Exact controllability of the 1-d wave equation from a moving interior point Carlos CASTRO



Abstract We consider the linear wave equation with Dirichlet boundary conditions in a bounded interval, and with a control acting on a moving point. We give sufficient conditions on the trajectory of the control in order to have the exact controllability property. Keywords: Exact controllability, wave equation, pointwise control AMS Subject Classification: 93B05; 93B07; 35L05

1

Introduction

We consider the one-dimensional linear wave equation, on a finite interval domain (0, L), with an interior control f which acts on a single moving point x = γ(t),  in 0 < x < L, 0 < t < T,  utt − uxx = f (t)δγ(t) (x), (1.1) u(0, t) = u(L, t) = 0, in 0 < t < T,  u(x, 0) = u0 (x), ut (x, 0) = u1 (x) in 0 < x < L. Here, δγ(t) represents the Dirac measure on x = γ(t) and the function γ describes the trajectory in time of the location of the control. We assume that the function γ : [0, T ] → (0, L) is piecewise C 1 in [0, T ]. We are interested in the following exact controllability problem: Given T > 0, some initial data (u0 , u1 ) and final data (v 0 , v 1 ), find a control f such that the solution u of (1.1) satisfies u(x, T ) = v 0 (x),

ut (x, T ) = v 1 (x),

∀x ∈ (0, L).

(1.2)

Let us briefly describe some related results and the main motivation of this problem. When the control acts in an interior open set ω or one of the extremes of the domain (0, L), it is well known that the corresponding exact controllability property of the wave equation holds, for some sufficiently large time T (see [12]). On the other hand, in most practical situations, ∗

Dep. Matem´ atica e Inform´ atica, ETSI Caminos, Canales y Puertos, Universidad Polit´ecnica de Madrid, 28040 Madrid, Spain. [email protected]

1

the support of the control is required to be very small compared to the total size of the domain (0, L) and therefore it is very natural to consider a limit situation where the subinterval ω is reduced to a single point γ ∈ (0, L). It turns out that the controllability property of system (1.1) depends on the location of γ. Indeed, it can be shown that this property holds if and only if the only eigenfunction of the Laplacian with homogeneous Dirichlet boundary conditions and vanishing on x = γ is the identically zero one (see [13], [14], [1] or the more recent reference [6], for example). In the sequel, the points γ for which this spectral property is satisfied will be referred to as strategic points. The property of γ being strategic is difficult to establish in practice since it is extremely unstable. In fact γ ∈ (0, L) is strategic if and only if it is irrational with respect to the length of the interval L. Consequently, controllability properties over points are hard to use in practice. To overcome this difficulty one may consider controls supported on moving points {γ(t)}0≤t≤T , as suggested in [14]. The main advantage of moving controls is that it is easy to construct trajectories {γ(t)}0≤t≤T for which the strategic property holds for γ(t) ∈ (0, L) a.e. in t ∈ [0, T ]. For example, this is the case when we assume that the control is located at a point that moves in time with constant velocity. In this case, γ(t) is irrational, and therefore strategic, a.e. in t ∈ [0, T ]. Therefore, the exact controllability is likely to hold for such moving controls. The aim of this work is to show that this is indeed the case under suitable conditions on the function γ(t). This problem has been previously studied in [9], [10] where some sufficient conditions on the curve γ are given. Here we introduce a new approach that provides the exact controllability property for a larger class of curves. In contrast, our approach provides controls in H 1 (0, T )0 , the dual space of H 1 (0, T ). We do not know if it can be adapted to obtain L2 −controls, at least in an easy way, due to the fact that the system is not autonomous. It is worth noting that in the context of parabolic equations a similar situation appears. In this case, the approximate controllability fails if the control is located on a single non-strategic point, but can be recovered if the control acts on moving trayectories under certain hypothesis. We refer to [4], [8], [1], [14] and the references therein for a detailed analysis of this related problem. The rest of this paper is divided as follows: in section 2 we state the main results, namely the existence of solutions for system (1.1) in suitable functional spaces and the exact controllability property. Both results can be reduced, by classical duality arguments, to some suitable regularity and observability estimates for the uncontrolled wave equation respectively. In section 3 we give some examples of curves for which the exact controllability property holds. In section 4 we give the proof of the observability estimates. Finally, in section 6 we study the existence of weak solutions of (1.1) and in section 7 we briefly comment the HUM method that allows to obtain the controllability result from the observability inequality.

2

2

Main results

We first state the well-posedness of system (1.1). The control f (t) in (1.1) is assumed to belong to (H 1 (0, T ))0 , the dual space of H 1 (0, T ), and the initial data (u0 , u1 ) in the class (u0 , u1 ) ∈ L2 (0, L) × H −1 (0, L). We define the weak solutions of system (1.1) by transposition (see [11]). To do that let ψ ∈ L1 (0, T ; L2 (0, L)) be a function and consider the non-homogeneous adjoint wave equation  in 0 < x < L, 0 < t < T,  ϕtt − ϕxx = ψ(x, t) ϕ(0, t) = ϕ(L, t) = 0 in 0 < t < T, (2.3)  ϕ(x, T ) = ϕt (x, T ) = 0, in 0 < x < L, 0 < t < T. It is well known that system (2.3) admits a unique solution ϕ of (2.3) in the class ϕ ∈ C([0, T ]; H01 (0, L)) ∩ C 1 ([0, T ]; L2 (0, L)).

(2.4)

Multiplying the equations in (1.1) by ϕ and integrating by parts we easily obtain, at least formally, the following identity: Z

1

L

< u (x), ϕ(x, 0) >1 − u0 (x)ϕt (x, 0) dx+ < f, ϕ(γ(t), t) >t1 0 Z TZ L = ψ(x, t)u(x, t) dx dt, for all ψ ∈ L1 (0, T ; L2 (0, L)), 0

(2.5)

0

where < ·, · >1 and < ·, · >t1 denote the duality products in H01 (0, L) and H 1 (0, T ) respectively. We adopt identity (2.5) as the definition of solutions of (1.1), in the sense of transposition. The following result establishes the existence of solutions for system (1.1). Theorem 2.1 Assume that γ : [0, T ] → (0, L) is in the class γ ∈ C 1 ([0, T ]) piecewise, i.e. there exists a partition of [0, T ], 0 = t0 < t1 < ... < tn = T , such that γ ∈ C 1 ([ti , ti+1 ]) for all i = 0, ..., n − 1. Assume also that this partition can be chosen in such a way that 1 − |γ 0 (t)| does not change the sign in t ∈ [ti , ti+1 ], for all i = 0, 1, ..., n − 1. Given any initial data (u0 , u1 ) ∈ L2 (0, L) × H −1 (0, 1) and f ∈ (H 1 (0, T ))0 , there exists an unique solution u of (1.1), in the sense of transposition, in the class ut ∈ L2 (0, T ; H −1 (0, L)).

u ∈ C([0, T ]; L2 (0, L)),

Moreover, there exists a one-to-one correspondence between the data and the solution in the given spaces. Concerning the exact controllability problem of system (1.1) we have to consider further hypothesis on the curve γ. 3

Definition 2.1 We say that a curve γ : [0, T ] → (0, L) is an ’admissible trajectory’ if system (1.1) is exactly controllable with controls in (H 1 (0, T ))0 , i.e. for any initial data (u0 , u1 ) ∈ L2 (0, L) × H −1 (0, L) and final data (v 0 , v 1 ) ∈ L2 (0, L) × H −1 (0, L), there exists a control f ∈ (H 1 (0, T ))0 such that the solution u of (1.1) satisfies (1.2). Remark 2.1 When dealing with control problems for the wave equation, the finite speed of propagation affects to the minimal time interval necessary for controllability [0, Tmin ]. The value Tmin is usually given by an optical geometric condition requiring that any ray, starting anywhere in the domain and with any initial direction, must meet the controllability zone before the time Tmin . According to this, an admissible trajectory requires both geometric constraints and a condition on the length of the time interval [0, T ], that must be sufficiently large. Theorem 2.2 Assume that γ ∈ C 1 ([0, T ]) piecewise and it satisfies the following hypothesis: 1. There exists constants c1 , c2 > 0 and a finite number of open subintervals Ij ⊂ [0, T ] with j = 0, ..., J such that, for each subinterval Ij , γ ∈ C 1 (Ij ) and it satisfies the following two conditions • c1 ≤ |γ 0 (t)| ≤ c2 for all t ∈ Ij , • 1 − |γ 0 (t)| does not change the sign in t ∈ Ij . We assume, without loss of generality, that there exists j1 with −1 ≤ j1 ≤ J such that γ(t) is decreasing in Ij for 0 ≤ j ≤ j1 , and γ(t) is increasing in Ij for j1 < j ≤ J. Here, j1 = −1 corresponds to the case where γ(t) is increasing in all the subintervals Ij . Analogously, j1 = J corresponds to the case where γ(t) is decreasing in all the subintervals Ij . 2. For each j = 0, ..., J, let Uj be the subintervals defined as follows:  {s − γ(s) with s ∈ Ij } if j ≤ j1 , Uj = {s + γ(s) with s ∈ Ij } if j > j1 . Then, there exists an interval W1 with length (W1 ) > 2L such that W1 ⊂

J [

Uj .

j=0

3. For each j = 0, ...J, let Vj be the subintervals defined as follows:  {s + γ(s) with s ∈ Ij } if j ≤ j1 , Vj = {s − γ(s) with s ∈ Ij } if j > j1 .

4

(2.6)

Then, there exists a subinterval W2 , with length (W2 ) ≤length (W1 ), such that J [

V j ⊂ W2 ,

(2.7)

j=0

and Vj ∩ Vk = ∅, if j 6= k. Then γ is admissible in the sense of Definition 2.1. The proof of the existence result above (Theorem 2.1) can be obtained from a suitable regularity property stated below (estimate (4.11)) by a duality argument. We refer to [11] for a general description, and [7] or [3] where this is done for similar problems. The novelty in the present situation is that the system (1.1) is not autonomous. We give the proof of Theorem 2.1 in section 6 below. The exact controllability property (Theorem 2.2) is also a straightforward consequence of the observability inequality (4.12) below and the Hilbert Uniqueness Method introduced by J.-L. Lions in [12]. We give the proof in section 7. We also refer to [7] and [3] where this method is applied for similar problems.

3

Admissible trajectories

In this section we show some examples of curves γ that satisfy the hypothesis of Theorem 2.2 and therefore are admissible trajectories for the exact controllability of system (1.1), in the sense of Definition 2.1. Example 1: Slowly decreasing trajectories: Assume that γ ∈ C 1 ([0, T ]) satisfies the following: 1. −1 ≤ γ 0 (t) ≤ −c1 for some constant c1 > 0 and for all t ∈ [0, T ], 2. T − γ(T ) + γ(0) > 2L. Then γ is an admissible trajectory in the sense of Definition 2.1. In fact we easily check that the hypothesis of Theorem 2.2 are satisfied with the interval I0 = (0, T ). In this case, U0 = (−γ(0), −γ(T ) + T ) = W1 and V0 ⊂ [γ(0), T + γ(T )] = W2 . The condition −1 ≤ γ 0 (t) guarantees that V0 is a proper interval, i.e. T + γ(T ) − γ(0) ≥ 0. In fact, this inequality is easily obtained by integrating −1 ≤ γ 0 (t) in t ∈ (0, T ). Note also that, by the hypothesis 2 above, the length of U0 is greater than 2L and the second hypothesis in Theorem 2.2 holds. On the other hand, the hypothesis 3 in Theorem 2.2 is easily checked since the length of U0 is also greater than the length of V0 due to the fact that γ is decreasing in [0, T ].

5

Figure 1: Slowly increasing (left) and decreasing trajectories that are admissible The second condition above, i.e. T − γ(T ) + γ(0) > 2L, provides an estimate of the minimal time for controllability which is probably not optimal. For example, if we take L ≥ 1, γ(t) = L/2 − t/4, and T > 8L/5 we have an admissible trajectory since it satisfies the conditions above. However, the time T = 8L/5 is strictly lower than the time given by the optical geometric condition T 0 = 4L/3, i.e. the time required for a ray starting at x0 = L/2 with initial velocity x0 (0) = 1, to meet the controllability zone, following the optical geometric laws (see Remark 2.1). Example 2: Slowly increasing trajectories: Assume that γ ∈ C 1 ([0, T ]) satisfies the following: 1. 0 < c1 ≤ γ 0 (t) ≤ 1 for all t ∈ [0, T ], 2. T + γ(T ) − γ(0) > 2L. Then γ is admissible. Again, we easily check that the hypothesis of Theorem are satisfied with the interval I0 = (0, T ).

Example 3: Rapidly decreasing trajectories: In this section we assume that γ 0 (t) ≤ −1. Obviously, such a trajectory will meet the boundary of the domain in short time. As the time for controllability is large in general, it is not expected to have an admissible trajectory γ ∈ C 1 ([0, T ]) with γ 0 (t) ≤ −1 for t ∈ (0, T ) (see Remark 2.1). Thus, we will assume that the trajectory is piecewise C 1 and it is decreasing only in some intervals, where it is also C 1 . More precisely, we assume that:

6

Figure 2: Rapidly decreasing trajectory that satisfies the conditions to be admissible. In particular we see that aj+1 − γ(aj+1 ) ≤ bj − γ(bj ) and bj+1 + γ(bj+1 ) ≥ aj + γ(aj ) (left). The result still holds if we complete the trajectory by piecewise C 1 curves in order to have a continuous trajectory (right). (a) There exists a family of open subintervals Ij = (aj , bj ) ⊂ (0, T ) with j = 0, ..., J, and the interval ω = (ω1 , ω2 ) ⊂ (0, L) such that γ : (aj , bj ) → (ω1 , ω2 ). (b) There exists c2 > 0 such that −c2 < γ 0 (t) ≤ −1 for all t ∈ (aj , bj ) and for all j = 0, 1, ..., J. (c) 0 = a0 < a1 < ... < aJ < T and 0 < b0 < b1 < ... < bJ = T . (d) aj+1 − γ(aj+1 ) ≤ bj − γ(bj ), for all j = 0, 1, ..., J − 1. (e) bj+1 + γ(bj+1 ) ≥ aj + γ(aj ), for all j = 0, 1, ..., J − 1. (f) γ(0) = ω2 , γ(T ) = ω1 . (g) T − γ(T ) + γ(0) > 2L. Then, γ is admissible. In figure 2 we show an example of such trajectory. It is interesting to note that the intervals Ij do not cover neccesarily the whole interval (0, T ). This is only the case if aj = bj−1 for all j = 1, ..., J. Obviously, this trajectory can be completed with piecisewise C 1 curves outside Ij and the result still holds.

Let us prove that γ satisfies the conditions of Theorem 2.2. The first condition is a direct consequence of the hypothesis (b) above. In order to check the second condition in Theorem 2.2

7

Figure 3: Rapidly decreasing trajectory that satisfies the conditions to be admissible. Here we show the geometrical interpretation of Uj and Vj . we define Uj = {s − γ(s) with s ∈ Ij } and W1 = (−ω2 , T − ω1 ). Then by hypothesis (d) and (f), W1 = (−γ(0), T − γ(T )) ⊂

J [

(aj − γ(aj ), bj − γ(bj )) =

j=0

J [

Uj ,

j=0

and by hypothesis (g) the length of W1 is equal or greater than 2L. Now we check the third condition in Theorem 2.2. Define Vj = {s + γ(s) with s ∈ Ij } and W2 = (b0 + ω1 , aJ + ω2 ). By hypothesis (c) and (e), Vj ∩ Vk = ∅ if j 6= k and J [

(bj + γ(bj ), aj + γ(aj )) ⊂ [b0 + γ(b0 ), aJ + γ(aJ )] ⊂ W2 .

(3.8)

j=0

The last inclusion is obtained from the fact that both γ(b0 ), γ(aJ ) ∈ (ω1 , ω2 ) which is deduced by hypothesis (a). On the other hand, W2 ⊂ (ω1 , T + ω2 ),

(3.9)

which has the same length as W1 . Thus, the third condition in Theorem 2.2 is a direct consequence of (3.8) and (3.9).

Example 4: Rapidly increasing trajectories: We assume that:

8

(a) There exists a family of open subintervals Ij = (aj , bj ) ⊂ (0, T ) with j = 0, ..., J, and the interval ω = (ω1 , ω2 ) ⊂ (0, L) such that γ : (aj , bj ) → ω. (b) There exists c2 > 0 such that 1 ≤ γ 0 (t) < c2 for all t ∈ (aj , bj ) and for all j = 0, 1, ..., J. (c) 0 = a0 < a1 < ... < aJ < T and 0 < b0 < b1 < ... < bJ = T . (d) aj+1 + γ(aj+1 ) ≤ bj + γ(bj ) . (e) bj+1 − γ(bj+1 ) ≥ aj − γ(aj ) for all j = 0, 1, ..., J − 1. (f) γ(0) = ω1 , γ(T ) = ω2 . (g) T + γ(T ) − γ(0) > 2L. Then, γ is admissible. The proof is analogous to the previous case. Example 5: Saw-teeth trajectories: As we said in Remark 2.1 above, when dealing with control problems for the wave equation, the minimal time for controllability Tmin is usually given by the a optical geometric condition. This example shows that this minimal time for controllability can be achieved with a particular class of saw-teeth trajectories, like in figure 4. To simplify we consider the case where γ is constituted by characteristic lines. We consider (ω1 , ω2 ) ⊂ (0, L), and L∗ > L to avoid the critical time for observability. We assume that there exists two natural numbers k, m such that L∗ = k(ω2 − ω1 ),

2ω2 = m(ω2 − ω1 ).

Let n = max{m − 1, 2k − m}. Consider the following periodic characteristic curve: 

−t + ω2 , if t ∈ [0, ω2 − ω1 ], t + 2ω1 − ω2 , if t ∈ [ω2 − ω1 , 2(ω2 − ω1 )], γ(t) = γ(t − j2(ω2 − ω1 )) for all t ∈ [(j − 1)2(ω2 − ω1 ), j2(ω2 − ω1 )],

γ(t) =

∀2 ≤ j ≤ n.

We define Ij

= ((j − 1)(ω2 − ω1 ), j(ω2 − ω1 )),

1 ≤ j ≤ n,

Uj

= {t − γ(t) with t ∈ Ij },

1 ≤ j ≤ n, and j odd.

Uj

= {t + γ(t) with t ∈ Ij },

1 ≤ j ≤ n, and j even.

Note that Vj is reduced to an unique point for all j ≥ 1, due to the fact that γ(t) is a characteristic curve in Ij . It is very easy to check that γ satisfies all the conditions in Theorem 2.2. The time of control is: T = max{2ω2 − (ω2 − ω1 ), 2(L∗ − ω2 )}. Note that, for L∗ = L we obtain the minimal time required for any ray to cross the curve γ with an angle larger than π. Here we consider any L∗ larger than L.

9

Figure 4: Saw-teeth trajectory that satisfies the conditions to be admissible with minimal time. Here ω1 = L∗ /4, ω2 = L∗ /2, and the time T = L∗ > L.

4

Observability

The main results in this paper can be obtained from some inequalities for the uncontrolled wave equation. In this section we prove these inequalities. Consider the system  in 0 < x < L, 0 < t < T,  ϕtt − ϕxx = 0, ϕ(0, t) = ϕ(L, t) = 0, in 0 < t < T, (4.10)  0 1 ϕ(x, 0) = ϕ (x), ϕt (x, 0) = ϕ (x), in 0 < x < L. We assume that (ϕ0 , ϕ1 ) ∈ H01 (0, L) × L2 (0, L). The following result holds: Proposition 4.1 Assume that γ satisfies the hypothesis of Theorem 2.1. Then, there exists a constant c(γ) > 0 such that the solution ϕ of (4.10) satisfies Z 0

T

2 d

ϕ(γ(t), t) dt ≤ c(γ) (ϕ0 , ϕ1 ) 2 1 2 . dt H0 ×L

(4.11)

Moreover, if γ satisfies the hypothesis of Theorem 2.2, then there exists a constant C(γ) > 0 such that 2 Z T

0 1 2 d

(ϕ , ϕ ) 1 2 ≤ C(γ) ϕ(γ(t), t) dt. (4.12) H0 ×L dt 0 Remark 4.1 Estimate (4.11) is a regularity result for the trace of the solution of the wave equation ϕ(x, t) on the curve defined by the trajectory γ. This result cannot be obtained from classical arguments or semigroup theory. 10

Estimate (4.12) is an observability inequality which establishes that the total energy of the solutions of the wave equation can be estimated from the value of the solution ϕ at γ(t) for a large enough time interval t ∈ (0, T ). The constant C(γ) in (4.12) depends on the bounds for γ 0 , c1 and c2 . As we show in the proof, it blows up as c1 → 0 or c2 → ∞. Proof. Note that it is enough to consider smooth solutions since for low regular ones we can argue by density. We first prove the estimate (4.11). Note also that it suffices to prove this inequality in the particular case in which γ ∈ C 1 ([0, T ]) and 1−|γ 0 (t)| does not change the sign in t ∈ [0, T ]. In the general case, for piecewise C 1 functions satisfying the hypothesis of the theorem, we can argue on each subinterval where the function is C 1 and then add the resulting inequalities. We observe that, in the one-dimensional wave equation, it is possible to change the variables x by t and t by x without altering the equation. Thus, D’Alembert formula can be used to obtain the solution ϕ(x, t) in terms of the solution at one extreme, say x = 0, instead of the data at t = 0 as usual. Indeed, we have Z 1 1 t+x ϕ(x, t) = [ϕ(0, t − x) + ϕ(0, t + x)] + ϕx (0, s) ds. (4.13) 2 2 t−x If ϕ is defined on (x, t) ∈ [0, L] × [0, T ] this formula holds only for those values (x, t) for which 0 ≤ t − x ≤ t + x ≤ T . However, we can extend the solution of the wave equation ϕ to (x, t) ∈ (0, L) × (−∞, ∞) and formula (4.13) is still valid for the whole domain (x, t) ∈ (0, L) × (0, T ). This is always possible because the wave equation with Cauchy data at t = 0 is well-posed for t ≥ 0 and t ≤ 0. In particular, taking into account the homogeneous Dirichlet boundary conditions in (4.10) we have Z 1 t+γ(t) ϕ(γ(t), t) = ϕx (0, s) ds. 2 t−γ(t) Therefore, 2

d ϕ(γ(t), t) = (1 + γ 0 (t))ϕx (0, t + γ(t)) − (1 − γ 0 (t))ϕx (0, t − γ(t)), dt

(4.14)

and we have, T

Z 4 0

Z

2 d ϕ(γ(t), t) dt dt

T

Z

2

T

(1 + γ (t)) |ϕx (0, t + γ(t))| dt + 2 (1 − γ 0 (t))2 |ϕx (0, t − γ(t))|2 dt 0 0 Z T |ϕx (0, t + γ(t))|2 |1 + γ 0 (t)| dt ≤ 2 sup |1 + γ 0 (t)| ≤ 2

0

2

0

t∈[0,T ] 0

Z

+2 sup |1 − γ (t)| t∈[0,T ]

T

|ϕx (0, t − γ(t))|2 |1 − γ 0 (t)| dt.

0

11

(4.15)

In order to simplify the two integrals on the right hand side we introduce a suitable change of variables. As we show below, this requires that 1 − |γ 0 (t)| does not change its sign in t ∈ [0, T ]. To fix ideas we assume that 1−|γ 0 (t)| ≥ 0 in t ∈ [0, T ]. The other case, i.e. when 1−|γ 0 (t)| ≤ 0 in t ∈ [0, T ], requires a similar analysis and we omit it. Let us introduce the set I1 ⊂ [0, T ] defined by I1 = {t ∈ [0, T ], such that 1 + γ 0 (t) 6= 0}. Note that the function s(t) = t + γ(t) is injective in I1 and therefore it can be used as a change of variables for the first integral in the right hand side of (4.15). In this way, we have T

Z

2

Z

0

|ϕx (0, t + γ(t))| |1 + γ (t)| dt = 0

Z

|ϕx (0, s)|2 ds ≤

= I1 +γ(I1 )

Z

I1 T +γ(T )

|ϕx (0, t + γ(t))|2 |1 + γ 0 (t)| dt |ϕx (0, s)|2 ds.

(4.16)

γ(0)

Analogously, if we define I2 ⊂ [0, T ] as I2 = {t ∈ [0, T ], such that 1 − γ 0 (t) 6= 0}, then s(t) = t − γ(t) is injective in I2 and therefore the second integral in the right hand side of (4.15) can be estimated as follows, T

Z

|ϕx (0, t − γ(t))|2 |1 − γ 0 (t)| dt =

0

Z

|ϕx (0, s)|2 ds ≤

=

Z

Z

I2 T −γ(T )

|ϕx (0, t − γ(t))|2 |1 − γ 0 (t)| dt |ϕx (0, s)|2 ds.

(4.17)

−γ(0)

I2 −γ(I2 )

Combining (4.15)-(4.17) we obtain, 2 d dt 4 ϕ(γ(t), t) dt 0 Z T +γ(T ) Z 2 0 0 ≤ 2 max |1 + γ (t)| |ϕx (0, s)| ds + 2 max |1 − γ (t)| Z

T

t∈[0,T ]

Z

≤ 2(1 + max |γ 0 (t)|) t∈[0,T ]

T +γ(T )

|ϕx (0, s)|2 ds +

Z

T +γ(T )

−γ(0)

Z

T −γ(T )

−γ(0)

|ϕx (0, s)|2 ds

−γ(0)

γ(0)

≤ 4(1 + max |γ 0 (t)|) t∈[0,T ]

t∈[0,T ]

γ(0)

T −γ(T )

2 |ϕx (0, s)|2 ds ≤ C (ϕ0 , ϕ1 ) H 1 ×L2 , 0

where C = 4(1 + max |γ 0 (t)|)(T + γ(T ) + γ(0) + 2L). t∈[0,T ]

12

|ϕx (0, s)|2 ds !

Here, the last inequality is the usual boundary direct inequality that can be obtained by classical multipliers techniques (see [12]). In fact, for any time interval (t1 , t2 ) the following holds: Z t2

2

|ϕx (0, t)|2 dt ≤ (t2 − t1 + 2L) (ϕ0 , ϕ1 ) H 1 ×L2 . 0

t1

To prove this inequality, we multiply the equation (4.10) by (L−x)ϕx and integrate in x ∈ (0, L). We easily obtain, 1 0= 2

Z

t2

Z

L 2

2

Z

(|ϕt | + |ϕx | ) dx dt − t1

0

t2

L

t2



(L − x)ϕx ϕt dx 0

Z

t1

t1

|ϕx (0, t)|2 dt. 2

Therefore, taking into account the conservation of the energy, i.e. k(ϕ(·, t), ϕt (·, t))kH 1 ×L2 is 0 constant in time, we have Z t2 Z t2 Z L Z L |ϕx (0, t)|2 dt = (|ϕt |2 + |ϕx |2 ) dx dt + 2 (L − x)ϕx (x, T )ϕt (x, T ) dx t1

t1

Z

0 L

0

−2

(L − x)ϕx (x, 0)ϕt (x, 0) dx Z L

0 1 2 

|ϕx (x, T )|2 + |ϕt (x, T )|2 dx ≤ (t2 − t1 ) (ϕ , ϕ ) H 1 ×L2 + L 0 0 Z L  |ϕx (x, 0)|2 + |ϕt (x, 0)|2 dx +L 0

2 = (t2 − t1 + 2L) (ϕ0 , ϕ1 ) H 1 ×L2 . 0

0

Now, we prove the estimate (4.12). To clarify the proof we divide it in several steps: Step 1 Our first objective is to prove the following inequality: There exist C > 0 and r < 1, independent of j = 0, ..., J, such that 2 Z Z Z d 2 2 ϕ(γ(t), t) dt. |ϕx (0, t)| dt − r |ϕx (0, t)| dt ≤ C (4.18) dt Ij Uj Vj The analysis is slightly different depending on the sign of γ 0 and we consider separately the two different cases. Case A: Assume that t ∈ Ij with j ≤ j1 . In this case −c2 ≤ γ 0 (t) ≤ −c1 < 0. From identity (4.14), we can estimate |ϕx (0, t − γ(t))| as follows: |ϕx (0, t − γ(t))|2  2 1 + γ 0 (t) 2 d = ϕ (0, t + γ(t)) − ϕ(γ(t), t) x 1 − γ 0 (t) 1 − γ 0 (t) dt 2    2 2 d 1 + γ 0 (t) 2 2 = |ϕ (0, t + γ(t))| + ϕ(γ(t), t) x 1 − γ 0 (t) 1 − γ 0 (t) dt 13

1 + γ 0 (t) d 2 ϕx (0, t + γ(t)) ϕ(γ(t), t) 0 0 1 − γ (t) 1 − γ (t) dt 2   2   d 2 1 1 + γ 0 (t) 2 2 |ϕx (0, t + γ(t))| + 1 + ϕ(γ(t), t) , ≤ (1 + a) 0 0 1 − γ (t) a 1 − γ (t) dt −2

for any a > 0 to be chosen later. Here we have used Young’s inequality. Multiplying by 1 − γ 0 (t) and integrating in t ∈ Ij we obtain, Z  |ϕx (0, t − γ(t))|2 1 − γ 0 (t) dt Ij Z 1 + γ 0 (t) 0 2 ≤ (1 + a) 1 − γ 0 (t) |ϕx (0, t + γ(t))| 1 + γ (t) dt Ij 2  Z d 1 4 + 1+ ϕ(γ(t), t) dt 0 a I 1 − γ (t) dt j 0 Z 1 + γ (t) |ϕx (0, t + γ(t))|2 1 + γ 0 (t) dt ≤ (1 + a) sup 0 t∈Ij 1 − γ (t) Ij 2   Z d 1 4 dt. + 1+ ϕ(γ(t), t) a 1 + c1 Ij dt

(4.19)

In order to simplify the first integral on the right hand side we follow the argument used for the first integral on the right hand side of (4.15). Again, this requires that 1 − |γ 0 (t)| does not change its sign in Ij . In this way, we easily obtain, Z Z 2 0 |ϕx (0, t + γ(t))| 1 + γ (t) dt ≤ |ϕx (0, s)|2 ds. (4.20) Ij

Vj

On the other hand, we simplify the integral on the left hand side of (4.19) with the change of variables s = t − γ 0 (t), Z Z  2 0 |ϕx (0, t − γ(t))| 1 − γ (t) dt = |ϕx (0, s)|2 ds. (4.21) Uj

Ij

Therefore, combining (4.19)-(4.21) we obtain, Z Z 1 + γ 0 (t) 2 |ϕx (0, t)|2 dt |ϕx (0, t)| dt − (1 + a) sup 0 t∈Ij 1 − γ (t) Uj Vj 2   Z T d 1 4 ϕ(γ(t), t) dt. ≤ 1+ a 1 + c1 0 dt In order to have the inequality (4.18) we only have to check the uniform bound 1 + γ 0 (t) ≤ c < 1, sup 0 t∈Ij 1 − γ (t) 14

(4.22)

(4.23)

for some constant c > 0, for all j = 0, ..., j1 . In fact, this estimate is easily checked from the hypothesis −c2 ≤ γ 0 (t) ≤ −c1 < 0. Once we have estimate (4.23), we choose 0 < a < 1/c − 1 in (4.22) and we obtain (4.18) with r = (1 + a)c. The constant r satisfies r = (1 + a)c → 1 if c1 → 0 or c2 → ∞. As we show below, the observability constant in (4.12) depends on 1/(1 − r) and therefore it blows up as c1 → 0 or c2 → ∞. Case B: Now, assume that t ∈ Ij with j > j1 , i.e. 0 < c1 ≤ γ 0 (t) ≤ c2 < ∞ for all t ∈ Ij . From identity (4.14), we estimate now |ϕx (0, t + γ(t))| as follows, |ϕx (0, t + γ(t))|2  2 1 − γ 0 (t) 2 d = ϕx (0, t − γ(t)) + ϕ(γ(t), t) 1 + γ 0 (t) 1 + γ 0 (t) dt 2    2 d 1 − γ 0 (t) 2 2 2 = |ϕx (0, t − γ(t))| + ϕ(γ(t), t) 0 0 1 + γ (t) 1 + γ (t) dt 0 1 − γ (t) 2 d +2 ϕx (0, t − γ(t)) ϕ(γ(t), t) 1 + γ 0 (t) 1 + γ 0 (t) dt 2    2  d 1 2 1 − γ 0 (t) 2 2 , |ϕ (0, t − γ(t))| + 1 + ϕ(γ(t), t) ≤ (1 + a) x 1 + γ 0 (t) a 1 + γ 0 (t) dt for any a > 0 to be chosen later. Multiplying by 1 + γ 0 (t) and integrating in t ∈ Ij we obtain, Z  |ϕx (0, t + γ(t))|2 1 + γ 0 (t) dt Ij Z 1 − γ 0 (t) 0 2 ≤ (1 + a) 1 + γ 0 (t) |ϕx (0, t − γ(t))| 1 − γ (t) dt Ij 2  Z d 1 4 ϕ(γ(t), t) dt. + 1+ 0 a Ij 1 + γ (t) dt

(4.24)

Therefore, arguing as in the previous case, we easily obtain, Z Z 1 − γ 0 (t) 2 |ϕx (0, t)| dt − (1 + a) sup |ϕx (0, t)|2 dt 0 t∈Ij 1 + γ (t) Vj Uj 2   Z d 1 4 dt. ≤ 1+ ϕ(γ(t), t) a 1 + c1 Ij dt 0 (t) Note that, in this case, we have the uniform bound supt∈Ij 1−γ 0 1+γ (t) ≤ c < 1 for some constant c > 0 and for all j > j1 . Then, as in the previous case, we can choose a such that (4.18) holds.

15

Step 2. Here we obtain the following inequality Z

2L

T

Z

2

|ϕx (0, t)| dt ≤ C

(1 − r)

0

0

2 d ϕ(γ(t), t) dt. dt

(4.25)

From (4.18) we obtain J Z X j=1

2

|ϕx (0, t)| dt − r

Uj

J Z X

2 J Z X d dt ϕ(γ(t), t) |ϕx (0, t)| dt ≤ C Vj Ij dt

j=1

Z

T

≤C 0

2

j=1

2 d ϕ(γ(t), t) dt. dt

(4.26)

On the other hand, the hypothesis (2.6) and (2.7) allow us to obtain, J Z X j=1

2

|ϕx (0, t)| dt − r

Uj

J Z X j=1

Z

2

2

|ϕx (0, t)| dt ≥

Z

|ϕx (0, t)| dt − r W1

Vj

|ϕx (0, t)|2 dt.

W2

Then, taking into account the 2L-periodicity of the solutions of the wave equation (4.10) and the fact that the length of W2 is lower than the length of W1 we have Z Z Z Z 2L 2 2 2 |ϕx (0, t)| dt − r |ϕx (0, t)| dt ≥ (1 − r) |ϕx (0, t)| dt ≥ (1 − r) |ϕx (0, t)|2 dt. W1

W2

W1

0

(4.27) This inequality together with (4.26) allows us to obtain (4.25). Step 3. Finally, inequality (4.12) is just a consequence of (4.25) and the following classical boundary inverse inequality for the wave equation (see [12]): For T ≥ 2L there exists a constant C > 0 such that Z 2L

k(ϕ0 , ϕ1 )k2H 1 ×L2 ≤ C 0

|ϕx (0, t)|2 dt.

0

This concludes the proof of Proposition 4.1.

5

Existence of weak solutions

In this section we prove Theorem 2.1. This requires some estimates on the nonhomogeneous problem (2.3) that we state first. Proposition 5.1 Assume that ψ ∈ L1 (0, T ; H −1 (0, L)) and γ a curve satisfying the hypothesis of Theorem 2.1. There exists a constant c(γ) such that the solution ϕ ∈ C([0, T ]; H01 (0, L)) ∩ C 1 ([0, T ]; L2 (0, L)) of (2.3) satisfies the folowing estimate, 2 Z T d ϕ(γ(t), t) dt ≤ c(γ) kψ(x, t)k2 1 (5.28) L (0,T ;L2 (0,L)) . dt 0 16

Assume now that ψ(x, t) = Ψt (x, t), with Ψ ∈ L2 (0, T ; L2 (0, L)), then the solution ϕ of (2.3) satisfies ϕ ∈ C(0, T ; H01 (0, L)) ∩ C 1 (0, T ; L2 (0, L)). Moreover, there exists a constant c˜(γ) > 0 such that 2 Z T d ϕ(γ(t), t) dt ≤ c˜(γ) kΨ(x, t)k2 2 (5.29) L (0,T ;H01 (0,L)) . dt 0 Let us complete the proof of Theorem 2.1 before proving this result. Let L(ψ) be the linear operator defined by the left hand side of (2.5), i.e. L(ψ) = < u1 (x), ϕ(x, 0) >1 −

Z

L

u0 (x)ϕt (x, 0) dx+ < f (t), ϕ(γ(t), t) >t1 .

(5.30)

0

From estimate (5.28), L(ψ) is continuous in L1 (0, T ; L2 (0, L)) and it defines a unique u ∈ L∞ (0, T ; H −1 ) satisfying Z

T

Z

L

for all ψ(x, t) ∈ L1 (0, T ; L2 (0, L)).

u(x, t)ψ(x, t) dx dt = L(ψ), 0

(5.31)

0

Therefore u solves (2.5). Moreover, for some constant C > 0, kukL∞ (0,T ;L2 (0,L)) ≤ C(ku0 kL2 (0,L) + ku1 kH −1 (0,L) + kf k(H 1 (0,L))0 ). The continuity t → u(t, x) from [0, T ] → L2 (0, L) is easily obtained by a density argument since the solutions of (1.1) with L2 −controls and smooth initial data satisfy u ∈ C([0, T ]; L2 (0, L)) (see [14]). Now we focus on the regularity for ut . We consider in (5.31) ψ(x, t) = Ψt (x, t) where Ψ ∈ D((0, T ) × (0, L)), the space of C ∞ functions with compact support. From Proposition 5.1 the map Ψ → L(Ψt ) is linear and continuous in L2 (0, T ; L2 (0, L)). Thus, we deduce that ut ∈ L2 (0, T ; L2 (0, L)). This concludes the proof of Theorem 2.1. Proof of Proposition 5.1. The main idea is to apply Duhamel’s principle to reduce the nonhomogeneous problem (2.3) to a suitable homogeneous one with nonzero final data for which estimate (4.11) holds. We observe that ϕ(x, t) can be written as T

Z ϕ(x, t) =

U (x, t.s)ds, t

where U (x, t, s) is the solution of  Utt − Uxx = 0, 0 < x < L,    U (x, s, s) = 0, 0 < x < L, U (x, s, s) = ψ(x, s), 0 < x < L,  t   U (0, t, s) = U (L, t, s) = 0,

17

0 < t < s < T, 0 < s < T, 0 < s < T, 0 < t < s < T.

From Proposition 4.1 and the conservation of the energy, there exists a constant c(γ) > 0 such that, 2 Z s   d 2 U (γ(t), t, s) dt ≤ c(γ) ||U (·, 0, s)||2 1 + ||U (·, 0, s)|| t L2 (0,L) H0 (0,L) 0 dt = c(γ)||ψ(·, s)||2L2 (0,L) .

(5.32)

On the other hand, we observe that Z Z T Z T d d T d d ϕ(γ(t), t) = U (γ(t), t, s)ds = U (γ(t), t, s)ds = χ(t,T ) (s) U (γ(t), t, s)ds, dt dt t dt t dt 0 where χ(t,T ) (s) is the characteristic function of the interval (t, T ). Therefore, from Minkowski inequality we have

Z T

d

d

ϕ(γ(t), t)

≤ U (γ(t), t, s) ds χ (s) (t,T )

dt

2

2 dt 0 L (0,T ) L (0,T )

Z T

d

= ds. (5.33)

dt U (γ(t), t, s) 2 0 L (0,s) Combining (5.32) and (5.33) we easily obtain the estimate (5.28). Now we prove the second part of Proposition 5.1. Assume that ψ(x, t) = Ψt (x, t), with Ψ ∈ L2 (0, T ; L2 (0, L)). Following the previous argument based on Duhamel’s principle we write Z T ϕ(x, t) = V (x, t.s)ds t

where V (x, t, s) is the solution of  Vtt − Vxx = 0, 0 < x < L, 0 < t < s < T,    V (x, s, s) = −Ψ(x, s), 0 < x < L, V (x, s, s) = 0, 0 < x < L, 0 < s < T,  t   V (0, t, s) = V (L, t, s) = 0, 0 < t < s < T. Note that V ∈ C(0, T ; H01 (0, L)) ∩ C 1 (0, T ; L2 (0, L)) and therefore the same is true for ϕ. From Proposition 4.1 and the conservation of energy we deduce 2 Z s d V (γ(t), t, s) dt ≤ c(γ)||ψ(·, s)||2 1 (5.34) H0 (0,L) . 0 dt On the other hand, we observe that Z Z T d d d T ϕ(γ(t), t) = V (γ(t), t, s)ds = V (γ(t), t, s)ds + Ψ(γ(t), t) dt dt t t dt Z T d = χ(t,T ) (s) V (γ(t), t, s)ds + Ψ(γ(t), t). dt 0 18

Therefore,

Z T

d

ϕ(γ(t), t) ≤

dt

2 0 L (0,T )



d

V (γ(t), t, s) ds + kΨ(γ(t), t)kL2 (0,T ) .

2

dt L (0,s)

(5.35)

The first term on the right hand side can be bounded using (5.34). On the other hand, taking into account that L∞ (0, L) ⊂ H01 (0, L), we have kΨ(γ(t), t)kL2 (0,T ) ≤ kΨkL2 (0,T ;L∞ (0,L)) ≤ kΨkL2 (0,T ;H 1 (0,L)) . 0

(5.36)

From (5.34), (5.35) and (5.36) we easily obtain the inequality (5.29).

6

Controllability

In this section we briefly describe how to obtain the controllability result stated in Theorem 2.2 from the observability result in Proposition 4.1. This is a straightforward application of HUM-method. The first step is to characterize the controls of (1.1). Let us introduce the adjoint system  in x ∈ (0, L), t ∈ (0, T )  ϕtt − ϕxx = 0 ϕ(0, t) = ϕ(L, t) = 0 in t ∈ (0, T ), (6.37)  ϕ(x, T ) = ϕ0 , ϕt (x, T ) = ϕ1 , in x ∈ (0, L). with (ϕ0 , ϕ1 ) ∈ H01 (0, L) × L2 (0, L). Multiplying the equations in (1.1) by ϕ and integrating by parts we easily obtain, at least formally, the following identity: Z L 1 < u (x), ϕ(x, 0) >1 − u0 (x)ϕt (x, 0) dx+ < f (t), ϕ(γ(t), t) >t1 0 Z L = < ut (x, T ), ϕ0 (x) >1 − u(x, T )ϕ1 (x) dx, (6.38) 0

where < ·, · >1 denotes the duality product in H01 (0, L) and < ·, · >t1 the duality product in H 1 (0, T ). Note that weak solutions must satisfy this identity. Therefore, a control f can be characterized as a function f (t) which satisfies the identity Z L 1 < u (x), ϕ(x, 0) >1 − u0 (x)ϕt (x, 0) dx+ < f (t), ϕ(γ(t), t) >t1 dt = 0, 0

(ϕ0 , ϕ1 )

H01 (0, L)

L2 (0, L).

for all ∈ × Let us now introduce the functional J : H01 (0, L) × L2 (0, L) → IR given by 2 Z Z L 1 T d dt+ < u1 (x), ϕ(x, 0) >1 − ϕ(γ(t), t) u0 (x)ϕt (x, 0) dx, J(ϕ0 , ϕ1 ) = 2 0 dt 0 19

where ϕ is the solution of (6.37) associated to the final data (ϕ0 , ϕ1 ). For T sufficiently large, this functional is continuous convex and coercive, due to the inverse inequality stated in (4.12). Therefore it has an unique minimizer (ϕ0m , ϕ1m ) ∈ H01 (0, L) × L2 (0, L). The minimizer satisfies Z 0

T

d d ϕm (γ(t), t) ϕ(γ(t), t)dt+ < u1 (x), ϕ(x, 0) >1 − dt dt

Z

L

u0 (x)ϕt (x, 0) dx = 0,

0

for all (ϕ0 , ϕ1 ) ∈ H01 (0, L) × L2 (0, L). Thus, a control is given by f ∈ (H01 (0, T ))0 such that < f (t), η(t)

>t1 =

Z 0

T

d dη ϕm (γ(t), t) (t) dt, dt dt

for all η ∈ H 1 (0, T ).

Acknowledgments Supported by the Grant MTM2008-03541 of the MICINN.

References [1] S. Avdonin and S. Ivanov, Families of exponentials: The method of moments in controllability problems for distributed paramenter systems, Cambridge University Press, 1995. [2] A. Bamberger, J. Jaffre and J.P. Yvon, Punctual control of a vibrating string: Numerical analysis, Comp. and Maths with appls. 4 113-138. [3] C. Castro, Boundary controllability of the one-dimensional wave equation with rapidly oscillating density. Asymptotic Analysis, 20, 317-350, 1999. [4] C. Castro and E. Zuazua, Unique continuation and control for the heat equation from a lower dimensional manifold, SIAM J. Cont. Optim, 42 (4), 1400-1434. [5] S. Cox and E. Zuazua, The rate at which energy decays in a string damped at one end, Indiana University Mathematics Journal, 44 (2) (1995), 545-573. [6] R. D´ager and E. Zuazua, Wave propagation observation and control in 1-d flexible multistructures, Mathmatiques et Applications, Vol. 50, Springer Verlag, 2006. [7] S. Hansen and E. Zuazua, Exact Controllability and Stabilization of a Vibrating String with an interior Point Mass, SIAM J. Control Optim., 33, 5 (1995), 1357-1391. [8] A. Khapalov, Mobile point controls versus locally distributed ones for the controllability of the semilinear parabolic equation, SIAM J. Cont. Optim., 40 (1), (2001) 231-252. [9] A. Khapalov, Controllability of the wave equation with moving point control, Applied Mathematics and Optimization, 31 (2), (1995) 155-175.

20

[10] A. Khapalov, Observability and stabilization of the vibrating string equipped with bouncing point sensors and actuators, Mathematical Methods in the Applied Sciences, 44 (14), (2001) 1055-1072. [11] J.-L. Lions and E. Magenes, Non-homogeneous boundary value problems and applications I, Springer-Verlag, 1972. [12] J.-L. Lions, Contrˆ olabilit´e exacte, stabilisation et perturbations de syst`emes distribu´es. Tomes 1 & 2. Masson, RMA 8 & 9, Paris, 1988. [13] J.-L. Lions, Some methods in the mathematical analysis of systems and their control, Gordon and Breach, 1981. [14] J.-L. Lions, Pointwise control for distributed systems, in Control and estimation in distributed parameter systems, edited by H.T. Banks, SIAM, 1992.

21

Exact controllability of the 1-d wave equation from a ...

data (u0,u1) and final data (v0,v1), find a control f such that the solution u of (1.1) satisfies ...... operator defined by the left hand side of (2.5), i.e. ..... [8] A. Khapalov, Mobile point controls versus locally distributed ones for the controllability of.

278KB Sizes 1 Downloads 168 Views

Recommend Documents

Exact boundary controllability of two Euler-Bernoulli ...
Dec 20, 2002 - For the system (1.3), when γ > 0, it turns that the space of controllable initial data can not be found among the family of energy spaces but it is ...

Boundary controllability of the one-dimensional wave ...
of the solutions is of the order of the size of the microstructure. More precisely, as it was proved in 2], there exist stationary solutions which concentrate most of its ...

Cylindrical wave, wave equation, and method of ...
Sep 19, 2007 - Schrodinger wave equation in which the method of separation of variables is used for obtaining the general spreading of the wave.) 2 Invalidity of the separation of variables for obtaining cylindrical wave function from the wave equati

Solving the tensorial 3D acoustic wave equation: A ...
finite-difference time-domain approach. Jeffrey Shragge1 and Benjamin Tapley2. ABSTRACT. Generating accurate numerical solutions of the acoustic wave.

New Exact Solution of Dirac-Coulomb Equation with ... - Springer Link
Sep 19, 2007 - brings on that the solutions of the Klein-Gordon equation and the Dirac ... and its magnetic moment in a completely natural way and so on.

Universal Wave Equation Worksheet.pdf
... dock measure about 1.5 m. from crest to crest and appear to be traveling at about 3.0 m/s. Determine the frequency and period of. these waves. Page 1 of 2 ...

Universal Wave Equation Worksheet.pdf
Whoops! There was a problem loading this page. Universal ... ksheet.pdf. Universal ... ksheet.pdf. Open. Extract. Open with. Sign In. Details. Comments. General Info. Type. Dimensions. Size. Duration. Location. Modified. Created. Opened by me. Sharin

THREE-COMPONENT 1D VISCOELASTIC FDM FOR PLANE-WAVE ...
May 10, 2010 - It may be an efficient tool for pre- or post-analysis of local structural ... calculated by semi-analytical methods such as the propagator matrix method [e.g. ..... J. M. Carcione, D. Kosloff and R. Kosloff, Geophys. J. Roy. Astron. So

three-component 1d viscoelastic fdm for plane- wave ...
Our FD scheme uses a 1D grid, which leads to a significant ... is a multi-component 1D viscoelastic finite-difference scheme for plane-wave simulation, which ...

Solving the 3D acoustic wave equation on generalized ...
difference time-domain (FDTD) methods represent the most com- mon numerical approach for generation solutions of the 3D acoustic wave equation, and have ...

exact solution to time-dependent Schrödinger Equation - Nature
Jan 28, 2016 - EF and W are the Fermi energy and the work function of the metal ... the conductivity σ= 4.1 × 107 S/m, for 800 nm laser, δm = 4.06 nm).

CHECKERBOARD MODES AND WAVE EQUATION 1 ...
mesh, the checkerboard modes do not exist at low Mach number when the compressible Navier-Stokes ...... [14] Guillard H. and Murrone A. – Behavior of upwind scheme in the low Mach number limit: III. Preconditioned dissipation for a five equation tw

Least-squares shot-profile wave-equation migration
example, allowing for migration velocity models that vary in depth only (Gazdag, ... for example, generalized inversion can compensate for incomplete data.

Least-squares shot-profile wave-equation migration
Department of Physics, University of Alberta, Edmonton, Alberta, Canada .... (with a chosen parameterization) are migrated images, c is the Earth's velocity, and ...

Nelson, Derivation of the Schrodinger Equation from Newtonian ...
Nelson, Derivation of the Schrodinger Equation from Newtonian Mechanics.pdf. Nelson, Derivation of the Schrodinger Equation from Newtonian Mechanics.pdf.

Extended wave-equation imaging conditions for ...
Sep 7, 2015 - The extended image volume is robust to sparsely sampled data and high levels ... 1The University of Western Australia, School of Earth and Environment and School of Physics, Centre for Energy Geoscience, Crawley, Australia. E-mail: ...

Controllability of complex networks.pdf
Page 1 of 7. ARTICLE. doi:10.1038/nature10011. Controllability of complex networks. Yang-Yu Liu1,2, Jean-Jacques Slotine3,4 & Albert-La ́szlo ́ Baraba ́si1,2,5. The ultimate proof of our understanding of natural or technological systems is reflect

The EQUATION of STATE of a PERFECT GAS
i u m p e n t n r e d d t m i n & i un c o r p emtreteou bi uoe tpmphtbw moindre, et ... lure I - dl du Nrpr B. Par suite de eel ahisremen1 de temperature ,. In presioo ...

C3-L4 - Writing the equation from a graph and Families of Polynomials ...
C3-L4 - Families of Polynomials - Practice Questions. Page 3 of 4. C3-L4 - Writing the equation from a graph and Families of Polynomials.pdf. C3-L4 - Writing the ...

Warm Up: Wave Characteristics Wave A Wave B
A sound wave is shown with a solid line. Draw in a new wave that would have lower volume and a higher pitch on the same graph. Wave A. Wave B.

More finding the equation of a graph practice
More finding the equation of a graph practice. Page 2. More finding the equation of a graph practice. Page 3. More finding the equation of a graph practice.

of the Hamilton-Jacobi Equation
Tbe second author wishes to express his indebtedness to the Office of Naval Research for a research contract at Indiana University. 939. Journal of Mathematics ...

The Equation of a Line in Standard Form
There is more than one way to write the equation of a line. We have already seen slope y-intercept form, and today we will look at the standard form of.