> P 3 schemes <strong>for</strong> differentvalues of λ, with respectively N = 25 (50 nodes) <strong>and</strong> N = 17 (51 nodes).The P 2 case We treat separately the P 2 case, <strong>for</strong> which the basis functionsin (6.33) have the <strong>for</strong>m:l 0 (y) = 1 (y − 1)(y − 2)2l 1 (y) = y(2 − y) (6.39)l 2 (y) = 1 y(y − 1)2<strong>and</strong> allow <strong>for</strong> an explicit computation. More precisely, we have the followingTheorem 6.5. Consider the problem (6.2). Assume that the scheme S is in theblock <strong>for</strong>m (6.33), with r = 2 <strong>and</strong> l 0 , l 1 , l 2 given by (6.39), that the space step ∆xis constant, <strong>and</strong> that 0 < λ < 1. Then, the scheme S satisfies condition (6.38).Proof. We refer to [P06] <strong>for</strong> the detailed computations, while sketching here onlythe main steps. Carrying out the computations outlined above, we obtain <strong>for</strong> (6.35)the expression:⎡⎤v k−2 = ⎣ ρ − l 0(λ) + l1(λ)l0(λ+1)ρ+l 1(λ+1)⎦ vl 2 (λ) − l1(λ)l2(λ+1) kρ+l 1(λ+1)


✐✐6.1. Semi-Lagrangian schemes <strong>for</strong> the advection equation 1590.80.60.40.20!0.2!0.4!0.6!0.8!1 !0.5 0 0.5 10.80.60.40.20!0.2!0.4!0.6!0.8!1 !0.5 0 0.5 1Figure 6.2. Eigenvalues of the P 2 (upper) <strong>and</strong> P 3 (lower) scheme <strong>for</strong>λ = 0.1, 0.2, . . . , 1.which, via some algebra, gives the condition:[ 2ρ 2 ] N+ (λ + 4)(λ − 1)ρ + (λ − 2)(λ − 1)= 1.λ(λ − 1)ρ + λ(λ + 1)Passing to the modulus, <strong>and</strong> writing ρ = z + iw ∈ C, we obtain the equationw 4 + [ 2z 2 + (λ 2 + 3λ − 4)z + (2λ 3 − λ 2 − 3λ + 2) ] w 2 ++ [ z 4 + (λ 2 + 3λ − 4)z 3 + (2λ 3 + λ 2 − 9λ + 6)z 2 ++(−5λ 2 + 9λ − 4)z − 2λ 3 + 3λ 2 − 3λ + 1 ] = 0.


✐✐160 Chapter 6. High-order SL approximation schemesBy working in the auxiliary variable w 2 we obtain, <strong>for</strong> z ∈ [1 − 2λ, 1], w 2 as thefollowing function of z <strong>and</strong> λ:withw 2 = −[ 2z 2 + (λ + 4)(λ − 1)z + (2λ 2 + λ − 2)(λ − 1) ] + √ D(z, λ)2D(z, λ) = λ 2 (λ − 1)(λ + 7)z 2 + 2λ 2 (λ − 1)(2λ 2 + 7λ − 7)z ++λ 2 (4λ 4 − 4λ 3 − 11λ 2 + 22λ − 7).Finally, <strong>for</strong> z ∈ [1 − 2λ, 1], it is easy to verify that w 2 < 1 − z 2 , which amounts toprove that eigenvalues are in the unit disc of C.Nonuni<strong>for</strong>m space grids In this situation, if H j = r∆x j is the measure ofthe j–th subinterval, we can define local Courant numbersλ j = c∆t∆x j.We set again k = jr with j = 1, . . . , N, so that the j–th block of the scheme readsnow ⎧⎪ u n+1⎨ k= l 0 (λ j )u n k + · · · + l r(λ j )u n k−r.(6.40)⎪ ⎩u n+1k−r+1 = l 0(λ j + r − 1)u n k + · · · + l r(λ j + r − 1)u n k−r .Theorem 6.6. Assume that, <strong>for</strong> any j, 0 < λ min ≤ λ j ≤ λ max < 1. Assumemoreover that the constant step scheme (6.33) satisfies condition (6.38) <strong>for</strong> anyλ ∈ [λ min , λ max ]. Then condition (6.38) is also satisfied by the variable step scheme(6.40).Proof. Following the same procedure as above, condition (6.36) is now replaced byN∏h(ρ, λ j ) = 1. (6.41)j=1Keeping the analogy with the uni<strong>for</strong>m mesh case, the condition which characterizesthe curve containing the eigenvalues may be rewritten asN∏|h(ρ, λ j )| = 1. (6.42)j=1Now, all the term in the product satisfy the boundsminj|h(ρ, λ j )| ≤ |h(ρ, λ j )| ≤ max |h(ρ, λ j )| (6.43)jso thatmin |h(ρ, λ j )| N ≤jN∏j=1|h(ρ, λ j )| ≤ max |h(ρ, λ j )| N (6.44)j


✐✐6.1. Semi-Lagrangian schemes <strong>for</strong> the advection equation 1610.80.60.40.20!0.2!0.4!0.6!0.8!1 !0.5 0 0.5 1Figure 6.3. Eigenvalues of the P 2 scheme <strong>for</strong> 0.014 ≤ λ j ≤ 0.77 <strong>and</strong> 50 nodes.<strong>and</strong> we can conclude that if condition (6.38) is satisfied with a uni<strong>for</strong>m grid <strong>for</strong> anyλ ∈ [λ min , λ max ], then it is also satisfied in the nonuni<strong>for</strong>m case.We show in Fig. 6.3 (in crosses) the eigenvalues of the P 2 scheme, obtainedwith 25 subintervals of r<strong>and</strong>om measure <strong>and</strong> 0.014 ≤ λ j ≤ 0.77, compared with (indots) the eigenvalues associated to the extreme values of the local Courant numbers.Remark 6.7. The stability analysis <strong>for</strong> the case of a nonuni<strong>for</strong>m grid also coversthe case of variable coefficient equations, although in the situation of low Courantnumbers. The Von Neumann condition |ρ| ≤ 1 should be regarded in general asa necessary condition, since we have no in<strong>for</strong>mations on the normality of the matrixB. However, there is a strong evidence that this matrix is quasi-normal, thismeaning that in its Jordan decomposition (4.40), the condition number K(T ) =‖T −1 ‖ · ‖T ‖, although not unity, still remains uni<strong>for</strong>mly bounded with respect to ∆.The general case <strong>for</strong> translation invariant reconstructionsWhen trying to extend to variable coefficient equations the stability analysis basedon the equivalence with LG schemes, it is clear that the relationship (6.21) canjustify replacing Galerkin projection with interpolation, but cannot account <strong>for</strong> thede<strong>for</strong>mation of the basis functions. On the other h<strong>and</strong>, if in (6.9) the approximateadvection X ∆ (ξ, t n+1 ; t n ) is replaced with a rigid translation ξ−x i +X ∆ (x i , t n+1 ; t n )along the approximate characteristic passing through x i , we obtain an approximate


✐✐162 Chapter 6. High-order SL approximation schemesLG scheme of the <strong>for</strong>m∑∫v n+1iv n i∫φ i(ξ − xi + X ∆ (x i , t n+1 ; t n ) ) φ j (ξ)dξ.φ i (ξ)φ j (ξ)dξ = ∑iR i R(6.45)With some abuse of notation, we will refer to a scheme in the <strong>for</strong>m (6.45) as anArea-weighted Lagrange–Galerkin (ALG) scheme. <strong>Schemes</strong> of this kind are used tocircumvent the problem of computing the integrals of de<strong>for</strong>med basis functions inexact LG schemes, <strong>and</strong> their stability analysis is relatively easy if the basis functionsφ k are piecewise polynomial <strong>and</strong> compactly supported, as usual in the finite elementsetting. Here, on the contrary, the φ k are obtained through a reference function φsolving (6.21), which is characterized in a somewhat weaker way by means of itsFourier trans<strong>for</strong>m.Thus, while the equivalence between (6.45) <strong>and</strong> (6.3) can be proved by thesame ideas of the constant coefficient case, an ad hoc proof is necessary to showthat (6.45) represents a “small” perturbation of an exact LG scheme, that is∥ B SL (∆; t n ) ∥ 2= ∥ B ALG (∆; t n ) ∥ 2≤ ∥ B LG (∆; t n ) ∥ 2+ C∆t. (6.46)In turn, this will be proved in the <strong>for</strong>m∥ B ALG − B LG∥ ∥2≤ (∥ ∥ B ALG − B LG∥ ∥1 · ∥∥ B ALG − B LG∥ ∥∞) 1/2≤ C∆t, (6.47)which is satisfied if both the 1-norm <strong>and</strong> the ∞-norm of B ALG − B LG are boundedby an O(∆t).In order to suit the main situation of interest, i.e., Lagrange interpolation ofodd degree, the LG reference basis function will be assumed to be piecewise smooth,with at most a countable set of isolated discontinuities y k . More precisely, we willassume the following:• The function φ(y) satisfies the decay condition• Its derivative is in the <strong>for</strong>m|φ(y)| ≤φ ′ (y) = φ ′ s(y) +C φ1 + |y| 3 (6.48)∞∑k=−∞w k δ(y − y k ), (6.49)where the regular <strong>and</strong> the singular part satisfy the further bounds<strong>and</strong> the singularities y k have the expression|φ ′ s(y)| ≤ C s1 + y 2 , (6.50)|w k | ≤ C w1 + k 2 , (6.51)y k = αk + β. (6.52)


✐✐6.1. Semi-Lagrangian schemes <strong>for</strong> the advection equation 163• The vectorfield f(·, t) is uni<strong>for</strong>mly Lipschitz continuous, that is<strong>for</strong> any t ∈ R.|f(x 1 , t) − f(x 2 , t)| ≤ L f |x 1 − x 2 | (6.53)• The approximation X ∆ (·) is consistent, that isz i = X ∆ (x i , t n+1 ; t n ) = x i − ∆t f(x i , t n+1 ) + O(∆t 2 ). (6.54)We give a couple of preliminary lemmas, which will be useful in order toestimate the perturbation appearing in (6.47).Lemma 6.8. For any a ∈ R + , b ∈ R,∫dyπ(1 + a)(1 + y 2 )(1 + (ay + b) 2 =) (1 + a) 2 + b 2 ; (6.55)R∞∑k=−∞<strong>for</strong> some positive constant C.1Cπ(1 + a)(1 + k 2 )(1 + (ak + b) 2 ≤) (1 + a) 2 + b 2 . (6.56)Proof. The integral (6.55) can be computed explicitly (the result has been obtainedby symbolic integration with Mathematica, see [Fe11]). The inequality (6.56) canbe obtained by observing that all terms of the series are bounded, <strong>and</strong> there<strong>for</strong>e itssum can be given an upper bound, up to a mutiplicative constant, by comparisonwith the corresponding integral (6.55).Lemma 6.9. Let assumptions (6.53), (6.54) hold. Then, <strong>for</strong> any c ∈ R + , d ∈ R,both series∞∑ 1c + (j − zi∆x + (6.57)d)2j=−∞∞∑i=−∞1c + (j − zi∆x + (6.58)d)2are uni<strong>for</strong>mly bounded as functions of respectively the index i <strong>and</strong> the index j.Proof. Assume first that the smallest value attained by the term (j − zi∆x + d)2 isexactly zero.Consider the series (6.57). Let i (<strong>and</strong> there<strong>for</strong>e z i ) be fixed, <strong>and</strong> denote by j 0the index <strong>for</strong> which (j 0 − zi∆x + d)2 = 0 (that is, j 0 = zi∆x− d). Accordingly, we have∞∑j=−∞1c + (j − zi∆x + = d)2=∞∑j=−∞∞∑j=−∞1c + (j − j 0 ) 2 =1c + j 2


✐✐164 Chapter 6. High-order SL approximation schemeswhere the last <strong>for</strong>m is obtained by an obvious shift in the summation index, <strong>and</strong> isindependent of the index i.Consider now (6.58), <strong>and</strong> let j be fixed. Denote by i 0 the index <strong>for</strong> which(j − zi 0∆x + d)2 = 0 (that is, zi 0∆x= j + d). By the consistency assumption (6.54), wehavez i = x i − ∆tf(x i ) + O(∆t 2 ) (6.59)z i0 = x i0 − ∆tf(x i0 ) + O(∆t 2 ) (6.60)so that using the Lipschitz continuity of f <strong>and</strong> the triangular inequality (in the <strong>for</strong>mof a difference), we obtain<strong>and</strong>, <strong>for</strong> ∆t small enough,|z i − z i0 | = ∣ ∣x i − x i0 − ∆t(f(x i ) − f(x i0 )) + O(∆t 2 ) ∣ ∣ ≥i=−∞≥ |x i − x i0 | − ∆t|f(x i ) − f(x i0 )|(1 + O(∆t)) ≥≥ (1 − ∆tL f )|x i − x i0 |(1 + O(∆t)) (6.61)|z i − z i0 | ≥ 1 2 |x i − x i0 |. (6.62)Turning back to (6.58), we have, using (6.62), <strong>and</strong> again a shift in the summationindex:∞∑ 1c + (j − zi∆x + =∑∞ 1d)2 c + ( zi 0∆x − ≤ zi∆x )2≤==i=−∞∞∑i=−∞∞∑i=−∞∞∑i=−∞1c + 1 4 ( xi 0∆x − = xi∆x )21c + 1 4 (i 0 − i) 2 =1c + 1 4 i2 (6.63)which is independent of j.Lastly, if the smallest value <strong>for</strong> the term (j − zi∆x + d)2 is nonzero, then thesame arguments may be applied, unless <strong>for</strong> a fractional perturbation of the constantd. We note here that the sum of both series is continuous with respect to such aperturbation, so that it is possible to pass to the supremum, obtaining a uni<strong>for</strong>mbound in this case as well. The technical details are left to the reader.We present now the main result. It states the stability of the Area-weightedLG scheme under assumptions which generalize the parallel result in [MPS88]. Thiswill allow to apply the theory to LG basis functions obtained by solving (6.21).Theorem 6.10. Let the basic assumptions (6.17) <strong>and</strong> (6.48)–(6.54) hold, <strong>and</strong>assume x j = j∆x. Then, there exists a positive constant C independent of n, ∆x<strong>and</strong> ∆t such that, at a generic step n ∈ N <strong>and</strong> <strong>for</strong> ∆t small enough,∥ B ALG (∆; t n ) − B LG (∆; t n ) ∥ 2≤ C∆t.Proof. The proof uses the bound (6.47), <strong>and</strong> is split in several steps.


✐✐6.1. Semi-Lagrangian schemes <strong>for</strong> the advection equation 165Step 1. We first need to estimate the difference |X ∆ (ξ)−(ξ−x i +X ∆ (x i , t n+1 ; t n ))|.Denoting <strong>for</strong> shortness X ∆ (ξ, t n+1 ; t n ) as X ∆ (ξ) <strong>and</strong> again X ∆ (x i , t n+1 ; t n ) as z i ,by (6.53) <strong>and</strong> (6.54) we obtain:|X ∆ (ξ) − (ξ − x i + z i )| ≤ C X |ξ − x i |∆t. (6.64)At a later time, we will use again the change of variables η = ξ/∆x − i. In this newvariable, (6.64) is rewritten asStep 2.∣ bALGij|X ∆ ((η + i)∆x) − (η∆x + z i )| ≤ C X |η|∆x∆t. (6.65)We proceed to estimate |b ALGij (∆; t n )−b LGij (∆; t n)|. By definition, we have∣∫∫∣∣∣ (∣ij = φ j (ξ − x i + z i ))φ i (ξ)dξ − φ j X ∆ (ξ) ) φ i (ξ)dξ∣ ≤− b LGR≤ 1∆x∫R( ξ −∣ φ xi + z i∆xR)− j − φ( )∣ X ∆ ( )∣(ξ) ∣∣∣∆x − j ·ξ ∣∣∣∣ φ ∆x − i dξ.Now, introducing the variable η = ξ/∆x − i, we obtain∫( ) ( ∣ bALGij − b LG ∣ij ≤ η∆x +∣ φ ziX ∆ ((η + i)∆x) ∣∣∣− j − φ− j)∣· |φ(η)| dη ≤R ∆x∆x∫∫ X ∆ ((η+i)∆x)∆x −j≤φ ′ (y)dyη∆x+zR ∣i· |φ(η)| dη ≤∆x −j∣∫≤ |φ ′ (y)| · |φ(η)| dydηDwhere D is defined by<strong>and</strong> y(η), y(η) byD = { (η, y) : η ∈ R, y ∈ [ y(η), y(η) ]}( η∆x + ziy(η) = min∆x), X∆ ((η + i)∆x)− j∆x( η∆x + ziy(η) = max, X∆ ((η + i)∆x)∆x ∆x)− j.Note that the integration domain is modified so as to have a y-projection of positivemeasure <strong>for</strong> any η.Since the integr<strong>and</strong> is positive, we can give a further bound on this integral byenclosing the integration domain in a larger domain on which the integr<strong>and</strong> could bebounded in a more straight<strong>for</strong>ward way. By (6.65), the domain D can be includedin a set of the <strong>for</strong>mE ={(η, y) : η ∈ R, y ∈[η − C X ∆t|η| + z i∆x − j, η + C X∆t|η| + z i∆x − j ]}Figure 6.4 shows the inclusion of the integration domain D in the set E definedabove. In the figure, the liney = η∆x + z i∆x− j


✐✐166 Chapter 6. High-order SL approximation schemesyDE1C X t z i x − j1−C X t z i x − jz i x − jFigure 6.4. Inclusion of the integration domain in the set E.<strong>and</strong> the curvey = X∆ ((η + i)∆x)∆x− jof the η−y plane, enclose the dark-shaded domain D, which in turn appears includedin the light-shaded set E.We have there<strong>for</strong>e:∣ bALGij∫− b LG ∣ ≤ij∫≤∫=D|φ ′ (y)| · |φ(η)| dydη ≤∫ η+CX ∆t|η|+ z i∆x −jR η−C X ∆t|η|+ z i∫ η+CX ∆t|η|+ z i∆x −jR∫+η−C X ∆t|η|+ z i∫ η+CX ∆t|η|+ z i∆x −jRη−C X ∆t|η|+ z i∆x −j |φ ′ (y)| · |φ(η)| dydη =∆x −j |φ ′ s(y)| · |φ(η)| dydη +k=−∞∆x −j [ ∞∑|w k |δ(y − y k )]· |φ(η)| dydη= A ij + B ij . (6.66)


✐✐6.1. Semi-Lagrangian schemes <strong>for</strong> the advection equation 167Step 3. Estimation of A ij . Using the decay in<strong>for</strong>mations on φ <strong>and</strong> φ ′ s, we canestimate A ij as:∫A ij =R≤ C s C φ∫∫ η+CX ∆t|η|+ z i∆x −jη−C X ∆t|η|+ z i∆x∫ −jη+CX ∆t|η|+ z i∆x −j 1R η−C X ∆t|η|+ z i∆x −j∫≤ 2C s C φ C X ∆tR|φ ′ s(y)|dy |φ(η)| dη ≤1 + y 2 dy 11 + |η| 3 dη ≤11 + ( (1 − C X ∆t)η + zi∆x − j) 2|η|dη, (6.67)1 + |η|3where the inner integral has been estimated by multiplying the sup of the functionto be integrated by the measure of the interval. Now, it is easy to see that we canbound the second term of the integr<strong>and</strong> in (6.67) as|η|1 + |η| 3 ≤ 21 + η 2 ,so that, applying lemma 6.8 <strong>and</strong> collecting all multiplicative constants in a singleconstant C A , we obtain∫1A ij ≤ 4C s C φ C X ∆tR 1 + ( 1(1 − C X ∆t)η + zi∆x − j) 21 + η 2 dη =π(2 − C X ∆t)= 4C s C φ C X ∆t(2 − C X ∆t) 2 + ( z i∆x − j) 2 ≤≤C A ∆t1 + ( z i∆x − j) 2(6.68)where the last inequality has been obtained assuming also that ∆t is small enough.Step 4.Estimation of B ij . The integral B ij may be splitted as∫ ∫ [η+CX ∆t|η|+ z i∆x −j ∞]∑B ij =|w k |δ(y − y k ) · |φ(η)| dydη =R η−C X ∆t|η|+ z i∆x −j k=−∞∞∑∫ ∫ η+CX ∆t|η|+ z i∆x −j= |w k |δ(y − y k ) · |φ(η)| dydη. (6.69)R η−C X ∆t|η|+ z i∆x −jk=−∞Note that the measure to be integrated has its support on the segment of intersectionbetween the line y = y k <strong>and</strong> the set E. Denoting by η − kthe abscissa of theintersection of the line y = y k with the boundary y = (1 + C X ∆t)η + zi∆x − j (thisrepresents the extremum of smallest magnitude <strong>for</strong> this segment), <strong>and</strong> vice versaη + k<strong>for</strong> the boundary y = (1 − C X∆t)η + zi∆x− j (extremum of largest magnitude),we have (see Figure 6.5)|η + k − η− k | = 2C X∆t1 − C 2 X ∆t2 ∣ ∣∣yk− z i∆x + j ∣ ∣∣ == 2C X∆t1 − C 2 X ∆t2 ∣ ∣∣αk + β −z i∆x + j ∣ ∣∣


✐✐168 Chapter 6. High-order SL approximation schemesFigure 6.5. Integration of the singular part of φ ′ on the set E.<strong>and</strong>|η − k | = 1∣∣y k − z ∣i ∣∣1 + C X ∆t ∆x + j ==11 + C X ∆tso that <strong>for</strong> ∆t small enough we also have∣∣αk + β − z i∆x + j ∣ ∣∣|η + k − η− k | ≤ 4C ∣X∆t ∣αk + β − z ∣i ∣∣∆x + j ,|η − k | ≥ 1 2∣∣αk + β − z i∆x + j ∣ ∣∣ .Now, using such bounds <strong>and</strong> again the decay estimate <strong>for</strong> φ as in step 3, we obtain∫R∫ η+CX ∆t|η|+ z i∆x −jη−C X ∆t|η|+ z i∆x −j∣ αk + β −z i∆xδ(y − y k ) · |φ(η)| dydη ≤ 2C X ∆t+ j∣ ∣1 + ∣ αk + β −z i∆x + j∣ ∣ 3 ≤≤ 4C X ∆t11 + ( αk + β − zi∆x + j) 2


✐✐6.1. Semi-Lagrangian schemes <strong>for</strong> the advection equation 169<strong>and</strong> there<strong>for</strong>e∞∑1B ij ≤ 4C X ∆t |w k |k=−∞ 1 + ( αk + β − zi∆x + j) 2 ≤∞∑ C w1≤ 4C X ∆t1 + k 2 k=−∞ 1 + ( αk + β − zi∆x + j) 2 ≤C B ∆t≤(1 + α) 2 + ( β − zi∆x + j) 2(6.70)where we have applied againg Lemma 6.8 <strong>and</strong> collected all multiplicative constantinto C B .Conclusions. To sum up, recalling (6.66) <strong>and</strong> applying Lemma 6.9 to the estimates(6.68) <strong>and</strong> (6.70), we obtain that∥∥B ALG (∆; t n ) − B LG (∆; t n ) ∥ ∑∞≤ sup (A ij + B ij ) ≤ C∆t,i∥ B ALG (∆; t n ) − B LG (∆; t n ) ∥ ∑1≤ sup (A ij + B ij ) ≤ C∆t,jso that the bound (6.47) holds.ijRemark 6.11. It is immediate to see that, according to this result, the class ofbasis functions <strong>for</strong> which the Area-weighted LG scheme is stable includes all bounded,piecewise Lipschitz continuous functions with bounded support. In this respect, thisresult includes <strong>and</strong> generalizes the corresponding theorem in [MPS88].Application to Lagrange interpolation We turn back to the case of Lagrangeinterpolation of odd order. As we said, the solution of (6.21) is not explicitly known,but rather characterized in terms of its Fourier trans<strong>for</strong>m. So what is needed is torelate properties (6.48)–(6.52) to properties of the Fourier trans<strong>for</strong>m ˆφ(ω).• Concerning the rate of decay of φ(y), it is known that φ decays like |y| −kprovided ˆφ (k) ∈ L 1 (R). There<strong>for</strong>e, assumption (6.48) is satisfied if ˆφ ′′′ ∈L 1 (R).• Checking assumption (6.49) requires to single out a periodic component iniω ˆφ(ω) (that is, in the trans<strong>for</strong>m F[φ ′ ]). This component generates the singularpart (sum of evenly spaced Dirac distributions) in the inverse Fouriertrans<strong>for</strong>m. Once defined iω ˆφ s as the difference between iω ˆφ <strong>and</strong> its periodiccomponent, the decay assumption (6.50) requires its second derivative withrespect to ω to be in L 1 (R).• Assumption (6.51) is related to the decay of Fourier coefficients <strong>for</strong> periodicfunctions. It turns out that this assumption is satisfied provided the periodiccomponent of iω ˆφ(ω) has locally L 1 second derivative with respect to ω.


✐✐170 Chapter 6. High-order SL approximation schemesExample: P 1 interpolation We re-read this example, already solved in theconstant coefficient case, by discussing the use of the two different solutions of theintegral equation (6.21). First, using (6.23), we get∣ sinω ∣ˆφ [1] 2(ω) =∣ ω ∣(6.71)<strong>and</strong> the solution could be defined accordingly as the inverse trans<strong>for</strong>m of (6.71) (seethe upper line of figure 6.1). Although this solution would be fine <strong>for</strong> the constantcoefficient case, it does not satisfy the required decay assumptions. For example, thetrans<strong>for</strong>m (6.71) is not three times differentiable, even in the sense of distributions,so it is not expected to satisfy (6.48).On the other h<strong>and</strong>, using the other possibility, that is,<strong>and</strong> taking the inverse trans<strong>for</strong>m as2ˆφ [1] (ω) = sin ω 2ω2{ sinω}φ [1] (y) = F −1 2ω=2, (6.72){1 if − 1/2 ≤ y ≤ 1/20 elsewhere.we obtain a solution which satisfies all the basic assumptions, as it is easy to see.Note that, even in lack of an explicit inverse trans<strong>for</strong>m, this could be extractedfrom the properties of the Fourier trans<strong>for</strong>m (6.72). First, (6.72) defines an analyticfunction, so the inverse trans<strong>for</strong>m φ [1] (y) decays faster than any negative power ofy <strong>and</strong> (6.48) is satisfied (actually, it has bounded support). Second, the trans<strong>for</strong>mof φ [1]′ readsF{φ [1]′} = 2i sin ω 2 .Here, we only have a periodic component <strong>and</strong> φ [1]′s ≡ 0, so (6.50) is satisfied. In turn,the periodic component is also analytic, so its Fourier coefficients have a fast decay<strong>and</strong> (6.51) holds (in fact, the only nonzero coefficient is related to the “harmonic”sin(ω/2), <strong>and</strong> φ [1]′ (y) = δ(y + 1/2) − δ(y − 1/2)).Lagrange interpolation of higher (odd) order Although, <strong>for</strong> r ≥ 3, theproblem of differentiability of ˆφ [r] outlined <strong>for</strong> r = 1 does not appear any longer,defining a solution through (6.23) would not allow to single out a periodic componentin F[φ ′ ], <strong>and</strong> as a result the solution in the y-domain would have a slower decay. Away to circumvent this problem is to define a solution in the ω-domain as√ˆφ [r] a0 + a 2 ω(ω) =2 + · · · + a r−1 ω r−1· sin ω ∣∣sin ω r−12∣ =2 2ω|ω| r−12= A r (ω) · B r (ω). (6.73)In (6.73), A r <strong>and</strong> B r denote respectively the algebraic <strong>and</strong> the trigonometric termof ˆφ [r] . Note that in (6.73) all derivatives are continuous at ω = 0, whereas due tothe structure of B r (ω) it occurs that derivatives are continuous only up to a certainorder (depending on r) <strong>for</strong> ω coinciding with multiples of 2π.


✐✐6.1. Semi-Lagrangian schemes <strong>for</strong> the advection equation 171FTo check the basic assumptions, we also need to compute F{√φ [r]′} (ω) = iω ˆφ [r] a0 + a 2 ω(ω) = i2 + · · · + a r−1 ω r−1· sin ω|ω| r−122{φ [r]′} :∣∣sin ω 2(6.74)which contains again an algebraic <strong>and</strong> a periodic term. The first one has a finiteimaginary limit i √ a r−1 <strong>for</strong> ω → ±∞, so that we can rewrite (6.74) as the sum of aterm vanishing <strong>for</strong> ω → ±∞, plus a periodic term giving the asymptotic behaviour,that isF{φ [r]′} (ω) = i(√ )a0 + a 2 ω 2 + · · · + a r−1 ω r−1− √ a r−1 sin ω ∣∣sin ω ∣2 2|ω| r−12+i √ a r−1 sin ω ∣∣sin ω ∣2 2= iC r (ω) + iD r (ω).r−12As it has been explained at the start of the section, in order to satisfy the basicassumptions we have to check that(i) The trans<strong>for</strong>m (6.73) has its third derivative (w.r.t. ω) in L 1 (R);(ii) The term C r (ω) has its second derivative in L 1 (R);(iii) The term D r (ω) has a locally L 1 second derivative.Let us start with point (i). We have, from elementary differentiation rules:=d 3dω 3 ˆφ [r] (ω) = A ′′′r B r + 3A ′′r B ′ r + 3A ′ rB ′′r + A r B ′′′r . (6.75)Now, the terms B r , B r ′ <strong>and</strong> B r ′′ are always bounded. The term B r′′′ is bounded <strong>for</strong>all r > 3, whereas, <strong>for</strong> r = 3, it has the <strong>for</strong>m of a sequence of Dirac distributions of= O(1/ω 2 ), so thatthe right-h<strong>and</strong> side of (6.75) is in L 1 (R) <strong>and</strong> point (i) is satisfied <strong>for</strong> any r ≥ 3.Concerning point (ii), using the same arguments of point (i), it suffices that thealgebraic term in brackets, along with its first <strong>and</strong> second derivative, be O(1/ω 2 ).This is in fact the case, so the requirement of point 2 is also satisfied. Lastly, pointconstant weight. On the other h<strong>and</strong>, we have that A r , . . . , A ′′′r(iii) is also trivially satisfied since D r ′′ is bounded on R. All symbolic computationsrequired <strong>for</strong> this subsection, along with the Fourier trans<strong>for</strong>ms, have been carriedout with Mathematica <strong>for</strong> r ≤ 13 (a sample of the related code is listed in [Fe11]).We show in Figure 6.6 the reference LG basis functions obtained by solvingnumerically (6.21) by Fast Fourier Trans<strong>for</strong>m, both in the <strong>for</strong>m (6.23) (left column)<strong>and</strong> in the <strong>for</strong>m (6.73) (right column), <strong>for</strong> the linear, cubic <strong>and</strong> quintic Lagrangeinterpolation. Note that, since D r has a period of 4π in the ω-domain, the discontinuitiesin φ [r] (that is, Dirac distributions in φ [r]′ ) appear in the y-domain atmultiples of 1/2, so that (6.52) holds with α = 1/2, β = 0 (in fact, even harmonicsare missing in D r , so w k = 0 <strong>for</strong> k even). If r−12is even (e.g., in the linear <strong>and</strong>quintic case), thensin ω ∣∣sin ω r−1 (2∣ = sin ω ) r+12,2 22so that D r (ω) contains a finite number of harmonics, <strong>and</strong> as a result the LG referencebasis function has a finite number of discontinuities.∣r−12r−12,+


✐✐172 Chapter 6. High-order SL approximation schemeslinearcubicquinticFigure 6.6. Equivalent LG reference basis functions obtained via (6.23)(left) <strong>and</strong> (6.73) (right), <strong>for</strong> linear (upper), cubic (middle) <strong>and</strong> quintic (lower) reconstruction6.1.4 ConvergenceWe sum up the convergence analysis in a theorem. Note that the technical assumptions(6.53)–(6.54) are already included as assumptions on smoothness of f <strong>and</strong>consistency rate of time discretization.Theorem 6.12. Let f, g ∈ W p,∞ , u be the solution of (6.1) <strong>and</strong> v n j be definedby (6.3), with an interpolation basis ψ i satisfying the assumptions of Theorem 6.1.


✐✐6.1. Semi-Lagrangian schemes <strong>for</strong> the advection equation 173Assume moreover that (5.55), (5.56), (6.4) hold, <strong>and</strong> that the function φ, solutionof (6.21), satisfies (6.48)–(6.52). Then, <strong>for</strong> any j ∈ Z <strong>and</strong> n ∈ [1, T/∆t],∣ vnj − u(x j , t n ) ∣ ∣ → 0 (6.76)as ∆t → 0, ∆x = o ( ∆t 1/r) .Moreover, if u ∈ L ∞ ([0, T ], W s,∞ (R)), then)‖V n − U(t n )‖ 2≤ C(∆t p + ∆xmin(s,r+1). (6.77)∆tWe also discuss here modified equation <strong>and</strong> dispersive behaviour of the SLscheme, along with the interactions between discretization steps at the level oftuning of the scheme.Numerical dispersionRepeating the analysis of chapter 5, we consider again the equation in the <strong>for</strong>m (6.2),a scheme in the <strong>for</strong>m (6.10), <strong>and</strong> assume v to be a regular extension of the numericalsolution of the scheme. If we replace the error estimate <strong>for</strong> P 1 interpolation with thecomplete estimate <strong>for</strong> Lagrange interpolation of a generic order r, we can expressthe value I r [V n ](x j − c∆t) asI r [V n ](x j − c∆t) = v(x j − c∆t, t n ) −∏(x j − c∆t − x k )x k ∈S(r + 1)!∂ r+1∂x r+1 v(ξ j),where S is the stencil of points involved in the reconstruction, <strong>and</strong> ξ j is an unknownpoint located between the extreme points of of S. Next, retracing the computationsof the first-order case, we obtain from(6.10):[ ]∂r+1v(x j , t n+1 ) = v(x j − c∆t, t n ) + ν(∆)∂x r+1 v(x j − c∆t) + O(∆x)which is in the <strong>for</strong>m (4.85), with k = r + 1, once set the dispersivity coefficient ν asν(∆) = −∏(x j − c∆t − x k )x k ∈S(r + 1)! ∆t(note that sgn(ν(∆)) = (−1) r−12 , <strong>and</strong> that all terms of the product are O(∆x)). Weobtain there<strong>for</strong>e a modified equation in the <strong>for</strong>m (4.84), <strong>and</strong> more preciselyv t + cv x = (−1) r−12 cr∆x r+1∆t∂ r+1 ( ) ∆xr+1∂x r+1 v + o , (6.78)∆twith c r a bounded nonnegative function of λ ′ . Then, the dispersivity of the schemecan be analysed by applying st<strong>and</strong>ard self-similarity arguments to the modifiedequation. This procedure shows that the resolution of discontinuities is of orderO(∆x/∆t 1/(r+1) ).


✐✐174 Chapter 6. High-order SL approximation schemesInterplay between discretization stepsAs it has been noticed treating the CIR scheme, the convergence estimate (6.77)shows an interaction between discretization steps, which requires to relate ∆x <strong>and</strong>∆t in a less trivial <strong>for</strong>m with respect to usual difference schemes. Once expressed∆t as a function of ∆x in the <strong>for</strong>m ∆t = ∆x α , we briefly discuss the optimizationof the convergence rate in (6.77), as well as the resolution of discontinuities, withrespect to the parameter α.Maximization of the convergence ratesteps in (6.77), we obtainUsing the relationship between the two(‖V n − U(t n )‖ 2≤ C ∆x αp + ∆x min(s,r+1)−α) .Since the first exponent increases with α, whereas the second decreases, the convergencerate is maximized when they coincide, that is, <strong>for</strong>α =min(s, r + 1). (6.79)p + 1Note that the ∆t/∆x relationship depends on the regularity.smooth (more precisely, if s > r + 1), then we obtainIf the solution isα = r + 1p + 1 ,a choice which clearly maximizes the consistency rate. Note that, in principle, ifr > p (e.g. coupling a second-order time integration with a cubic reconstruction),the relationship optimizing the consistency rate entails a vanishing Courant number.As the regularity decreases, α decreases, which means that less regular solutions arebetter approximated with larger time steps.Resolution of singularities In analysing the behaviour of the scheme on discontinuoussolutions, still two opposite effects appear: on one h<strong>and</strong>, large Courantnumbers cause a lesser number of reconstructions to be per<strong>for</strong>med (this causing asmaller numerical dispersion), on the other h<strong>and</strong>, they cause larger errors in computingcharacteristics (this causing a less precise location of singularities).Assuming as be<strong>for</strong>e that ∆t = ∆x α , the analysis of the modified equationshows that the dispersion of discontinuities is of the order of ∆x 1−α/(r+1) . Hence,taking into account that characteristics are computed with an error of ∆x αp , it turnsout that the resolution of the scheme is maximized when the two orders coincide,that is, with the choicer + 1α =p(r + 1) + 1 . (6.80)Note that this relationship implies α < 1/p, so that ∆x = o(∆t p ), <strong>and</strong> in practiceα ≈ 1/p (<strong>for</strong> example, <strong>for</strong> a second-order time <strong>and</strong> space discretization p = r = 2,<strong>and</strong> we would obtain α = 3/7). In general, this choice could lead to relatively largetime discretization errors. However, it confirms the indication that high Courantnumbers are more suitable to treat nonsmooth solutions.


✐✐6.1. Semi-Lagrangian schemes <strong>for</strong> the advection equation 1756.1.5 ExamplesIn order to compare the per<strong>for</strong>mances of the various schemes, we consider again thesimple one-dimensional problem (5.66), which is rewritten here:{u t (x, t) − (x − ¯x)u x (x, t) = 0 (x, t) ∈ (0, 1) × (0, 1)u(x, 0) = u 0 (x).(6.81)As in Chapter 5, we have set ¯x = 1.1 <strong>and</strong> used three initial conditions u 0 of differentregularity with bounded support. The first is the characteristic function (5.67), thesecond is a W 1,∞ function in the <strong>for</strong>m:u 0 (x) = max(sin(2πx), 0), (6.82)<strong>and</strong> the third is in W 2,∞ :u 0 (x) = max(sin(2πx), 0) 2 . (6.83)Note that, in addition to what has been remarked in Chapter 5, the initial conditionshave been changed to have a non-polynomial structure, so as to evaluate theefficiency of the reconstruction procedure.The test has been carried out combining a second-order Runge–Kutta schemeto follow characteristics with P 2 finite element, cubic <strong>and</strong> 3rd-order ENO reconstructions.∆t/∆x relationships in the <strong>for</strong>m ∆t = ∆x α have been considered <strong>for</strong>three different values of α: the first, α = 1/2, is close to the value optimizing theresolution of singularities, the second is the classical linear relationship α = 1 (moreprecisely, ∆t = 5∆x following the choice made <strong>for</strong> the monotone case) <strong>and</strong> the thirdis the value of α optimizing the consistency rate (in the P 2 case, this criterion givesagain α = 1). The ENO scheme has been tested on the most efficient relationshipused in the cubic case.We list in Table 6.1 numerical errors <strong>and</strong> overall convergence rates in the 2-norm <strong>for</strong> the various schemes. Figure 6.7 compares the schemes on the advectionof the characteristic function (5.67). Here, the first three figures are obtained withα = 1/2 which (nearly) optimizes the resolution of discontinuities, whereas thefourth is obtained with α = 4/3 which optimizes the consistency rate <strong>for</strong> the cubicreconstruction.As a first general remark, note that the error tables confirm the general indicationthat less regular solutions are better approximated with large Courantnumbers, i.e., with lower values of α. In cubic approximation, the relationshipmaximizing the consistency rate proves to be relatively inefficient, partly because itwould require a smoother solution (in W 4,∞ ) to achieve its best per<strong>for</strong>mance. Thisis also apparent from Figure 6.7, which shows the remarkable increase in numericaldispersion caused by the choice α = 4/3. Note also that the difference in convergencerates should be considered against complexity. Under the relationships shownin the tables, a test with 100 nodes requires 10 time steps with α = 1/2, 20 withα = 1 <strong>and</strong> 464 with α = 4/3. In terms of absolute error, the best per<strong>for</strong>mancesare obtained by the P 2 scheme – but we point out once more that the cubic schemewould require a smoother solution to increase its convergence rate.


✐✐176 Chapter 6. High-order SL approximation schemesL ∞ solutionP 2 cubic ENO3n n α = 1/2 α = 1 α = 1/2 α = 1 α = 4/3 α = 1/225 1.44 · 10 −1 1.44 · 10 −1 1.43 · 10 −1 1.43 · 10 −1 2.35 · 10 −1 2.1 · 10 −150 6.82 · 10 −2 5.96 · 10 −2 1.01 · 10 −1 1.04 · 10 −1 1.29 · 10 −1 1.06 · 10 −1100 5.38 · 10 −2 5.52 · 10 −2 6.6 · 10 −2 7.74 · 10 −2 9.39 · 10 −2 8.22 · 10 −2200 4.09 · 10 −2 4.09 · 10 −2 5.36 · 10 −2 6.06 · 10 −2 7.65 · 10 −2 5.94 · 10 −2400 2.57 · 10 −2 3.56 · 10 −2 3.72 · 10 −2 4.63 · 10 −2 5.83 · 10 −2 4.39 · 10 −2rate 0.62 0.50 0.49 0.41 0.50 0.56W 1,∞ solutionP 2 cubic ENO3n n α = 1/2 α = 1 α = 1/2 α = 1 α = 4/3 α = 1/225 2.04 · 10 −2 2.04 · 10 −2 8.37 · 10 −2 8.37 · 10 −2 1.30 · 10 −1 8.94 · 10 −250 9.92 · 10 −3 9.62 · 10 −3 2.03 · 10 −2 2.44 · 10 −2 4.99 · 10 −2 2.58 · 10 −2100 3.44 · 10 −3 3.15 · 10 −3 7.21 · 10 −3 9.11 · 10 −3 1.83 · 10 −2 1.01 · 10 −2200 1.82 · 10 −3 1.98 · 10 −3 3.26 · 10 −3 4.39 · 10 −3 8.18 · 10 −3 4.47 · 10 −3400 7.96 · 10 −4 7.92 · 10 −4 1.29 · 10 −3 1.87 · 10 −3 3.59 · 10 −3 1.86 · 10 −3rate 1.17 1.17 1.50 1.37 1.29 1.40W 2,∞ solutionP 2 cubic ENO3n n α = 1/2 α = 1 α = 1/2 α = 1 α = 4/3 α = 125 1.80 · 10 −2 1.80 · 10 −2 1.06 · 10 −1 1.06 · 10 −1 1.40 · 10 −1 1.06 · 10 −150 5.76 · 10 −3 3.70 · 10 −3 1.96 · 10 −2 2.01 · 10 −2 6.47 · 10 −2 2.56 · 10 −2100 2.72 · 10 −3 1.19 · 10 −3 3.84 · 10 −3 4.32 · 10 −3 1.65 · 10 −2 5.99 · 10 −3200 1.35 · 10 −3 2.96 · 10 −4 1.41 · 10 −3 9.83 · 10 −4 3.38 · 10 −3 1.64 · 10 −3400 6.63 · 10 −4 8.42 · 10 −5 6.67 · 10 −4 2.49 · 10 −4 7.65 · 10 −4 4.73 · 10 −4rate 1.19 1.93 1.83 2.18 1.88 1.95Table 6.1. Errors in the 2-norm, P 2 , cubic <strong>and</strong> 3rd-order ENO schemes,second-order time discretization.6.2 Semi-Lagrangian schemes <strong>for</strong> the convex HJequationIn order to illustrate the convergence theory of high-order Semi-Lagrangian schemes<strong>for</strong> <strong>Hamilton</strong>–Jacobi equations, we go back to our one-dimensional model problem:{u t (x, t) + H(u x (x, t)) = 0 (x, t) ∈ R × [0, T ]u(x, 0) = u 0 (x) x ∈ R.(6.84)Throughout this section, we will make the st<strong>and</strong>ing assumption that H is a strictlyconvex function, <strong>and</strong> more precisely:H ′′ (p) ≥ m H . (6.85)


✐✐6.2. Semi-Lagrangian schemes <strong>for</strong> the convex HJ equation 177Figure 6.7. Numerical results <strong>for</strong> the advection of a characteristic functionobtained via P 2 (upper left), cubic (upper right), 3rd-order ENO (lower left)<strong>and</strong> cubic with maximum consistency rate (lower right) schemes, second-order timediscretization, 200 nodes.We recall that this implies that, if u 0 is a Lipschitz continuous function, then u(x, t)is semiconcave <strong>for</strong> all t > 0, <strong>and</strong> if u 0 is semiconcave itself, then u(x, t) is uni<strong>for</strong>mlysemiconcave <strong>for</strong> t ≥ 0. Moreover, in terms of the Legendre trans<strong>for</strong>m, condition(6.85) implies thatH ∗′′ (α) ≤ 1m H. (6.86)6.2.1 Construction of the schemeAs it has been shown <strong>for</strong> the linear case, turning to high-order the first-order SLscheme (5.92) only requires to replace the monotone P 1 interpolation I 1 with areconstruction of degree r > 1:{vn+1j= minα∈R [∆tH∗ (α) + I r [V n ](x j − α∆t)]v 0 j = u 0(x j ).(6.87)In this section, we will develop a convergence theory which allows to treat the casesof Lagrange, finite elements, ENO <strong>and</strong> WENO reconstructions. Strictly speaking,the scheme (6.87) is not monotone, but in fact Barles–Souganidis theorem requiresa weaker condition, that is, that (6.87) should be monotone up to a term o(∆t).As we will show, this generalized <strong>for</strong>m of monotonicity can occur under increasing


✐✐178 Chapter 6. High-order SL approximation schemesCourant numbers if the scheme is Lipschitz stable, so this will be the main stabilityissue.6.2.2 ConsistencyConsistency analysis clearly follows the same lines of the low-order scheme, once theP 1 reconstruction is replaced by a high-order version, <strong>and</strong> the definition of S SL ischanged accordingly. We can repeat the steps of the corresponding estimate, unless<strong>for</strong> using (6.4) in place of (5.43), obtaining at last <strong>for</strong> the high-order SL scheme(6.87) the consistency estimate∥∥L SL (∆; t, U(t)) ∥ ≤ C ∆xr+1. (6.88)∆tRemark 6.13. As it has already been noticed about the first-order SL scheme,due to the fact that characteristics are straight lines, the bound (6.88) does nottake into account the accuracy of time discretization. This is related to the specificstructure of the HJ equation (6.84), but if a transport term appears in the equation,<strong>and</strong> is discretized with order p, then the consistency estimate coincides with the oneobtained in the linear case, that is,∥ L SL (∆; t, U(t)) ∥ ()≤ C ∆t p + ∆xr+1.∆t6.2.3 Lipschitz stabilityIn order to be as general as possible, we start with some basic assumption on thereconstruction operator. First, we define the stencil of reconstruction asS(x) = (x − h − ∆x, x + h + ∆x), (6.89)so that the reconstruction at the point x uses the nodes inside S(x). For example,a quadratic Lagrange reconstruction can be per<strong>for</strong>med taking one node on the left<strong>and</strong> two nodes on the right of the point x (in this case, h − = 1, h + = 2), or twonodes on the left <strong>and</strong> one on the right (<strong>and</strong> in this case, h − = 2, h + = 1). In asecond-order finite element or ENO/WENO reconstruction, both cases are possible,thus h − = h + = 2. A third-order Lagrange reconstruction is per<strong>for</strong>med using twonodes on the left <strong>and</strong> two on the right, so that h − = h + = 2. In the third-orderfinite element or ENO/WENO case, h − = h + = 3, <strong>and</strong> so <strong>for</strong>th.Then, we assume that, given a Lipschitz continuous function v(x) <strong>and</strong> thesequence V = {v j } j∈I = {v(x j )} j∈I , the operator I r [V ] satisfies:|I r [V ](x) − I 1 [V ](x)| ≤ C r maxx j−1,x j,x j+1∈S(x) |v j+1 − 2v j + v j−1 | (C r < 1). (6.90)The key point of assumption (6.90) consists in requiring that C r < 1. In fact,as it is immediate to check, any polynomial reconstruction satisfies (6.90) with aconstant depending on the interpolation degree r. Besides a technical reason whichwill become clear in the proof of Lemma 6.14, the bound on C r amounts to assumethat I r is not “too sensitive” with respect to large second increments which canoccur on singularities of the solution.


✐✐6.2. Semi-Lagrangian schemes <strong>for</strong> the convex HJ equation 179It will also be useful to define the functionF j (α) := ∆tH ∗ (α) + I r [V n ](x j − α∆t), (6.91)<strong>and</strong> denote again by ᾱ j the value of α achieving the minimum in (6.87) <strong>and</strong> in F j .In order to stress the “locality” in the above definition, the notation explicitly showsthe dependence of F on the node index j, although <strong>for</strong> simplicity the dependenceon the time step n has been dropped.We give now a bound on the second increment of the numerical solution. Thebound is globally one-sided, but becomes two-sided at the foot of characteristics,that is, in a neighborhood U(x j + ᾱ j ∆t), with U defined asU(x) = (x − h∆x, x + h∆x), (6.92)with a fixed h > max(h + , h − ), so that S(x) ⊂ U(x). More precisely, we have thefollowing technical result:Lemma 6.14. Let V n be defined by the scheme (6.87). If (6.86) holds, then, <strong>for</strong>any j ∈ Z <strong>and</strong> n ≥ 1,vj+1 n − 2vj n + vj−1 n ≤∆x2m H ∆t . (6.93)Moreover, assuming, in addition, that (6.90) holds, thenmax |vi+1 n − 2vi n + vi−1| n ≤ C ∆x2x i−1,x i,x i+1∈U(x j+ᾱ j∆t)∆t(6.94)with U defined by (6.92) <strong>and</strong> <strong>for</strong> some positive constant C depending on C r , h, <strong>and</strong>m H .Proof. We start by proving (6.93). By (6.87) we have, <strong>for</strong> n ≥ 1,v n j= ∆tH ∗ (ᾱ j ) + I r[Vn−1 ] (x j − ᾱ j ∆t),vj−1 n = ∆tH ∗ [(ᾱ j−1 ) + I r Vn−1 ] (x j−1 − ᾱ j−1 ∆t) ≤(≤ ∆tH ∗ ᾱ j − ∆x )[+ I r Vn−1 ] (x j − ᾱ j ∆t),∆tvj+1 n = ∆tH ∗ [(ᾱ j+1 ) + I r Vn−1 ] (x j+1 − ᾱ j+1 ∆t) ≤(≤ ∆tH ∗ ᾱ j + ∆x )[+ I r Vn−1 ] (x j − ᾱ j ∆t)∆t(where the inequalities come from the use of ᾱ j ± ∆x/∆t instead of the actualminimizers ᾱ j±1 ), so that(vj+1 n − 2vj n + vj−1 n ≤ ∆t[H ∗ ᾱ j + ∆x ) (− 2H ∗ (ᾱ j ) + H ∗ ᾱ j − ∆x )]≤∆t∆t( ) 2 ∆x≤ ∆t sup H ∗′′ , (6.95)∆twhich, using (6.86), gives (6.93).


✐✐180 Chapter 6. High-order SL approximation schemesIn order to prove (6.94), we take into consideration the values of F j (α) <strong>for</strong>x j − α∆t coinciding with a grid node, that is <strong>for</strong> α = k∆x/∆t. Let k j be definedas:⌊k j = sgn (ᾱ j ) |ᾱ j | ∆t ⌋∆xso that{x j − ᾱ j ∆t ∈ [x j−kj , x j−kj+1) if ᾱ j ≤ 0x j − ᾱ j ∆t ∈ (x j−kj−1, x j−kj ] if ᾱ j ≥ 0Since the two cases are perfectly symmetric, we examine in detail the second, ᾱ j ≥ 0.In this case, the point x j − ᾱ j ∆t is on the left of x j , <strong>and</strong> more preciselyMoreover, the minimizer ᾱ j satisfiesx j−kj−1 < x j − ᾱ j ∆t ≤ x j−kj ≤ x j .k j∆x∆t ≤ ᾱ j < (k j + 1) ∆x∆t .We will denote by δ j the (signed) maximal second increment of the numerical solutionin the neighborhood U(x j − ᾱ j ∆t) <strong>and</strong> by j − k j + l j the index of the node atwhich this second increment occurs. If δ j ≥ 0, then (6.94) follows from (6.93) <strong>and</strong>we have nothing else to prove. We will assume, there<strong>for</strong>e, that δ j < 0, so thatδ j = −maxx i−1,x i,x i+1∈U(x j−ᾱ j∆t)|v n i+1 − 2v n i + v n i−1| == v n j−k j+l j+1 − 2v n j−k j+l j+ v n j−k j+l j−1. (6.96)We recall that |l j | < h, <strong>and</strong> will also assume in what follows that l j ≥ 0, the reversecase being similar to prove.Step 1.Upper bound of the second increment of the function F j . We have:(F j (k + 1) ∆x∆t= ∆t≤) (− 2F j k ∆x ) (+ F j (k − 1) ∆x )=∆t∆t( [H ∗ (k + 1) ∆x ) (− 2H ∗ k ∆x∆t∆t+ v n j−k−1 − 2v n j−k + v n j−k+1 ≤)+ H ∗ ((k − 1) ∆x∆t)]+∆x2m H ∆t + vn j−k−1 − 2vj−k n + vj−k+1 n ≤ 2∆x2m H ∆t , (6.97)where the second increment of H ∗ has been estimated as in (6.95), <strong>and</strong> the lastinequality holds <strong>for</strong> n ≥ 1.Step 2.In the second step, we prove that( ) (∆xF j (ᾱ j ) ≥ min[F j k j , F j (k j + 1) ∆x )]− ∆x2∆t∆t 8m H ∆t − C r|δ j |. (6.98)


✐✐6.2. Semi-Lagrangian schemes <strong>for</strong> the convex HJ equation 181In fact, set ᾱ j = (1 − ¯θ)k j ∆x/∆t + ¯θ(k j + 1)∆x/∆t, with ¯θ ∈ [0, 1]. Taking intoaccount the uni<strong>for</strong>m convexity of H ∗ implied by (6.86), we have( ) (H ∗ (ᾱ j ) ≥ (1 − ¯θ)H ∗ ∆xk j +∆t¯θH ∗ (k j + 1) ∆x )−∆t− 1 ( ) 22 ¯θ(1 − ¯θ) ∆xsup H ∗′′ ≥∆t(≥ (1 − ¯θ)H ∗ ∆xk j∆t) (+ ¯θH ∗ (k j + 1) ∆x )−∆t−∆x28m H ∆t 2 (6.99)in which the third term has been minimized with respect to ¯θ.In a similar way, we also get, using (6.90) <strong>and</strong> the inclusion of S(x) in U(x),I r [V n ](x j − ᾱ j ∆t) ≥ I 1 [V n ](x j − ᾱ j ∆t) −−C r maxx |vn i+1 − 2vi n + vi−1| n ≥i−1,x i,x i+1∈S(x j−ᾱ j∆t)≥ I 1 [V n ](x j − ᾱ j ∆t) −−C rmaxx i−1,x i,x i+1∈U(x j−ᾱ j∆t)= (1 − ¯θ)v n j−k j+ ¯θv n j−k j−1 − C r |δ j | =(= (1 − ¯θ)I r [V n ]+¯θI r [V n ]x j − k j∆x∆t ∆t )+|v n i+1 − 2v n i + v n i−1| =(x j − (k j + 1) ∆x∆t ∆t )− C r |δ j |,which, combined with (6.99), gives( ) (F j (ᾱ j ) ≥ (1 − ¯θ)F ∆xj k j +∆t¯θF j (k j + 1) ∆x )− ∆x2∆t 8m H ∆t − C r|δ j |,which in turn implies (6.98).Step 3. Upper bound of the increment of F j between k j ∆x/∆t <strong>and</strong> (k j +1)∆x/∆t.Assume first that F j (k j ∆x/∆t) < F j ((k j + 1)∆x/∆t). Using (6.98) <strong>and</strong> the optimalityof ᾱ j , we obtain(F j (k j − 1) ∆x ) ( )∆x≥ F j (ᾱ j ) ≥ F j k j − ∆x2∆t∆t 8m H ∆t − C r|δ j |,that is, considering the extreme terms(∆xF j k j∆t)− F j((k j − 1) ∆x∆t)≤ ∆x28m H ∆t + C r|δ j |.On the other h<strong>and</strong>, from (6.97) written with k = k j , we also have(F j (k j + 1) ∆x ) ( ) ( ) (∆x∆x− F j k j ≤ F j k j − F j (k j − 1) ∆x )+ 2∆x2∆t∆t∆t∆t m H ∆t


✐✐182 Chapter 6. High-order SL approximation schemes<strong>and</strong> there<strong>for</strong>e, combining the last two inequalities, we get the desired bound on thefirst increment:(F j (k j + 1) ∆x ) ( )∆x− F j k j ≤ 17∆x2∆t∆t 8m H ∆t + C r|δ j |. (6.100)Moreover, if F j ((k j + 1)∆x/∆t) < F j (k j ∆x/∆t), then(F j (k j + 1) ∆x ) ( )∆x− F j k j ≤ 0∆t∆t<strong>and</strong> (6.100) is also trivially satisfied.Step 4. We show that, in order <strong>for</strong> ᾱ j to be a minimizer <strong>for</strong> F j , δ j must satisfy(6.94). To this end, we first assume that F j (k j ∆x/∆t) ≤ F j ((k j + 1)∆x/∆t) <strong>and</strong>bound the values F j (k∆x/∆t) from above using a function ˜F j (k∆x/∆t) constructedso as to coincide with F j at k j ∆x/∆t <strong>and</strong> to have first <strong>and</strong> second increments greateror equal to the corresponding increments of F j . Taking into account the boundson the first <strong>and</strong> second increment of F j obtained in the previous steps of the proof,we could define ˜F j so that its first increment between k j ∆x/∆t <strong>and</strong> (k j + 1)∆x/∆twould be given by the right-h<strong>and</strong> side of (6.100), <strong>and</strong> the second increment wouldbe obtained in view of (6.96), (6.97) as(˜F j (k + 1) ∆x ) (− 2∆t˜F j k ∆x ) (+∆t˜F j (k − 1) ∆x )=∆t⎧2∆x 2⎪⎨if k ≠ k j + l j ,m H ∆t=∆x ⎪⎩2m H ∆t + δ j if k = k j + l j .(6.101)We recall that, as it is easy to prove by induction, if a sequence f i has aconstant second increment,f i+2 − 2f i+1 + f i ≡ d,then the values of the elements <strong>and</strong> of the first increments are given byf k+l = f k + l(f k+1 − f k ) + (1 + 2 + · · · + (l − 1))d,f k+l+1 − f k+l = (f k+1 − f k ) + ld.Using the previous equalities, a function ˜F j suitable <strong>for</strong> our purpose could be


✐✐6.2. Semi-Lagrangian schemes <strong>for</strong> the convex HJ equation 183defined more explicitly as(∆x˜F j k j∆t) (∆x= F j k j∆t),(˜F j (k j + 1) ∆x ) ( )∆x= F j k j + 17∆x2∆t∆t 8m H ∆t + C r|δ j |,˜F j((k j + 2) ∆x∆t)= F j(k j∆x∆t)+ 2( ) 17∆x28m H ∆t + C r|δ j | + 2∆x2m H ∆t ,(.˜F j (k j + l j ) ∆x ) ( ) ( )∆x 17∆x2= F j k j + l j∆t∆t 8m H ∆t + C r|δ j |+(1 + · · · + (l j − 1)) 2∆x2m H ∆t ,(˜F j (k j + l j + 1) ∆x ) ( ) ( )∆x 17∆x2= F j k j + l j∆t∆t 8m H ∆t + C r|δ j | ++ (1 + · · · + (l j − 1)) 2∆x2m H ∆t +( ) 17∆x2+8m H ∆t + C 2∆x 2r|δ j | + l jm H ∆t + δ j(note that the computation of ˜Fj ((k j + l j + 1)∆x/∆t) is “restarted” because of(6.101)) <strong>and</strong> last, <strong>for</strong> any integer m > 0,(˜F j (k j + l j + m) ∆x ) ( )∆x= F j k j +∆t∆t( )17∆x2+ (l j + m)8m H ∆t + C r|δ j | ++ (1 + · · · + (l j + m − 1)) 2∆x2m H ∆t + mδ j ≤( ) ( )∆x17∆x2≤ F j k j + (h + m)∆t8m H ∆t + C r|δ j | ++ (1 + · · · + (h + m − 1)) 2∆x2m H ∆t + mδ j. (6.102)On the other h<strong>and</strong>, by (6.98) <strong>and</strong> the optimality of ᾱ j , we also have, <strong>for</strong> any m > 0,(˜F j (k j + l j + m) ∆x ) (≥ F j (k j + l j + m) ∆x )≥∆t∆t≥ F j (ᾱ j ) ≥( ) (∆x≥ min[F j k j , F j (k j + 1) ∆x )]−∆t∆t−∆x28m H ∆t − C r|δ j |. (6.103)We explicitly note that in (6.103) the first inequality follows from the construction+


✐✐184 Chapter 6. High-order SL approximation schemesof ˜F j as an upper bound <strong>for</strong> F j , the second one from the optimality of ᾱ j in F j ,<strong>and</strong> the third one from the lower bound (6.98).Recalling that we have assumed F j (k j ∆x/∆t) ≤ F j ((k j + 1)∆x/∆t) <strong>and</strong> δ j 0 of the left-h<strong>and</strong> side(which is in fact a maximum) is O(∆x 2 /∆t), provided C r < 1. Using the reverseestimate (6.93), we get at last (6.94).If otherwise F j ((k j +1)∆x/∆t) < F j (k j ∆x/∆t), then it is possible to redefinethe function ˜F j so that ˜F j ((k j + 1)∆x/∆t) = F j ((k j + 1)∆x/∆t), <strong>and</strong> the sameupper bounds are used on the first <strong>and</strong> second increments. In this way, we replace(6.102) with(˜F j (k j + l j + m) ∆x ) (= F j (k j + 1) ∆x∆t∆t)+( )17∆x2+ (l j + m − 1)8m H ∆t + C r|δ j | ++ (1 + · · · + (l j + m − 1)) 2∆x2m H ∆t + mδ j,which again implies (6.104). This construction completely parallels the previousone <strong>and</strong> is there<strong>for</strong>e left to the reader.Be<strong>for</strong>e proving Lipschitz continuity, we derive a couple of useful consequencesof Lemma 6.14. First, given a Lipschitz continuous function v(x) <strong>and</strong> the sequenceV = {v j } j∈I = {v(x j )} j∈I , the condition|I r [V ](x) − I 1 [V ](x)| = O(∆x),typical in the interpolation of a function with bounded derivative, is written by(6.90) as|I r [V ](x) − I 1 [V ](x)| ≤ C r maxx i−1,x i,x i+1∈S(x) |v i+1 − 2v i + v i−1 | ≤≤ C r (sup |v i+1 − v i | + sup |v i−1 − v i |) ≤≤ 2L∆x. (6.105)


✐✐6.2. Semi-Lagrangian schemes <strong>for</strong> the convex HJ equation 185Also, at the foot of a characteristic the stronger condition|I r [V n ](x i + ᾱ j ∆t) − I 1 [V n ](x i + ᾱ j ∆t)| ≤ C r C ∆x2∆t(6.106)holds <strong>for</strong> any j ∈ Z, n ≥ 1 <strong>and</strong> <strong>for</strong> any node i = j ± 1. In fact, (6.106) follows from(6.90) <strong>and</strong> Lemma 6.14, since U(x j + ᾱ j ∆t) contains all the nodes involved in thereconstructions I r [V n ](x j±1 + ᾱ j ∆t).We can now state the result of Lipschitz stability.Theorem 6.15. Let V n be defined by the scheme (6.87). Assume that (6.86),(6.90) hold, that ∆x = O(∆t 2 ), <strong>and</strong> that u 0 is Lipschitz continuous with Lipschitzconstant L 0 . Then, V n satisfy, <strong>for</strong> any i <strong>and</strong> j, the discrete Lipschitz estimate|v n i − vn j ||x i − x j | ≤ L<strong>for</strong> a constant L independent of ∆x <strong>and</strong> ∆t, <strong>and</strong> <strong>for</strong> 0 ≤ n ≤ T/∆t, as ∆t → 0.Proof. It clearly suffices to prove the claim <strong>for</strong> i, j such that i = j ± 1. Assumethat at the previous step the discrete solution satisfies|v n−1i− v n−1j |∆x≤ L n−1 .Making the argmin explicit <strong>and</strong> using (6.106), we havev n j= ∆tH ∗ (ᾱ j ) + I r[Vn−1 ] (x j + ᾱ j ∆t) ≥≥ ∆tH ∗ (ᾱ j ) + I 1[Vn−1 ] (x j + ᾱ j ∆t) − C r C ∆x2∆t . (6.107)In order to estimate the discrete incremental ratio of V n , we give on v n iv n i= ∆tH ∗ (ᾱ i ) + I r[Vn−1 ] (x i + ᾱ i ∆t) ≤≤ ∆tH ∗ (ᾱ j ) + I r[Vn−1 ] (x i + ᾱ j ∆t) ≤the bound≤ ∆tH ∗ (ᾱ j ) + I 1[Vn−1 ] (x i + ᾱ j ∆t) + C r C ∆x2∆t , (6.108)which results from both the optimality of ᾱ i <strong>and</strong> (6.106), <strong>and</strong> holds <strong>for</strong> any n ≥ 2.If n = 1, applying (6.105) instead of (6.106), we obtainv 1 i = ∆tH ∗ (ᾱ i ) + I r[V0 ] (x i + ᾱ i ∆t) ≤≤ ∆tH ∗ (ᾱ j ) + I r[V0 ] (x i + ᾱ j ∆t) ≤≤ ∆tH ∗ (ᾱ j ) + I 1[V0 ] (x i + ᾱ j ∆t) + 2L 0 ∆x. (6.109)From (6.107) <strong>and</strong> (6.108) we obtain, <strong>for</strong> n ≥ 2, the unilateral estimatev n i − vn j∆x≤ 1∆x(I1[Vn−1 ] (x i + ᾱ j ∆t) − I 1[Vn−1 ] (x j + ᾱ j ∆t) ) + 2C r C ∆x2∆t≤≤ L n−1 + 2C r C ∆x∆t , (6.110)


✐✐186 Chapter 6. High-order SL approximation schemesin which we have used the fact that the first-order reconstruction I 1 at step n − 1has also Lipschitz constant L n−1 . Interchanging the role of ᾱ j <strong>and</strong> ᾱ j+1 , we get thereverse estimatev n j − vn i∆x≤ L n−1 + 2C r C ∆x∆t ,<strong>and</strong> there<strong>for</strong>e|vj n − vn i | ≤ L n−1 + 2C r C ∆x∆x∆t . (6.111)A similar computation yields, <strong>for</strong> n = 1,|vj 1 − v1 i | ≤ L 0 + 4L 0 = 5L 0 , (6.112)∆xso that combining (6.111) with (6.112) <strong>and</strong> iterating back, we haveL n ≤ L n−1 + 2C r C ∆x∆t ≤ · · · ≤≤ L 1 + T − ∆t 2C r C ∆x∆t ∆t ≤≤ 5L 0 + 2T C r C ∆x∆t 2 . (6.113)Last, it is possible to get a finite limit in (6.113), if <strong>and</strong> only if ∆x =O(∆t 2 ).Applications to various reconstruction operatorsIn order to check the applicability of this theory to reconstruction operators of practicalinterest, we first test assumption (6.86) on the situation of an interpolatingpolynomial with stencil including the reconstruction point x (this is directly applicableto the case of finite element <strong>and</strong> ENO interpolations). Then, starting withthis result, we discuss the cases of symmetric Lagrange <strong>and</strong> WENO interpolations.Finite element <strong>and</strong> ENO interpolations Here, we make no assumptions on thestructure of the stencil S(x), unless that it must include one node on each side ofx. The reconstruction is assumed to be in the Newton <strong>for</strong>mI r [V ](x) = V [x j0 ] + V [x j0 , x j1 ](x − x j0 ) + · · · ++ V [x j0 , . . . , x jr ](x − x j0 ) · · · (x − x jr−1 ), (6.114)where x j0 , . . . , x jr are r+1 adjacent nodes so that max(x j0 , . . . , x jr )−min(x j0 , . . . , x jr )= r∆x <strong>and</strong>, moreover,x ∈ (min(x j0 , . . . , x jr ), max(x j0 , . . . , x jr )) ⊂ S(x). (6.115)This definition includes both finite element interpolations, <strong>for</strong> which the reconstructionstencil is fixed (but not necessarily symmetric around x), <strong>and</strong> ENO reconstructions,<strong>for</strong> which it depends on the solution itself. The divided differences are defined,as usual, byV [x j0 ] = v j0 ,..V [x j0 , . . . , x jk ] = V [x j 1, . . . , x jk ] − V [x j0 , . . . , x jk−1 ]x jk − x j0(k = 1, . . . , r).


✐✐6.2. Semi-Lagrangian schemes <strong>for</strong> the convex HJ equation 187Note that, although in principle the nodes x j0 , . . . , x jk need neither to be adjacentnor to satisfy x ∈ (min x ji , max x ji ), it is possible to reorder the nodes so that bothconditions would be satisfied.According to its definition, we can bound a generic k-th order divided differenceas follows:|V [x j0 , . . . , x jk ]| ≤ 2 max |V [x j i, . . . , x jk+i−1 ]|k∆x(6.116)in which the max is per<strong>for</strong>med <strong>for</strong> x ji , . . . , x jk+i−1 ∈ S(x). To prove (6.90), we startfrom the second divided difference<strong>and</strong>, hence,|V [x j0 , x j1 , x j2 ]| = |v j 0− 2v j1 + v j2 |2∆x 2 ,|V [x j0 , x j1 , x j2 , x j3 ]| ≤ 2 max |v j i− 2v ji+1 + v ji+2 |3!∆x 3.|V [x j0 , . . . , x jr ]| ≤ 2r−2 max |v ji − 2v ji+1 + v ji+2 |r!∆x r .Plugging such bounds into (6.114), we get an estimate in the <strong>for</strong>m (6.90); that is,where|I r [V ](x) − I 1 [V ](x)| ≤ |V [x j0 , x j1 , x j2 ](x − x j0 )(x − x j1 ) + · · · ++V [x j0 , . . . , x jr ](x − x j0 ) · · · (x − x jr−1 )| ≤≤ |v j 0− 2v j1 + v j2 |2∆x 2 M 2 ∆x 2 + · · · ++ 2r−2 max |v ji − 2v ji+1 + v ji+2 |r!∆x rM r ∆x r ≤r∑ M k 2 k−2≤ max |v ji − 2v ji+1 + v ji+2 |, (6.117)k!k=2M k := 1∆x k max |(x − x j 0) · · · (x − x jk−1 )| =x∈(x j0 ,x jk−1 )= max |t(t − 1) · · · (t − k + 1)|.t∈(0,k−1)It remains to check that C r < 1; that is,r∑ M k 2 k−2< 1. (6.118)k!k=2Indeed, the first values M k may be either computed by algebraic manipulations(up to M 5 ), or estimated by simply plotting the polynomials t(t − 1) · · · (t − k + 1)on the interval [0, k − 1]. It turns out that M 2 = 1/4, M 3 = 2 √ 3/9 ≈ 0.3849,M 4 = 1, M 5 ≈ 3.6314, M 6 < 17, <strong>and</strong> so on. Accordingly, the computation of theleft-h<strong>and</strong> side of (6.118) <strong>for</strong> various values of r gives C 2 = 1/8 = 0.125, C 3 ≈ 0.2533,C 4 ≈ 0.42, C 5 ≈ 0.6621, C 6 ≈ 1.04. We can conclude that by this technique, finiteelement <strong>and</strong> ENO reconstructions up to the fifth order can be proved to satisfy(6.90).


✐✐188 Chapter 6. High-order SL approximation schemesSymmetric Lagrange <strong>and</strong> WENO interpolations It is possible to extend thistheory to Lagrange <strong>and</strong> WENO interpolations, which work with a symmetric reconstructionstencil, that is, h + = h − . Assume that the interpolation I r [V ](x) isobtained as a sumI r [V ] = W 1 (x)P 1 (x) + · · · + W q (x)P q (x) (6.119)in which the P i are interpolating polynomials constructed on smaller stencils whichinclude the point x <strong>and</strong>, once x is fixed, the W i (x) are coefficients of a convexcombination (W i (x) ≥ 0, ∑ i W i(x) = 1). Then, we havemin(P 1 (x), . . . , P q (x)) ≤ I r [V ](x) ≤ max(P 1 (x), . . . , P q (x)).If deg P i ≤ 5, all the polynomials P i satisfy (6.90) <strong>and</strong> there<strong>for</strong>e, being satisfied byall P i , (6.90) is also satisfied by I r [V ].Note that (6.119) is the <strong>for</strong>m <strong>for</strong> both symmetric Lagrange <strong>and</strong> WENO interpolation.In the first case, the functions W i play the role of linear weights asin (3.24), <strong>and</strong> it has been proved in Theorem 3.10 that they are positive <strong>and</strong> haveunity sum. In the second case, the <strong>for</strong>m of I r is given by (3.27) <strong>and</strong>, although thenonlinear weights are no longer polynomials, they still per<strong>for</strong>m a convex combinationdue to the positivity of the linear weights. There<strong>for</strong>e, (6.90) is again satisfiedprovided deg P i ≤ 5. In both cases, this means a reconstruction of order r ≤ 9.6.2.4 ConvergenceLast, we present the main convergence result <strong>for</strong> the scheme (6.87).Theorem 6.16. Let u be the solution of (6.84), <strong>and</strong> V n be defined by the scheme(6.87). Assume that (6.105), (6.106) hold, that ∆x = O(∆t 2 ), <strong>and</strong> that u 0 isLipschitz continuous. Then,<strong>for</strong> 0 ≤ n ≤ T/∆t, as ∆t → 0.‖I r [V n ] − U(t n )‖ ∞ → 0Proof. First, note that the scheme is invariant <strong>for</strong> the addition of constants <strong>and</strong>consistent. Then, retracing the proof of (5.103) <strong>and</strong> using (6.105) <strong>and</strong> Theorem 6.15,we obtain that the scheme is monotone up to a term O(∆x) (there<strong>for</strong>e, o(∆t)), <strong>and</strong>the generalized monotonicity condition (4.58)–(4.59) is satisfied. Then, convergencefollows from Barles–Souganidis theorem.Remark 6.17. In this theorem, the condition ∆x = O(∆t 2 ) is only required toensure Lipschitz stability. In fact (<strong>and</strong> numerical experiments confirm this idea)this could be an overly restrictive assumption, possibly owing to some intrinsiclimitations in the technique of proof. Once ensured uni<strong>for</strong>m Lipschitz continuityof numerical solutions, the generalized monotonicity condition requires instead theweaker constraint ∆x = o(∆t).


✐✐6.3. Commented references 189Be<strong>for</strong>e singularityn n cubic ... WENO3 WENO525 2.52 · 10 −3 1.29 · 10 −350 8.77 · 10 −5 1.87 · 10 −5100 1.53 · 10 −5 9.13 · 10 −7200 9.63 · 10 −7 2.01 · 10 −8rate 3.78 5.32After singularityn n cubic ... WENO3 WENO525 2.88 · 10 −3 3.05 · 10 −350 5.12 · 10 −5 5.83 · 10 −6100 2.19 · 10 −6 7.25 · 10 −8200 2.39 · 10 −7 1.89 · 10 −9rate 4.52 6.87Table 6.2. Errors be<strong>for</strong>e <strong>and</strong> after the singularity <strong>for</strong> Test (6.120),WENO3 <strong>and</strong> WENO5 schemes.6.2.5 ExamplesUni<strong>for</strong>mly semiconcave solution This test (also considered, <strong>for</strong> example, in [LT00],[JP00]) concerns the HJ equation:{u t (x, t) + 1 2 (u x(x, t) + 1) 2 = 0(6.120)u(x, 0) = u 0 (x) = − cos(πx),on the interval [0, 2], with periodic boundary conditions. The solution eventuallydevelops a singularity in the derivative, <strong>and</strong> the approximate solution is computedbe<strong>for</strong>e the singularity, at T = 0.8/π 2 , with 4 time steps, <strong>and</strong> after the singularity, atT = 1.5/π 2 , with 5 time steps. Plots of exact vs. approximate solutions are shownin Figure 6.8 <strong>and</strong> L ∞ errors obtained with different schemes are compared in Table6.2.Note that, according to the consistency estimate (6.88), the convergence ratesshould be r + 1 = 4 <strong>for</strong> the WENO3 scheme <strong>and</strong> r + 1 = 6 <strong>for</strong> the WENO5 scheme,since the time step has been kept constant. Roughly speaking, this behavior isconfirmed by numerical tests. Due to semiconcavity, even after the onset of the singularitythe computation of the approximate solution remains remarkably accurate,more than what could be deduced from consistency analysis <strong>for</strong> a nondifferentiablesolution. The influence of semiconcavity on the accuracy of numerical schemes hasalready been discussed in the presentation of numerical examples of Chapter 5.6.3 Commented referencesThe first attempts of a convergence analysis <strong>for</strong> high-order Semi-Lagrangian schemeshave appeared in a somewhat fragmentary <strong>for</strong>m. As far as we can say, the first papercollecting convergence issues <strong>for</strong> (linear) SL schemes is [FF98]. Even this paper,


✐✐190 Chapter 6. High-order SL approximation schemes1’Initial_condition_test_1’10.8’Exact_solution’’3rd_order’0.60.50.40.200-0.2-0.4-0.5-0.6-0.8-1-10 0.5 1 1.5 20.80.6-1.20 0.5 1 1.5 2’Exact_solution’’3rd_order’0.40.20-0.2-0.4-0.6-0.8-1-1.20 0.5 1 1.5 2Figure 6.8. Approximate vs. exact solution <strong>for</strong> Test (6.120): initial condition(upper left), solution be<strong>for</strong>e singularity (upper right) <strong>and</strong> solution after singularity(lower).however, does not solve the point of l 2 stability in a completely rigorous way: aFourier analysis is per<strong>for</strong>med, but Von Neumann condition is only en<strong>for</strong>ced graphically,whereas an explicit (<strong>and</strong> very technical) solution of the Von Neumann analysis<strong>for</strong> Lagrange reconstructions has been given later in [BM08].Equivalence between SL <strong>and</strong> Lagrange–Galerkin schemes has proved to be a simplertool <strong>for</strong> proving l 2 stability. This approach has been first studied in [Fe10]<strong>for</strong> constant coefficient equations, <strong>and</strong> extended to variable coefficient equations in[Fe11]. Basic references on the Lagrange–Galerkin technique are the seminal papers[DR82] <strong>and</strong> [Pi82], while the area-weighted version is proposed <strong>and</strong> studied in[MPS88]. The simplified analysis <strong>for</strong> finite element reconstructions has been developedin [P06] (the paper [FP07] gives a synthesis of this work), whereas an extensiveanalysis of the interaction between discretization steps is presented in [FFM01].High-order SL schemes <strong>for</strong> HJ equations have been first considered in [FF02],<strong>and</strong> a convergence analysis based on the condition ∆x = O(∆t 2 ) is carried out in[Fe03]. Convergence is directly proved in this paper starting from Lipschitz stability,whereas we have chosen here to prove it via the Barles–Souganidis Theorem. Theadaptation of the theory to weighted ENO reconstructions is given in [CFR05], alongwith a number of numerical tests comparing the various high-order versions of thescheme. Other numerical tests, mostly in higher dimension <strong>and</strong> concerned withapplications to front propagation <strong>and</strong> optimal control, are presented in [CFF04].


✐✐Chapter 7Control <strong>and</strong> GamesOne of the most typical applications of the theory of <strong>Hamilton</strong>–Jacobi equations isin the field of optimal control problems <strong>and</strong> differential games. Via the DynamicProgramming Principle (DPP in the sequel), <strong>for</strong>mulated by R. Bellman in the 60s,many optimal control problems can be characterized by means of the associatedvalue function, which can be shown in turn to be the unique viscosity solution of apartial differential equation of convex <strong>Hamilton</strong>–Jacobi type, usually called Bellmanequation or Dynamic Programming equation.Although the direct numerical solution of the Bellman equation presents seriouscomplexity problems related to the dimension of the state space (the so-calledcurse of dimensionality), it also allows to recover a sharper in<strong>for</strong>mation. Actually,in comparison with the more conventional technique given by the Pontryagin MaximumPrinciple (which characterizes the optimal control related to a given initialstate as a function of time), the Bellman equation provides a synthesis of feedbackcontrols, i.e. of controls expressed as a function of the state variable (instead oftime).In this chapter we present algorithms of Semi-Lagrangian type <strong>for</strong> the approximatesolution of Dynamic Programming equations related to optimal controlproblems <strong>and</strong> differential games. In fact, the same approach can be applied to theanalysis <strong>and</strong> approximation of zero-sum differential games where, under suitableconditions, the value function of the game solves a non-convex <strong>Hamilton</strong>–Jacobitype equation, usually called Isaacs equation.The Semi-Lagrangian numerical methods presented here are strongly connectedto Dynamic Programing <strong>for</strong> two main reasons. The first is that both Bellman<strong>and</strong> Isaacs equations are derived themselves by passing to the limit in the DynamicProgramming Principle, the second is that the SL schemes are obtained by a discreteversion of the DPP, which plays the role of the Hopf–Lax <strong>for</strong>mula <strong>for</strong> thisspecific case. This strategy of construction gives a nice control interpretation of theschemes <strong>and</strong> also helps in their theoretical analysis. Moreover, it allows to constructat the same time the approximations of both value function <strong>and</strong> optimal feedback,the latter being a practical solution to the original optimization problem.While the construction <strong>and</strong> theoretical analysis of SL schemes <strong>for</strong> DynamicProgramming equations allows <strong>for</strong> an elegant <strong>and</strong> general setting, their applicationto real-life industrial problems still suffers from an impractical computationalcomplexity. However, the continuous improvement of computers, as well as the191


✐✐192 Chapter 7. Control <strong>and</strong> Gamesdevelopment of faster <strong>and</strong> more accurate algorithms, make this field an active <strong>and</strong>promising research area.7.1 Optimal control problems: first examplesIn order to present the main ideas, we start with the classical infinite horizonproblem. Let a controlled dynamical system be given by{ẏ(s) = f(y(s), s, α(s))(7.1)y(t 0 ) = x 0 ,where x 0 , y(s) ∈ R d , α : [t 0 , T ] → A, A ⊆ R m , T finite or +∞. In the sequel, wewill possibly use the less general <strong>for</strong>mulation{ẏ(s) = f(y(s), α(s))(7.2)y(t 0 ) = x 0 .In the applications, it is unreasonable to assume a priori any regularity of the controlα. We will rather assume that the control is measurable, in which case the existence<strong>and</strong> uniqueness <strong>for</strong> the solution of (7.1) is ensured by the Carathèodory theorem:Theorem 7.1. (Carathèodory) Assume that:(i) f(·, ·, ·) is continuous;(ii) there exists a positive constant L f > 0 such that|f(x, t, a) − f(y, t, a))| ≤ L f |x − y|,<strong>for</strong> all x, y ∈ R d , t ∈ R + <strong>and</strong> a ∈ A;(iii) f(x, t, α(t)) is measurable with respect to t.Then,is the unique solution of (7.1)∫ sy(s) = x 0 + f(y(τ), τ, α(τ))dτ (7.3)t 0Note that the solution is continuous, but only a.e. differentiable, so it mustbe regarded as a weak solution of (7.1). Note also that we are not interested hereto the most general version of this theorem, in which a condition of local Lipschitzcontinuity of f still allows to define a local solution y. This generalization is not ofgreat interest in Control Theory, in which a controlled dynamical system is usuallyrequired to have a solution <strong>for</strong> all s ≥ t 0 .By the theorem above, once fixed a control in the set of admissible controlsα ∈ A := {a : [t 0 , T ] → A, measurable} (7.4)there exists a unique trajectory y x0,t 0(s; α) of (7.1). Changing the control policy thetrajectory will change, <strong>and</strong> we will have a family of solutions of the controlled system(7.1) depending on α. To simplify notations, when considering the autonomousdynamics (7.2) <strong>and</strong> the initial time t 0 = 0, we will denote this family as y x0 (s; α)


✐✐7.1. Optimal control problems: first examples 193(or even with y(s) if no ambiguity occurs). Moreover, it is customary in DynamicProgramming to use the notations x <strong>and</strong> t instead of x 0 <strong>and</strong> t 0 (x <strong>and</strong> t will appearas variables in the <strong>Hamilton</strong>–Jacobi–Bellman equation), <strong>and</strong> there<strong>for</strong>e we willpermanently adopt this latter notation in the sequel.The second ingredient in the definition of an optimal control problem is thecost functional to be minimized. We introduce the problem by describing a coupleof classical examples.7.1.1 The infinite horizon problemOptimal control problems require the introduction of a cost functional J : A → Rwhich is used to select the “optimal trajectory” <strong>for</strong> (7.2). In the case of the infinitehorizon problem, we set t 0 = 0, x 0 = x, <strong>and</strong> this functional is defined asJ x (α) =∫ ∞0g(y x (s, α), α(s))e −λs ds (7.5)<strong>for</strong> a given λ > 0. The function g represents the running cost <strong>and</strong> λ is the discountfactor which allows to compare the costs at different times by rescaling the costs atthe initial time. From a technical point of view, the presence of the discount factorensures that the integral is finite whenever g is bounded.The goal of optimal control theory is to find an optimal pair (y ∗ , α ∗ ) whichminimizes the cost functional. If we look <strong>for</strong> optimal controls in open-loop <strong>for</strong>m, i.e.as functions of t, then the main tool is given by the Pontryagin Maximum Principle,which gives the necessary conditions to be satisfied by an optimal couple (y ∗ , α ∗ ).A major drawback of an open-loop control is that, being constructed as a functionof time, it cannot take into account errors in the real state of the system, due <strong>for</strong>example to model errors or external disturbances, which may bring the evolutionfar from the optimal <strong>for</strong>ecasted trajectory. Another limitation of this approach isthat it requires to compute again the optimal control <strong>for</strong> any different initial state,so that it becomes impossible to design a static controller <strong>for</strong> the system.For these reasons, we are interested in the so-called feedback controls, thatis, controls expressed as functions of the state of the system. Under an optimalfeedback control, if the system trajectory is modified by an external perturbation,then the system reacts by changing its control strategy according to the change inthe state. One of the main motivations to use the Dynamic Programming approachis precisely that it allows to characterize the optimal feedback, as we will see laterin this chapter.The starting point of Dynamic Programming is to introduce an auxiliary function,the value function, which in the case of the infinite horizon problem is definedasv(x) = infα∈A J x(α), (7.6)where, as above, x is the initial position of the system. The value function has aclear meaning: it is the optimal cost associated to the initial position x. This isa reference value which can be useful to evaluate the efficiency of a control – <strong>for</strong>example, if J x (ᾱ) is close to v(x), this means that ᾱ is “efficient”.Bellman’s Dynamic Programming Principle gives a first characterization ofthe value function.


✐✐194 Chapter 7. Control <strong>and</strong> GamesProposition 7.2 (DPP <strong>for</strong> the infinite horizon problem). Under the assumptionsof Theorem 7.1, <strong>for</strong> all x ∈ R d <strong>and</strong> τ > 0,{∫ τ}v(x) = inf g(y x (s; α), α(s))e −λs ds + e −λτ v(y x (τ; α)) . (7.7)α∈A0Proof. Denote by ¯v(x) the right-h<strong>and</strong> side of (7.7). First, we remark that, <strong>for</strong> anyx ∈ R d <strong>and</strong> ᾱ ∈ A,J x (ᾱ) ===≥∫ ∞0∫ τ0∫ τ0∫ τ0g(ȳ(s), ᾱ(s))e −λs ds =g(ȳ(s), ᾱ(s))e −λs +∫ ∞τg(ȳ(s), ᾱ(s))e −λs ds =∫ ∞g(ȳ(s), ᾱ(s))e −λs + e −λτ g(ȳ(s + τ), ᾱ(s + τ))e −λs ds ≥g(ȳ(s), ᾱ(s))e −λs + e −λτ v(ȳ(τ))(here, y x (s, ᾱ) is denoted <strong>for</strong> shortness as ȳ(s)).extreme terms of the inequality, we get0Passing to the infimum in thev(x) ≥ ¯v(x) (7.8)To prove the opposite inequality, we recall that ¯v is defined as an infimum, so that,<strong>for</strong> any x ∈ R d <strong>and</strong> ε > 0, there exists a control ᾱ ε (<strong>and</strong> the corresponding evolutionȳ ε ) such that¯v(x) + ε ≥∫ τ0g(ȳ ε (s), ᾱ ε (s))e −λs ds + e −λτ v(ȳ ε (τ)). (7.9)On the other h<strong>and</strong>, the value function v being also defined as an infimum, <strong>for</strong> anyx ∈ R d <strong>and</strong> ε > 0 there exists a control ˜α ε such thatusing (7.10) into (7.9), we get¯v(x) ≥∫ τ0v(ȳ ε (τ)) + ε ≥ Jȳε(τ)(˜α ε ) (7.10)g(ȳ ε (s), ᾱ ε (s))e −λs ds + e −λτ Jȳε(τ)(˜α ε ) − (1 + e −λτ )ε ≥≥ J x (̂α) − (1 + e −λτ )ε ≥≥ v(x) − (1 + e −λτ )ε, (7.11)where ̂α is a control defined bŷα(s) ={ᾱ ε (s) 0 ≤ s < τ˜α ε (s − τ) s ≥ τ.(7.12)Since ε is arbitrary, (7.11) finally yields ¯v(x) ≥ v(x).We note that this proof crucially relies on the fact that the control defined by(7.12) still belongs to A, being a measurable control. The possibility of obtaining


✐✐7.1. Optimal control problems: first examples 195an admissible control by joining together two different measurable controls is knownas concatenation property.The DPP can be used to characterize the value function in terms of a nonlinearpartial differential equation. In fact, let α ∗ ∈ A be the optimal control, <strong>and</strong> y ∗ theassociated evolution (to simplify, we are assuming that the infimum is a minimum).Then,that is,v(x) =∫ τ0g(y ∗ (s), α ∗ (s))e −λs ds + e −λτ v(y ∗ (τ)), (7.13)v(x) − e −λτ v(y ∗ (τ)) =∫ τ0g(y ∗ (s), α ∗ (s))e −λs ds (7.14)so that adding <strong>and</strong> subtracting e −λτ v(x) <strong>and</strong> dividing by τ, we gete −λτ (v(x) − v(y∗ (τ)))τ+ v(x)(1 − e−λτ )τ= 1 τ∫ τ0g(y ∗ (s), α ∗ (s))e −λs ds. (7.15)Assume now that v is regular. By passing to the limit <strong>for</strong> τ → 0 + , we havelim −v (y∗ (τ)) − v(x)= −Dv(x) · ẏ ∗ (x) = −Dv(x) · f(x, α ∗ (0))τ→0 + τlim v(x)(1 − e−λτ )= λv(x)τ→0 + τ∫1 τlim g(y ∗ (s), α ∗ (s))e −λs ds = g(x, α ∗ (0))τ→0 + τThen, we can conclude0λv(x) − Dv(x) · f(x, a ∗ ) − g(x, a ∗ ) = 0where a ∗ = α ∗ (0). Rewriting the DPP in the equivalent <strong>for</strong>m{ ∫ τ}v(x) + supα∈A− g(y(s), α(s))e −λs ds − e −λτ v(y(τ))0= 0,we obtain the <strong>Hamilton</strong>–Jacobi–Bellman equation (or Dynamic Programming equation)corresponding to the infinite horizon problem,λu(x) + sup {−f(x, a) · Du(x) − g(x, a)} = 0. (7.16)a∈ANote that, given x, the value of a achieving the max (assuming it exists) correspondsto the control a ∗ = α ∗ (0), <strong>and</strong> this makes it natural to interpret the argmax in (7.16)as the optimal feedback at x. We will get back to this point in the sequel.In short, (7.16) can be written aswith x ∈ R d , <strong>and</strong>H(x, u, Du) = 0 (7.17)H(x, u, p) = λu(x) + sup {−f(x, a) · p − g(x, a)} . (7.18)a∈ANote that H(x, u, ·) is convex (being the sup of a family of linear functions) <strong>and</strong>that, H(x, ·, p) is monotone (since λ > 0), so that we are in the framework of theexistence <strong>and</strong> uniqueness results presented in Chapter 2 (see Theorem 2.13).


✐✐196 Chapter 7. Control <strong>and</strong> Games7.1.2 Riccati equation <strong>and</strong> feedback controls <strong>for</strong> the<strong>Linear</strong>-Quadratic Regulator problemIn this problem, a linear (or linearized) controlled dynamical system must be stabilizedminimizing a linear combination of respectively the state <strong>and</strong> the controlenergy. The dynamics is given by:{ẏ(s) = Ay(s) + Bα(s)(7.19)y(0) = x,where A ∈ R d×d <strong>and</strong> B ∈ R d×m , x, y(s) ∈ R d <strong>and</strong> α(s) ∈ R m . The cost functionalis defined asJ x (α) = 1 2∫ +∞0(y(s; α) t Qy(s; α) + α(s) t Rα(s) ) e −λs ds (7.20)where Q ∈ R d×d <strong>and</strong> R ∈ R m×m are symmetric, strictly positive definite matrices,<strong>and</strong> λ ≥ 0. We are not interested in giving the most general framework of theproblem, but simply remark that it should be assumed that the system (7.19) iscontrollable, or that λ is large enough.We show now that this special structure of the problem allows to derive an explicitexpression <strong>for</strong> the optimal control in feedback <strong>for</strong>m, as α(s) = Ky(s) where K isan m × d matrix defined via the solution of the so-called Riccati equation.To write the Bellman equation of the <strong>Linear</strong>-Quadratic Regulator (LQR), weassume that the value function has a quadratic structure:v(x) = 1 2 xt P x,so that we also have Dv(x) = P x, <strong>for</strong> some matrix P ∈ R d×d to be determined.With this choice, (7.16) takes the <strong>for</strong>m{λ2 xt P x + max −(Ax + Ba) t P x − 1a2 xt Qx − 1 }2 at Ra = 0, (7.21)in which, given x, the argmax can be explicitly computed asUsing (7.22) into (7.21), we geta ∗ = −R −1 B t P x. (7.22)λ2 xt P x − x t A t P x − 1 2 xt Qx + 1 2 xt P t BR −1 B t P x = 0,that is,1 [( 2 xt λI − 2A t) P − Q + P t BR −1 B t P ] x = 0, (7.23)<strong>for</strong> any x ∈ R d . There<strong>for</strong>e, the unknown matrix P should solve the quadratic RiccatiequationP t BR −1 B t P + ( λI − 2A t) P − Q = 0.Once computed the matrix P , the optimal control is obtained as a function of thecurrent state y(s) by means of (7.22), <strong>and</strong> more preciselyα ∗ (s) = −R −1 B t P y ∗ (s). (7.24)


✐✐7.2. Dynamic Programming <strong>for</strong> other classical control problems 197To check that (7.24) is indeed the optimal control, define the functionG(s; α) = e−λs2(y(s) t Qy(s) + α(s) t Rα(s) ) + e −λs y(s) t P (Ay(s) + Bα(s)) −− λe−λsy(s) t P y(s).2Now, substituting the expression (7.24) <strong>for</strong> α ∗ , <strong>and</strong> using the Riccati equation <strong>and</strong>some elementary algebra, we can show thatOn the other h<strong>and</strong>, we also haveG(s; α) = e−λs2G(s; α ∗ ) = 0, <strong>for</strong> any s ∈ R + . (7.25)(y(s) t Qy(s) + α(s) t Rα(s) ) + d ds[e −λs v(y(s)) ] ,so that, using the control α ∗ <strong>and</strong> integrating on [0, T ], we get∫1 T(y ∗ (s) t Qy ∗ (s) + α ∗ (s) t Rα ∗ (s) ) e −λs ds + e −λT v(y ∗ (T )) − v(x).2 0Passing to the limit <strong>for</strong> T → ∞ <strong>and</strong> taking into account (7.25), we obtain that, inthe class of controls <strong>for</strong> which |y(s)| 2 = o(e λs ),12∫ ∞0(y ∗ (s) t Qy ∗ (s) + α ∗ (s) t Rα ∗ (s) ) e −λs ds = v(x),that is, the pair (y ∗ , α ∗ ) is optimal.The results above show the relationship between the max operator appearingin the Bellman equation, <strong>and</strong> the feedback <strong>for</strong>m of the optimal control. Theyalso explain the popularity of the <strong>Linear</strong>-Quadratic Regulator problem in controlliterature: the optimal feedback control can be obtained via the solution of a matrixequation <strong>and</strong> has an explicit <strong>for</strong>m. Actually, in many real problems of stabilization,the first try is to linearize the dynamics <strong>and</strong> approximate the cost by quadraticterms in order to directly apply the solution of the Riccati equation to the linearizedproblem.7.2 Dynamic Programming <strong>for</strong> other classical controlproblemsGiven a controlled dynamical system of the <strong>for</strong>m (7.1)–(7.2), it is possible to considerother optimal control problems with different <strong>for</strong>ms of the cost functional.We briefly review other classical problems <strong>and</strong> give their corresponding Bellmanequations.7.2.1 The finite horizon problemAssume the control system has the <strong>for</strong>m (7.2), with t 0 = t, x 0 = x. In the Bolza<strong>for</strong>mulation, the functional of the finite horizon problem has the <strong>for</strong>mJ x,t (α) =∫ Ttg(y x,t (s; α), α(s))e −λs ds + e −λ(T −t) ψ(y x,t (T ; α)) (7.26)


✐✐198 Chapter 7. Control <strong>and</strong> Gamesin which t ∈ [0, T ) <strong>and</strong> λ ≥ 0. For the finite horizon problem we can follow thesame arguments already seen <strong>for</strong> the infinite horizon problem. The first step is thedefinition of the value functionv(x, t) = infα∈A J x,t(α) (7.27)which in turn can be proved to satisfy the following DPP:Proposition 7.3 (DPP <strong>for</strong> the finite horizon problem). Under the assumptionsof Theorem 7.1, <strong>for</strong> all x ∈ R d <strong>and</strong> t < τ ≤ T ,{∫ τv(x, t) = infα∈A t}g(y x,t (s; α), α(s))e −λs ds + e −λ(τ−t) v(y x,t (τ; α), τ) (7.28)Then, in a similar way, it is possible to derive the Bellman equation in the<strong>for</strong>m of an evolutive <strong>Hamilton</strong>–Jacobi equation with a terminal condition:⎧⎨−u t + λu + sup{−f(x, a) · Du − g(x, a)} = 0 (x, t) ∈ R d × (0, T ),a∈A(7.29)⎩u(x, T ) = ψ(x) x ∈ R d .We refer the interested reader to [E98] p. 557 <strong>for</strong> a complete proof.7.2.2 The optimal stopping problemIn this problem, beside the usual control α, it is possible to stop the dynamics atany intermediate time θ ∈ [0, +∞] (respectively, θ ∈ [t, T ] in the finite horizon case),paying a corresponding final cost given by ψ(y(θ)) (respectively, ψ(y(T ∧ θ))). Inthe infinite horizon case, the cost functional defining the optimal stopping problemis there<strong>for</strong>e given byJ x (θ, α) =∫ θ0g(y x (s; α), α(s))e −λs ds + e −λθ ψ(y(θ; α)). (7.30)Then, <strong>for</strong> x ∈ R d , θ ∈ [0, +∞] <strong>and</strong> α ∈ A, the value function is defined asv(x) = infθ,α J x(θ, α). (7.31)This problem can be rephrased as a st<strong>and</strong>ard infinite horizon (respectively, finitehorizon) problem by adding to the control set A a new control ā, such that f(y, ā) =0 <strong>and</strong> g(y, ā) ≡ 0 <strong>for</strong> any y ∈ R d . A more explicit way of treating the problem is toderive a DPP which takes into account the option of stopping at a finite time, thenobtain the Bellman equation, which has the <strong>for</strong>m of an obstacle problem:max(H(x, u, Du), u − ψ) = 0, (7.32)where the <strong>Hamilton</strong>ian function H is defined as in (7.16). In (7.32), the set on whichu = ψ corresponds to the so-called stopping set, in which the optimal strategy isprecisely to stop the system.


✐✐7.2. Dynamic Programming <strong>for</strong> other classical control problems 1997.2.3 The minimum time problemAnother classical example is the minimum time problem. We assume a dynamicsin the <strong>for</strong>m (7.2), which has to be steered in the shortest possible time to a targetset T ⊂ R d . This set is given <strong>and</strong> we will always assume thatThe target set T is closed <strong>and</strong> int(T ) ≠ ∅, ∂T is sufficiently regular. (7.33)For every control α ∈ A we can define the first time of arrival on the target T as{t x (α) = inf s ∈ R + : y x (s; α) ∈ T }sprovided the set at the right-h<strong>and</strong> side in nonempty. If the set is empty, that is, ify x (s; α) /∈ T <strong>for</strong> any s > 0, it is natural to define t x (α) = +∞. This leads to thefollowing definition:{+∞ if yx (s; α) /∈ T ∀s > 0t x (α) = {inf s ∈ R + : y x (s; α) ∈ T } (7.34)elsesSince we want to minimize the time of arrival on the target, the value function <strong>for</strong>this problem is given byT (x) = inf t x(α) (7.35)α∈AA basic example, discussed in Chapter 1, is the case in whichẏ(s) = α(s)with α(s) ∈ B(0, 1). In this case, the minimum time T (x) equals the distanced(x, T ) <strong>and</strong> can be a nonsmooth function whenever T is nonconvex.Dynamic Programming <strong>for</strong> the minimum time problem Consider a system drivenby (7.2). Since the the minimum time function (i.e. the value function <strong>for</strong> thisproblem) may not be finite everywhere, we need to define its domain.Definition 7.4 (Reachable set). The reachable set is defined as the set of initialstates from which it is possible to reach the target, that is,R := { x ∈ R d : T (x) < +∞ } .The reachable set clearly depends on the target, on the dynamics <strong>and</strong> on theset of admissible controls. However, this dependence is complex even in simplesituations, so that the reachable set cannot be regarded as a datum in our problem.This means that we have to determine the pair (T, R) as the solution of a freeboundary problem.In order to derive the HJB equation <strong>for</strong> the minimum time problem, we state therelated Dynamic Programming Principle.Proposition 7.5 (DPP <strong>for</strong> the minimum time problem). For all x ∈ R <strong>and</strong>τ ∈ [0, T (x)) (so that x /∈ T ),T (x) = infα∈A {τ + T (y x(τ; α))}. (7.36)


✐✐200 Chapter 7. Control <strong>and</strong> GamesThe proof of this DPP relies basically on the same arguments used <strong>for</strong> theinfinite horizon case, <strong>and</strong> uses the possibility of generating new admissible controlsby concatenation <strong>and</strong>/or translation. To derive the <strong>Hamilton</strong>–Jacobi–Bellmanequation from the DPP, we rewrite (7.36) <strong>for</strong> t ∈ (0, T (x)) as<strong>and</strong> divide by τ, obtainingsupα∈AT (x) − infα∈A T (y x(τ; α)) = τ{ }T (x) − T (yx (τ; α))= 1.τIn order to pass to the limit as τ → 0 + we assume, <strong>for</strong> the moment, that T isdifferentiable at x <strong>and</strong> that lim τ→0 + commute with sup α . Provided ẏ x (0; a) exists,we getsup {−DT (x) · ẏ x (0, α)} = 1,α∈Aso that, if lim τ→0 + α(τ) = a, we obtainsup {−DT (x) · f(x, a)} = 1, (7.37)a∈Awhich is the <strong>Hamilton</strong>–Jacobi–Bellman equation associated to the minimum timeproblem. This derivation of the HJB equation basically shows that if T is regular,then it is a classical solution of (7.37), i.e., proves the following proposition:Proposition 7.6. If T ∈ C 1 in a neighborhood of x ∈ R\T , then T satisfies (7.37)at x.However, (7.37) needs to be complemented with a boundary condition, the“natural” condition beingThere<strong>for</strong>e, once defined the <strong>Hamilton</strong>ianT (x) = 0 (x ∈ ∂T )H(x, p) := sup{−p · f(x, a)} − 1,a∈Awe can rewrite (7.37) as a boundary value problem in the compact <strong>for</strong>m{H(x, DT (x)) = 0 x ∈ R \ T ,T (x) = 0 x ∈ ∂T .(7.38)Note again that H(x, ·) is convex since it is the sup of linear operators.We have already seen that T may not be differentiable even in the simplestcases. In addition, the continuity of T around the target may also fail to hold,unless the following property of Small-Time Local Controllability (STLC) aroundT is assumed.Definition 7.7 (Small Time Local Controllability). Assume ∂T is smooth.Then, the property of STLC is satisfied if, <strong>for</strong> any x ∈ ∂T , there exists a controlvector â ∈ A such that:f(x, â) · η(x) < 0,


✐✐7.2. Dynamic Programming <strong>for</strong> other classical control problems 201where η(x) is the exterior normal to ∂T at x.The STLC ensures that R is an open subset of R d <strong>and</strong> that T is continuousup to ∂T . Moreover, <strong>for</strong> all z ∈ ∂R,lim T (x) = +∞.x→zWith this last assumption, it is possible to prove that T ( · ) is a viscosity solutionof the boundary value problem (7.38).Theorem 7.8. If R \ T is open <strong>and</strong> T is continuous, then T is a viscosity solutionof (7.38).Proof. For τ small enough, by the DPP we have that, <strong>for</strong> any ᾱ ∈ A:<strong>and</strong> that, given ε > 0, there exists an α ε ∈ A such that:T (x) − T (ȳ(τ)) ≤ τ (7.39)T (x) − T (y ε (τ)) ≥ t(1 − ε), (7.40)where, as usual, ȳ(τ) = y x (τ; ᾱ) <strong>and</strong> y ε (τ) = y x (τ; α ε ).Consider now a test function φ ∈ C 1 (R d \ T ) such that T − φ has a local maximumat x. Then, <strong>for</strong> τ small enough, we have<strong>and</strong> by (7.39):T (x) − φ(x) ≥ T (ȳ(τ)) − φ(ȳ(τ)),φ(x) − φ(ȳ(τ)) ≤ T (x) − T (ȳ(τ)) ≤ τ,so that, dividing by τ <strong>and</strong> passing to the limit <strong>for</strong> τ → 0 + (as it has been done inthe derivation of (7.37)), we getH(x, ∇φ(x)) ≤ 0.Similarly, if φ ∈ C 1 <strong>and</strong> T − φ has a local minimum at x, <strong>for</strong> τ small enough wehaveT (x) − φ(x) ≤ T (y ε (τ)) − φ(y ε (τ)),<strong>and</strong> by (7.40):φ(x) − φ(y ε (τ)) ≥ T (x) − T (y ε (τ)) ≥ τ(1 − ε).Since ε is arbitrary, dividing by τ <strong>and</strong> passing to the limit, we obtainH(x, ∇φ(x)) ≥ 0,which fulfills the definition of viscosity solution.Using STLC property, it is also possible to prove that the boundary condition issatisfied on ∂T . This part of the proof will be skipped.A different approach to the solution of (7.37) is to per<strong>for</strong>m a suitable rescalingof T , obtained via the auxiliary function{1 if T (x) = +∞v(x) :=1 − e −µT (x) (7.41)else,


✐✐202 Chapter 7. Control <strong>and</strong> Games<strong>for</strong> some positive µ. It is easy to check that v is itself a value function, that is,where J x is given byv(x) = infα∈A J x(α),J x (α) =∫ tx(α)0e −µs ds<strong>and</strong> t x (α) is defined by (7.34). This control problem is nothing but an infinitehorizon problem (with constant running cost g(x, a) ≡ 1) in which the state of thesystem is stopped as soon as it reaches the target T , with zero stopping cost. TheDynamic Programming equation is in the <strong>for</strong>m of a boundary value problem:⎧⎨µv(x) + sup{−Dv · f(x, a) − 1} = 0 x ∈ R d \ Ta∈A(7.42)⎩v(x) = 0 x ∈ ∂T .The change of variable (7.41) is called Kružkov trans<strong>for</strong>mation <strong>and</strong> gives severaladvantages. First, v(x) takes values in [0, 1] (whereas T is generally unbounded)<strong>and</strong> this helps both in the analysis <strong>and</strong> in the numerical approximation. Second,the <strong>for</strong>mulation (7.42) avoids any reference to the reachable set. Once obtained v,the minimum time T <strong>and</strong> reachable set R can be recovered by the relationshipsT (x) = −ln(1 − v(x)), R = {x ∈ R d : v(x) < 1}.µIn addition, the <strong>for</strong>mulation as an infinite horizon problem allows the time-marchingnumerical scheme to be a contraction mapping. We will come back to this point inthe section on numerical schemes.7.3 The link between Bellman equation <strong>and</strong>Pontryagin Maximum PrincipleThere exists an important link between the DPP <strong>and</strong> the Pontryagin MaximumPrinciple (PMP in the sequel), which gives a set of necessary conditions to characterizethe optimal open-loop solution (y ∗ (·), α ∗ (·)). Using this relationship, wederive here the PMP <strong>for</strong> the finite horizon problem (7.1), (7.26) under additionalsimplifying assumptions on the data. In particular, we will assume that ψ ∈ C 2 (R d ),that Df is continuous, <strong>and</strong>, above all, that v ∈ C 2 (R d ) so that the <strong>Hamilton</strong>–Jacobiequation (7.29) is verified in a classical sense. We also assume, <strong>for</strong> simplicity, thatg ≡ 0 <strong>and</strong> λ = 0 so that, by the DPP, the functionh(s) := v(y(s; α), s)is nondecreasing <strong>for</strong> every α, <strong>and</strong> is constant only <strong>for</strong> the optimal trajectory. Then,an optimality condition <strong>for</strong> a pair (y ∗ , α ∗ ) is that, <strong>for</strong> any s ∈ [0, T ]:h ′ (s) = f(y ∗ (s), α ∗ (s)) · Dv(y ∗ (s), s) + v s (y ∗ (s), s) = 0. (7.43)On the other h<strong>and</strong>, since at any point (z, s) ∈ R d × [0, T ] the <strong>Hamilton</strong>–Jacobi–Bellman equation gives−v t (z, s) − f(z, a) · Dv(z, s) ≤ 0 <strong>for</strong> every a ∈ A, (7.44)


✐✐7.3. The link between Bellman equation <strong>and</strong> Pontryagin Maximum Principle 203setting z = y ∗ (s), by (7.43)–(7.44) we obtain an equivalent condition known as theMinimum Principle:f(y ∗ (s), α ∗ (s)) · Dv(y ∗ (s), s) = mina∈A f(y∗ (s), a) · Dv(y ∗ (s), s). (7.45)Now, in order to prove the PMP, we recall that, by (7.43), the left-h<strong>and</strong> side of (7.44)vanishes at (z, s) = (y ∗ (s), s) <strong>for</strong> a = α ∗ (s) if the control α ∗ is optimal. This impliesthat the trajectory y ∗ (s) maximizes the map z → −v t (z, s) − Dv(z, s) · f(z, α ∗ (s))<strong>and</strong> that, due to our regularity assumptions, we can differentiate this map withrespect to z obtaining−Dv t − J f Dv − D 2 vf = 0, (7.46)in which J f is the d × d Jacobian matrix <strong>for</strong> f <strong>and</strong> D 2 v is the d × d Hessian matrix<strong>for</strong> v. Let us now define the co-state or adjoint vector p(t) as the vectorp(s) := Dv(y(s)), s). (7.47)Differentiating with respect to s we obtain, by (7.46) <strong>and</strong> the regularity of v,p ′ = D 2 vf − Dv t = −J f p. (7.48)This means that the co-state associated with (y(s), s) <strong>and</strong> α ∗ can be redefined,without making reference to the value function, as the unique solution on [0, T ] ofthe system of linear differential equationswith the terminal conditionp ′ (s) = −J f (y(s), α(s))p(s), (7.49)p(T ) = Dψ(y(T )) (7.50)The necessary optimality conditions on the couple (y ∗ (t), α ∗ (t)) are then written as−p ∗ (t) · f(y ∗ (t), α ∗ (t)) = max{−p(t) · f(y(t), a)} (7.51)a∈Av t (y ∗ (t), T − t) = p ∗ (t) · f(y ∗ (t), α ∗ (t)) = −H(y ∗ (t), p ∗ (t)) ?? (7.52)In a remarkably technical way, Pontryagin’s Maximum Principle can also be derivedin a more general setting which, in particular, does not require the regularityassumptions on the data <strong>and</strong> on the value function.Theorem 7.9. Assume v ∈ C 2 (R d ), A compact, ψ ∈ C 2 (R d ), f differentiable withrespect to x <strong>and</strong> Df continuous, Let y(·) = y x0 (·; α) <strong>and</strong> let the co-state p(·) be thecorresponding solution of (7.49)-(7.50). Then, the open-loop control α ∗ is optimal<strong>for</strong> (x, T ) if <strong>and</strong> only if <strong>for</strong> almost any t ∈ (0, T ]:−p ∗ (t) · f(y ∗ (t), α ∗ (t)) = max{−p(t) · f(y(t), a)} := H(y(t), p(t))a∈A(p ∗ (t), H(y ∗ (t), p ∗ (t))) ∈ D + v(y(t), T − t)We refer the interested reader to [BCD97] <strong>for</strong> a more general proof of thisresult.


✐✐204 Chapter 7. Control <strong>and</strong> Games7.4 SL schemes <strong>for</strong> optimal control problemsThe construction of SL schemes <strong>for</strong> Dynamic Programming equations follows thesame guidelines shown in Chapters 5–6 <strong>for</strong> the simpler model problems. In fact,equations (5.70) or (5.128) may be seen as special cases of HJB equations <strong>for</strong> controlproblems in which⎧⎪⎨ f(x, a) = a,g(x, a) = H⎪⎩∗ (a),(7.53)ψ(x) = u 0 (x)(with the minor difference that (5.70) is a <strong>for</strong>ward, instead of a backward, equation).When considering more general DP equations, it should be clear that optimaltrajectories play the role of characteristics, but they are no longer expected to bestraight lines. There<strong>for</strong>e, in a sense, time discretization must combine strategiesused in both advection <strong>and</strong> convex HJ equations.For all these reasons, as well as <strong>for</strong> the sake of obtaining explicit error estimates,we are led to consider time discretization as an intermediate (semi-)discretizationstep. This step also has the meaning of approximating the original controlproblem with a problem in discrete time, <strong>and</strong> is in fact the situation in which theDP Principle has been first <strong>for</strong>mulated <strong>and</strong> applied.Next, the space discretization is per<strong>for</strong>med as in Chapter 5, unless <strong>for</strong> replacing theHopf–Lax <strong>for</strong>mula with the discrete version of the Dynamic Programming Principle.This section will review the construction <strong>and</strong> convergence analysis of SLschemes <strong>for</strong> HJB equations, with a special emphasis on the control problems examinedin Sections 7.1–7.2. For simplicity, we will consider schemes on the whole ofR d , <strong>and</strong> low-order (monotone) implementations. The treatment of boundary conditions,along with the construction of high-order versions, can be per<strong>for</strong>med withthe techniques discussed in Chapters 5–6.7.4.1 <strong>Approximation</strong> of the infinite horizon problemWe sketch the basic ideas in the case of the infinite horizon problem. The plan isto construct a time-marching scheme, starting with time discretization.Time discretization In order to give a discrete time approximation of the controlproblem, we fix a time step ∆t <strong>and</strong> set t m = m∆t (m ∈ N). The simplest wayto discretize (7.2) is by using the explicit Euler scheme, which corresponds to thefollowing discrete dynamical system:{y m+1 = y m + ∆tf(y m , a m )(7.54)y 0 = x.Here, the sequence of vectors a m ∈ A has the role of a (discrete) control. We willdenote by α ∆t = {a m } the sequence as a whole, <strong>and</strong> possibly identify this sequence(with a slight abuse of notation) with the continuous, piecewise constant controldefined byα ∆t (s) = a m , s ∈ [t m , t m + 1). (7.55)Whenever we will need to emphasize the dependence of the trajectory on x <strong>and</strong> α ∆twe will use the notation y m (x; α ∆t ).


✐✐7.4. SL schemes <strong>for</strong> optimal control problems 205Remark 7.10. It is also possible to implement higher order time discretizationsby using, e.g., Runge–Kutta type schemes in (7.54). We refer to [FF94] <strong>for</strong> a moreextensive study of this technique, whereas an example of application will be given inthe numerical tests section.Given the time discretization (7.54) <strong>for</strong> the controlled system, the correspondingdiscrete version of the cost functional J x may be obtained by a rectangle quadrature:Jx∆t (α∆t ) +∞∑:= ∆t g(y m , a m )e −λtm , (7.56)m=0in which the dependence on α ∆t also appears via the trajectory points y m .The discrete dynamics (7.54) <strong>and</strong> cost functional (7.56) define a discrete time controlproblem, whose value function isv ∆t (x) := inf J ∆tα ∆t x(α∆t ) . (7.57)Adapting the arguments of the continuous case, it is possible to prove a DiscreteDPP <strong>for</strong> the optimal control problem (7.54)–(7.56):Proposition 7.11 (DDPP <strong>for</strong> the infinite horizon problem). Fix ∆t > 0.Then, <strong>for</strong> all x ∈ R d <strong>and</strong> any positive integer ¯n:{}∑¯n−1v ∆t (x) = inf ∆t g(y m , a m )e −λtm + e −λt¯n v ∆t (y¯n ) . (7.58)α ∆t m=0The discrete time version of the SL scheme can be obtained by choosing ¯n = 1in (7.58). This gives an approximation in the fixed point <strong>for</strong>mv ∆t {(x) = min ∆tg(x, a) + e −λ∆t v ∆t (x + ∆tf(x, a)) } (7.59)a∈Ain which e −λ∆t is sometimes replaced by its first order Taylor approximation 1−λ∆t(this has no influence on the consistency rate, but introduces a constraint on ∆t toget a contraction in the right-h<strong>and</strong> side of (7.59)).Space discretization The space discretization of the semi-discrete approximation(7.59) is per<strong>for</strong>med in a completely st<strong>and</strong>ard way by replacing the value v ∆t (x +∆f(x, a)) by a P 1 or Q 1 interpolation. We end up with the fully discrete scheme{v j = min ∆tg(xj , a) + e −λ∆t I 1 [V ](x j + ∆tf(x j , a)) } . (7.60)a∈ADenoting the right-h<strong>and</strong> side of (7.60) as S(∆, V ), the scheme may be written asos also, in iterative <strong>for</strong>m, asV = S(∆, V )V (k+1) = S(∆, V (k) ) (7.61)where k is the iteration index. An analysis of the scheme along the guidelines ofChapter 5 shows that:


✐✐206 Chapter 7. Control <strong>and</strong> Games• S(∆, ·) is monotone;• S(∆, ·) is a contraction in l ∞ , <strong>and</strong> there<strong>for</strong>e V (k) converges towards a uniquesolution V ∈ l ∞ <strong>for</strong> any initial guess V (0) . The contraction coefficient is easilyshown to be L S = e −λ∆t <strong>and</strong> hence, as usual in time-marching schemes, convergencemay become very slow as ∆t → 0. Suitable acceleration techniques(exploiting monotonicity of the operator S) have been constructed to h<strong>and</strong>lethis problem (see [F87, F97] <strong>for</strong> more details).• The scheme is l ∞ stable if the running cost g is bounded. In fact, since theinterpolation I 1 is nonexpansive, from (7.60) we have, assuming that V (0) = 0:∥ ∥ ∥∥V (k+1) ∥∥∞≤ ∆t‖g‖ ∞ + e −λ∆t ∥∥V (k)∥ ≤∞≤ · · · ≤ ∆t‖g‖ ∞(1 + e −λ∆t + e −2λ∆t + · · · ) ≤≤∆t‖g‖ ∞1 − e −λ∆t = ‖g‖ ∞(1 + O(∆t))λConvergence analysis First, we point out that a first convergence analysis mightbe per<strong>for</strong>med by means of Barles–Souganidis theorem, the scheme being monotone.When passing from the Hopf–Lax <strong>for</strong>mula to the Dynamic Programming Principle,however, consistency analysis (if carried out with the arguments of Chapter 5)requires more caution.Rather than duplicating the analysis per<strong>for</strong>med in Chapter 5, we will applyhere the same arguments to estimate the rate of convergence of the approximatesolution to the value function, i.e., to the solution of the Bellman equation (7.16).At this level, it can be convenient to separate time from space discretization.Concerning time discretization, we first assume that v is Lipschitz continuous,<strong>and</strong> in addition, that(i) ‖f‖ is uni<strong>for</strong>mly bounded. Moreover, f(x, a) is linear with respect to a, i.e.,it has the structuref(x, a) = f 1 (x) + F 2 (x)awith f 1 : R d → R d <strong>and</strong> F 2 (x) is a d × m matrix;(ii) g(·, a) is Lipschitz continuous, <strong>and</strong> g(x, ·) is convex;(iii) There exists an optimal control α ∗ ∈ A (this could be derived from the twoassumptions above, but we state it explicitly).First, making explicit the minimum in (7.7) <strong>and</strong> (7.59), we can rewrite v <strong>and</strong>v ∆t asv(x) =∫ ∆t0g(y x (s; α ∗ ), α ∗ (s))e −λs ds + e −λ∆t v(y x (∆t; α ∗ )),v ∆t (x) = ∆tg(x, a ∗ ) + e −λ∆t v ∆t (x + ∆tf(x, a ∗ )).Then, we can bound v ∆t (x) − v(x) from above by using in the continuous problemthe constant control α(s) ≡ a ∗ (which is suboptimal), obtainingv(x) − v ∆t (x) ≤∫ ∆t0g(y x (s; a ∗ ), a ∗ )e −λs ds + e −λ∆t v(y x (∆t; a ∗ )) −−∆tg(x, a ∗ ) − e −λ∆t v ∆t (x + ∆tf(x, a ∗ )).


✐✐7.4. SL schemes <strong>for</strong> optimal control problems 207Now, by elementary approximation arguments,whereas∫ ∆t0g(y x (s; a ∗ ), a ∗ )e −λs ds − ∆tg(x, a ∗ ) = O ( ∆t 2) ,v(y x (∆t; a ∗ )) − v ∆t (x + ∆tf(x, a ∗ )) ≤ v(y x (∆t; a ∗ )) − v(x + ∆tf(x, a ∗ )) ++v(x + ∆tf(x, a ∗ )) − v ∆t (x + ∆tf(x, a ∗ )) ≤≤ O ( ∆t 2) + ∥ ∥ v − v∆t ∥ ∥∞,where the first term follows from the Lipschitz continuity of v <strong>and</strong> elementary results<strong>for</strong> the Euler scheme. We can there<strong>for</strong>e deduce thatv(x) − v ∆t (x) ≤ O ( ∆t 2) + e −λ∆t ∥ ∥ v − v∆t ∥ ∥∞. (7.62)Note that this unilateral estimation does not require assumptions (i) <strong>and</strong> (ii) above.To per<strong>for</strong>m the reverse estimate, we should replace the discrete optimal control a ∗by a suboptimal ā, so thatv ∆t (x) − v(x) ≤ ∆tg(x, ā) + e −λ∆t v ∆t (x + ∆tf(x, ā)) − (7.63)−∫ ∆t0g(y x (s; α ∗ ), α ∗ (s))e −λs ds − e −λ∆t v(y x (∆t; α ∗ )).A suitable choice <strong>for</strong> ā is to define it as the integral mean of α ∗ over [0, ∆t]:ā = 1 ∆t∫ ∆t0α ∗ (s)ds.In fact, with this choice we can write y ∗ (∆t) = y x (∆t; α ∗ ) asy ∗ (∆t) = x += x +On the other h<strong>and</strong>,∫ ∆t0∫ ∆t0∫ ∆t0f 1 (y ∗ (s))ds +∫ ∆t[f 1 (x) + O(∆t)]ds += x + ∆tf 1 (x) + F 2 (x)0∫ ∆t0F 2 (y ∗ (s))α ∗ (s)ds =∫ ∆t= x + ∆t[f 1 (x) + F 2 (x)ā] + O ( ∆t 2) =0[F 2 (x) + O(∆t)]α ∗ (s)ds =α ∗ (s)ds + O ( ∆t 2) == x + ∆tf(x, ā) + O ( ∆t 2) . (7.64)g(y ∗ (s), α ∗ (s))e −λs ds ==∫ ∆t0∫ ∆t0[g(x, α ∗ (s)) + O(∆t)](1 + O(∆t))ds =g(x, α ∗ (s))ds + O ( ∆t 2) ≥≥ ∆tg(x, ā) + O ( ∆t 2) , (7.65)where the last inequality follows from Jensen’s inequality, due to the convexity ofg(x, ·). Using (7.64)–(7.65) in (7.63), <strong>and</strong> retracing the proof of (7.62), we getv ∆t (x) − v(x) ≤ O ( ∆t 2) + e −λ∆t ∥ ∥ v − v∆t ∥ ∥∞, (7.66)


✐✐208 Chapter 7. Control <strong>and</strong> Gameswhich gives, taking into account (7.62) itself:∣ v ∆t (x) − v(x) ∣ ∣ ≤ O(∆t2 ) + e −λ∆t ∥ ∥ v − v∆t ∥ ∥∞. (7.67)Due to the assumptions on f <strong>and</strong> g, the O(∆t 2 ) term is uni<strong>for</strong>m with respect to x,so we can pass to the ∞-norm in (7.67), obtaining that, <strong>for</strong> some positive constantC: (1 − e−λ∆t ) ∥ ∥v ∆t − v ∥ ∥∞≤ C∆t 2 ,<strong>and</strong> there<strong>for</strong>e ∥ ∥v ∆t − v ∥ ∥∞≤ C∆t. (7.68)This estimate has been proved (with different techniques) in [CDI84], under theassumption that v is semiconcave. If v is only Lipschitz continuous, <strong>and</strong> assumptions(i)–(iii) above are dropped, then∥ v ∆t − v ∥ ∥∞≤ C∆t 1/2 . (7.69)Note that a sufficient condition <strong>for</strong> the Lipschitz continuity of v is to have Lipschitzcontinuous data (f <strong>and</strong> g) <strong>and</strong> λ > L f .For the fully discrete scheme (7.60), we estimate the space discretization error comparingwith the same technique above v j <strong>and</strong> v ∆t (x j ). Using the control a ∗ , optimal<strong>for</strong> v ∆t , <strong>and</strong> denoting z ∗ j = x j + ∆tf(x j , a ∗ ), we obtain the one-sided estimate:v j − v ∆t (x j ) ≤ ∆tg(x j , a ∗ ) + e −λ∆t I 1 [V ](z ∗ j )) −−∆tg(x j , a ∗ ) + e −λ∆t v ∆t (z ∗ j )) ≤≤ e −λ∆t ∣ ∣ I1 [V ](z ∗ j ) − I 1[V∆t ] (z ∗ j ) ∣ ∣ ++e −λ∆t ∣ ∣ I1[V∆t ] (z ∗ j ) − v ∆t (z ∗ j ) ∣ ∣ ≤≤ e −λ∆t ∥ ∥V − V ∆t∥ ∥∞+ O(∆x),where V ∆t is the vector of samples of v ∆t <strong>and</strong>, by the Lipschitz continuity of v ∆t ,we have bounded its interpolation error with O(∆x). Moreover, we have used themonotonicity of the reconstruction I 1 to bound the error at zj∗ with the error onthe nodes.Giving the opposite bound <strong>and</strong> passing to the ∞-norm, we obtain∥ V ∆t − V ∥ ∞≤ C ∆x∆t . (7.70)Combining time <strong>and</strong> space discretization error estimates, we have the completeestimate:|v j − v(x j )| ≤ C 1 ∆t γ ∆x+ C 2 (7.71)∆twhere γ = 1/2 or γ = 1, depending on the assumptions. In the <strong>for</strong>mer case, thebest coupling is obtained <strong>for</strong> ∆x = ∆t 3/2 , in the latter <strong>for</strong> ∆x = ∆t 2 .7.4.2 <strong>Approximation</strong> of the finite horizon problemLet us briefly review the results (proved in [FG98]) <strong>for</strong> the finite horizon problem.Time discretization uses the time step ∆t = T/N, the time horizon of the problemis [t n , t N ] = [n∆t, T ] <strong>and</strong> the discrete dynamical system is defined by{y m+1 = y m + ∆tf(y m , a m )(7.72)y n = x,


✐✐7.4. SL schemes <strong>for</strong> optimal control problems 209whereas the cost functional is discretized byJx,t ∆t (nα∆t ) N−1∑:= ∆t g(y m , a m )e −λtm + e −λ(t N −t n) ψ(y N ), (7.73)m=nin which again the dependence on the discrete control appears also through thepoints y m . The value function <strong>for</strong> the discrete time problem is naturally defined as<strong>and</strong> satisfies the corresponding DDPP:v ∆t,n (x) := infα ∆t J ∆tx,t n(α∆t ) (7.74)Proposition 7.12 (Discrete DPP <strong>for</strong> the finite horizon problem).∆t = T/N. Then, <strong>for</strong> all x ∈ R d , ¯n ∈ {n + 1, . . . , N},{}∑¯n−1v ∆t,n (x) = inf ∆t g(y m , a m )e −λtm + e −λ(t¯n−tn) v ∆t,¯n (y¯n ) . (7.75)α ∆t m=nSetThe discrete time version of the SL scheme is obtained setting ¯n = n + 1 in(7.75) <strong>and</strong> en<strong>for</strong>cing the final condition at time T , that is{v ∆t,n (x) = min a∈A{∆tg(x, a) + e −λ∆t v ∆t,n+1 (x + ∆tf(x, a)) }v ∆t,N (x) = ψ(x),(7.76)so that, as <strong>for</strong> the continuous problem, the solution is obtained backward in timefrom t N = T to t 0 = 0. Last, a P 1 or Q 1 interpolation is applied to reconstruct thevalue v ∆t,n+1 at the right-h<strong>and</strong> side of (7.76), giving the fully discrete scheme{ {vnj = min ∆tg(xj , a) + e −λ∆t [I 1 Vn+1 ] (x j + ∆tf(x j , a)) } ,a∈Avj N = ψ(x j ).(7.77)Convergence analysis We outline the main results of convergence <strong>for</strong> the finitehorizon problem. First, the scheme is monotone <strong>and</strong> consistent, <strong>and</strong> hence convergentby the Barles–Souganidis theorem. Concerning error bounds, if the solutionv of the Bellman equation (7.28) is Lipschitz continuous, then the discrete timeapproximation satisfy the estimate‖v ∆t − v‖ ∞ ≤ C∆t 1/2 (7.78)where C is a positive constant which in general depends on T . Note that Lipschitzcontinuity of the data f <strong>and</strong> g imply again Lipschitz continuity of the value function.For the fully discrete scheme (7.77) we have (see [FG98]):|vj n − v(x j , t n )| ≤ C 1 ∆t 1/2 ∆x+ C 2 , (7.79)∆t1/2 <strong>and</strong> in this case the best coupling is obtained by choosing ∆x = ∆t.


✐✐210 Chapter 7. Control <strong>and</strong> Games7.4.3 <strong>Approximation</strong> of the optimal stopping problemIn the (infinite horizon) optimal stopping problem, it is possible to use the argumentsgiven in the continuous <strong>for</strong>mulation to derive the time discrete approximation(v ∆t {(x) = min ∆tg(x, a) + e −λ∆t v ∆t (x + ∆tf(x, a)) } ), ψ(x) . (7.80)mina∈AIn (7.80) it is possible to recognize that, in addition to the continuous control, thecontroller has the further choice of stopping the system, in which case a price ψis paid. The optimal strategy clearly corresponds to the lower payoff, this beingselected by the outer min operator.In the fully discrete version, the value of v ∆t is replaced as usual with a P 1 or Q 1interpolation, <strong>and</strong> there<strong>for</strong>e the scheme is written as()v j = minmina∈A{∆tg(xj , a) + e −λ∆t I 1 [V ](x j + ∆tf(x j , a)) } , ψ(x j ). (7.81)Denoting again by S(∆, V ) the right-h<strong>and</strong> side of (7.60), <strong>and</strong> by Ψ the vector ofsamples of the function ψ, the scheme can be put in the compact <strong>for</strong>mwhich allows <strong>for</strong> the iterative versionV = min (S(∆, V ), Ψ) , (7.82)V (k+1) = minAn equivalent <strong>for</strong>m of the scheme (7.82) is also()S(∆, V (k) ), Ψ .max (V − S(∆, V ), V − Ψ) = 0,which, dividing the first argument by ∆t, may be rewritten as( )V − S(∆, V )max, V − Ψ = 0.∆tIn this <strong>for</strong>m, it is immediate to recognize that the scheme is consistent with (7.32).To check that (7.82) is monotone, assume that U ≥ W component by component.Then, the scheme is monotone ifNow, in the caseas well as in the casemin (S(∆, U), Ψ) − min (S(∆, W ), Ψ) ≥ 0. (7.83)min(S(∆, U), Ψ) = min(S(∆, W ), Ψ) = Ψ,min(S(∆, U), Ψ) = S(∆, U)min(S(∆, W ), Ψ) = S(∆, W ),condition (7.83) is clearly satisfied, S being monotone. On the other h<strong>and</strong>, ifmin(S(∆, U), Ψ) = S(∆, U)min(S(∆, W ), Ψ) = Ψ,


✐✐7.4. SL schemes <strong>for</strong> optimal control problems 211we can write<strong>and</strong> henceS(∆, W ) ≥ min(S(∆, W ), Ψ) = Ψ,min (S(∆, U), Ψ) − min (S(∆, W ), Ψ) = S(∆, U) − Ψ ≥≥ S(∆, U) − S(∆, W ) ≥ 0so that (7.83) is satisfied again. Since the same argument proves the symmetriccase, we can conclude that the scheme is also monotone <strong>and</strong> there<strong>for</strong>e convergentby the Barles–Souganidis theorem.The finite horizon optimal stopping can be treated in a similar way.7.4.4 <strong>Approximation</strong> of the minimum time problemIn the previous section, we have shown how the Kružkov change of variable (7.41)allows to work on the (rescaled) value function as the unique viscosity solution of(7.42). This will be the st<strong>and</strong>ing line of work <strong>for</strong> the numerical approximation ofthe minimum time problem.Time discretization In order to introduce the time discrete minimum time problem<strong>for</strong> the dynamics (7.54), we start by defining the time discrete analogue of thereachable set,R ∆t := { x ∈ R d : ∃α ∆t <strong>and</strong> m ∈ N such that y m(x; α∆t ) ∈ T } (7.84)<strong>and</strong> of the number of steps of first arrival,N ( {x, α ∆t) +∞ x /∈ R ∆t=(min{m ∈ N : y ) m x; α∆t∈ T } x ∈ R ∆t (7.85)The role of the minimum time is played now by the minimal number of stepsnecessary to reach the target starting at xN(x) := minα ∆t N ( x, α ∆t) , (7.86)so that the discrete approximation of the minimum time function T (x) is now∆tN(x). By an adaptation of the usual arguments, it is possible to prove thefollowingProposition 7.13 (DDPP <strong>for</strong> the minimum time problem). Let ∆t > 0 befixed. For all x ∈ R ∆t <strong>and</strong> 0 ≤ ¯n < N(x),N(x) = infα ∆t {¯n + N(yx(¯n; α∆t ))} . (7.87)However, rather then per<strong>for</strong>ming a time discretization on the basis of (7.87),we further apply the Kružkov change of variable (<strong>for</strong> simplicity, with µ = 1):v ∆t (x) = 1 − e −∆tN(x) . (7.88)


✐✐212 Chapter 7. Control <strong>and</strong> GamesNote that 0 ≤ v ∆t ≤ 1, as in the continuous case, <strong>and</strong> that v ∆t has constant valueon the set of initial points x which can be driven to T by the discrete dynamicalsystem in the same number of steps.Writing (7.87) <strong>for</strong> ¯n = 1, <strong>and</strong> changing variable, we characterize v ∆t as the solutionof{v ∆t {(x) = min e −∆t v ∆t (x + hf(x, a)) } + 1 − e −∆t x ∈ R d \ Ta∈A(7.89)v ∆t (x) = 0x ∈ Twhere the boundary condition on T stems from the definition of N(x, α ∆t ).Space discretization The space discretization of (7.89) is per<strong>for</strong>med as above bya P 1 or Q 1 interpolation. The iterative fully discrete scheme reads⎧[⎨v (k+1)j = e −∆t min{I 1 V (k)] }(x j + hf(x j , a)) + 1 − e −∆t x j ∈ R d \ Ta∈A⎩v (k+1)j = 0 x j ∈ T(7.90)where k is the iteration index. Unless <strong>for</strong> the second condition, which works as aboundary condition, the scheme has the same structure of (7.61), so that denotingthe right-h<strong>and</strong> side of (7.90) as S(∆, V (k) ), it can easily be shown that:• S(∆, ·) is monotone;• S(∆, ·) is a contraction in l ∞ (with contraction coefficient L S = e −∆t ), <strong>and</strong>there<strong>for</strong>e V (k) converges towards a unique solution V ∈ l ∞ <strong>for</strong> any initialguess V (0) . If V (0) is chosen as a numerical supersolution, then the sequenceis monotonically decreasing, <strong>and</strong> if in particularv (0)j ={0 x j ∈ T1 x j ∉ T ,then at the step k the scheme updates only the nodes which are driven to thetarget in precisely k steps. This lends itself to a fast implementation (thispoint will be further discussed in Chapter ??).Clearly, the presence of a boundary condition requires some extra ef<strong>for</strong>t intreating possibly complex geometries (e.g., by means of unstructured meshes asdiscussed in Chapter 3), as well as in implementing a variable step technique (asshown in Chapter 5). Moreover, when restricting to a bounded computationaldomain, a further, suitable boundary condition must be imposed on the externalboundary. We skip the problem here, <strong>and</strong> refer the reader to [F97] <strong>for</strong> a discussionof this point.Convergence analysis In order to prove an error bound <strong>for</strong> the discrete time approximationwe need to introduce a discrete analogue of the Small Time LocalControllability assumption. We define the δ-neighbourhood of ∂T asT δ := ∂T + δB(0, 1)(where, <strong>for</strong> shortness, d(x) = d(x, ∂T )). A first technical result states the followinglocal upper bound <strong>for</strong> the discrete time approximation around T :


✐✐7.4. SL schemes <strong>for</strong> optimal control problems 213Lemma 7.14. Let the assumptions of Theorem 7.1 be satisfied <strong>and</strong> let STLC holdtrue. Then, there exist positive constants ∆t, δ such that, <strong>for</strong> all ∆t < ∆t <strong>and</strong>x ∈ T δ :v ∆t (x) ≤ C d(x) + ∆t.We are now able to prove a convergence result:Theorem 7.15. Let the assumptions of Lemma 7.14 be satisfied <strong>and</strong> let T becompact with nonempty interior.Then, v ∆t converges to v locally uni<strong>for</strong>mly in R d <strong>for</strong> ∆ → 0 +Proof. Since v ∆t is a discontinuous function, we define the two semicontinuousenvelopesv = lim inf v ∆t (y), v = lim sup v ∆t (y)h→0 +h→0y→x+y→xNote that v (respectively, v) is lower (respectively, upper) semicontinuous. The firststep is to show that1. v is a viscosity subsolution <strong>for</strong> (7.37)2. v is a viscosity supersolution <strong>for</strong> (7.37)Then, we want show that both the envelopes satisfy the boundary condition on T .In fact, by Lemma 7.14,v ∆t (x) ≤ Cd(x) + ∆twhich impliesThere<strong>for</strong>e, we have|v| ≤ Cd(x) (7.91)|v| ≤ Cd(x). (7.92)v = v = 0 on ∂T .Since the two envelopes coincide on ∂T we can apply the comparison theorem <strong>for</strong>semicontinuous sub <strong>and</strong> supersolutions obtainingv = v = v on R d .For the discrete time approximation, it is also possible to give an error estimate,restricted to a compact set Ω ⊂ R in which the following condition issatisfied:∃ C 0 > 0 : ∀ x ∈ Ω there exists a time optimal control withtotal variation less than C 0 bringing the system to T .The typical choice <strong>for</strong> Ω is to take an hypercube containing the target set T .(7.93)Theorem 7.16. Let the assumptions of Lemma 7.14 be satisfied, <strong>and</strong> let Ω be acompact subset of R where assumption (7.93) holds. Then, there exists two positiveconstants ∆t <strong>and</strong> C such that, <strong>for</strong> all ∆t ≤ ∆t <strong>and</strong> x ∈ Ω,|v ∆t (x) − v(x)| ≤ C∆t. (7.94)


✐✐214 Chapter 7. Control <strong>and</strong> GamesThe above results show that the rate of convergence <strong>for</strong> the scheme based onthe Euler discretization is 1, which is exactly what we expected. A complete proofcan be found in [BF90b]. The analysis of the fully discrete scheme leads to ana-priori error estimate hat we will not prove here( ) ) 2 ∆x‖v j − v(x j )‖ ∞ ≤ C∆t(1 1/2 +(7.95)∆tA sketch of that proof will be given in the section on differential games, since <strong>for</strong> asingle player that problem reduces to the minimum time problem.7.4.5 <strong>Approximation</strong> of optimal feedback <strong>and</strong> trajectories <strong>for</strong>control problemsOne of the main goals of numerical schemes <strong>for</strong> control problems is to computeapproximate optimal controls. As it has already been pointed out, the algorithmsproposed <strong>and</strong> analyzed in this chapter inherently compute an approximate optimalcontrol at every point of the grid. This in<strong>for</strong>mation will not require extra ef<strong>for</strong>t.However, this is not sufficient to compute accurately the optimal feedback in thecomputational domain Ω <strong>and</strong> to derive the corresponding optimal trajectories. Thenumerical solution v, which is computed at the nodes, is extended to to the wholedomain by interpolation, so that an approximate optimal feedback can be computedat every point x, i.e. <strong>for</strong> the control problem we can define the feedback mapF : Ω → A. In order to construct F , we first introduce the notationI ∆ (x, a) := e −λ∆t I[V ](x + hf(x, a)) + ∆tg(x, a). (7.96)where the suffix ∆ indicates that this function also depends on the discretization.Note that I ∆ (x, ·) has a minimum over A, but the minimum point may not beunique.The possible occurrence of multiple minimizers requires to construct a selection,e.g. to take a strictly convex φ <strong>and</strong> defineThe selection is defined byA ∗ (x) = {a ∗ ∈ A : I ∆ x (x, a ∗ ) = minA I∆ (x, a)} (7.97)a ∗ x = arg min A ∗ (x) φ(a) (7.98)We are now able to compute our approximate optimal trajectories, by defining thepiecewise constant controla ∆ (s) = a ∗ x n,hs ∈ [t n , t n+1 ) (7.99)where y n; ∆t is the state of the Euler scheme, at the iteration n. Error estimatesof the approximation of feedbacks <strong>and</strong> optimal trajectories are available <strong>for</strong> controlproblems in [BCD97, F01].7.4.6 Numerical tests <strong>for</strong> control problemsWhile one-dimensional, academic examples are intended to complement the theoreticalpresentation of chapters 4 <strong>and</strong> 5, this chapter rather collects examples <strong>and</strong> testsrelated either to applications or to more challenging benchmarks. Among them, wemention the treatment of singular solutions in more than one space dimension, <strong>and</strong>Dynamic Programming in dimension three <strong>and</strong> four.


✐✐7.4. SL schemes <strong>for</strong> optimal control problems 215W 1,∞ initial conditionn n L ∞ error L 1 error25 5.75 · 10 −2 6.57 · 10 −250 1.85 · 10 −2 2.76 · 10 −2100 6.17 · 10 −3 8.32 · 10 −3rate 1.61 1.49Table 7.1. Numerical errors <strong>for</strong> Test 1.Test 1: Quadratic <strong>Hamilton</strong>ian, Lipschitz continuous initial data This test refersto the HJ equation:{u t (x, t) + 1 2 |Du(x, t)|2 = 0 (x, t) ∈ [−2, 2] 2 × [0, 1]v(x, 0) = v 0 (x) = max(0, 1 − |x| 2 ) x ∈ [−2, 2] 2 (7.100)<strong>and</strong> is (along with the next example) a two-dimensional version of the test presentedin Chapter 5, which may be recast as a control problem via the definitions (7.53).The solution of this test problem generates a singularity in the gradient <strong>for</strong> t > 1/2.The exact solution of (7.100) can be explicitly computed <strong>and</strong> <strong>for</strong> t > 1/2 itreads{(|x|−1)2v(t, x) =2tif |x| ≤ 10 if |x| ≥ 1The solution, computed with ∆t = 0.1 <strong>and</strong> a cubic reconstruction, is shown in Figure7.1 at even time steps between t = 0 <strong>and</strong> t = 1. The graphs show in particularthat the singularity in the gradient is resolved isotropically <strong>and</strong> without undesirednumerical dispersion. Table 7.1 shows L ∞ <strong>and</strong> L 1 errors at T = 1 <strong>for</strong> differentnumbers of nodes n n on the edge of the computational domain. The solution iscomputed with good accuracy even with a low number of nodes, although therate of convergence is limited by the lack of semiconcavity, as it has already beenremarked in the one-dimensional examples of Chapter 5.Test 2: Quadratic <strong>Hamilton</strong>ian, semiconcave initial data We consider again theequation with quadratic <strong>Hamilton</strong>ian, but reversing the sign of the initial condition:{u t (x, t) + 1 2 |Du(x, t)|2 = 0 (x, t) ∈ [−2, 2] 2 × [0, 1]u(x, 0) = u 0 (x) = min(0, |x| 2 − 1) x ∈ [−2, 2] 2 (7.101).The exact solution of this test problem isu(t, x) ={|x|22t+1 − 1 if |x| ≤ √ 2t + 10 if |x| ≥ √ 2t + 1<strong>and</strong> generates no further singularity in the gradient. Figure 7.2 shows the numericalsolution computed <strong>for</strong> t ∈ [0, 1] with ∆t = 0.1 <strong>and</strong> a cubic reconstruction, <strong>and</strong> table7.2 shows L ∞ <strong>and</strong> L 1 errors at T = 1. In this case, due to the uni<strong>for</strong>m semiconcavityof the solution, the improvement in L ∞ error is of one order of magnitude, <strong>and</strong> inL 1 error (which is clearly measured in a weaker norm) of more then two orders ofmagnitude.


✐✐216 Chapter 7. Control <strong>and</strong> Games110.50.500-2-1.5-1-0.500.511.52 -2 -1.5-1 -0.50 0.5 1 1.5 2-2-1.5-1-0.500.511.52 -2 -1.5-1 -0.50 0.5 1 1.5 2(a)(b)110.50.500-2-1.5-1-0.500.511.52 -2 -1.5-1 -0.50 0.5 1 1.5 2-2-1.5-1-0.500.511.52 -2 -1.5-1 -0.50 0.5 1 1.5 2(c)(d)110.50.500-2-1.5-1-0.500.511.52 -2 -1.5-1 -0.50 0.5 1 1.5 2-2-1.5-1-0.500.511.52 -2 -1.5-1 -0.50 0.5 1 1.5 2(e)(f)Figure 7.1. Numerical solutions <strong>for</strong> Test 1Semiconcave initial conditionn n L ∞ error L 1 error25 1.57 · 10 −2 2.88 · 10 −250 4.03 · 10 −3 8.42 · 10 −4100 6.46 · 10 −4 3.29 · 10 −5rate 2.30 4.89Table 7.2. Numerical errors <strong>for</strong> Test 2.


✐✐7.4. SL schemes <strong>for</strong> optimal control problems 21700-0.5-0.5-1-1-2-1.5-1-0.500.511.52 -2 -1.5-1 -0.50 0.5 1 1.5 2-2-1.5-1-0.500.511.52 -2 -1.5-1 -0.50 0.5 1 1.5 2(a)(b)00-0.5-0.5-1-1-2-1.5-1-0.500.511.52 -2 -1.5-1 -0.50 0.5 1 1.5 2-2-1.5-1-0.500.511.52 -2 -1.5-1 -0.50 0.5 1 1.5 2(c)(d)00-0.5-0.5-1-1-2-1.5-1-0.500.511.52 -2 -1.5-1 -0.50 0.5 1 1.5 2-2-1.5-1-0.500.511.52 -2 -1.5-1 -0.50 0.5 1 1.5 2(e)(f)Figure 7.2. Numerical solutions <strong>for</strong> Test 2Test 3: the moon l<strong>and</strong>ing problem This is a classical control problem which istaken from [FR75]. A spacecraft is attempting to l<strong>and</strong> smoothly on a surface usingthe minimum amount of fuel. The gravity acceleration g is assumed to be constantnear the surface. The motion is described by the equations⎧⎪⎨ ḣ = v˙v = −g + m⎪⎩−1 α(7.102)ṁ = −kαwhere h <strong>and</strong> v denote respectively the height <strong>and</strong> vertical velocity of the spacecraft,m is the mass of the spacecraft (including the variable mass of the fuel), <strong>and</strong> the


✐✐218 Chapter 7. Control <strong>and</strong> Gamescontrol α ∈ [0, ̂α] denotes the thrust of the spacecraft’s engine. The initial time ist 0 = 0 <strong>and</strong> the final time T is the first time the spacecraft touches the moon. Theinitial <strong>and</strong> final conditions <strong>for</strong> this problem readh(0) = h 0 , v(0) = v 0 , m(0) = M + F (7.103)h(T ) = 0, v(T ) = 0 (7.104)where h 0 <strong>and</strong> v 0 are the initial height <strong>and</strong> velocity of the spacecraft, M <strong>and</strong> Fare mass of the spacecraft without fuel <strong>and</strong> the initial mass of the fuel. Setting(x 1 , x 2 , x 3 ) = (h, v, m), the problem of minimizing the fuel is equivalent to minimize−m(T ) = −x 3 (T ) over the class of measurable controls which drive the system tothe point x 1 (T ) = x 2 (T ) = 0 <strong>and</strong> satisfy α(t) ∈ A = [0, ̂α].Treating the final condition by penalization to remove the target constraint,the control problem is in the Mayer <strong>for</strong>mwhereJ x,t (α) = ψ(y x (T ; α)),ψ(x) = −x 3 + x2 1 + x 2 2(ɛ ≪ 1),ɛ<strong>and</strong> its value function solves the Bellman equation⎧⎨−v t (x, t) + sup{−f(x, α) · Dv(x, t)} = 0α∈A⎩v(x, T ) = ψ(x).The test is per<strong>for</strong>med using a P 2 reconstruction in the interpolation routine, <strong>and</strong>(following [FF94]) the Heun scheme to approximate the dynamics. In this versionof (7.77), the point x j + ∆tf(x j , a) is replaced by x j + ∆tΦ(∆t; x j , a 1 , a 2 ), whereΦ(∆t; x j , a 1 , a 2 ) = 1 2 [f(x j, a 1 ) + f(x j + ∆tf(x j , a 1 ), a 2 )] ,<strong>and</strong> the minimization is per<strong>for</strong>med with respect to both a 1 <strong>and</strong> a 2 . For each pointof the space-time grid the approximate optimal feedback is defined by taking thearg min a1,a 2in (7.77) <strong>and</strong> extending the two discrete controls to all the computationaldomain by interpolation.The test uses the initial condition (x 1 , x 2 , x 3 ) = (0.3, 0, 18), final time T = 1,space steps of (0.03, 0.15, 0.26) <strong>and</strong> a time step of ∆t = 0.03. Figure 7.3 shows theoptimal trajectory in the phase space (velocity, position), along with the discreteoptimal controls a 1 <strong>and</strong> a 2 during the time interval [0, 1]. The spacecraft actuallyl<strong>and</strong>s at t = 0.99, <strong>and</strong> the end point (x 1 , x 2 ) = (0, 0) is reached with some numericalerror, due to the penalization strategy. All figures show that the optimal trajectoryis basically obtained falling down by gravity up to a switching time t ∗ , <strong>and</strong> thenbraking until the l<strong>and</strong>ing with maximum thrust. This is in perfect agreement withtheory.Test 4: a Ramsey model <strong>for</strong> a multi-sectors economy Consider an economywhich has three sectors S i , i = 1, . . . , 3 (the model can be easily extended to manysectors) <strong>and</strong> denote by Y i (s), K i (s), L i (s) the output rate, the capital <strong>and</strong> the labourof sector S i at time s. Capital <strong>and</strong> labour are connected to the production through


✐✐7.4. SL schemes <strong>for</strong> optimal control problems 21930"controls1"30"controls2"25252020controls15controls15101055000 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1time(a)0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1(b)time0.3’trajectory’0.250.20.15height0.10.050-0.05-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0 0.1velocity(c)Figure 7.3. Moon L<strong>and</strong>ing, (a): optimal control a 1 , (b): optimal controla 2 , (c): optimal trajectory.a production function Y i = F i (K i , L i ). A classical example is the Cobb–DouglasfunctionF (K, L) = e λs K α L 1−α , (0 ≤ α ≤ 1) (7.105)where λ is a coefficient connected to the technical innovation (see [1]). For everysector, we can divide by L i <strong>and</strong> define y i = YiL i, k i = KiL i, i = 1, . . . , 3 obtaining thepro-capite production functionsy i = e λis k αii , (0 ≤ α i ≤ 1) (7.106)A typical choice is <strong>for</strong> the parameters λ i is λ i = m(1 − α i ) (see [1]), where m is aconstant rate of technical innovation.The rate of consumption c(s) is a fraction u (the control) of a function g(k(s))of the capital, i.e., <strong>for</strong> every sector the consumption rate is given byc i (s) = u(s)g i (k i (s)), (u min ≤ u(s) ≤ 1). (7.107)Then, the rate of change of the capital in every sector (assuming that the sectorsare separated) is˙k i = f(k i ) − ug(k i ), (u min ≤ u ≤ 1). (7.108)


✐✐220 Chapter 7. Control <strong>and</strong> Games’controls’10.80.60.40.200 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8Figure 7.4. Ramsey model, optimal control.Let now U be a utility function which represents the utility of consuming at ratec = c 1 + c 2 + c 3 > 0. This function U is typically monotone increasing <strong>and</strong> boundedon the positive axis. The goal is to maximize the functional∫ T0U(c(s))ds, (7.109)with respect to the choice of the control u under the constraints k i (0) = k 0 i , k i(T ) =k 1 i .Figures 7.4 <strong>and</strong> 7.5 show respectively the optimal control u <strong>and</strong> the evolutionin time of the variables k i <strong>and</strong> c i (i = 1, 2, 3). The various parameters have beenset as α 1 = 0.2, α 2 = 0.5 <strong>and</strong> α 3 = 0.8, u min = 0.1, m = 1, whereas g is defined as<strong>and</strong> the utility function U asg(k) = k − k min , (7.110)U(c) = Mc21 + c 2 , (7.111)with k min = k 0 /3 <strong>and</strong> M = 10. The initial condition <strong>for</strong> the three components ofthe capital is k 0 = (6, 4, 2) <strong>and</strong> the final objective <strong>for</strong> T = 1.8 is k 1 = (8, 6, 4). Asin the previous test, we treat the final constraint by penalization, using the finalcondition∣ ∣ k − k1 2ψ(k) = −ɛFor the computation, we have used ∆t = 0.05, ɛ = 10 −4 <strong>and</strong> 30 nodes in space <strong>for</strong>every dimension.Figures 7.4 <strong>and</strong> 7.5 depict a situation in which the capital increase, with consumptionat its minimum, up to a time t ≈ 1.2. After t, the control switches from0.1 to 1, the consumption go to its maximum <strong>and</strong> the first two component of the capitalstart decreasing. The final position of the trajectory is k(T ) ≈ (7.91, 7.41, 5.68),<strong>and</strong> significantly differs from the target, due to both penalization <strong>and</strong> lack of controllability(a one-dimensional control has been used <strong>for</strong> a system in R 3 ).Test 5: control of the HIV-1 dynamics In the paper [PN99] Perelson <strong>and</strong> Nelsonhave described the dynamics of the HIV-1 virus infection (the virus that causes


✐✐7.4. SL schemes <strong>for</strong> optimal control problems 221’trajectory_k_1’7’consumption_c_1’8.56857.543726.5160 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8(a)00 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8(d)8’trajectory_k_2’7’consumption_c_2’7.56756.54635.5524.5140 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8(b)00 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8(e)6’trajectory_k_3’5.5’consumption_c_3’5.554.5544.53.5342.53.5231.512.50.520 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8(c)00 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8(f)Figure 7.5. Ramsey model, optimal solution, (a)–(c): k i , (d)–(f): c i .AIDS). In this test, we add a control term in their model in order to describe thedrug therapy effect.The model considers three populations: the uninfected cells C, the productivelyinfected cells I <strong>and</strong> the virus V . All the populations are described by their density


✐✐222 Chapter 7. Control <strong>and</strong> Gamesper volume unit. The dynamics is described by the system⎧(⎪⎨ Ċ = s + pC 1 − C )− d C C − (1 − u)kV C,C maxI ˙ = (1 − u)kV C − d I I,⎪⎩˙V = Nd I I − d V V,(7.112)where s represents the rate at which new C cells are created from sources withinthe body, p is the maximum proliferation rate <strong>for</strong> the C cells, C max is the maximumcell population density, d C , d I <strong>and</strong> d V are the death rates <strong>for</strong> respectively the C,I <strong>and</strong> V populations, Nd I is the average rate of virion production, u is the controldriven by the drug therapy (usually, the constraint d C C max > s is added). Theterm kV C tells that the rate of creation of new infected cell is proportional to theproduct V C (at least one virion should enter a cell to infect it).A typical drug therapy is a treatment based on RT (reverse transcriptase) or proteaseinhibitors. The control u measures the effectiveness of the inhibitor: <strong>for</strong> u = 1the inhibition is 100% effective, whereas <strong>for</strong> u = 0 there is no inhibition. Typically,0 ≤ u ≤ u max < 1. The functional we want to minimize has the <strong>for</strong>m∫ T0(M(I 2 + V 2) + u ) dt (7.113)(<strong>for</strong> a positive constant M) <strong>and</strong> takes into account the dimension of the populationsI <strong>and</strong> V <strong>and</strong> the fact that a drug excess can have negative global effects on thepatient.In the test, parameters have been set as s = 0.2, p = 0.05, k = 0.08, d C = 0.01,d I = 0.3, d V = 0.009, N = 0.005 <strong>and</strong> M = 8 (a biologically realistic choice ofsome of the parameters can be found in [2]). Figure 7.6 shows the optimal controlu <strong>and</strong> the time evolution of the state variables C, I, V . The virus populationdecreases monotonically under the effect of the drug therapy, but since the use oflarge amounts of drug is penalized, the optimal therapy suggested by the model isto use the drug at its maximum level <strong>for</strong> the first period <strong>and</strong> then stop the druginjection. After this first time interval, the evolution of infected cells <strong>and</strong> virusesnaturally decreases by the uncontrolled dynamics.Tests 6–7: the minimum time problem In the first test, the system has a variablevelocity defined byf(x, a) = c(x) a = |x 1 + x 2 | a,with a ∈ B(0, 1). It has to be driven to the origin, <strong>and</strong> more precisely to the targetT = B(0, ε),where, in the numerical computation, we have set ε = ∆x/2.Figure 7.7 shows the approximate minimum time function T (x) <strong>and</strong> its levelscurves, whereas the optimal trajectories (not shown in the figure) are orthogonal tothe level curves. On the line x = −y the velocity drops to zero <strong>and</strong> we would haveT = +∞. Note that <strong>for</strong> points close to this line, the optimal trajectory first movesto a region in which the velocity is higher, then turns towards the target.


✐✐7.4. SL schemes <strong>for</strong> optimal control problems 223’controls’9.25’trajectory_C’19.20.80.69.150.49.10.29.0500 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1(a)90 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1(b)0.82’trajectory_I’0.202’trajectory_V’0.80.20.780.1980.760.1960.740.1940.720.1920.70.190.680.1880.660.1860.640 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1(c)0.1840 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1(d)Figure 7.6. HIV-1 dynamics, (a): optimal control, (b): uninfected cells,(c): infected cells, (d): virus.281.571650.5403220.510210122011.522 1.5 1 0.5 0 0.5 1 1.5 2Figure 7.7. Minimum time function T , graph (left) <strong>and</strong> contours (right).


✐✐224 Chapter 7. Control <strong>and</strong> Games1.51.5110.50.5000.50.511.51.5 1 0.5 0 0.5 1 1.5griglia= 100x100110griglia= 100x10011.5 1 0.5 0 0.5 1 1.5Figure 7.8. Level sets of T (left) <strong>and</strong> an optimal trajectory on the surfacez (right).The second test <strong>for</strong> the minimum time problem is related to the computationof geodesics on a nonsmooth surface. In this test, we set a point target:on the pyramidal surfacez(x) =T = (0, −0.6, 0.4){1 − (|x 1 | + |x 2 |) if |x 1 | + |x 2 | < 10 elsewhere.The system moves on the surface with intrinsic velocity in B(0, 1), so that theminimum time function (restricted to the surface) is in fact the intrinsic distance,<strong>and</strong> optimal trajectories are geodesics of the surface. In this test, we compute thegeodesic reaching the target from the starting point (0, 0.5, 0.5).It can be shown that the 3D problem can be reduced to a 2D problem bymodifying the velocity field according to the function z. In fact, in case of a unityintrinsic velocity on the surface, it can be shown (see [Se96, Se96b]) that the velocityof the corresponding 2D problem becomesc(x, a) =1√1 + (Dz · a)2 .Figure 7.8 shows the level sets of T <strong>and</strong> its surface with the optimal trajectory on it.7.5 Problems with state constraintsMany relevant applications in control theory require the state of the system to satisfyadditional constraints – typically, to remain within some prescribed set of the statespace R d . We turn now our attention to this case, trying to give a presentation ofthe general ideas while skipping the most delicate technical points. The interestedreader will find suitable references <strong>for</strong> state constrained problems in the referencesection at the end of this chapter.


✐✐7.5. Problems with state constraints 225Consider our st<strong>and</strong>ard dynamics (7.2) coupled with the infinite horizon costfunctional (7.5). In their most typical <strong>for</strong>m, the state constraints require that, <strong>for</strong>all s > 0,y x (s; α) ∈ Ω, (7.114)where we assume thatΩ is a bounded, open subset of R d . (7.115)While, in the case of free control problems, the cost J x is minimized over the setA of measurable controls taking values in A, condition (7.114) restricts admissiblecontrols to a new set which will depend on both the constraint Ω <strong>and</strong> the initialposition x of the system:A x = {α ∈ A : y x (s; α) ∈ Ω <strong>for</strong> all s > 0}. (7.116)This set will be assumed to be nonempty <strong>for</strong> all x ∈ Ω.Moreover, in general the “state constrained” value functionv(x) = infα∈A x∫ ∞0g(y x (s; α))e −λs ds (7.117)can happen to be discontinuous, due to the complex relationship between the dataof the problem (dynamics <strong>and</strong> constraints) <strong>and</strong> the set A x . Let us denote by A x ,<strong>for</strong> x ∈ ∂Ω, the subset of controls such that the corresponding vectorfield f pointsinside the constraint, i.e.{Aif x ∈ ΩA x =(7.118){a ∈ A : f(x, a) · ν(x) < 0} if x ∈ ∂Ω.Clearly, this set coincides with the usual control set A at interior points, where theconstraint is not active. Note that the map x → A x is not regular in general, as thefollowing simple example shows.Consider the dynamics in R 2 :ẏ(s) = α(s) ∈ A = B(0, 1),<strong>and</strong> Ω = (−1, 1) 2 . The system can move in every direction at internal points,whereas on the right h<strong>and</strong> side of the square the admissible controls can only bechosen in the half ball B(0, 1)∩{a 1 < 0}, on the upper side of the square in the halfball B(0, 1) ∩ {a 2 < 0} <strong>and</strong> so on. At the vertices of Ω, admissible controls must bechosen in a quadrant of B(0, 1).It has been shown by Soner [So86a, So86b] that the value function is continuous (<strong>and</strong>there<strong>for</strong>e, uni<strong>for</strong>mly continuous) on Ω if <strong>for</strong> some positive constant β the followingboundary condition on the vector field f is satisfied:For all x ∈ ∂Ω there exists a ∈ A such that f(x, a) · ν(x) ≤ −β < 0, (7.119)where ν(x) is the outward normal to Ω at the point x. Condition (7.119) is satisfiedin the example above. However, if the state constraint Ω is cut along the x 2 axisto obtain the new constraint ˜Ω = Ω \ {(0, x 2 ) : −1 ≤ x 2 ≤ 0} <strong>and</strong> the dynamics isdefined by ẏ = (a 1 , 0) with a 1 > 0, it is easy to see that starting at an initial pointon the left of the cut the dynamics will remain trapped on the left, <strong>and</strong> even <strong>for</strong> a


✐✐226 Chapter 7. Control <strong>and</strong> Gameslinear running cost such as g(x 1 , x 2 ) = 1 − x 1 we will have a discontinuity of thevalue function inside ˜Ω. Note that condition (7.119) is not satisfied in this latterexample, although the set A x is always nonempty.Under the condition (7.119), by the Dynamic Programming Principle, Sonerhas shown that the value function v is the unique “state constrained” viscositysolution ofH(x, u(x), Du(x)) = 0, x ∈ Ω, (7.120)which in turn means that v satisfies (in the viscosity sense) the inequalitiesH(x, u(x), Du(x)) ≤ 0 x ∈ Ω (7.121)H(x, u(x), Du(x)) ≥ 0 x ∈ Ω (7.122)with H defined by (7.18). In other terms, the value function v is a solution insideΩ, but only a supersolution on the boundary ∂Ω.Theorem 7.17. Let the st<strong>and</strong>ard assumptions on f <strong>and</strong> g be satisfied, <strong>and</strong> letv ∈ C(Ω). Then, v is the unique “state constrained” viscosity solution of (7.120)on Ω.Rather than giving the proof of this theorem, we show the underlying heuristicargument. In our previous examples we have seen that at all internal point we haveA x = A since the dynamics can move without restrictions. This implies that vshould satisfy the same equation as in the unconstrained case. On the other h<strong>and</strong>,if x ∈ ∂Ω, then the constraint reduces the set of admissible directions, so thatA x ⊂ A <strong>and</strong> we haveH(x, v(x), Dv(x)) = λv(x) + max{−f(x, a) · Dv(x) − g(x, a)} ≥a∈A≥ λv(x) + max{−f(x, a) · Dv(x) − g(x, a)} = 0. (7.123)a∈A xThe value function should there<strong>for</strong>e be a supersolution <strong>for</strong> (7.120).Remark 7.18 (Necessary <strong>and</strong> sufficient conditions). Condition (7.119) isknown to be only a sufficient condition <strong>for</strong> the existence of trajectories which remainwithin Ω. However, necessary <strong>and</strong> sufficient condition <strong>for</strong> the existence of solutionsin Ω have been extensively studied in viability theory (see [A91].In order to simplify the presentation, let Ω be an open convex subset of R d . Atrajectory is called viable wheny(s) ∈ Ω, ∀s ≥ 0. (7.124)Let F : Ω → R d be a multivalued map, assumed to be lower semicontinuous <strong>and</strong> withcompact convex images (we refer to [AC84] <strong>for</strong> the theory <strong>and</strong> definitions related tomultivalued maps). Define the tangent cone to Ω at the point x, as( )⋃T K (x) := µ(K − x) . (7.125)µ>0A result due to Haddad [Ha81] shows that the conditionF (x) ∩ T K (x) ≠ ∅, ∀x ∈ Ω, (7.126)


✐✐7.5. Problems with state constraints 227is necessary <strong>and</strong> sufficient to have viable trajectories <strong>for</strong> the multivalued Cauchyproblem{ẏ(s) ∈ F (y(s)) s ≥ 0(7.127)y(0) = x.This result has also been extended to more general sets <strong>and</strong> more general tangentcones.Time <strong>and</strong> space discretization In order to build a discretization of (7.120) we usethe st<strong>and</strong>ard discretization in time (7.54), (7.56) <strong>for</strong> respectively the dynamics <strong>and</strong>the cost functional. For x ∈ Ω, the corresponding value function <strong>for</strong> the discretetime problem isv ∆t (x) = inf J ∆tα ∆t x (α ∆t ), (7.128)where α ∆t ∈ A ∆tx , the set of discrete control sequences {a m } such thata m ∈ A ∆tx = {a ∈ A : x + ∆tf(x, a) ∈ Ω}. (7.129)By st<strong>and</strong>ard arguments, we can obtain also <strong>for</strong> the state constrained problem aDynamic Programming Principle, in the <strong>for</strong>m{}∑¯n−1v ∆t (x) = inf ∆t g(y m , a m )e −λtm + e −λt¯n v ∆t (y¯n ) . (7.130)α ∆t m=0<strong>for</strong> all x ∈ R d <strong>and</strong> any positive integer ¯n. Setting as usual ¯n = 1, this gives thetime-discrete schemev ∆t (x) =inf{∆tg(x, a) + e −λ∆t v ∆t (x + ∆tf(x, a)) } (7.131)a∈A ∆t x<strong>and</strong> finally, replacing the computation of v ∆t (x + ∆tf(x, a)) with an interpolation,the fully discrete schemev j = mina∈A ∆t x j{∆tg(xj , a) + e −λ∆t I 1 [V ](x j + ∆tf(x j , a)) } (7.132)written at the grid node x j . An iterative version can also be derived, as in theunconstrained case.Convergence analysis It can be proved that <strong>for</strong> v ∈ C 0,γ (Ω) <strong>and</strong> bounded variationoptimal controls the following estimate holds true‖v − v ∆t ‖ ∞ ≤ C∆t γ (7.133)Finally, <strong>for</strong> the fully discrete scheme the following a-priori estimate holds true‖V ∆t − V ‖ ∞ ≤where ω(∆t) denotes the modulus of continuity of v ∆t .ω(∆x) . (7.134)1 − e−λ∆t


✐✐228 Chapter 7. Control <strong>and</strong> Games7.6 Dynamic Programming <strong>for</strong> differential gamesThe Dynamic Programming approach can be also be applied applied to the analysis<strong>and</strong> approximation of differential games. . In the general setting, we consider thenonlinear system { ẏ(t) = f(y(t), a(t), b(t)), t > 0,(7.135)y(0) = xwhere y(t) ∈ R d is the statea( · ) ∈ A is the control of player 1 (player a)b( · ) ∈ B is the control of player 2 (player b),A{ a : [0, +∞[ → A, measurable } (7.136)B = { b : [0, +∞[ → B, measurable }, (7.137)A, B ⊂ R M are given compact sets. A typical choice is to take as admissible controlfunction <strong>for</strong> the two players piecewise constant functions respectively with valuesin A or B. Assume f is continuous <strong>and</strong>|f(x, a, b) − f(y, a, b)| ≤ L |x − y| ∀x, y ∈ R N , a ∈ A, b ∈ B.By Caratheodory’s theorem the choice of measurable controls guarantees that <strong>for</strong>any given a( · ) ∈ A <strong>and</strong> b( · ) ∈ B, there is a unique trajectory of (7.135) which wewill denote by y x (t; a, b). The payoff can be defined as in control problems. Forexample, <strong>for</strong> the infinite horizon problem we will haveJ x (a(·), b(·)) :=∫ +∞0g(y x (s; a, b))e −λs ds (7.138)The typical situation described by differential games is when player-a wants tominimize the payoff <strong>and</strong> player-b wants to maximize the payoff. So it is natural todefine the value of the game asv(x) := infsupa(·) b(·))J x (a(·), b(·)). (7.139)This is the case <strong>for</strong> 0 − sum differential games, where the gain of one player correspondsto the loss of the other player. This class of games is simpler with respect tothe general class of noncooperative N-players games where Nash theory of equilibriacan be applied.However, note that due to the presence of two players it is delicate to definewho is choosing first since, in general, we know thatinf supa(·) b(·)J x (a(·), b(·)) does not coincide with sup inf J x(a(·), b(·)) (7.140)b(·) a(·)We will present the main ideas of this approach in the simplified case of pursuitevasiongames .Example 7.19 (Pursuit-Evasion Games) These are particular differential gameswhere, given a closed target T ⊆ R N we define the payoff ast x (a( · ), b( · )) = min{ t : y x (t; a, b) ∈ T } ≤ +∞, (7.141)


✐✐7.6. Dynamic Programming <strong>for</strong> differential games 229xx yy EE PPFigure 7.9. Zermelo navigation problemNaturally, t x {a(·), b(·)} will be finite only under additional assumptions on thetarget <strong>and</strong> on the dynamics. The two players are opponents since player a wants tominimize the payoff (he is called the pursuer) whereas player b wants to maximizethe payoff (he is called the evader). The dynamics of these games can be splittedsince each each player is just controlling its own dynamics. So we assume that thedimension d of the state space is even <strong>and</strong> write the dynamics as{ẏ1 = f 1 (y 1 , a), y i ∈ R d/2 , i = 1, 2(7.142)ẏ 2 = f 2 (y 2 , b)For this game, the target will be given byT ɛ ≡ { |y 1 − y 2 | ≤ ɛ }, <strong>for</strong> ɛ > 0, or T 0 ≡ { (y 1 , y 2 ) : y 1 = y 2 } .Then, t x (a( · ), b( · )) is the capture time corresponding to the strategies a(·) <strong>and</strong>b(·).Example 7.20 (Zermelo navigation problem) A boat B moves with constantvelocity in a river <strong>and</strong> it can change its direction istantaneously. The water of theriver flows with a velocity σ <strong>and</strong> the boat tries to reach an isl<strong>and</strong> in the middle ofthe river (the target) maneuvering against water current. We choose a system ofcoordinates such that the velocity of the current is (σ, 0) (see Figure 7.9). In thenew system, the dynamics of the boat is described by{ẋ = σ + v B cos a,ẏ = v B sin a,where a ∈ [−π, π] is the control over the boat direction. Let z(t) = ( x(t), y(t) ) . Itis easy to see that v B > σ is a sufficient condition <strong>for</strong> the boat to reach the isl<strong>and</strong>from any initial condition in the river. This is not the case if the opposite conditionholds true (see below). Let us introduce the reachable setR ≡ { x 0 : ∃t > 0, a(·) ∈ A such that y(t; x 0 , a(·)) ∈ T } . (7.143)This is the set of initial positions from which the boat can reach the isl<strong>and</strong>. Let uschoose T = { (0, 0) } , then a simple geometric argument shows that⎧⎪⎨ R 2 se v B > σ,{ }R = (x, y) : x < 0 or x = y = 0 if v B = σ,⎪⎩ { }(x, y) : x < 0, |y| ≤ −x vB (σ 2 − vB 2 )− 1 2 if 0 ≤ vB < σ.(7.144)The result is obvious <strong>for</strong> v B ≥ σ. Now assume 0 < v B < σ. The motion of theboat at every point is determined by the (vector) sum of its velocity v B <strong>and</strong> thecurrent velocity σ. (Figure 7.10). The maximum angle which is allowed to the boat


✐✐230 Chapter 7. Control <strong>and</strong> Gamesdirection is≡ π 2 + arctan (v B√σ2 − v 2 B)<strong>and</strong> the equation of the line with this slope passing from the origin isy = −xv B√σ2 − v 2 B,which explains (7.144).7.6.1 Dynamic Programming <strong>for</strong> gamesThe first question is: how can we define the value function <strong>for</strong> the 2-players game? Certainly it is notinf sup J x (a, b)a∈A b∈Bbecause a would choose his control function with the in<strong>for</strong>mation of the wholefuture response of player b to any control function a( · ) <strong>and</strong> this will give him a bigadvantage. A more unbiased in<strong>for</strong>mation pattern can be modeled by means of thenotion of nonanticipating strategies (see [EK72] <strong>and</strong> the references therein),∆ ≡ { α : B → A : b(t) = ˜b(t) ∀t ≤ t ′ ⇒ α[b](t) = α[˜b](t) ∀t ≤ t ′ }, (7.145)Γ ≡ { β : A → B : a(t) = ã(t) ∀t ≤ t ′ ⇒ β[a](t) = β[ã](t) ∀t ≤ t ′ } . (7.146)The above definition is fair with respect to the two players. In fact, if player achooses his control in ∆ he will not be influenced by the future choices of player b(Γ has the same role <strong>for</strong> player b). Now we can define the lower value of the gameorT (x) ≡ infv(x) ≡ infsupα∈∆ b∈Bsupα∈∆ b∈Bt x (α[b], b),J x (α[b], b)when the payoff is J x (a, b) = ∫ t x(a,b)e −t dt. Similarly the upper value of the game0is˜T (x) ≡ sup inf t x(a, β[a]),a∈Aorβ∈Γṽ(x) ≡ sup inf J x(a, β[a]) .a∈Aβ∈ΓWe say that the game has a value if the upper <strong>and</strong> lower values coincide, i.e. ifT = ˜T or v = ṽ.Lemma 7.21 (DPP <strong>for</strong> games). For all 0 ≤ t < T (x)T (x) = infsupα∈∆ b∈B{ t + T (y x (t; α[b], b)) }, ∀x ∈ R \ T ,<strong>and</strong>v(x) = infα∈∆ supb∈B{ ∫ t0}e −s ds + e −t v(y x (t; α[b], b)) , ∀x ∈ T c .


✐✐7.6. Dynamic Programming <strong>for</strong> differential games 231The proof is similar to the 1-player case but more technical due to the use ofnon-anticipating strategies . Note that the upper values ˜T <strong>and</strong> Ṽ satisfy a similarDPP. Let us introduce the two <strong>Hamilton</strong>ians <strong>for</strong> games Isaacs’ Lower <strong>Hamilton</strong>ianIsaacs’ Upper <strong>Hamilton</strong>ianH(x, p) ≡ minb∈B˜H(x, p) ≡ maxa∈AThe following theorem can be found in [ES84].max { −p · f(x, a, b) } − 1 .a∈Amin { −p · f(x, a, b) } − 1 .b∈BTheorem 7.22. 1. If R\T is open <strong>and</strong> T ( · ) is continuous, then T ( · ) is a viscositysolution ofH(x, ∇T ) = 0 in R \ T . (7.147)2. If v( · ) is continuous, then it is a viscosity solution ofv + H(x, ∇v) = 0 in T c .As we already said, the main issue in this theory of weak solutions is to proveuniqueness results. This point is very important also <strong>for</strong> numerical purposes sincethe fact that we have a unique solution allow to prove convergence results <strong>for</strong> theapproximation schemes. To this end, let us consider the Dirichlet boundary valueproblem { u + H(x, ∇u) = 0 in Ω(7.148)u = gon ∂Ω<strong>and</strong> prove a uniqueness result under assumptions on H including Bellman’s <strong>Hamilton</strong>ianH 1 . This new boundary value problem is connected to (7.37) because thenew solution of (7.148) is a rescaling of T . In fact, introducing the new variablev(x) ≡{1 − e−T (x)if T (x) < +∞, i.e. x ∈ R1 if T (x) = +∞, (x /∈ R)(7.149)it is easy to check that, by the DPP,v(x) =inf J x(a)a( · )∈AwhereMoreover, V is a solution ofJ x (a) ≡∫ tx(a)0e −t dt .{v(x) + max { −∇v(x) · f(x, a) − 1 } = 0 ina∈Av(x) = 0RN \ Tx ∈ ∂T(7.150)which is a special case of (7.148), with H(x, p) = H 1 (x, p) ≡ max a∈A { −p · f(x, a) −1 } <strong>and</strong> Ω = T c ≡ R N \ T . The change of variable (7.41) is called Kružkov trans<strong>for</strong>mation<strong>and</strong> has several advantages. First of all V takes values in [0, 1] whereas T


✐✐232 Chapter 7. Control <strong>and</strong> Gamesis generally unbounded <strong>and</strong> this helps in the numerical approximation. Moreover,one can always reconstruct T <strong>and</strong> R from V by the relationsT (x) = − log(1 − v(x)), R = { x : v(x) < 1 }.Lemma 7.23. The Mininimum Time <strong>Hamilton</strong>ian H 1 satisfies the “structuralcondition”|H(x, p) − H(y, q)| ≤ K(1 + |x|)|p − q| + |q| L |x − y| ∀x, y, p, q, (7.151)where K <strong>and</strong> L are two positive constants.The following theorem has been proved in [CL83].Theorem 7.24. Assume H satisfies (SH), u, w ∈ BUC(¯Ω), u subsolution, wsupersolution of v + H(x, ∇v) = 0 in Ω (open), u ≤ w on ∂Ω. Then, u ≤ w in Ω.Definition 7.25. We call subsolution (respectively supersolution) of (7.42) a subsolution(respectively supersolution) u of the differential equation such that u ≤ 0on ∂Ω (respectively ≥ 0 on ∂Ω).Corollary 7.26. If the value function V ( · ) ∈ BUC(T c ), then V is the maximalsubsolution <strong>and</strong> the minimal supersolution of (7.42) (we say it is the completesolution). Thus V is the unique viscosity solution.If the system is STLC around T (i.e. T ( · ) is continuous at each point of ∂T )then V ∈ BUC(T c ) <strong>and</strong> we can apply the Corollary.The structural condition (7.151) plays an important role <strong>for</strong> uniqueness.Lemma 7.27. Isaacs’ <strong>Hamilton</strong>ians H, ˜H satisfy the structural condition (7.151)Then Comparison Theorem 7.24 applies <strong>and</strong> we getTheorem 7.28. If the lower value function v( · ) ∈ BUC(T c ), then v is the completesolution (maximal subsolution <strong>and</strong> minimal supersolution) of{ u + H(x, Du) = 0 in T c ,. (7.152)u = 0 on ∂T .Thus V is the unique viscosity solution.Note that <strong>for</strong> the upper value functions ˜T <strong>and</strong> ˜W the same results are validwith H = ˜H. We can give “capturability” conditions on the system ensuringv, ṽ ∈ BUC(T c ). However, those conditions are less studied <strong>for</strong> games becausethere are important pursuit-evasion games with discontinuous value, the games with“barriers” (see the monography [I65] <strong>for</strong> an extensive presentation of differentialgames). It is important to note that in general the upper <strong>and</strong> the lower values aredifferent. However, the Isaacs conditionH(x, p) = ˜H(x, p) ∀x, p, (7.153)


✐✐7.6. Dynamic Programming <strong>for</strong> differential games 233guarantees that they coincide.Corollary 7.29. If v, ṽ ∈ BUC(T c ), thenv ≤ ṽ, T ≤ ˜T .If Isaacs condition holds then v = ṽ <strong>and</strong> T = ˜T , (i.e. the game has a value).Proof. Immediate from the comparison <strong>and</strong> uniqueness <strong>for</strong> (7.148).For numerical purposes, one can decide to write down an approximationscheme <strong>for</strong> either the upper or the lower value using the techniques of the nextsection. Be<strong>for</strong>e going to it, let us give some in<strong>for</strong>mations about the characterizationof discontinuous value functions <strong>for</strong> games. This is an important issue becausediscontinuities appear even in classical pursuit-evasion games (e.g. in the homicidalchauffeur game that we will present later). We will denote by B(Ω) the set ofbounded real functions defined on Ω. Let us start with a definition which has beensuccessfully applied to convex <strong>Hamilton</strong>ians. As we have seen, to obtain uniquenessone should prove that a comparison principle holds, i.e. <strong>for</strong> every subsolution w<strong>and</strong> supersolution W we havew ≤ WAlthough this is sufficient to get uniqueness in the convex case the above definitionwill not guarantee uniqueness <strong>for</strong> nonconvex hamiltonians (e.g. min-max <strong>Hamilton</strong>ians).Two new definitions have been proposed. Let us denote by S the set of subsolutionsof our equation <strong>and</strong> by Z the set of supersolutions always satisfying the Dirichletboundary condition on ∂Ω.Definition 7.30 (minmax solutions [Su95]). u is a minmax solution if thereexists two sequences w n ∈ S <strong>and</strong> W n ∈ Z such thatw n = W n = 0 on ∂Ω (7.154)w n is continuous on ∂Ω (7.155)<strong>and</strong> limnw n (x) = u(x) = limnW n (x), x ∈ Ω. (7.156)Definition 7.31 (e-solutions, see e.g. [BCD97]). u is an e-solution (envelopesolution) if there exists two non empty subsetsS(u) ⊂ SZ(u) ⊂ Zsuch that ∀ x ∈ Ωu(x) =sup w(x) = inf W (x)w∈S(u)W ∈Z(u)By the comparison Lemma there exists a unique e-solution.In fact, if u <strong>and</strong> v are two e-solutions,u(x) =sup w(x) ≤ inf W (x) = v(x)w∈S(u)W ∈Z(v)


✐✐234 Chapter 7. Control <strong>and</strong> Games<strong>and</strong> alsov(x) =sup w(x) ≤ inf W (x) = u(x)w∈S(v)W ∈Z(u)It is interesting to note that in our problem the two definitions coincide.Theorem 7.32. Under our hypotheses, u is a minmax solution if <strong>and</strong> only if u isan e-solution.Time discretization <strong>for</strong> games The same time discretization can be written <strong>for</strong>the dynamics <strong>and</strong> natural extensions of N <strong>and</strong> v ∆t are easily obtained. The crucialpoint is to prove that the discrete DPP holds true <strong>and</strong> that the upper value of thediscrete game is the unique solution in L ∞ (R d ) of the external Dirichlet problemv ∆t (x) = S(v ∆t )(x) on R d \ T (7.157)v ∆t (x) = 0 on ∂T (7.158)where the fixed point operator now isS(v ∆t )(x) ≡ max min[e −∆t v ∆t (x + hf(x, a, b)) ] + 1 − e −∆tb∈B a∈AThe next step is to show that the discretization (7.157)–(7.158) is convergent tothe upper value of the game. A detailed presentation goes beyond the purposes ofthis paper, the interested reader will find these results in [BS91b]. We just give themain convergence result <strong>for</strong> the continuous case.Theorem 7.33. Let v ∆t be the solution of (7.157)–(7.158), Let T be compact withnonempty interior, the assumptions on f be verified, v be continuous.Then, v ∆tconverges to v locally uni<strong>for</strong>mly in R d <strong>for</strong> ∆t → 0 + .Space discretization <strong>for</strong> differential games Let us get back to zero-sum games.Using the change of variable v(x) ≡ 1 − e −T (x) we can set the Isaacs equation in R dobtaining{v(x) + minv(x) = 0b∈B maxa∈A[−f(x, a, b) · ∇v(x)] = 1 inRd \ T<strong>for</strong> x ∈ ∂T(7.159)Assume we want to solve the equation in Q, an hypercube in R N . As we have seen,in the case of a single player we need to impose boundary conditions on ∂Q or, atleast, on I out . However, the situation <strong>for</strong> games is much more complicated. In fact,setting the value of the solution outside Q equal to 1 (as in the single player case)will imply that the pursuer looses every time the evader drives the dynamics outsideQ. On the contrary, setting the value to 0 outside Q will give a great advantage tothe pursuer. One way to define more unbiased boundary conditions is the following.Assume that Q = Q 1 ∩ Q 2 , where Q i , i = 1, 2 are subsets of R N/2 which can beinterpreted as the set of constraints <strong>for</strong> the i-th player. For example, in R 2 we canconsider as Q 1 a vertical strip <strong>and</strong> as Q 2 an horizontal strip <strong>and</strong> compute in therectangle Q which is the intersection of those strips. According to this construction,we penalize the pursuer if the dynamics exits Q 1 <strong>and</strong> the evader if the dynamicsexits Q 2 . When the dynamics exits Q 1 <strong>and</strong> Q 2 we have assign a value, e.g. giving


✐✐7.6. Dynamic Programming <strong>for</strong> differential games 235an advantage to one of them (in the following scheme we are giving advantage tothe evader). The discretization in time <strong>and</strong> space leads to a fully discrete schemewhere β ≡ e −∆t <strong>and</strong>w(x i ) = max min[βw(x i + hf(x i , a, b))] + 1 − β <strong>for</strong> i ∈ I in (7.160)b aw(x i ) = 1 <strong>for</strong> i ∈ I out2 (7.161)w(x i ) = 0 <strong>for</strong> i ∈ I T ∪ I out1 (7.162)I in = {i : x i + hf(x i , a, b) ∈ Q \ T <strong>for</strong> any a ∈ A, b ∈ B} (7.163)I T = {i : x i ∈ T ∩ Q} (7.164)I out1 = {i : x i /∈ Q 2 } (7.165)I out2 = {i : x i /∈ Q 2 \ Q} (7.166)Theorem 7.34. The operator S defined in (7.160) has the following properties:i) S is monotone, i.e. U ≤ V implies S(U) ≤ S(V );ii) S : [0, 1] L → [0, 1] L ;iii) S is a contraction mapping in the max norm,‖S(U) − S(V )‖ ∞ ≤ β‖U − V ‖ ∞The proof of the above theorem is a generalization of that related to theminimum time problem <strong>and</strong> can be found in [BFS94]. The above result guaranteesthat there is a unique fixed point U ∗ <strong>for</strong> S. Naturally the numerical solution w willbe obtained extending by linear interpolation the values of U ∗ <strong>and</strong> it will dependon the discretization steps ∆t <strong>and</strong> ∆x. Let us state the first convergence result <strong>for</strong>continuous value functionsTheorem 7.35. Let T be the closure of an open set with Lipschitz boundary, “diamQ → +∞” <strong>and</strong> v be continuous. Then, <strong>for</strong> ∆t → 0 + <strong>and</strong> ∆x∆t → 0+ , w ∆ convergesto v locally uni<strong>for</strong>mly on the compact sets of R d .Note that the requirement “diam Q → +∞” is just a technical trick to avoidto deal with boundary conditions on Q (a similar statement can be written in thewhole space just working on an infinite mesh). We conclude this section quoting aconvergence result which holds also in presence of discontinuities (barriers) <strong>for</strong> thevalues function. Let w ɛ n be the sequence generated by the numerical scheme withtarget T ɛ = {x : d(x, T ) ≤ ɛ}.Theorem 7.36. For all x there exists the limitw(x) =lim wn(x)ɛε→0 +n→+∞n≥n(ɛ)<strong>and</strong> it coincides with the lower value V of the game with target T , i.e. w = V .Convergence is uni<strong>for</strong>m on every compact set where V is continuous.


✐✐236 Chapter 7. Control <strong>and</strong> GamesCan we know a-priori what is the accuracy of the method in the approximationof the value function? This result is necessary to underst<strong>and</strong> how far we are fromthe real solution when we compute our approximate solution. To simplify, let usassume that the Lipschitz constant <strong>for</strong> f L f ≤ 1 <strong>and</strong> that v is Lipschitz continuous.Then,( ) ) 2 ∆x‖w h,k − v‖ ∞ ≤ C∆t(1 1/2 +∆tThe proof of the above error estimate is rather technical <strong>and</strong> can be found in [Sor98].<strong>Approximation</strong> of optimal feedback <strong>and</strong> trajectories <strong>for</strong> games The numericalsolution w ∆ has been extended to Q by interpolation we can also compute anapproximate optimal feedback at every point of x ∈ Q, i.e. <strong>for</strong> the control problemwe can define the feedback map F : Q → A. For games, the algorithm computes anapproximate optimal control couple (a ∗ , b ∗ ) at every point of the grid. Again by wwe can also compute an approximate optimal feedback at every point x ∈ Q.(a ∗ (x), b ∗ (x)) ≡ argminmax{e −∆t w(x + hf(x, a, b))} + 1 − e −∆t (7.167)If that control is not unique then we can select a unique couple, e.g. minimizing twoconvex functionals. A typical choice is to introduce an inertial criterium to stabilizethe trajectories, i.e. if at step n + 1 the set of optimal couples contains (a ∗ n, b ∗ n) wekeep it.7.6.2 Numerical tests <strong>for</strong> pursuit-evasion gamesLet us examine some classical games <strong>and</strong> look at their numerical solutions. We willfocus our attention to the accuracy in the approximation of the value function aswell as to the accuracy in the approximation of optimal feedbacks <strong>and</strong> trajectories.In the previous sections we always assumed that the sets of controls A <strong>and</strong> B werecompact. In the algorithm <strong>and</strong> in the numerical tests we have used a discrete finiteapproximation <strong>for</strong> those sets which allows to compute the min-max by comparison.For example, we will consider the following discrete sets{A = a 1 + j a 2 − a}1, j = 0, . . . , c − 1;{B =c − 1b 1 + j b 2 − b 1c − 1}, j = 0, . . . , c − 1;where [a 1 , a 2 ] <strong>and</strong> [b 1 , b 2 ] represent the control sets respectively <strong>for</strong> the pursuer P<strong>and</strong> the evader E. Finally, note that all the value functions represented in thepictures have values in [0, 1] because we have computed the fixed point after theKružkov change of variable.7.6.3 The Tag-Chase GameTwo boys P <strong>and</strong> E are running one after the other in the plane R 2 . P wants tocatch E in minimal time whereas E wants to avoid the capture. Both of them arerunning with constant velocity <strong>and</strong> can change their direction instantaneously. Thismeans that the dynamics of the system isf P (y, a, b) = v P af E (y, a, b) = v E b


✐✐7.6. Dynamic Programming <strong>for</strong> differential games 237where v P <strong>and</strong> v E are two scalars representing the maximum speed <strong>for</strong> P <strong>and</strong> E <strong>and</strong>the admissible controls are taken in the setsA = B = B(0, 1).Let us give a more explicit version of the dynamics which is useful <strong>for</strong> the discretization.Let us denote by (x P , y P ) the position of P <strong>and</strong> by (x E , y E ) the position ofE, we can write the dynamics as⎧⎪⎨⎪⎩ẋ P = v P sin θ Pẏ P = v P cos θ Pẋ E = v E sin θ Eẏ E = v E cos θ E(7.168)where θ P ∈ [a 1 , a 2 ] ⊆ [−π, π] is the control <strong>for</strong> P <strong>and</strong> θ E ∈ [b 1 , b 2 ] ⊆ [−π, π] is thecontrol <strong>for</strong> E, θ P <strong>and</strong> θ E are the angles between the y axis <strong>and</strong> the velocities <strong>for</strong> P<strong>and</strong> E (see Figure 7.6.3).We say that E has been captured by P if their distance in the plane is lower than agiven threshold ɛ > 0. Introducing z ≡ (x P , y P , x E , y E ) we can say that the captureoccurs whenever z ∈ ˜T where{˜T ≡ z ∈ R 4 : √ }(x P − x E ) 2 + (y P − y E ) 2 < ɛ . (7.169)The Isaacs equation is set in R 4 since every player belongs to R 2 . However theresult of the game just depends on the relative positions of P <strong>and</strong> E, since theirdynamics are homogeneous. In order to reduce the amount of computations neededto compute the value function we describe the game in a new coordinate systemintroducing the variables˜x = (x E − x P ) cos θ − (y E − y P ) sin θ (7.170)ỹ = (x E − x P ) sin θ − (y E − y P ) cos θ (7.171)(7.172)The new system (called relative coordinates system) has the origin fixed on theposition of P <strong>and</strong> moves with this player (see Figure 7.6.3). Note that the y axis isoriented from P to E. In the new coordinates the dynamics (7.168) becomes{˙˜x = v E sin θ E − v P sin θ P˙˜y = v E cos θ E − v P cos θ P(7.173)<strong>and</strong> the target (7.169) isT ≡{(x, y) : √ }x 2 + y 2 < ɛ .It is important to note that the above change of variables greatly simplifies thenumerical solution of the problem <strong>for</strong> three different reasons. The first is that wenow solve the Isaacs equation in R 2 <strong>and</strong> we need a grid of just M 2 nodes insteadof M 4 nodes (here M denotes the number of nodes in one dimension). The secondreason is that we now have a compact target in R 2 whereas the original target (7.169)is unbounded. This is a major advantage since we can choose a fixed rectangulardomain Q in R 2 such that it contains T <strong>and</strong> compute the solution in it. Finally, we


✐✐238 Chapter 7. Control <strong>and</strong> Gamesget rid of the boundary conditions on ∂Q. It is easily seen that the game has alwaysa value <strong>and</strong> that the only interesting case is v P > v E (if the opposite inequalityholds true capture is impossible if E plays optimally). In this situation the beststrategy <strong>for</strong> E is to run at maximal velocity in the direction opposite to P along theline passing through the initial positions of P <strong>and</strong> E. The optimal strategy <strong>for</strong> P isto run after E at maximal velocity. The corresponding minimal time of capture is√(xE − x P )T (x P , y P , x E , y E ) =2 + (y E − y P ) 2v P − v Eor, in relative coordinates,T (x, y) =√x2 + y 2v P − v E.Let us comment some numerical experiments. We have chosen Q = [−1, 1] 2v P = 2, v E = 1, A = B = [−π, π].Figures 7.12 correspond to the following discretization# Nodes ∆t ɛ # Controls23 × 23 0.05 0.20 P=41 E=41The value function is represented in the relative coordinate system, so P is fixedat the origin <strong>and</strong> the value at every point is the minimal time of capture (afterKružkov trans<strong>for</strong>m). As one can see in Figure 7.12, the behaviour is correct since itcorrespond to a (rescaled) distance function. The optimal trajectories <strong>for</strong> the initialpositions P = (0.3, 0.3), E = (0.6, −0.3) are represented in Figure 7.12 (right).7.6.4 The Tag-Chase game with constraints on the directionsThis game has the dynamics (7.168). The only difference with respect to the Tag-Chase game is that now the pursuer P has a constraint on his displacement directions.He can choose his control in the set θ P ∈ [a 1 , a 2 ] ⊆ [−3/4π, 3/4π]. Theevader can still choose his control as θ E ∈ [b 1 , b 2 ] = [−π, π], i.e.A = [θ 1 , θ 2 ] <strong>and</strong> B = [−π, π].In the numerical experiment below we have chosen v P = 2, v E = 1, A = [ 3 4 π, 3 4 π]<strong>and</strong> B = [−π, π]. As one can see in Figure 7.13 the time of capture at points whichare below the origin <strong>and</strong> which cannot be reached by P in a direct way have a valuebigger than at the symmetric points (above the origin). This is clearly due to thefact that P has to zig-zag to those points because the directions pointing directlyto them are not allowed as one can see in Figure 7.13 (right).7.6.5 The Homicidal chauffeurThe presentation of this game follows [FS05]. Let us consider two players (P <strong>and</strong>E) <strong>and</strong> the following dynamics:⎧ẋ P = v P sin θ,⎪⎨ ẏ P = v P cos θ,ẋ E = v E sin b,(7.174)ẏ E = v E cos b,⎪⎩ ˙θ = R v Pa,


✐✐7.6. Dynamic Programming <strong>for</strong> differential games 239Figure 7.10. Zermelo problem: boat directionysyyPPtePxyEvPEteEvExPxExsFigure 7.11. The Tag-Chase gameTest 31Test 3: P=(0.3,0.3) E=(0.6,-0.3)0.8v(x1,x2)0.80.70.60.50.40.30.20.10-0.5-1x20.60.40.20-0.2PE10.5x10-0.5-110.50x2-0.4-0.6-0.8-1-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1x1Figure 7.12. Tag-Chase game: value function (relative coordinates) <strong>and</strong>optimal trajectoriesTest 4v(x1,x2)1Test 4: P=(-0.5,0.8) E=(-0.5,0.0)0.90.80.70.60.50.40.30.20.10x20.50-0.5PE-1-0.5x200.5110.50x1-0.5-1-1-1.5-2-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1x1Figure 7.13. Sector Tag-Chase game: value function <strong>and</strong> optimal trajectories


✐✐240 Chapter 7. Control <strong>and</strong> GamesysyyEyPPtevPEbvERaxPxxExsFigure 7.14. The Homicidal Chauffeur problemwhere a ∈ A ≡ [−1, 1] <strong>and</strong> b ∈ B ≡ [−π, π] are the two player’s controls. Thepursuer P is not free in his movements, he is constrained by a minimum curvatureradius R. The target is defined as in the Tag-Chase game. Also in this example wehave used the reduced coordinate system (7.173). We have considered the homicidalchauffeur game where Q = [−1, 1] 2 , v P = 1, v E = 0.5, R = 0.2 <strong>and</strong> the followingdiscretization parameters:# Nodes ∆t ɛ # Controls120 × 120 0.05 0.10 P=36 E=36Figure 7.15 shows the value function of the game. Note that when E is in front ofP the behaviour of the two players is analogous to the tag-chase game: in this case,indeed, the constraint on P ’s radius turn does not come into action (Figure 7.16(left)). However, on the P sides the value function has two higher lobes. In fact, toreach the corresponding points of the domain, the pursuer must first turn aroundhimself to be able to catch E following a straight line (see Figure 7.16 (right)).Finally, behind P there is a region where capture is impossible (v = 1) because theevader has the time to exit Q be<strong>for</strong>e the pursuer can catch him. Figure 7.16 (right)shows a set of optimal trajectories near a barrier in the relative coordinates system .Figure 7.17 (left) is taken from [M71] <strong>and</strong> shows the optimal trajectories which havebeen obtained by Merz via analytical methods. One can see that our results are quiteaccurate since the approximate trajectories (Figure 7.17 (right)) look very similarto the exact solutions sketched in Figure 7.17 (left). Moreover, in the numericalapproximation the barrier curve is clearly visible: that barrier cannot be crossedif both the players behave optimally. It divides the initial positions from whichthe trajectories point directly to the origin from those corresponding to trajectoriesreaching the origin after a round trip.7.7 Commented referencesAs we already mentioned, the application to deterministic control problems <strong>and</strong>games has been one of the main motivations <strong>for</strong> the development of the theory ofviscosity solutions. Dynamic Programming (DP) singles out the value function v(x)as a key tool in the analysis of optimal control problems <strong>and</strong> zero-sum differentialgames. This technique has been introduced in the classical works on calculus of


✐✐7.7. Commented references 241Test 5v(x1,x2)10.50-1-0.5x200.5110.50x1-0.5-1Figure 7.15. Homicidal Chauffeur, value function1Test 5: P=(-0.1,-0.3) E=(0.1,0.3)1Test 5: P=(0.0,0.2) E=(0.0,-0.2)0.80.80.60.60.40.4E0.20.2Px20x20-0.2-0.2EP-0.4-0.4-0.6-0.6-0.8-0.8-1-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1x1-1-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1x1Figure 7.16. Homicidal Chauffeur, optimal trajectories1Test 50.80.60.40.2x20-0.2-0.4-0.6-0.8-10 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1x1Figure 7.17. Homicidal Chauffeur optimal trajectories: Merz description,computed trajectories


✐✐242 Chapter 7. Control <strong>and</strong> Gamesvariations <strong>and</strong> extensively applied to optimal control problems by R. Bellman. Inthe sixties, a number of discrete-time control problems were investigated via theDP method, showing that v satisfies a functional (difference) equation <strong>and</strong> thatits knowledge allows to derive optimal feedbacks. We refer to the classical books[Be87] <strong>and</strong> [BeS78] <strong>for</strong> an extended presentation of these results. In continuous-timeproblems the DP equation is a nonlinear PDE, <strong>and</strong> in principle it would be necessaryto show first that the equation has a unique solution in order to identify it withthe value function. By the time DP techniques were introduced, this was a majoranalytical difficulty, since, in general, the value functions of deterministic optimalcontrol problems <strong>and</strong> games may fail to be differentiable <strong>and</strong> even continuous, sothe classical concept of solution could not be applied. Nevertheless, Isaacs usedextensively the DP principle <strong>and</strong> the first order PDE which is now associated to hisname in his book [I65] on two-persons zero-sum differential games, working mainlyon the explicit solution of several examples in which the value function is regulareverywhere except on some smooth surfaces.Only at the beginning of the eighties, M.C. Cr<strong>and</strong>all <strong>and</strong> P.L. Lions [CL83]introduced the notion of weak solution in the viscosity sense <strong>for</strong> a class of first-orderPDEs including Isaacs’ equations, <strong>and</strong> proved their existence, uniqueness <strong>and</strong> stability<strong>for</strong> the main boundary value problems. The theory was re<strong>for</strong>mulated by Cr<strong>and</strong>all,Evans <strong>and</strong> Lions in [CEL84], <strong>and</strong> Lions proved in [Li82] that continuous valuefunctions of optimal control problems are viscosity solutions of the correspondingDP equations. Similar results were obtained <strong>for</strong> the value function of some zerosumdifferential games by Barron, Evans, Jensen <strong>and</strong> Souganidis in [BEJ84], [ES84],[Sor93a] <strong>for</strong> various definitions of upper <strong>and</strong> lower value. Independently, Subbotin[Su80, Su84] found that the value functions in the Krasovskii–Subbotin sense [KS88]of some differential games satisfy certain inequalities <strong>for</strong> the directional derivativeswhich reduce to the Isaacs’ equation at points of differentiability. Moreover, heintroduced a different notion of weak solution <strong>for</strong> first order nonlinear PDEs, theminmax solution (where the name is related to the minmax operator appearing inthe DP equation of games). The book [Su80] presents this theory with a specialemphasis on the solution of differential games which motivated his study <strong>and</strong> thename <strong>for</strong> these solutions. The equivalence of this notion of weak solution with viscositysolutions has been addressed by several authors (see [EI84, LS85, SuT86]),<strong>and</strong> currently the two theories are essentially unified [Su95].Concerning control problems with state constraints, the first characterizationof the value function is due to Soner [So86a, So86b] under a controllability assumptionswhich ensure continuity. This result has been later extended by Ishii <strong>and</strong>Koike [IK96], whereas a recent <strong>for</strong>mulation due to Bokanowski, Forcadel <strong>and</strong> Zidani[BFZ10a] drops controllability assumptions still getting a characterization in termsof a DP equation. At the present state of development, the theory of viscosity solutionsapplies to a wide range of boundary <strong>and</strong> Cauchy problems <strong>for</strong> first <strong>and</strong> secondorder nonlinear PDEs, including discontinuous solutions (at least <strong>for</strong> a subclasswhich covers DP equations <strong>for</strong> control problems <strong>and</strong> games [B00, BJ89, Sor93a]).Two extensive surveys on its application to control problems <strong>and</strong> games are givenin the books [FS93] <strong>and</strong> [BCD97].We should also mention an alternative characterization of the value function,which is based on viability theory [A91]. In many classical problems, the graphof the value function is the boundary of the viability kernel <strong>for</strong> a new dynamicsassociated to the original control problem. This relationship has produced interestingresults, mainly <strong>for</strong> state-constrained problems where it is natural to consider


✐✐7.7. Commented references 243(lower) semicontinuous solutions. We refer to [Fr93] <strong>and</strong> [BJ90, BJ91] <strong>for</strong> two differentcharacterizations of semicontinuous solutions of HJB equations, <strong>and</strong> to [FP00]<strong>for</strong> an application to a state-constrained control problem.The approximation of the value function in the framework of viscosity solutionshas been investigated by several authors starting from Capuzzo Dolcetta [CD83,CDI84, F87]. Abstract convergence results, as well as a priori error estimates, havebeen proved <strong>for</strong> the approximation of classical control problems <strong>and</strong> games, see e.g.[BF90a], [BF90b], [CDF89], [Al91a], [Al91b], [BS91b]. Most of the material in thischapter originates from the survey papers [F97] <strong>and</strong> [BFS94], devoted respectivelyto control problems <strong>and</strong> games.The approximation of feedback controls is still a widely open problem, especiallyin the case of games. A first convergence result requiring regularity assumptionson the value function can be found in [DS01]. The synthesis of feedbackcontrols has also been addressed in [F01], obtaining L 1 error estimates <strong>for</strong> the approximationof optimal trajectories.Techniques <strong>for</strong> the numerical approximation of games (mainly pursuit–evasiongames) have been originated by the Krasovskii–Subbotin theory. The numericalschemes are based on a suitable construction of generalized gradients in the finitedifference operators used to approximate the value function. This approach hasbeen developed by the russian school (see [TUU95, PT00]) <strong>and</strong> allows to computean approximation of optimal feedback controls [T99]. A different technique <strong>for</strong> theapproximation of value <strong>and</strong> optimal policies of a dynamic zero-sum stochastic gamehas been proposed in [TA96], [TPA97] <strong>and</strong> is based on the approximation of thegame by a finite state approximation.The numerical counterpart of the viability approach is based on the approximationof the viability kernel. Details on the related techniques can be found in[CQS99] <strong>and</strong> [CQS00]. Finally, we mention that numerical methods based on theapproximation of open-loop controls have also been proposed, although this approachhas not been pursued here. The advantage is to replace the approximationof the DP equation (which can be difficult or impossible in high-dimensional problems)by a large system of ordinary differential equations exploiting the necessaryconditions <strong>for</strong> the optimal policy <strong>and</strong> trajectory. A general review on the relatedmethods can be found in [P94].


✐✐244 Chapter 7. Control <strong>and</strong> Games


✐✐Chapter 8Fluid DynamicsThis chapter collects some extensions of the previous algorithms to CFD applicationsas well as some examples <strong>and</strong> tests in R 2 related to more challenging benchmarks.8.1 The incompressible Euler equation in R 2The most classical model of inviscid fluid dynamics dates back to Euler <strong>and</strong> is namedafter him. In its simplest <strong>for</strong>mulation, the Euler equation describes the evolutionof an inviscid, incompressible, homogeneous fluid, <strong>and</strong> takes the <strong>for</strong>m of the system{u t + (u · ∇)u + ∇p = 0(8.1)∇ · u = 0,with suitable initial <strong>and</strong> boundary conditions. Here, <strong>for</strong> d ≤ 3, u = (u 1 , . . . , u d ) tdenotes the fluid velocity, <strong>and</strong> the second condition en<strong>for</strong>ces mass conservation.Writing the operator (u · ∇) more explicitly, (8.1) reads⎧∂u id∑⎪⎨ ∂t + j=1d∑⎪⎩i=1∂u i∂x i= 0.u j∂u i∂x j+ ∂p∂x i= 0 (i = 1, . . . , d)*** risultati di buona posizione? proprietá qualitative? paradossi? ****** mettere il rotore (curl) tra le notazioni generali ****** notazione ∇ ↔ D? ∇ 2 ↔ ∆? ***8.1.1 Vorticity–stream function <strong>for</strong>mulationAs usual in fluid dynamics, we define the vorticity as the curl of velocity:(8.2)ω = ∇ × u (8.3)In the special case of two space dimensions, the vorticity has only one nozero component(the one related to x 3 ) <strong>and</strong> can there<strong>for</strong>e be treated as a scalar. On the245


✐✐246 Chapter 8. Fluid Dynamicsother h<strong>and</strong>, once defined the operator⎛∇ ⊥ := ⎝− ∂∂x 2∂∂x 1the divergence-free vector field u may be written as⎞⎠u = −∇ ⊥ ψ (8.4)<strong>for</strong> a suitable function ψ, termed as stream function. Using (8.4) into (8.3), we get−∆ψ = ω. (8.5)Taking the curl of the momentum equation, <strong>and</strong> recalling that ∇p has a zero curl,we obtainω t + u · ∇ω = 0, (8.6)which states that vorticity is advected along the streamlines of the velocity u.Equations (8.4)–(8.6) represent the so-called vorticity–stream function <strong>for</strong>mulationof the two-dimensional incompressible Euler equation.8.1.2 SL schemes <strong>for</strong> the vorticity–stream function modelThe use of SL schemes <strong>for</strong> the VS model dates back to Wiin–Nielsen, who proposedwhat is usually recognized as the first SL scheme in Numerical Weather Prediction.The scheme was originally <strong>for</strong>mulated <strong>for</strong> the barotropic vorticity equation, but werewrite it here in a conceptually equivalent <strong>for</strong>m <strong>for</strong> the Euler equation.The general idea is to use a SL technique to implement the advection ofvorticity in (8.6). Once vorticity is computed at a new time step, stream function<strong>and</strong> velocity are updated via respectively (8.5) <strong>and</strong> (8.4). To keep a second-orderaccuracy in time, a midpoint scheme is used in the advective step as shown inChapter 5 (see (5.61)). More explicitly, we are given an initial condition u 0 , acorresponding stream function ψ 0 such thatu 0 = ∇ ⊥ ψ 0 ,<strong>and</strong> an initial vorticity ω 0 = ∇ × u 0 . Moreover, we assume that a further startupvalue of the vorticity ω −1 is available.A step <strong>for</strong>ward in time is carried out via four substeps. First, the feet ofcharacteristics are computed as in (5.61), that is,(X ∆ (x j , t n+1 ; t n−1 ) = x j − 2∆t I[U n xj + X ∆ )(x j , t n+1 ; t n−1 )]. (8.7)2Second, an advection step <strong>for</strong> the vorticity is per<strong>for</strong>med asω n+1j = I[Ω n−1 ] ( X ∆ (x j , t n+1 ; t n−1 ) ) . (8.8)Third, the updated stream function is computed by (numerically) solving the Poissonequation−∆Ψ n+1 = Ω n+1 , (8.9)


✐✐8.2. The Shallow Water Equation 247<strong>and</strong> last, the velocity is updated asU n+1 = ∇ ⊥ Ψ n+1 . (8.10)With some abuse of notation, in (8.9)–(8.10) we have used the same symbols ∆<strong>and</strong> ∇ ⊥ to denote approximate (e.g., finite difference) versions of the correspondingcontinuous operators.Note that, as it has been discussed in Chapter 5, to achieve second-orderaccuracy in time it suffices to use a P 1 or Q 1 interpolation in (8.7).8.2 The Shallow Water EquationA more complex <strong>and</strong> useful model is given by the shallow water equation (SWE),which models the flow of a relatively thin layer of fluid, with a free surface <strong>and</strong>subject to friction at the bottom. In compact <strong>for</strong>m, the system of equations reads{u t + (u · ∇)u + g∇z + γu − τ = 0(8.11)z t + ∇ · [(h + z)u] = 0.Here, u st<strong>and</strong>s again <strong>for</strong> the velocity of the fluid, h denotes the depth of the fluid atthe equilibrium, <strong>and</strong> z the difference between the actual <strong>and</strong> the equilibrium depth,so that h+z is the total depth of the fluid. The bottom friction coefficient has beendenoted by γ, <strong>and</strong> has the expressionγ =g|u|C 2 z (h + z)whereas τ denotes the stress operated by external <strong>for</strong>ces (typically, wind).8.2.1 The one-dimensional case: characteristic decompositionIn the particular case od a single space dimension, the SWE (8.11) takes the <strong>for</strong>m{u t + uu x + gz x = −γu + τz t + (h + z)u x + uz x = −uh x .Defining the solution vector W = (u, z) t , the one-dimensional SWE may be rewrittenasW t + A(W )W x = D(W ),with A(W ), D(W ) defined by( )u gA(W ) =, D(W ) =h + z u( )−γu + τ.−uh xIn this setting, it is known that the speeds of propagation of the solution are givenby the eigenvalues of A, which are easily computed asu ± √ g(h + z),in which the terms u <strong>and</strong> ± √ g(h + z) represent respectively the speed of advection<strong>and</strong> the motion of the gravitational waves. This latter component is usually muchfaster then the <strong>for</strong>mer, <strong>and</strong> causes there<strong>for</strong>e strong stability limits in conventionaleulerian schemes.


✐✐248 Chapter 8. Fluid DynamicsSL schemes based on characteristicsSemi-implicit SL schemes8.2.2 The two-dimensional case8.3 Some additional techniques8.3.1 Conservative SL schemesWhen applied to fluid dynamics problems, a major drawback of SL schemes is thatthey are not strictly mass conserving. This is due to their structure, which is notoriginally in conservative <strong>for</strong>m, nor can be recast as such. In order to overcomethis problem, <strong>and</strong> try to put together the advantages of both approaches, flux-<strong>for</strong>mversions of SL schemes have been proposed starting from the mid-90s.We briefly outline the underlying idea, working on the simplified model of theconstant-coefficient advection equation on R, which will be <strong>for</strong>mally put in the <strong>for</strong>mof a continuity equation:u t + (cu) x = 0. (8.12)First, as usual in conservative schemes, we divide the x-axis in cells of the <strong>for</strong>m[x j−1/2 , x j+1/2 ] around x j , as shown in Fig. ??. A numerical solution V n representsan approximation of the cell averages of the solution at time t n , so thatvj n ≈ 1 ∫ xj+1/2u(x, t n )dx,∆x x j−1/2<strong>and</strong>, in particular, the value vj 0 is the average over the j-th cell of the initial conditionu 0 . We also need a reconstruction operator R[V n ](x) to compute the numericalsolution V n at a generic x. The reconstruction operator R is assumed to preservethe cell averages of V n , <strong>and</strong> in the simplest case, it is implemented by taking thevalue constant <strong>and</strong> equal to vjn on the whole cell.In order to update the cell averages at the time step n + 1, we start from themass balance of the j-th cell. Assuming <strong>for</strong> example that c > 0, the mass enteringthe cell from the interface x j−1/2 in the time step [t n , t n+1 ] is∫ xj−1/2x ∗ j−1/2v(x, t n )dx,where x ∗ j−1/2= x j−1/2 − c∆t denotes the foot of the characteristic starting at(x j−1/2 , t n+1 ). Applying this general idea to a numerical solution, we define anumerical flux by averaging over a time step the mass of numerical solution crossingthe interface:F j−1/2 (V n ) := 1 ∆t∫ xj−1/2x ∗ j−1/2R[V n ](x)dx.With this definition, a conservative version of the SL scheme can be defined asv n+1j= vj n + ∆t [Fj−1/2 (V n ) − F j+1/2 (V n ) ]∆x


✐✐8.4. Numerical Tests 2498.3.2 Treatment of diffusive termsTo sketch the basic ideas, we consider here the constant-coefficient advection–diffusion equation on R:u t + cu x − νu xx = 0. (8.13)We claim that (8.13) can be approximated via the schemev n+1j = 1 (2 I[V n ] x j − c∆t − √ )2ν∆t + 1 (2 I[V n ] x j − c∆t + √ )2ν∆t . (8.14)First,√note that (8.14) per<strong>for</strong>ms an average of two different upwind points x j −c∆t±2ν∆t. This clearly preserves the stability properties of the inviscid version, e.g.,if the interpolation I[·] is monotone, then (8.14) is also monotone. To show that itis consistent, we neglect <strong>for</strong> simplicity the interpolation step, <strong>and</strong> writeu(x j − c∆t ± √ )(2ν∆t, t n = u(x j , t n ) + −c∆t ± √ )2ν∆t u x (x j , t n ) ++ 1 (−c∆t ± √ 22ν∆t)uxx (x j , t n ) +2!+ 1 (−c∆t ± √ 32ν∆t)uxxx (x j , t n ) + O(∆t 2 ) =3!(= u(x j , t n ) + −c∆t ± √ )2ν∆t u x (x j , t n ) ++ 1 (2ν∆t ∓ 2c∆t √ )2ν∆t u xx (x j , t n ) +2!+ 1 (± √ 2ν∆t 3) u xxx (x j , t n ) + O(∆t 2 ).3!We have there<strong>for</strong>e1(2 u x j − c∆t + √ )2ν∆t, t n + 1 (2 u x j − c∆t − √ )2ν∆t, t n == u(x j , t n ) + ∆t [ − cu x (x j , t n ) + νu xx (x j , t n ) ] + O(∆t 2 ).Recalling that the interpolation error is in general O(∆x r+1 ) <strong>for</strong> a smooth solution,so thatu(x j − c∆t ± √ ) (2ν∆t, t n = I[U n ] x j − c∆t ± √ )2ν∆t + O(∆x r+1 ),we can conclude that the consistency error is of order O(∆t) + O(∆x r+1 /∆t).8.4 Numerical Tests8.4.1 <strong>Linear</strong> advection8.4.2 Examples in REsempi da [FFM01] sull’ordine di convergenza e la localizzazione dell’errore (?)8.4.3 Examples in R 2Rotating cone


✐✐250 Chapter 8. Fluid Dynamics8.5 Commented referencesMarchioro–Pulvirenti (ref. generale sull’eq. di Eulero)Letteratura su SISLNavier–Stokes: Karniadakis, Milstein–TretyakovDiffusione nonlineare: TeixeiraSL conservativiVarie su problemi del secondo ordine


✐✐Bibliography[1] R.G.D. Allen, Macro-economic theory: a mathematical treatment, (Mac Millan,1967).[A91] J.P. Aubin, Viability theory,Birkhäuser, Boston-Basel-Berlin, 1991.[AC84] J. P. Aubin <strong>and</strong> A. Cellina, Differential Inclusions, Springer Verlag,Berlin-Heidelberg-New York,1984.[AF90] J.-P. Aubin <strong>and</strong> H. Frankowska, Set-Valued Analysis, Birkhäuser, Boston-Basel-Berlin, 1990.[A96] R. Abgrall, Numerical discretization of the first-order <strong>Hamilton</strong>-Jacobi equationon triangular meshes, Comm. Pure Appl. Math., 49 (1996), 1339–1373.[A03] R. Abgrall, Numerical discretization of boundary conditions <strong>for</strong> first order<strong>Hamilton</strong>-Jacobi equations, SIAM J. Numer. Anal., 41 (2003), 2233–2261.[AS64] M. Abramowitz, I.A. Stegun (eds.) - H<strong>and</strong>book of mathematical functionswith <strong>for</strong>mulas, graphs <strong>and</strong> mathematical tables - Washington, National Bureauof St<strong>and</strong>ard, 1964.[Al91a] B. Alziary de Roque<strong>for</strong>t, Jeux différentiels et approximation numérique defonctions valeur, 1re partie: étude théorique, RAIRO Math. Model. Numer.Anal., 25 (1991), 517–533.[Al91b] B. Alziary de Roque<strong>for</strong>t, Jeux différentiels et approximation numérique defonctions valuer, 2e partie: étude numérique, RAIRO Math. Model. Numer.Anal., 25 (1991), 535–560.[AL94] B. Alziary de Roque<strong>for</strong>t, P.L. Lions, A grid refinement method <strong>for</strong> deterministiccontrol <strong>and</strong> differential games, Math. Models Methods Appl. Sci, 4(1994), 899–910.[AA00] S. Augoula, R. Abgrall, High order numerical discretization <strong>for</strong> <strong>Hamilton</strong>-Jacobi equations on triangular meshes, J. Sci. Comput., 15 (2000), 197–229.[B89] M. Bardi, A boundary value problem <strong>for</strong> the minimum time function, SIAMJ. Control Optim., 26 (1989), 776–785.[B00] M. Bardi, Viscosity solutions of Isaacs’ equations <strong>and</strong> existence of a value,Università di Padova, preprint, 2000.251


✐✐252 Bibliography[BBF95] M. Bardi, S. Bottacin, M. Falcone, Convergence of discrete schemes <strong>for</strong>discontinuous value functions of pursuit-evasion games, in G.J. Olsder (ed.),”New Trends in Dynamic Games <strong>and</strong> Applications”, Birkhäuser, Boston, 1995,273-304.[BCD97] M. Bardi, I. Capuzzo Dolcetta, Optimal control <strong>and</strong> viscosity solutions of<strong>Hamilton</strong>–Jacobi–Bellman equations, Birkhäuser, 1997.[BF90a] M. Bardi, M. Falcone, An approximation scheme <strong>for</strong> the minimum timefunction, SIAM J. Control Optim., 28 (1990), 950–965.[BF90b] M. Bardi, M. Falcone, Discrete approximation of the minimal time function<strong>for</strong> systems with regular optimal trajectories, in A. Bensoussan, J.L. Lions(eds.) ”Analysis <strong>and</strong> optimization of systems (Antibes, 1990)”, Lecture Notesin Control <strong>and</strong> In<strong>for</strong>m. Sci., vol. 144, Springer, Berlin, 1990, 103–112.[BFS94] M. Bardi, M. Falcone, P. Soravia, Fully discrete schemes <strong>for</strong> the value functionof pursuit-evasion games, Advances in dynamic games <strong>and</strong> applications,T. Basar <strong>and</strong> A. Haurie eds., Annals of the International Society of DifferentialGames, Boston: Birkhäuser, (1994), vol. 1, 89–105.[BFS00] M. Bardi, M. Falcone, P. Soravia, Numerical Methods <strong>for</strong> Pursuit-EvasionGames via Viscosity Solutions, in M. Bardi, T. Parthasarathy <strong>and</strong> T.E.S.Raghavan (eds.) “Stochastic <strong>and</strong> differential games: theory <strong>and</strong> numericalmethods”, Annals of the International Society of Differential Games, Boston:Birkhäuser, (2000) vol. 4, 289-303.[BKS00] M. Bardi, S. Koike, P. Soravia, Pursuit-evasion games with state constraints:dynamic programming <strong>and</strong> discrete-time approximations, DiscreteContin. Dynam. Systems, 6 (2000), 361–380.[BRP00] M. Bardi, T.E.S. Raghavan, T. Parthasarathy (eds.), Stochastic <strong>and</strong> DifferentialGames: Theory <strong>and</strong> Numerical Methods, Annals of the InternationalSociety of Differential Games, Birkhäuser, Boston, 2000.[BS89] M. Bardi, P. Soravia, A PDE framework <strong>for</strong> differential games of pursuitevasiontype, in T. Basar <strong>and</strong> P. Bernhard, editors, Differential games <strong>and</strong> applications,Lecture Notes in Control <strong>and</strong> In<strong>for</strong>mation Sciences, 62–71. Springer-Verlag, 1989.[BS91a] M. Bardi, P. Soravia, <strong>Hamilton</strong>-jacobi equations with singular boundaryconditions on a free boundary <strong>and</strong> applications to differential games, Trans.Amer. Math. Soc., 325 (1991), 205–229.[BS91b] M. Bardi, P. Soravia, <strong>Approximation</strong> of differential games of pursuitevasionby discrete-time games, in R.P. Hamalainen <strong>and</strong> H.K. Ethamo, editors,volume 156 of Lecture Notes in Control <strong>and</strong> In<strong>for</strong>mation Sciences, 131–143.Springer-Verlag, 1991.[Ba93] G. Barles, Discontinuous viscosity solutions of first order <strong>Hamilton</strong>–Jacobiequations: A guided visit, Nonlinear Anal. T.M.A, 20 (1993), 1123-1134.[Ba98] G. Barles, Solutions de viscositè des equations d’<strong>Hamilton</strong>–Jacobi, Springer–Verlag, 1998.


✐✐Bibliography 253[BaS91] G. Barles, P.E. Souganidis, Convergence of approximation schemes <strong>for</strong> fullynonlinear second order equations, Asympt. Anal., 4 (1991), 271-283.[Bar90] E. N. Barron, Differential games with maximum cost, Nonlinear Anal. T.M. A., 14 (1990), 971–989.[BEJ84] E.N. Barron, L.C. Evans, R. Jensen, Viscosity solutions of Isaacs’ equations<strong>and</strong> differential games with Lipschitz controls, J. Differential Equations,53 (1984), 213–233.[BJ89] E. N. Barron <strong>and</strong> R. Jensen, Total risk aversion, stochastic optimal control,<strong>and</strong> differential games, Appl. Math. Optim., 19 (1989), 313–327.[BJ90] E. N. Barron <strong>and</strong> R. Jensen, Semicontinuous viscosity solutions <strong>for</strong><strong>Hamilton</strong>-Jacobi equations with convex <strong>Hamilton</strong>ian, Comm. Partial DifferentialEquations 15 (1990), 17131742.[BJ91] E. N. Barron, R. Jensen, Optimal control <strong>and</strong> semicontinuous viscosity solutions,Proc. Amer. Math. Soc., 113 (1991), 397–402.[BO82] T. Basar <strong>and</strong> G.J. Olsder, Dynamic non-cooperative game theory, AcademicPress, New York, 1982.[Be87] D. P. Bertsekas, Dynamic programming. Deterministic <strong>and</strong> stochastic models,Prentice Hall, Inc., Englewood Cliffs, NJ, 1987.[BeS78] D. P. Bertsekas, S.E. Shreve, Stochastic optimal control. The discrete timecase, Mathematics in Science <strong>and</strong> Engineering, 139. Academic Press, New York-London, 1978.[BM08] N. Besse, M. Mehrenberger, Convergence of classes of high-order semi-Lagrangian schemes <strong>for</strong> the Vlasov–Poisson system, Math. Comp., 77 (2008),93-123.[BCZ10] O. Bokanowski, E. Cristiani, H. Zidani, An efficient data structure <strong>and</strong>accurate scheme to solve front propagation problems, J. Sci. Comput. 42 (2010),251273.[BFZ10a] O. Bokanowski, N. Forcadel, H. Zidani, Reachability <strong>and</strong> minimal times<strong>for</strong> state constrained nonlinear problems without any controllability assumption.SIAM J. Control Optim. 48 (2010), 42924316.[BFZ10b] O. Bokanowski, N. Forcadel, H. Zidani, L 1 -error estimates <strong>for</strong> numericalapproximations of <strong>Hamilton</strong>-Jacobi-Bellman equations in dimension 1, Math.Comp. 79 (2010), 13951426.[BMZ09] O. Bokanowski, S. Maroso, H. Zidani,Some convergence results <strong>for</strong>Howard’s algorithm, SIAM J. Numer. Anal. 47 (2009), 30013026,[BZ07] O. Bokanowski, H. Zidani, Anti-dissipative schemes <strong>for</strong> advection <strong>and</strong> applicationto <strong>Hamilton</strong>-Jacobi-Bellmann equations, J. Sci. Comput. 30 (2007),133.[Br89] Y. Brenier, Un algorithme rapide pour le calcul de tras<strong>for</strong>mées de Legendre–Fenchel discretes, C.R. Acad, Sci. Paris Ser. I Math., 308 (1989), 587-589.


✐✐254 Bibliography[BS08] S.C. Brenner, L.R. Scott, The mathematical theory of finite element method,Springer-Verlag, New York, 2008.[Bre73] R. Brent, Algorithms <strong>for</strong> minimization without derivatives, Pentice–Hall,Englewood Cliffs, 1973.[BrL03] S. Bryson, D. Levy, Central <strong>Schemes</strong> <strong>for</strong> Multi-Dimensional <strong>Hamilton</strong>-Jacobi Equations, SIAM J. Sci. Comput., 25 (2003), 767–791.[BrL04] S. Bryson, D. Levy, High-Order Central <strong>Schemes</strong> <strong>for</strong> Multidimensional<strong>Hamilton</strong>-Jacobi Equations, SIAM J. Num. Anal., 41 (2004), 1339–1369.[Buh03] M.D. Buhmann, Radial Basis Functions, Cambridge University Press,Cambridge, 2003.[BG04] H.-J. Bungartz, M. Griebel, Sparse grids, Acta Numer., 13 (2004), 147269[But03] J.C. Butcher, Numerical analysis of Ordinary Differential Equations, Wiley,Chichester, 2003.[CFLS94] CFLS94 F. Camilli, M. Falcone, P. Lanucara e A. Seghini, A domain decompositionmethod <strong>for</strong> Bellman equations, in D.E. Keyes <strong>and</strong> J.Xu (eds), DomainDecomposition methods in Scientific <strong>and</strong> Engineering Computing, ContemporaryMathematics n.180, AMS, 1994, 477-483.[CF96] F. Camilli, M. Falcone, <strong>Approximation</strong> of optimal control problems withstate constraints: estimates <strong>and</strong> applications, B.S. Mordukhovic, H.J. Sussmaneds., Nonsmooth analysis <strong>and</strong> geometric methods in deterministic optimal control,IMA Volumes in Applied Mathematics 78, Springer Verlag, 1996, 23-57[CF97] F. Camilli, M. Falcone, Analysis <strong>and</strong> approximation of the infinite horizonproblem with impulsive controls, Automatika i Telemekanika, 7, 1997, 169-184.[CF99] CF99 F. Camilli, M. Falcone, <strong>Approximation</strong> of control problems involvingordinary <strong>and</strong> impulsive controls, Control, Optimisation <strong>and</strong> Calculus of Variation,4 (1999), 159-176.[CS04] P. Cannarsa, C. Sinestrari, Semiconcave functions. <strong>Hamilton</strong>-Jacobi equations<strong>and</strong> optimal control, Birkhäuser, 2004.[CSo87] P. Cannarsa, H.M. Soner, On the singularities of the viscosity solutions to<strong>Hamilton</strong>-Jacobi-Bellman equations,Indiana Univ. Math. J. 36 (1987), 501524.[CD83] I. Capuzzo Dolcetta, On a discrete approximation of the <strong>Hamilton</strong>-Jacobiequation of dynamic programming, Appl. Math. Optim., 10 (1983), 367–377.[CDF89] I. Capuzzo Dolcetta, M. Falcone Viscosity solutions <strong>and</strong> discrete dynamicprogramming, Annal. Inst. H. Poincaré Anal. Non Lineaire, 6 (supplement)(1989),161–183.[CDI84] I. Capuzzo Dolcetta, H. Ishii, Approximate solutions of the Bellman equationof deterministic control theory, Appl. Math. Optim., 11 (1984), 161–181.[CDL90] I. Capuzzo Dolcetta, P.L. Lions , <strong>Hamilton</strong>-Jacobi equations with stateconstraints, Trans. Amer. Math. Soc., 318 (1990), 643–683.


✐✐Bibliography 255[CQS99] P. Cardaliaguet, M. Quincampoix, P. Saint-Pierre, Set valued numericalanalysis <strong>for</strong> optimal control <strong>and</strong> differential games, in M. Bardi, T.Parthasarathy <strong>and</strong> T.E.S. Raghavan (eds.) “Stochastic <strong>and</strong> differential games:theory <strong>and</strong> numerical methods”, Annals of the International Society of DifferentialGames, Boston: Birkhäuser, vol. 4, 1999, 177-247.[CQS00] P. Cardaliaguet, M. Quincampoix, P. Saint-Pierre, Numerical schemes <strong>for</strong>discontinuous value functions of optimal control. Set-valued analysis in controltheory, Set-Valued Anal., 8 (2000), 111–126.[CFF10] E. Carlini, M. Falcone, R. Ferretti, Convergence of a large time-stepscheme <strong>for</strong> mean curvature motion, Interf. Free Bound., 12 (20010), 409-441.[CFF04] E. Carlini, M. Falcone, R. Ferretti, An efficient algorithm <strong>for</strong> <strong>Hamilton</strong>–Jacobi equations in high dimension, Comp. Visual. Sci., 7 (2004), 15-29.[CFR05] E. Carlini, R. Ferretti, G. Russo, A weighted essentially non–oscillatory,large time–step scheme <strong>for</strong> <strong>Hamilton</strong>–Jacobi equations, SIAM J. Sci. Comp.,27 (2005), 1071–1091.[CK76] C.Z. Cheng, G. Knorr, The integration of the Vlasov equation in configurationspace, J. Comp. Phys., 22 (1976), 330–351.[Ch05] D.L. Chopp, H<strong>and</strong>book of Biomedical Image Analysis, volume 1, chapterRecent Advances in the Level Set Method, Kluwer Academic, 2005.[Cia02] Ph.G. Ciarlet, The finite element method <strong>for</strong> elliptic problems, SIAM,Philadelphia, 2002.[Cl83] F.H. Clarke, Optimization <strong>and</strong> nonsmooth analysis, Wiley, New York, 1983.[CST97] A.R. Conn, K. Scheinberg, Ph.L. Toint, Recent progress in unconstrainednonlinear optimization without derivatives, Math. Programming, 79 (1997),397–414.[Co96] L. Corrias, Fast Legendre-Fenchel trans<strong>for</strong>m <strong>and</strong> applications to <strong>Hamilton</strong>–Jacobi equations <strong>and</strong> conservation laws, SIAM J. Numer. Anal., 33 (1996),1534–1558.[CFN95] L. Corrias, M. Falcone, R. Natalini, Numerical schemes <strong>for</strong> conservationlaws via <strong>Hamilton</strong>–Jacobi equations, Math. Comp., 64 (1995), 555-580.[CFL67] R. Courant, K. Friedrichs, H. Lewy, On the partial difference equations ofmathematical physics, IBM J. Res. Develop., 11 (1967), 215–234.[CIR52] R. Courant, E. Isaacson, M. Rees, On the solution of nonlinear hyperbolicdifferential equations by finite differences, Comm. Pure Appl. Math., 5 (1952),243–255.[CEL84] M.G. Cr<strong>and</strong>all, L.C. Evans, P.L. Lions, Some properties of viscosity solutionsof <strong>Hamilton</strong>-Jacobi equations, Trans. Amer. Math. Soc., 282 (1984),487–502.[CIL92] M.G. Cr<strong>and</strong>all, I. Ishii, P.L. Lions, User’s guide to viscosity solutions ofsecond order partial differential equations, Bull. Amer. Math. Soc., 27 (1992),1–67.


✐✐256 Bibliography[CL83] M.G. Cr<strong>and</strong>all, P.L. Lions, Viscosity solutions of <strong>Hamilton</strong>-Jacobi equations,Trans. Amer. Math. Soc., 277 (1983), 1–42.[CL84] M.G. Cr<strong>and</strong>all, P.L. Lions, Two approximations of solutions of <strong>Hamilton</strong>–Jacobi equations, Math. Comp., 43 (1984), 1–19.[CM80] M.G. Cr<strong>and</strong>all, A. Majda, Monotone difference approximations <strong>for</strong> scalarconservation laws, Math. Comp., 34 (1980), 1–21.[CT80] M.G. Cr<strong>and</strong>all, L. Tartar, Some relations between nonexpansive <strong>and</strong> orderpreserving mappings, Proc. Amer. Math. Soc., 78 (1980), 385–390.[CF07] E. Cristiani, M. Falcone, Fast semi-Lagrangian schemes <strong>for</strong> the eikonal equation<strong>and</strong> applications, SIAM J. Num. Anal., vol. 45, n. 5 (2007), 1979-2011.[CF08] E. Cristiani, M. Falcone, Numerical solution of the Isaacs equation <strong>for</strong> differentialgames with state constraints, Proceedings Conferenza IFAC 2008, Seul,2008.[D80] P. Danielsson, Euclidean distance mapping, Computer Graphics <strong>and</strong> ImageProcessing, 14 (1980), 227-248.[DZS10] C. De Zan, P. Soravia, Cauchy problems <strong>for</strong> noncoercive <strong>Hamilton</strong>-Jacobi-Isaacs equations with discontinuous coefficients, Interfaces Free Bound. 12(2010), 347368.[DR82] J. Douglas, T.F. Russell, Numerical methods <strong>for</strong> convection-dominated diffusionproblems based on combining the method of characteristics with finiteelement or finite difference procedures, SIAM J. Num. Anal., 19 (1982), 871-885.[DS01] P. Dupuis, A. Szpiro, Convergence of the optimal feedback policies in anumerical method <strong>for</strong> a class of deterministic optimal control problems, SIAMJ. Control Optim., 40 (2001), 393-420.[EK72] R.J. Elliott, N.J. Kalton, The existence of value in differential games, Mem.Amer. Math. Soc., 126, 1972.[E98] L. C. Evans, Partial Differential Equations, Graduate Studies in Mathematics,vol. 19, AMS, 1998.[EI84] L.C. Evans, H. Ishii. Differential games <strong>and</strong> nonlinear first order PDEs inbounded domains, Manuscripta Math., 49 (1984), 109–139.[ES84] L.C. Evans, P.E. Souganidis. Differential games <strong>and</strong> representation <strong>for</strong>mulas<strong>for</strong> solutions of <strong>Hamilton</strong>–Jacobi equations, Indiana Univ. Math. J., 33 (1984),773–797.[F87] M. Falcone, A numerical approach to the infinite horizon problem of deterministiccontrol theory, Applied Mathematics <strong>and</strong> Optimization, 15 (1987),1-13.[F94] M. Falcone, The minimum time problem <strong>and</strong> its applications to front propagation,in A. Visintin e G. Buttazzo (eds) , ”Motion by mean curvature <strong>and</strong>related topics”, De Gruyter Verlag, Berlino, 1994.


✐✐Bibliography 257[F97] M. Falcone, Numerical solution of Dynamic Programming equations, AppendixA in the volume M. Bardi <strong>and</strong> I. Capuzzo Dolcetta, Optimal control <strong>and</strong>viscosity solutions of <strong>Hamilton</strong>–Jacobi–Bellman equations, Birkhäuser, Boston,1997.[F01] M. Falcone, Some remarks on the synthesis of feedback controls via numericalmethods, in J.L. Menaldi, E. Rofman, A. Sulem (eds), Optimal Control <strong>and</strong>Partial Differential Equations, IOS Press, 2001, 456-465.[FF94] M. Falcone, R. Ferretti, Discrete-time high-order schemes <strong>for</strong> viscosity solutionsof <strong>Hamilton</strong>-Jacobi equations, Num. Math., 67 (1994), 315-344.[FF98] M. Falcone, R. Ferretti, Convergence analysis <strong>for</strong> a class of Semi–Lagrangian advection schemes, SIAM J. Num. Anal., 35 (1998), 909-940.[FF02] M. Falcone, R. Ferretti, Semi–Lagrangian schemes <strong>for</strong> <strong>Hamilton</strong>–Jacobiequations, discrete representation <strong>for</strong>mulae <strong>and</strong> Godunov methods, J. Comp.Phys., 175 (2002), 559-575.[FFM01] M. Falcone, R. Ferretti, T. Manfroni, Optimal discretization steps <strong>for</strong> aclass of semi–lagrangian schemes, in ”Numerical methods <strong>for</strong> viscosity solutions<strong>and</strong> applications (Heraklion, 1999)”, Ser. Adv. Math. Appl. Sci. 59, World Sci.Publishing, River Edge, NJ, 2001.[FG98] M. Falcone, T. Giorgi, An approximation scheme <strong>for</strong> evolutive <strong>Hamilton</strong>–Jacobi equations, in W.M. McEneaney, G. Yin <strong>and</strong> Q. Zhang (eds.), “StochasticAnalysis, Control, Optimization <strong>and</strong> Applications: A Volume in Honor of W.H.Fleming”, Birkhäuser, 1998.[FLM01] M. Falcone, P. Lanucara, M. Marinucci, Parallel Algorithms <strong>for</strong> the Isaacsequation, in O. Pourtallier <strong>and</strong> E. Altman (eds.), ”Advances in Dynamic Games<strong>and</strong> Applications”, Annals of the International Society of Differential Games,Birkhäuser (2001), 203–223.[FS05] M. Falcone, P. Stefani, Advances in Parallel Algorithms <strong>for</strong> the Isaacs equation,A.S. Nowak <strong>and</strong> K. Szajowski (eds.), ”Advances in Dynamic Games”,Annals of the International Society of Differential Games, Birkhäuser (2005),515–544.[FT09] M. Falcone, C. Truini, A level-set algorithm <strong>for</strong> front propagation in thepresence of obstacles, Rendiconti di Matematica e delle sue Applicazioni, vol.29 (2009), 1-19[Fe03] R. Ferretti, Convergence of Semi–Lagrangian approximations to convex<strong>Hamilton</strong>–Jacobi equations under (very) large Courant numbers, SIAM J. Num.Anal., 40 (2003), 2240–2253.[Fe10] R. Ferretti, Equivalence of Semi-Lagrangian <strong>and</strong> Lagrange–Galerkin schemesunder constant advection speed, J. Comput. Math., 28 (2010), 461–473.[Fe11] R. Ferretti, On the relationship between Semi-Lagrangian <strong>and</strong> Lagrange-Galerkin schemes, submitted, 2011.


✐✐258 Bibliography[FP07] R. Ferretti, G. Perrone, On the stability of semi-Lagrangian advectionschemes under finite element interpolation, in ”Applied <strong>and</strong> industrial mathematicsin Italy II”, Ser. Adv. Math. Appl. Sci., 75, World Sci. Publishing,Singapore, 2007.[FR75] W. H. Fleming. R.W. Rishel, Deterministic <strong>and</strong> stochastic optimal control,Springer-Verlag, New York, 1975.[FS93] W.H. Fleming, H.M. Soner, Controlled Markov processes <strong>and</strong> viscosity solutions,Springer–Verlag, New York, 1993.[FSo89] W.H. Fleming, P.E. Souganidis, On the existence of value function of twoplayers,zero-sum differential games, Indiana Univ. Math. J., 38 (1989), 293–314.[Fr93] H. Frankowska, Lower semicontinuous solutions of <strong>Hamilton</strong>-Jacobi-Bellmanequations, SIAM J. Control Optim., 31 (1993), 257–272.[FP00] H. Frankowska, S. Plaskacz, Semicontinuous solutions of <strong>Hamilton</strong>-Jacobi-Bellman equations with degenerate state constraints, J. Math. Anal. Appl. 251(2000), 818838.[Fr71] A. Friedman, Differential games, Wiley, New York, 1971.[GS06] M. Garavello, P. Soravia, Representation <strong>for</strong>mulas <strong>for</strong> solutions of the HJIequations with discontinuous coefficients <strong>and</strong> existence of value in differentialgames, J. Optim. Theory Appl. 130 (2006), 209229.[GBSJFF90] A. Ghizzo, P. Bertr<strong>and</strong>, M.M. Shoucri, T. W. Johnston, E. Fijalkow,M.R. Feix, A Vlasov code <strong>for</strong> the numerical simulation of stimulated Ramanscattering, J. Comput. Phys., 90 (1990), 431–457.[GR96] E. Godlewski, P.A. Raviart, Numerical approximation of hyperbolic systemsof conservation laws, Applied Mathematical Sciences, 118. Springer-Verlag, New York, 1996.[GR91] E. Godlewski, P.A. Raviart, Hyperbolic systems of conservation laws,Mathématiques & Applications (Paris), 3/4. Ellipses, Paris, 1991.[GRo85a] R. Gonzalez, E. Rofman, On deterministic control problems: an approximationprocedure <strong>for</strong> the optimal cost. I. The stationary problem, SIAM J.Control Optim. 23 (1985), 242–266.[GRo85b] R. Gonzalez, E. Rofman, On deterministic control problems: an approximationprocedure <strong>for</strong> the optimal cost. II. The nonstationary case, SIAM J.Control Optim. 23 (1985), 267–285.[Ha81] H,G. Haddad, Monotone viable trajectories <strong>for</strong> functional differential inclusions,J.Diff. Eq.,42 (1981),[HNW93] E. Hairer, S.P.Nørsett, G. Wanner, Solving Ordinary Differential Equations:nonstiff problems, Springer-Verlag, Berlin, 1993[Ho65] E. Hopf, Generalized solutions of non linear equations of first order, J. Math.Mech. 14 (1965), 951-973.


✐✐Bibliography 259[HB89] B.K.P. Horn, M.J. Brooks (eds.), Shape from Shading, The MIT Press,Cambridge. Massachussetts, 1989.[HS99] C. Hu, C.-W. Shu, A discontinuous Galerkin finite element method <strong>for</strong><strong>Hamilton</strong>-Jacobi equations, SIAM J. Sci. Comput., 21 (1999), 666–690.[I65] R. Isaacs, Differential games, Wiley, New York, 1965.[Is89] H. Ishii, A boundary value problem of the Dirichlet type <strong>for</strong> <strong>Hamilton</strong>–Jacobiequations, Ann. Sc. Norm. Sup. Pisa, (IV) 16 (1989), 105–135.[IK96] H. Ishii <strong>and</strong> S. Koike, A new <strong>for</strong>mulation of state constraint problem <strong>for</strong>first-order PDEs, SIAM J. Control Optim. 34 (1996), 554571.[Is03] A. Iske, Radial basis functions: basics, advanced topics <strong>and</strong> meshfree methods<strong>for</strong> transport problems, Rend. Sem. Mat. Univ. Pol. Torino, 61 (2003), 247–285.[JP00] G.S. Jiang, D. Peng, Weighted ENO schemes <strong>for</strong> <strong>Hamilton</strong>-Jacobi equations,SIAM J. Sci. Comput., 21 (2000), 2126-2143.[KLT03] T.G. Kolda, R.M. Lewis, V. Torczon, Optimization by Direct Search: newperspectives on some classical <strong>and</strong> modern methods, SIAM Review, 45 (2003),385–482.[KS88] N.N. Krasovskii, A.I. Subbotin, Game theoretical control problems.Springer, New York 1988.[Kr66] S.N. Kružkov, The method of finite differences <strong>for</strong> a first-order non linearequation with many independent variables, USSR Comput. Math. Math. Phys.,6 (1966), 136-151.[KuD01] H.J. Kushner <strong>and</strong> P. Dupuis, Numerical methods <strong>for</strong> stochastic controlproblems in continuous time, Springer–Verlag, Berlin, 2001.[LR56] P.D. Lax <strong>and</strong> R.D. Richtmeyer, Survey of the stability of linear finite differenceequations, Comm. Pure Appl. Math., 9 (1956), 267-293.[L92] R.J. Leveque, Numerical methods <strong>for</strong> conservation laws, Birkhäuser, Basel,1992.[LT00] C.T. Lin, E. Tadmor, High–resolution nonoscillatory central schemes <strong>for</strong><strong>Hamilton</strong>–Jacobi equations, SIAM J. Sci. Comput., 21 (2000), 2163-2186.[LT01] C.T. Lin, E. Tadmor, L 1 stability <strong>and</strong> error estimates <strong>for</strong> <strong>Hamilton</strong>–Jacobisolutions, Num. Math., 87 (2001), 701-735.[Li82] P.-L. Lions. Generalized Solutions of <strong>Hamilton</strong>Jacobi Equations. Number 69in Research Notes in Mathematics. Pitman Advanced Publishing Program,1982.[LS95] P.L. Lions, P.E. Souganidis, Convergence of MUSCL <strong>and</strong> filtered schemes<strong>for</strong> scalar conservation laws <strong>and</strong> <strong>Hamilton</strong>–Jacobi equations, Num. Math., 69(1995), 441-470.


✐✐260 Bibliography[LS85] P.L. Lions, P.E. Souganidis, Differential games <strong>and</strong> directional derivativesof viscosity solutions of Bellman’s <strong>and</strong> Isaacs’ equations I <strong>and</strong> II, SIAM J.Control Optim., 23 (1985), 566–583 <strong>and</strong> 24 (1986), 1086–1089.[Lo87] P. Loreti, Some properties of constrained viscosity solutions of <strong>Hamilton</strong>-Jacobi- Bellman equations, Siam J. Control <strong>and</strong> Optimization, 25,1987, ???-???.[Lu97] Y. Lucet, Faster than the fast Legendre trans<strong>for</strong>m, the linear-time Legendretrans<strong>for</strong>m, Numer. Algorithms, 16 (1997), 171–185.[M71] A.W. Merz, The Homicidal Chauffeur - a Differential Game, PhD Dissertation,Stan<strong>for</strong>d University, 1971.[MPS88] K.W. Morton, A. Priestley, E. Süli, Stability of the Lagrange–Galerkinmethod with non-exact integration, M 2 AN, 22 (1988), 625-653.[NW06] J. Nocedal, S.J. Wright, Numerical Optimization, Springer-Verlag, NewYork, 2006.[OF03] S. Osher, R.P. Fedkiw, Level Set Methods <strong>and</strong> Dynamic Implicit Surfaces,Springer-Verlag, New York, 2003.[OS88] S. Osher, J.A. Sethian, Fronts propagating with curvature-dependent speed:algorithms based on <strong>Hamilton</strong>–Jacobi <strong>for</strong>mulations, J. Comput. Phys., 79(1988), 12-49.[OS91] S. Osher, C.W. Shu, High-order Essentially Non-Oscillatory schemes <strong>for</strong><strong>Hamilton</strong>–Jacobi equation, SIAM J. Num. Anal., 28 (1991), 907-922.[P77] A. Papoulis, Signal analysis, MacGraw–Hill, New York, 1977.[PT00] V.S. Patsko, V.L. Turova Numerical study of differential games with thehomicidal chauffeur dynamics, Scientific report of the Russian Academy ofScience, Ural Branch, Ekaterinburg, 2000.[PN99] Perelson <strong>and</strong> Nelson, HIV model....[P06] G. Perrone, Sulla stabilitá dei metodi Semi-Lagrangiani di ordine alto, Tesidi laurea, Universitá di Roma Tre, 2006.[P94] H.J. Pesch, Solving optimal control <strong>and</strong> pursuit-evasion game problems of highcomplexity, in Computational optimal control (Munich, 1992), 43–61, Internat.Ser. Numer. Math., 115, Birkhuser, Basel, 1994.[PB79] M.L. Puterman, S.L. Brumelle, On the convergence of policy iteration in stationarydynamic programming, Mathematics of Operations Research, 4 (1979),60-69.[Pi82] O. Pironneau, On the transport–diffusion algorithm <strong>and</strong> its application to theNavier–Stokes equations, Num. Math., 38 (1982), 309-332.[2] A.S. Perelson, D.E. Kirschner <strong>and</strong> R. de Boer, Dynamics of HIV infection ofCD4 + cells, Math. Biosci., 114 (1993), 81-125.


✐✐Bibliography 261[Pow08] M.J.D. Powell, Developments of NEWUOA <strong>for</strong> minimization withoutderivatives, IMA J. Numer. Anal., 28 (2008), 649–664.[QV99a] A. Quarteroni, A. Valli, Domain decomposition methods <strong>for</strong> partial differentialequations, Ox<strong>for</strong>d University Press, 1999.[QV99b] A. Quarteroni, A. Valli, Numerical methods <strong>for</strong> partial differential equations,Ox<strong>for</strong>d University Press, 1999.[RW90] P.J. Rasch, D.L. Williamson, On shape–preserving interpolation <strong>and</strong> semi-Lagrangian transport, SIAM J. Sci. Stat. Comput., 11 (1990), 656–687[RM67] R.D. Richtmeyer, K.W. Morton, Difference methods <strong>for</strong> Initial Value Problems,Interscience, New York, 1967.[Ro81] A. Robert, A stable numerical integration scheme <strong>for</strong> the primitive meteorologicalequations, Atmos. Ocean., 19 (1981), 35-46.[Sa01] G. Sapiro, Geometric partial differential equations <strong>and</strong> image analysis, CambridgeUniversity Press, 2001.[Sch99] M. Schlather, Introduction to positive definite functions <strong>and</strong> to unconditionalsimulation of r<strong>and</strong>om fields, Technical Report ST-99-10, Department ofMathematics <strong>and</strong> Statistics, Lancaster University, Lancaster, UK (1999).[Se99a] D. Serre, Systems of conservation laws. 1. Hyperbolicity, entropies, shockwaves., Cambridge University Press, Cambridge, 1999.[Se99b] D. Serre, Systems of conservation laws. 2. Geometric structures, oscillations,<strong>and</strong> initial-boundary value problems., Cambridge University Press, Cambridge,1999.[Se96] J.A. Sethian, Level Set Method. Evolving interfaces in geometry, fluid mechanics,computer vision, <strong>and</strong> materials science, Cambridge Monographs onApplied <strong>and</strong> Computational Mathematics, vol. 3, Cambridge University Press,Cambridge, 1996.[Se96b] J. A. Sethian, A fast marching level set method <strong>for</strong> monotonically advancingfronts, Proc. Natl. Acad. Sci. USA, 93 (1996), 1591-1595.[SV01] J. A. Sethian, A. Vladimirsky, Ordered upwind methods <strong>for</strong> static <strong>Hamilton</strong>-Jacobi equations, Proc. Natl. Acad. Sci. USA, 98 (2001), 11069-11074.[SV03] J. A. Sethian, A. Vladimirsky, Ordered upwind methods <strong>for</strong> static <strong>Hamilton</strong>-Jacobi equations: theory <strong>and</strong> algorithms, SIAM J. Numer. Anal., 41 (2003),325-363.[Sh98] C.W. Shu, Essentially non-oscillatory <strong>and</strong> weighted essentially nonoscillatoryschemes <strong>for</strong> hyperbolic conservation laws, in ”Advanced Numerical<strong>Approximation</strong> of Nonlinear Hyperbolic Equations (Cetraro, 1997)”, LectureNotes in Math., 1697, Springer–Verlag, Berlin, 1998.[So86a] H.M. Soner, Optimal control with state-space constraint. I, SIAM J. ControlOptim. 24 (1986), 552561.


✐✐262 Bibliography[So86b] H.M. Soner, Optimal control with state-space constraint. II. SIAM J. ControlOptim. 24 (1986), 11101122.[ST02a] H.M. Soner, N. Touzi, A stochastic representation <strong>for</strong> the level set equations,Comm. Partial Differential Equations 27 (2002), 20312053.[ST02b] H.M. Soner, N. Touzi, Dynamic programming <strong>for</strong> stochastic target problems<strong>and</strong> geometric flows, J. Eur. Math. Soc. (JEMS) 4 (2002), 201236.[ST03] H.M. Soner, N. Touzi, A stochastic representation <strong>for</strong> mean curvature typegeometric flows, Ann. Probab. 31 (2003), 11451165.[SBG99] E. Sonnendrücker, J. Roche, P. Bertr<strong>and</strong>, A. Ghizzo, The Semi-Lagrangianmethod <strong>for</strong> the numerical resolution of the Vlasov equation, J. Comput. Phys.,149 (1999), 201–220.[Sor92] P. Soravia, The concept of value in differential games of survival <strong>and</strong> viscositysolutions of <strong>Hamilton</strong>–Jacobi equations, Differential Integral Equations,5 (1992), 1049–1068.[Sor93a] P. Soravia, Pursuit-evasion problems <strong>and</strong> viscosity solutions of Isaacs equations,SIAM J. Control Optim., 31 (1993), 604–623.[Sor93b] P. Soravia, Discontinuous viscosity solutions to Dirichlet problems <strong>for</strong><strong>Hamilton</strong>–Jacobi equations with convex <strong>Hamilton</strong>ians, Comm. Partial Diff.Equations, 18 (1993), 1493–1514.[Sor94] P. Soravia, Generalized motion of a front propagating along its normal direction:a differential games approach, Nonlinear Anal. 22 (1994), 12471262.[Sor98] P. Soravia, Estimates of convergence of fully discrete schemes <strong>for</strong> the Isaacsequation of pursuit-evasion games via maximum principle, SIAM J. ControlOptim., 36 (1998), 1-11.[Sou85a] P.E. Souganidis, <strong>Approximation</strong> schemes <strong>for</strong> viscosity solutions of<strong>Hamilton</strong>–Jacobi equations, J. Differential Equations, 57 (1985), 1–43.[Sou85b] P.E. Souganidis, Max-min representations <strong>and</strong> product <strong>for</strong>mulas <strong>for</strong> theviscosity solutions of <strong>Hamilton</strong>–Jacobi equations with applications to differentialgames, Nonlinear Anal. T.M.A., 9 (1985), 217–257.[SC91] A.N. Stani<strong>for</strong>th, J. Côtè, Semi–Lagrangian integration schemes <strong>for</strong> atmosphericmodels – A review, Mon. Wea. Rev., 119 (1991), 2206–2223.[St99] J. Strain, Semi-Lagrangian methods <strong>for</strong> level set equations, J. Comput. Phys.,151 (1999), 498–533.[St99a] J. Strain, Fast tree-based redistancing <strong>for</strong> level set computations, J. Comput.Phys. 152 (1999), no. 2, 664686.[St99b] J. Strain, Tree methods <strong>for</strong> moving interfaces, J. Comput. Phys., 151(1999), no. 2, 616648,[St00] J. Strain, A fast modular semi-Lagrangian method <strong>for</strong> moving interfaces, J.Comput. Phys., 161 (2000), 512–536.


✐✐Bibliography 263[St01] J. Strain, A fast semi-Lagrangian contouring method <strong>for</strong> moving interfaces,J. Comput. Phys., 170 (2001), 373–394.[Str89] J.C. Strickwerda, Finite difference schemes <strong>and</strong> partial differential equations,Pacific Grove, Cali<strong>for</strong>nia, Wadsworth <strong>and</strong> Brooks/Cole, 1989.[Su80] A. I. Subbotin, A generalization of the basic equation of the theory of differentialgames, Soviet Math. Dokl., 22 (1980), 358–362.[Su84] A. I. Subbotin. Generalization of the main equation of differential gametheory, J. Optim. Th. Appl. 43 (1984), 103–133.[Su91] A. I. Subbotin, Minimax <strong>and</strong> viscosity solutions of <strong>Hamilton</strong>-Jacobi equations,Nauka, Moscow, 1991 (In Russian.)[Su93] A. I. Subbotin, Discontinuous solutions of a Dirichlet type boundary valueproblem <strong>for</strong> first order partial differential equations, Russian J. Numer. Anal.Math. Modelling, 8 (1993), 145–164.[Su95] A. I. Subbotin, Generalized solutions of first-order PDEs, Birkhäuser,Boston, 1995.[SuT86] A. I. Subbotin, A.M. Tarasyev, Stability properties of the value function ofa differential game <strong>and</strong> viscosity solutions of <strong>Hamilton</strong>-Jacobi equations, Probl.Control Inf. Theory, 15 (1986), 451–463.[T91] E. Tadmor, Local error estimates <strong>for</strong> discontinuous solutions of nonlinearhyperbolic equations, SIAM J. Num. Anal., 28 (1991), 891–906.[T99] A.M. Tarasyev, Control synthesis in grid schemes <strong>for</strong> <strong>Hamilton</strong>-Jacobi equations.Optimal control <strong>and</strong> differential games (Vienna, 1997), Ann. Oper. Res.,88 (1999), 337–359.[TUU95] A.M.Tarasyev, A.A. Uspenskiy, V.N. Ushakov, <strong>Approximation</strong> schemas<strong>and</strong> finite-difference operators <strong>for</strong> constructing generalized solutions of<strong>Hamilton</strong>-Jacobi equations, J. Comput. Systems Sci. Internat., 33 (1995), no.6, 127–139; translated from Izv. Ross. Akad. Nauk Tekhn. Kibernet. 1994, ,no. 3, 173–185(Russian)[TA96] M.M. Tidball, E. Altman, <strong>Approximation</strong>s in dynamic zero-sum games. I,SIAM J. Control Optim. 34 (1996), no. 1, 311–328.[TG92] M.M. Tidball <strong>and</strong> R.L.V. Gonzalez, Fast solution of general nonlinear fixedpoint problems, in ”System modelling <strong>and</strong> optimization (Zurich, 1991)”, LectureNotes in Control <strong>and</strong> In<strong>for</strong>m. Sci., 180, Springer, Berlin, 1992, 35–44.[TPA97] M.M. Tidball, O. Pourtallier, E. Altman, <strong>Approximation</strong>s in dynamiczero-sum games. II, SIAM J. Control Optim., 35 (1997), 2101–2117.[Ts95] J. N. Tsitsiklis, Efficient algorithms <strong>for</strong> globally optimal trajectories, IEEETran. Automatic. Control, 40 (1995), 1528-1538.[VR50] J. Von Neumann, R.D. Richtmeyer, A method <strong>for</strong> the numerical calculationof hydrodynamic shocks, J. Appl. Phys., 21 (1950), 232–237.


✐✐264 Bibliography[W59] A. Wiin-Nielsen, On the application of trajectory methods in numerical <strong>for</strong>ecasting,Tellus, 11 (1959), 180–196.[ZS02] Y.-T. Zhang, C.-W. Shu, High-order WENO schemes <strong>for</strong> <strong>Hamilton</strong>-Jacobiequations on triangular meshes, SIAM J. Sci. Comput., 24 (2002), 1005–1030.[3] H. Zhao, A fast sweeping method <strong>for</strong> eikonal equations, Math. Comp., 74(2005), 603-627.


✐✐Indexadvection–diffusion equation, 249SL scheme, 249amplification factor, 90autocorrelation integral, 153Barles–Souganidis theorem, 93barotropic vorticity equation, 246Bellman equation, 191boundary conditions, 36Dirichlet, 36, 125, 139Neumann, 36, 139periodic, 125, 139state constraints, 36, 140boundary layer effect, 18change of topology, 5circulating matrices, 90coercivity condition, 29comparison principle, 26conservation, 248consistency, 79, 84generalized <strong>for</strong>m, 94of multistep schemes, 46of one-step schemes, 44of the Courant–Isaacson–Rees scheme,115of the high-order SL scheme, 150,178of the Lax–Friedrichs scheme, 109,131of the SL scheme, 133, 134of the upwind scheme, 105, 129constructionof the Courant–Isaacson–Rees scheme,113of the first-order SL scheme, 133of the high-order SL scheme, 149,177of the Lax–Friedrichs scheme, 109,131of the upwind scheme, 104, 129convergence, 81, 84of multistep schemes, 47of one-step schemes, 45of the Courant–Isaacson–Rees scheme,117of the first-order SL scheme, 138of the high-order SL scheme, 172,188of the Lax–Friedrichs scheme, 112,132of the upwind scheme, 107, 131cost functionalinfinite horizon problem, 193Courant–Friedrichs–Lewy (CFL) condition,85Courant–Isaacson–Rees scheme, 113,144Cr<strong>and</strong>all–Lions theorem, 92differenced <strong>for</strong>m, 92differential games, 191, 228non-anticipating strategies, 231pursuit-evasion games, 228diffusive schemes, 101discount factor, 193dispersive schemes, 101distance function, 7domain of dependence, 85dynamic programming, 6eikonal equation, 7ENO interpolationin R 1 , 52in R d , 64feedback controls, 193finite elementdefinition, 66finite element interpolation, 64fixed-point iteration, 83flux-<strong>for</strong>m SL schemes, 248265


✐✐266 Index<strong>for</strong>mulaHopf–Lax, 28, 30front propagation, 4Galerkin projection, 150Gibb’s oscillations, 48, 52gravitational waves, 247Hölder norms, 78image irradiance equation, 10interpolation error<strong>for</strong> ENO interpolation, 54<strong>for</strong> finite element interpolation, 68<strong>for</strong> Lagrange interpolation, 52<strong>for</strong> WENO interpolation, 62invarianceby translation, 88with respect to the addition ofconstants, 86Isaacs equation, 191Kružkov trans<strong>for</strong>mation, 202, 211, 231Lagrange finite elements, 66Lagrange interpolationin R 1 , 48in R d , 63Lagrange–Galerkin schemes, 150Lax–Friedrichs scheme, 108, 131, 144Lax–Richtmeyer theorem, 81Legendre trans<strong>for</strong>m, 28level set method, 4Lin–Tadmor theorem, 98linear weights, 56Lipschitz stability, 178marginal functions, 32matrix norms, 78merging, 5methodlevel set, 4of characteristics, 28of lines, 104, 109minimizationdescent methods, 71direct search methods, 70Powell method, 72trust-region methods, 73modified equation, 101of the Courant–Isaacson–Rees scheme,119of the high-order SL scheme, 173of the Lax–Friedrichs scheme, 112of the upwind scheme, 107monotonicity, 87<strong>and</strong> nonexpansivity in ∞–norm,88<strong>and</strong> nonexpansivity in Lipschitznorm, 88generalized <strong>for</strong>m, 94of the Courant–Isaacson–Rees scheme,116of the first-order SL scheme, 137of the Lax–Friedrichs scheme, 111,132of the upwind scheme, 106, 130narrow b<strong>and</strong>, 6normal matrix, 89numerical dispersionof high-order SL schemes, 173numerical flux, 248numerical <strong>Hamilton</strong>ian, 92numerical viscosity, 101of the Courant–Isaacson–Rees scheme,118of the Lax–Friedrichs scheme, 112of the upwind scheme, 107open-loop controls, 193optimal control, 6, 191concatenation property, 195Dynamic Programming Principle,193feedback map, 214finite horizon problem, 6, 197infinite horizon problem, 193<strong>Linear</strong> Quadratic Regulator problem,196minimum time function, 199minimum time problem, 199optimal stopping problem, 198Pontryagin Maximum Principle,193synthesis of feedback controls, 214target set, 199value function, 6Ordinary Differential Equationsapproximation, 43


✐✐Index 267multistep schemes, 45, 122one-step schemes, 44, 120Pontryagin Maximum Principle, 202positive semi-definite functions, 154problemminimum time, 7Shape-from-Shading, 9proper representation, 4WENO interpolationin R 1 , 55in R d , 64reachable set, 199reference basis functions, 48, 49, 51representation <strong>for</strong>mulae, 28Riccati equation, 196running cost, 193semi-concave stability, 98discrete <strong>for</strong>mulation, 100semiconcavity, 32, 177shallow water equation, 247small time local controllability, 200smoothness indicators, 56stability, 80of multistep schemes, 46of one-step schemes, 45of the Courant–Isaacson–Rees scheme,116of the high-order SL scheme, 150,178of the Lagrange–Galerkin scheme,151of the Lax–Friedrichs scheme, 110of the upwind scheme, 106sufficient conditions, 80stationary advection equation, 142stream function, 246sub <strong>and</strong> super-differentials, 24system of base characteristics, 2time-marching schemes, 83, 142, 204upwind scheme, 104, 129, 143value function, 193, 196, 198, 199, 202,205, 209viscosity solution, 18, 25viscosity solutions<strong>for</strong> the evolutive case, 27vorticity, 245vorticity–stream function <strong>for</strong>mulation,246

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!