This investigative study explored the field of electrical discharge machining (EDM), with a particular focus on the machining of Ti6Al4V, a titanium alloy that finds widespread application in aerospace, airframes, engine components, and non-aerospace applications such as power generation and marine and offshore environments. Ti6Al4V presents difficulties for conventional metal cutting techniques because of high cutting forces, poor surface integrity, and tool wear. This has led to the adoption of unconventional techniques like EDM. However, problems like high electrode wear rates, low material removal rates, long machining times, and less-than-ideal surface finishes still exist, especially in large-scale applications. By addressing the particular difficulties associated with large-scale electrical discharge machining and by putting forth a multi-objective optimization strategy, this research makes a substantial contribution to the field. With an emphasis on the optimization of input parameters like pulse on time (Ton), pulse off time (Toff), voltage (HV), and current (LV), which are critical in large-scale industrial applications, the study attempts to evaluate the optimal parameter states that simultaneously accomplish multiple goals during the machining process. This work is the first to simultaneously optimize all relevant output responses, such as material removal rate (MRR), electrode wear rate (EWR), machining time (Tm), surface roughness (Ra), and base radius. Previous studies have concentrated on one or two output responses. To optimize MRR, EWR, Tm, Ra, and base radius, the experiments were carefully planned using design of experiment (DOE) and response Surface methodology (RSM). Regression analysis and ANOVA are two statistical techniques that were used with Minitab 15 to help interpret experimental data and build a solid regression model specifically for Ti6Al4V. Throughout the experiment, a variety of input factor settings were employed, and the responses to those were noted. The following parameters were used to obtain the experimental data: current (LV) at 30 and 50 A, voltage (HV) at 0.3 and 0.7 V, pulse on time (Ton) at 4 and 6.5 µs, and pulse off time (Toff) at 5.5 and 6.5 µs. Ton and current are the most significant variables that influence most of the output responses. By addressing the simultaneous optimization of multiple output responses, this investigative study not only sets a new standard in the field but also identifies current bottlenecks and offers solutions.
Citation: Muhammad Mansoor Uz Zaman Siddiqui, Syed Amir Iqbal, Ali Zulqarnain, Adeel Tabassum. An investigative study on the parameters optimization of the electric discharge machining of Ti6Al4V[J]. Clean Technologies and Recycling, 2024, 4(1): 43-60. doi: 10.3934/ctr.2024003
[1] | Federico Cluni, Vittorio Gusella, Dimitri Mugnai, Edoardo Proietti Lippi, Patrizia Pucci . A mixed operator approach to peridynamics. Mathematics in Engineering, 2023, 5(5): 1-22. doi: 10.3934/mine.2023082 |
[2] | Serena Federico, Gigliola Staffilani . Sharp Strichartz estimates for some variable coefficient Schrödinger operators on R×T2. Mathematics in Engineering, 2022, 4(4): 1-23. doi: 10.3934/mine.2022033 |
[3] | Annalisa Cesaroni, Matteo Novaga . Second-order asymptotics of the fractional perimeter as s → 1. Mathematics in Engineering, 2020, 2(3): 512-526. doi: 10.3934/mine.2020023 |
[4] | Alessandra De Luca, Veronica Felli . Unique continuation from the edge of a crack. Mathematics in Engineering, 2021, 3(3): 1-40. doi: 10.3934/mine.2021023 |
[5] | Plamen Stefanov . Conditionally stable unique continuation and applications to thermoacoustic tomography. Mathematics in Engineering, 2019, 1(4): 789-799. doi: 10.3934/mine.2019.4.789 |
[6] | Bin Deng, Xinan Ma . Gradient estimates for the solutions of higher order curvature equations with prescribed contact angle. Mathematics in Engineering, 2023, 5(6): 1-13. doi: 10.3934/mine.2023093 |
[7] | Zaffar Mehdi Dar, M. Arrutselvi, Chandru Muthusamy, Sundararajan Natarajan, Gianmarco Manzini . Virtual element approximations of the time-fractional nonlinear convection-diffusion equation on polygonal meshes. Mathematics in Engineering, 2025, 7(2): 96-129. doi: 10.3934/mine.2025005 |
[8] | José Antonio Vélez-Pérez, Panayotis Panayotaros . Wannier functions and discrete NLS equations for nematicons. Mathematics in Engineering, 2019, 1(2): 309-326. doi: 10.3934/mine.2019.2.309 |
[9] | Catharine W. K. Lo, José Francisco Rodrigues . On the obstacle problem in fractional generalised Orlicz spaces. Mathematics in Engineering, 2024, 6(5): 676-704. doi: 10.3934/mine.2024026 |
[10] | Prashanta Garain, Kaj Nyström . On regularity and existence of weak solutions to nonlinear Kolmogorov-Fokker-Planck type equations with rough coefficients. Mathematics in Engineering, 2023, 5(2): 1-37. doi: 10.3934/mine.2023043 |
This investigative study explored the field of electrical discharge machining (EDM), with a particular focus on the machining of Ti6Al4V, a titanium alloy that finds widespread application in aerospace, airframes, engine components, and non-aerospace applications such as power generation and marine and offshore environments. Ti6Al4V presents difficulties for conventional metal cutting techniques because of high cutting forces, poor surface integrity, and tool wear. This has led to the adoption of unconventional techniques like EDM. However, problems like high electrode wear rates, low material removal rates, long machining times, and less-than-ideal surface finishes still exist, especially in large-scale applications. By addressing the particular difficulties associated with large-scale electrical discharge machining and by putting forth a multi-objective optimization strategy, this research makes a substantial contribution to the field. With an emphasis on the optimization of input parameters like pulse on time (Ton), pulse off time (Toff), voltage (HV), and current (LV), which are critical in large-scale industrial applications, the study attempts to evaluate the optimal parameter states that simultaneously accomplish multiple goals during the machining process. This work is the first to simultaneously optimize all relevant output responses, such as material removal rate (MRR), electrode wear rate (EWR), machining time (Tm), surface roughness (Ra), and base radius. Previous studies have concentrated on one or two output responses. To optimize MRR, EWR, Tm, Ra, and base radius, the experiments were carefully planned using design of experiment (DOE) and response Surface methodology (RSM). Regression analysis and ANOVA are two statistical techniques that were used with Minitab 15 to help interpret experimental data and build a solid regression model specifically for Ti6Al4V. Throughout the experiment, a variety of input factor settings were employed, and the responses to those were noted. The following parameters were used to obtain the experimental data: current (LV) at 30 and 50 A, voltage (HV) at 0.3 and 0.7 V, pulse on time (Ton) at 4 and 6.5 µs, and pulse off time (Toff) at 5.5 and 6.5 µs. Ton and current are the most significant variables that influence most of the output responses. By addressing the simultaneous optimization of multiple output responses, this investigative study not only sets a new standard in the field but also identifies current bottlenecks and offers solutions.
Higher order local and nonlocal elliptic equations arise naturally in problems from conformal geometry and scattering theory [1,2], (nonlinear) higher order elliptic PDEs and free boundary value problems [3,4], control theory [5,6] and inverse problems [7,8,9]. Motivated by these applications, in this article we study the strong unique continuation property for higher order fractional Schrödinger equations. More precisely, here we are concerned with equations of the form
(−Δ)γu+qu=0 in Rn, | (1.1) |
where γ∈R+∖N with suitable, possibly singular (critical and subcritical) potentials q. Here we say that a solution u to (1.1) satisfies the strong unique continuation property if the condition that u vanishes of infinite order at a point x0∈Rn, i.e., if for all m∈N
limr→0r−m‖u‖2L2(Br(x0))=0, |
already implies that u≡0 in Rn. The strong unique continuation property can hence be viewed as a generalisation of analyticity to rougher equations.
Apart from dealing with the model setting of equation (1.1), we also address the corresponding differential inequalities and the setting of variable coefficient fractional Schrödinger operators with coefficients which might be only of low regularity. This provides new proofs for the results from [10] and [11], where respectively C2 and C0,1 regular coefficients had been treated in the case γ∈(0,1), and extends appropriately adapted versions of these low regularity results to higher order equations. Further, we discuss possible applications of the unique continuation results to inverse and control theoretic problems.
Let us outline our main results: As a model situation we deal with the strong unique continuation property (SUCP) for Schrödinger equations of the form (1.1). Without loss of generality, here we normalise our set-up such that x0=0.
Theorem 1 (SUCP). Let γ∈R+∖N and let u∈H2γ(Rn) be a solution to (1.1), where the potential q satisfies the following bounds
|q(x)|≤{Cq|x|−2γ,if γ>12,c0|x|−2γ,if γ=12,Cq|x|−2γ+ϵ,if γ∈(14,12). |
Here c0>0 is a sufficiently small constant, and Cq>0 is an arbitrarily large, finite constant. Assume that u vanishes of infinite order at x0=0, i.e., for all m∈N
limr→0r−m‖u‖2L2(Br(0))=0. |
Then u≡0 in Rn.
Remark 1.1. We remark that the limitation of the result to γ>14 arises naturally and was already present in [10]: Relying on Carleman estimates with weights which only have a radial dependence, we do not directly obtain positive boundary contributions but have to derive these through boundary-bulk interpolation estimates (see Lemma 2.2). With respect to bulk estimates, L2 Carleman estimates are however subelliptic in the large parameter τ. Through the boundary-bulk estimates this is propagated to the boundary which is then reflected in the loss of a quarter derivative in τ on the boundary. In the case that one only considers radial Carleman weights this loss seems unavoidable. In order to extend the unique continuation results to the regime γ∈(0,14) in a setting where only radial Carleman weights are used, the loss in τ has thus to be compensated by regularity of the potential (see [10] for corresponding results). In this case the lower order contributions would be included in the main part of the operator in the Carleman estimates (this then allows one to treat exact Hardy type potentials, but any type of perturbation of such potentials will need to obey regularity assumptions).
One could hope to avoid this loss of derivatives by considering Carleman weights which are not only of a radial structure but also depend on the normal directions. However, due to the weighted form of the inequalities, the exact Lopatinskii type conditions necessary for this are not immediate. We do not pursue this further in this article but postpone this to future work.
Remark 1.2. It would also have been possible to treat additional nonlinear terms in the the equations. As we are mainly interested in the associated differential inequalities, we do not consider them here.
Motivated by the work on the strong unique continuation properties on higher order elliptic equations (see [12] and the references therein), it is natural to wonder whether it is possible to extend the unique continuation property to (Hardy type higher) gradient potentials. Using iterative applications of our main Carleman estimate, we note that this is indeed the case:
Theorem 2 (SUCP with gradient potentials). Let γ∈R+∖N and let u∈H2γ(Rn) be a weak solution to the differential inequality
|(−Δ)γu(x)|≤⌊γ⌋∑j=0|qj(x)||∇ju(x)|in Rn, |
where the potentials qj satisfy the following bounds
|qj(x)|≤Cqj|x|−2γ+jif j<⌊γ⌋,|q⌊γ⌋(x)|≤{Cq⌊γ⌋|x|−2γ+⌊γ⌋,if γ−⌊γ⌋>12,c0|x|−2γ+j,if γ−⌊γ⌋=12,Cq⌊γ⌋|x|−2γ+⌊γ⌋+ϵ,if {γ∈(14,12),⌊γ⌋≥1 and γ−⌊γ⌋∈(0,12), |
Here c0>0 is a sufficiently small constant, and Cqj>0 are arbitrarily large, finite constants. Assume that u vanishes of infinite order at x0=0. Then u≡0 in Rn.
Remark 1.3. As in [12] it would have been possible to extend this result even further to (slightly subcritical) gradient potentials involving also contributions |qj(x)||∇ju(x)| with j∈(⌊γ⌋,32⌊γ⌋). As this requires some extra care, we do not present the details of this here but refer to the ideas in [12].
Further, relying on the methods from [10,13] as well as the spliting argument from [8], we also address the case with variable coefficient metrics and study the unique continuation properties at a point x0∈Rn. In the sequel, without loss of generality, we will normalise our set-up such that x0=0. Under this assumption, we consider the operator
L=−∇⋅˜a∇ |
for Lipschitz metrics with the following structural conditions:
(A1) ˜a:Rn→Rn×n is symmetric, (strictly) positive definite, bounded.
(A2) ˜a∈Cμ,1loc(Rn,Rn×nsym) with [˜aij]˙Cμ,1(B′4)+[˜aij]˙C0,1(B′4)≪δ, where B′4:={x∈Rn: |x|≤4}, for some small parameter δ>0. The constant μ>0 is specified below.
(A3) We assume that ˜aij(0)=δij.
For this class of coefficients, we can prove the analogue of Theorem 2:
Theorem 3 (SUCP with variable coefficients). Let γ∈R+∖N and let u∈Dom(Lγ) be a solution to
|(−∇⋅˜a∇)γu(x)|≤⌊γ⌋∑j=0|qj(x)||∇ju(x)|in Rn, | (1.2) |
where
● the metric ˜a satisfies the conditions (A1)–(A3) with μ=2⌊γ⌋,
● the potentials qj satisfy the following bounds
|qj(x)|≤Cqj|x|−2γ+jif j<⌊γ⌋,|q⌊γ⌋(x)|≤{Cq⌊γ⌋|x|−2γ+⌊γ⌋,if γ−⌊γ⌋>12,c0|x|−2γ+j,if γ−⌊γ⌋=12,Cq⌊γ⌋|x|−2γ+⌊γ⌋+ϵ,if {γ∈(14,12),⌊γ⌋≥1 and γ−⌊γ⌋∈(0,12), |
where c0>0 is a sufficiently small constant, and Cqj>0 are arbitrarily large, finite constants.
Then the strong unique continuation property holds at x0=0, i.e., if u vanishes of infinite order at x0=0, then u≡0 in Rn.
Remark 1.4. As explained in the Appendix (Section A.2), we interpret the variable coefficient fractional Laplacian through its spectral decomposition as directly related to a generalised Caffarelli-Silvestre extension. Relying on spectral theory in order to establish this, we restrict to functions u∈Dom(Lγ). Using ideas as outlined in [14] and [15], it would also have been possible to lower the required regularity of u in this discussion.
Remark 1.5. Based on counterexamples to the weak unique continuation property with metrics of any C0,α Hölder regularity with α∈(0,1) due to Miller [16] and Mandache [17], it is expected that the coefficient regularity stated in condition (A2) in our variable coefficient strong unique continuation result is optimal in the case ⌊γ⌋=0. This strengthens the results from [10,Section 7] and [11].
The condition (A3) is to be read as a normalisation condition which can be assumed without loss of generality. We remark that condition (A2) together with interpolation estimates implies that [aij]˙Cℓ,1(B′4)≤˜Cδ for any ℓ∈{1,…,μ}.
In both the settings of Theorems 1 and 3 also the unique continuation property from measurable sets (MUCP) holds:
Theorem 4 (MUCP). Let γ∈R+∖N and let u∈Dom(Lγ) be a solution to
|(−∇⋅˜a∇)γu(x)|≤|q(x)||u(x)|in Rn, | (1.3) |
where ˜aij satisfies the conditions (A1)–(A3) with μ=2⌊γ⌋ and q∈L∞(Rn). If there exists a measurable set E⊂Rn with |E|>0 and density one at x0=0 such that u|E=0, then u≡0 in Rn.
Let us comment on the results of Theorems 1–4 in the context of the literature on fractional Schrödinger equations: The weak unique continuation property, i.e., the question whether for solutions u of (1.1) the condition that u=0 on an open set in Rn already implies that u≡0 in the whole of Rn is rather well understood for constant coefficient fractional Schrödinger equations, even for potentials in very rough, non-L2-based function spaces (see [18]). In contrast, the understanding of the strong unique continuation properties of solutions to (higher order) fractional Schrödinger equations is still much less developed (for non-fractional higher order Schrödinger operators we refer to [12] and the references therein). The main known results are here given in the regime γ∈(0,1) and can be summarised as the following statements:
● Strong unique continuation for constant coefficient fractional Schrödinger equations with essentially L∞ potentials. In the regime γ∈(0,1) the articles [10,19] deal with the strong unique continuation property for L∞ as well as "Hardy type" critical and subcritical potentials. Both results crucially exploit the possibility of rephrasing the fractional Schrödinger operator in terms of the Caffarelli-Silvestre extension (see [20]), i.e., in terms of a (degenerate) Dirichlet-to-Neumann map associated with a (degenerate) elliptic, local equation in the upper half-plane. Technically, this allowed the authors of [19] to rely on frequency function methods for local equations, while, similarly, the key tool in [10] consisted of several Carleman inequalities for the Caffarelli-Silvestre extension.
● Unique continuation property from measurable sets. Relying on the arguments from [10,19] also unique continuation results from measureable sets can be proved in the regime γ∈(0,1). Indeed, in [19] this is formulated as one of the main results. For rougher equations this is deduced in [8] based on variants of the Carleman estimates from [10].
● Variable coefficient operators. Using more refined frequency function or Carleman estimates, also the case of variable coefficient fractional Schrödinger equations has been treated in [10,Section 7] (C2 regular coefficients) and in [11] (Lipschitz regular coefficients).
In contrast, the situation for higher order fractional Schrödinger operators is much less studied. Here the main known properties are:
● Representation of the equation through a Caffarelli-Silvestre type extension problem. In [15] and in [1] and later also in [21] it was observed that the higher order fractional Laplacian can be realised as a Dirichlet-to-Neumann map of a Caffarelli-Silvestre type extension problem. This can either take the form of a scalar equation (however with a weight which is no longer in the Muckenhoupt class) or a system of Caffarelli-Silvestre extensions.
● Strong unique continuation for fractional harmonic functions. Exploiting the systems characterisation from [15], Yang also sketches the proof of the strong unique continuation property for fractional harmonic functions based on frequency function methods for systems of equations. In the regime γ∈(1,2) this was further detailed in [22], where the authors also obtained precise asymptotics of the solutions under consideration.
● Strong unique continuation for fractional Schrödinger equations. In the case γ=32 the strong unique continuation property for Schrödinger equations with Hardy type potentials was recently proved in [23]. In this context, the regime γ=32 is special, since the degeneracy of the problem disappears and the result can be viewed as a boundary unique continuation result for the Bilaplacian. The authors again rely on frequency function methods.
In contrast, in this article we study Carleman estimates to deduce the desired unique continuation property. Also relying on the systems Caffarelli-Silvestre extension of the higher order fractional Laplacian (which is recalled in the Appendix), we view the various unique continuation properties from above as boundary unique continuation results. In order to deduce these, we hence derive Carleman estimates for the associated systems. This is inspired by the work in [12] in which unique continuation in the interior is discussed for higher order equations (or equivalently systems). As in [10] we combine these Carleman estimates with careful compactness and blow-up arguments.
We emphasise that the results from Theorems 1–4 improve significantly on the known strong unique continuation results for the fractional Laplacian by for instance including (Hardy type) potentials and variable coefficients of low regularity. The strength of these aspects are even novel for the regime γ∈(0,1).
Remark 1.6. Using a characterisation of the higher order fractional Laplacian through a system and a bootstrap argument, for aij=δij we here impose H2γ regularity on the solutions to the equations at hand. We remark that even in the setting of fractional Schrödinger equations in bounded domains, it would be possible to apply our arguments: Indeed, starting from Hγ solutions, it would be possible to bootstrap the regularity properties of the solutions by means of the estimates in [24] (away from the boundary).
We however emphasise that by pseudolocality of the fractional Laplacian our results could also be formulated under only local regularity assumptions: Assuming that we had a weak notion of a generalised Caffarelli-Silvestre extension (which is the case for any Hr(Rn), r∈R, boundary datum, see Section A.1) as well as only local H2γ(B′1) regularity for u, it would have been possible to invoke our unique continuation arguments. We refer to Proposition 1.9 and Remark 1.10 for more on this.
Approaching the problem by means of the generalised systems Caffarelli-Silvestre extension, our arguments for unique continuation rely on several Carleman estimates for the localised equation in combination with a careful blow-up analysis. To this end, we rely on similar ideas as in [8,10].
A crucial technical tool thus is the derivation of higher order Carleman estimates, which we address by iteration of second order estimates. As a consequence, we obtain the following bounds:
Proposition 1.7. Let m∈N∪{0} and b∈(−1,1). Let a∈C2m,1(B+4,R(n+1)×(n+1))∩C0,1(B+4,R(n+1)×(n+1)) be of a block form
a(x)=(˜a(x′)001), | (1.4) |
where the metric ˜a satisfies the conditions (A1)–(A3) from above with μ=2m. Assume that ˜u0,…,˜um∈H1(B+4,xbn+1) with supp(˜uj)⊂B+4∖{0}, j∈{0,…,m}, are solutions to the bulk system
x−bn+1∇⋅xbn+1a∇˜u0=˜u1+f0,x−bn+1∇⋅xbn+1a∇˜u1=˜u2+f1,⋮x−bn+1∇⋅xbn+1a∇˜um=fm, | (1.5) |
in B+4 with f0,…,fm∈H1(B+4,xbn+1). Further suppose that on B′4 we have
limxn+1→0xbn+1∂n+1˜uj=0 for j∈{0,…,m−1},limxn+1→0xbn+1∂n+1˜um=g,limxn+1→0˜u0=u, | (1.6) |
where all limits are considered with respect to the L2 topology. Then, there exists τ0>1 such that for all τ>τ0 there is a weight h such that
τm+1−b2‖eh(−ln(|x|))(1+¯h)m+12|x|−2m+b−12˜u0‖L2(B′4)+m∑j=0(τm+1−j‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−j2|x|−2m−1+2j˜uj‖L2(B+4)+τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−j2|x|−2m+2j∇˜uj‖L2(B+4))≤C(m∑j=0τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)m−j2|x|−2m+1+2jfj‖L2(B+4)+τ1+b2‖eh(−ln(|x|))|x|1−b2g‖L2(B′4)). | (1.7) |
Here ¯h(x):=h″(t)|t=−ln(|x|).
This estimate improves previous results even in the case m=0 by allowing for only Lipschitz continuous metrics a. It includes strengthened bounds which exploit the spectral gap of the fractional Laplacian on the sphere with Neumann (or Dirichlet) conditions (see the Appendix A, Section 8.3 in [25] for these spectral gap properties). A relevant ingredient in the derivation of this Carleman estimate involves the use of a splitting technique in a similar way as in [8].
Estimating the commutators, the bounds from Proposition 1.7 can be further improved to include higher order tangential derivatives on the left hand side of (1.7):
Proposition 1.8. Let m∈N∪{0}, b∈(−1,1) and a:Rn+1+→R(n+1)×(n+1) as in Proposition 1.7. Assume that ˜u0,…,˜um∈H1(B+4,xbn+1) with supp(˜uj)⊂B+4∖{0}, j∈{0,…,m}, are solutions to the bulk system (1.5) in B+4 with fj∈Hm−j(B+4,xbn+1) for j∈{0,…,m}. Further suppose that on B′4 the boundary conditions (1.6) hold, where all limits are considered with respect to the L2 topology. Then, there exists τ0>1 such that for all τ>τ0 there is a weight h such that
m∑j=0τm−j+1−b2‖eh(−ln(|x|))(1+¯h)m+12|x|−2m+j+b−12(∇′)j˜u0‖L2(B′4)+m∑j=0j∑k=0(τm+1−j‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−k2|x|−2m−1+j+k(∇′)j−k˜uk‖L2(B+4)+τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−k2|x|−2m+j+k∇(∇′)j−k˜uk‖L2(B+4))≤C(m∑j=0j∑k=0τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)m−k2|x|−2m+1+j+k(∇′)j−kfk‖L2(B+4)+τ1+b2‖eh(−ln(|x|))|x|1−b2g‖L2(B′4)). | (1.8) |
Here ¯h(x):=h″(t)|t=−ln(|x|).
It is in this form that we exploit the Carleman estimates to infer our main results.
Motivated by the recent introduction of the fractional Calderón problem [7,8,9], we here discuss applications of our unique continuation results in inverse problems. As a first main property, we deduce the antilocality of the higher order fractional Laplacian:
Proposition 1.9 (Antilocality). Let γ∈R+∖N and let L=−∇⋅˜a∇, where ˜a satisfies the conditions (A1)–(A3) from above with μ=2⌊γ⌋. Let u∈Dom(Lγ). Assume that for some open set W⊂Rn containing the (n-dimensional) unit ball B1⊂Rn we have
u=0 and Lγu=0in W. |
Then, u≡0 in Rn.
We emphasise that in this result the function u is not assumed to satisfy any equation globally. This result thus provides a strong global rigidity property in which the nonlocality of the equation under consideration plays a major role. Originally, the notion of antilocality appeared in the context of quantum field theory as the Reeh-Schlieder theorem [26], but has recently found various applications in inverse problems [7,8,9,27,28] and control theory [5,6].
Remark 1.10. We remark that a result of the form stated in Proposition 1.9 also holds in a large range of less regular spaces. The only ingredient needed is the presence of a Caffarelli-Silvestre extension at the given regularity (but this holds under very weak assumptions, see Proposition A.6 in the constant coefficient setting). The pseudolocality of the associated operators then allows one to locally deduce the vanishing of u from which it is possible to propagate the deduced information to an arbitrary point through the upper half plane (in which the extension problem is formulated).
As a property dual to the antilocality of the higher order fractional Laplacian, we further obtain approximation properties for these operators:
Proposition 1.11 (Runge approximation). Let Ω⊂Rn be open, bounded and non-empty and let ˜Ω⊂Rn be open with Ω⊂˜Ω and B′1⊂˜Ω∖Ω. Let v∈Hγ(Ω) with γ∈R+∖N. Assume that L=−∇⋅˜a∇, where ˜a satisfies the conditions (A1)–(A3) from above with μ=2⌊γ⌋. Suppose that q∈L∞(Ω) is such that zero is not a Dirichlet eigenvalue of the operator Lγ+q. Then, for any ϵ>0 there exists a solution to
Lγu+qu=0 in Ω, supp(u)⊂˜Ω, |
such that ‖v−u‖L2(Ω)≤ϵ.
Remark 1.12. The assumptions on the sets W and ˜Ω∖Ω with respect to the unit ball B′1={x∈Rn:|x|<1} are taken without loss of generality after normalising the set-up in such a way that the assumptions (A2), (A3) on ˜a are satisfied.
These type of approximation properties again crucially exploit the nonlocality of the operator. They were first observed in [29] and generalised to larger classes of equations in [28,30,31,32,33,34]. As first highlighted in [7] in the context of nonlocal inverse problems they play an important role in deducing injectivity. In [9] these properties were strengthened to quantitative estimates which were applied in proving stability of the associated inverse problem.
In addition to the rigidity and flexibility properties from Propositions 1.9 and 1.11, it is possible to make use of our unique continuation results in many further contexts. As in [22,23,35] one could for instance study the associated problems more quantitatively and derive vanishing order or nodal domain estimates. Using Carleman estimates, one could here proceed similarly as in [36]. Also control theoretic questions similar to for instance [6] could be addressed with our results. We postpone such a discussion to future work.
The remainder of the article is organised as follows: After first recalling a number of auxiliary results (including the generalised Caffarelli-Silvestre extension) in Section 2, in Section 3 we then deduce the Carleman estimates which form the basis of our unique continuation results. In Section 4 we derive compactness results for the systems which are associated with the SUCP for fractional Schrödinger operators. Here we reduce the SUCP to the weak unique continuation property (WUCP) for the associated systems. This is complemented by a bootstrap argument to derive the WUCP in Section 5. In Section 6 we combine all the previous results and deduce the statements of Theorems 1–4 and Propositions 1.9 and 1.11. Finally, in the Appendix, we present a sketch of the derivation of the generalised Caffarelli-Silvestre extension for the higher order fractional Laplacian which had been introduced in [15] and which we here discuss at low regularity.
In this section we recall several auxiliary results that will be used frequently throughout the text: On the one hand, we recall a higher order Caffarelli-Silvestre extension result. On the other hand, we discuss appropriate boundary-bulk estimates.
We summarise the notation that we will use in the sequel:
Working in Rn+1+:={x∈Rn+1: xn+1≥0}, we will always use the convention that x=(x′,xn+1) with x′∈Rn and xn+1≥0. For x0∈Rn×{0} we will denote (half) balls in Rn+1+ and Rn×{0} by
B+r(x0)={x∈Rn+1+: |x−x0|≤r}, B′r(x0)={x∈Rn×{0}: |x−x0|≤r}. |
If x0=0, we will simply write B+r and B′r.
In the sequel and in particular in the proofs of the Carleman inequalities, we will often use conformal polar coordinates in the upper half space: x=e−tθ, where t=−ln(|x|) and θ=x|x|. Here θ∈Sn+, where Sn+ denotes the upper half (unit) sphere in Rn+1+.
We will frequently use the operator
Lb:=x−bn+1(∂xn+1xbn+1∂xn+1−xbn+1L), | (2.1) |
where L=−∇′⋅˜a∇′ and ˜a satisfies the conditions from (A1)–(A3) with μ=2⌊γ⌋. Usually, we will be thinking of b=1−2⌊γ⌋+2γ, where ⌊t=max{k∈N: k≤t}.
We define the variable coefficient fractional Laplacian through functional calculus, see Section A.2. The set Dom(Lγ) is then defined as the space in which this functional calculus can be directly applied.
In analogy to our convention for the space variables, we often use the notation
∇′, ∂′ℓ, ℓ∈{1,…,n}, |
to denote the corresponding tangential gradient and partial derivatives.
Studying operators of the form (2.1), in the sequel, we will often work in function spaces of the form
H1(Ω,xbn+1):={u:Ω→R: ‖xb2n+1u‖L2(Ω)+‖xb2n+1∇u‖L2(Ω)<∞}, |
where Ω⊂Rn+1+ is a relatively open set. Similarly, we also deal with the function spaces
˙H1(Ω,xbn+1):={u:Ω→R: ‖xb2n+1∇u‖L2(Ω)<∞},L2(Ω,xbn+1):={u:Ω→R: ‖xb2n+1u‖L2(Ω)<∞}. |
We denote the homogeneous Hölder norms by [⋅]˙Ck,α(Ω) for k∈N∪{0}, α∈(0,1) and Ω⊂Rn or Ω⊂Rn+1+.
For a sequence {ak}k∈N⊂ℓ1, we set
‖ak‖ℓ1:=∞∑k=1|ak|. |
We first recall that the higher order fractional Laplacian can be interpreted by means of a Caffarelli-Silvestre-type extension [1,15,20,21]:
Proposition 2.1. Let γ>0, f∈Dom(Lγ) and L:=−∇′⋅˜a∇′, where ˜a satisfies the conditions (A1)–(A3) with μ=2⌊γ⌋. Then, there exists an extension operator
Eγ:Dom(Lγ)→C2,1loc(Rn×(0,∞))∩H1loc(Rn+1+,xbn+1), f↦u:=Eγ(f), |
such that u=Eγ(f) is a weak solution to the scalar higher order problem
L⌊γ⌋+1bu=0 in Rn+1+,limxn+1→0u=f on Rn×{0},limxn+1→0Lkbu=cn,γ,kLkf on Rn×{0} for k∈{1,…,⌊γ⌋},limxn+1→0x1−2γ+2⌊γ⌋n+1∂xn+1L⌊γ⌋bu=cn,γLγf on Rn×{0},limxn+1→0x1−2γ+2⌊γ⌋n+1∂xn+1Lkbu=0 on Rn×{0} for k∈{0,⋯,⌊γ⌋−1}. | (2.2) |
Here Lb:=x−bn+1(∂xn+1xbn+1∂xn+1+xbn+1∇′⋅˜a∇′) and b=1−2⌊γ⌋+2γ. All boundary conditions in (2.2) are attained as L2(Rn) limits.
Moreover, setting u0:=u and uj+1=Lbuj for j∈{0,…,⌊γ⌋−1}, the functions uj are in C2,1(Rn×(0,1))∩H1loc(Rn+1+,xbn+1). They are weak solutions of the following system of second order equations
Lbum=0 in Rn+1+,Lbuj=uj+1 in Rn+1+ for j∈{0,…,m−1},limxn+1→0uj=cn,γ,jLjf on Rn×{0} for j∈{0,…,m},limxn+1→0xbn+1∂xn+1um=cn,γLγf on Rn×{0},limxn+1→0xbn+1∂xn+1uj=0 on Rn×{0} for j∈{0,…,m−1}, | (2.3) |
where m=⌊γ⌋. All boundary conditions hold in an L2(Rn) sense.
In order to keep our presentation self-contained, we provide a short sketch of the proof of this statement in the Appendix.
Compared to the original nonlocal Eqs (1.1) and (1.2), the Eqs (2.2) and (2.3) arising from the generalised Cafferelli-Silvestre extension have the advantage that they can be approached with tools from the analysis of unique continuation properties of local elliptic equations. As in [10], [37] the price to pay for this localisation is the introduction of the additional dimension in which the solutions u have to be controlled through the corresponding equations. This gives our problem and our argument the flavour of boundary unique continuation results (see for instance [38] and the references therein).
We recall the boundary-bulk interpolation inequality from [10] for fractional Sobolev spaces on the sphere:
Lemma 2.2. Let u:Sn+→R and let b∈(−1,1). Then, identifying ∂Sn+ with Sn−1, there exists a constant C>0 such that for any τ>1
‖u‖L2(Sn−1)≤C(τ1+b2‖θb2nu‖L2(Sn+)+τb−12‖θb2n∇Snu‖L2(Sn+)). |
Proof. The proof follows as in [37] by using the trace inequality in the associated weighted Sobolev spaces. We discuss the details:
First, by the trace inequality, any function w∈H1(Rn+1+,xbn+1) with b∈(−1,1) satisfies
‖w‖L2(Rn)≤C(‖xb2n+1w‖L2(Rn+1+)+‖xb2n+1∇w‖L2(Rn+1+)). | (2.4) |
Indeed, for w∈C∞(Rn+1+) with compact support in the tangential directions this follows from the fundamental theorem of calculus: For t∈(0,1)
|w(x′,0)−w(x′,t)|=|t∫0∂zw(x′,z)dz|≤1∫0|∂zw(x′,z)|dz≤1∫0z−b2zb2|∂zw(x′,z)|dz≤Cb(1∫0zb|∂zw(x′,z)|2dz)12, |
where we used that b∈(−1,1). Thus, applying the triangle inequality, integrating in t∈[0,1] and applying the Cauchy-Schwarz inequality, we infer
|w(x′,0)|≤Cb‖xb2n+1∂xn+1w(x′,⋅)‖L2((0,1))+‖x−b2n+1xb2n+1w(x′,⋅)‖L1((0,1))≤Cb(‖xb2n+1∂xn+1w(x′,⋅)‖L2((0,1))+‖xb2n+1w(x′,⋅)‖L2((0,1))). |
Taking squares and integrating in x′∈Rn then yields
‖w‖2L2(Rn)≤Cb(‖xb2n+1∂xn+1w‖2L2(Rn×(0,1))+‖xb2n+1w‖2L2(Rn×(0,1)))≤Cb(‖xb2n+1∂xn+1w‖2L2(Rn+1+)+‖xb2n+1w‖2L2(Rn+1+)). |
By density considerations, this concludes the proof of the trace estimate (2.4).
Rescaling (2.4), we then infer
‖w‖L2(Rn)≤C(τ1+b2‖xb2n+1w‖L2(Rn+1+)+τb−12‖xb2n+1∇w‖L2(Rn+1+)). |
Finally, we apply this to w(x)=η(|x|)u(x|x|), where η is a smooth, positive cut-off function supported on B+2∖B+1/2 and which is identically one on B+3/2∖B+3/4. This yields the desired claim.
We derive a Caccioppoli type inequality for tangential derivatives of a solution to a variable coefficient equation associated with the operator Lb from above.
Lemma 2.3. Let b∈(−1,1) and a:B+4→R(n+1)×(n+1) be of the block form (1.4) with ˜a satisfying the conditions (A1)–(A3) with μ=2m. Let u∈H1(B+4,xbn+1) be a solution to
x−bn+1∇⋅axbn+1∇u=f in B+4,limxn+1→0xbn+1∂xn+1u=g on B′4, |
with f,|∇′f|,⋯,|(∇′)kf|∈L2(B+4,xbn+1) and g∈Hk(B′4) for some k∈N∪{0}. Then there exists a constant C>0 such that for any J∈N∪{0} with J≤min{2m,k} and for any r∈(0,2)
J∑j=0rj‖xb2n+1∇(∇′)ju‖L2(B+r/2)≤CJ∑j=0(rj−1‖xb2n+1(∇′)ju‖L2(B+r)+rj+1‖xb2n+1(∇′)jf‖L2(B+r)+rj(∫B′r|(∇′)jg||(∇′)ju|dx′)12). | (2.5) |
Proof. First of all we note that by scaling it suffices to prove the estimate for r=1. Next, we observe that by the block structure of the metric a for any j∈{1,…,2m}
x−bn+1∇⋅axbn+1∇(∇′)ju=(∇′)j(x−bn+1∇⋅axbn+1∇u)−j−1∑k=0∇′⋅((∇′)j−k˜a)∇′(∇′)ku, |
provided that these expressions are j-times differentiable in the tangential directions. Hence, formally (∇′)ju is a weak solution to
Lb(∇′)ju=(∇′)jf−j−1∑k=0∇′⋅((∇′)j−k˜a)∇′(∇′)ku in B+3,limxn+1→0xbn+1∂xn+1(∇′)ju=(∇′)jg on B′3, |
Using difference quotient arguments together with the regularity of f and g, these formal differentiations in the tangential directions can be justified. As a consequence, elliptic regularity implies that (∇′)ju∈H1(B+3,xbn+1). In particular, for any test function φ it holds
∫B+1xbn+1∇φ⋅a∇(∇′)judx=−∫B+1xbn+1φ(∇′)jfdx−j−1∑k=0∫B+1xbn+1∇′φ⋅((∇′)j−k˜a)∇′(∇′)kudx+∫B′1φ(∇′)jgdx′. |
Let η be a radial cut-off function equal to one on B+1/2 which vanishes outside of B+1 and satisfies |∇η|≤C. We remark that the function φ=η2(∇′)ju is an admissible test function {since by the above considerations as an H1(B+1,xbn+1) function it is sufficiently regular}. Using that by condition (A2) it holds |(∇′)α˜a(x)|≤Cα for x∈B+4 and |α|≤2m, we obtain the following estimate
‖xb2n+1η∇(∇′)ju‖2L2(B+1)≤C(∫B+1xbn+1η|∇(∇′)ju||∇η||(∇′)ju|dx+∫B+1xbn+1η2|(∇′)ju||(∇′)jf|dx+j−1∑k=0∫B+1xbn+1(η2|∇(∇′)ju|+η|∇η||(∇′)ju|)|(∇′)k+1u|dx+∫B′1η2|(∇′)jg||(∇′)ju|dx′). |
By virtue of Young's inequality and absorbing terms into the left hand side we infer
‖xb2n+1η∇(∇′)ju‖L2(B+1)≤C(‖xb2n+1|∇η|(∇′)ju‖L2(B+1)+‖xb2n+1η(∇′)ju‖L2(B+1)+‖xb2n+1η(∇′)jf‖L2(B+1)+j−1∑k=0‖xb2n+1η(∇′)k+1u‖L2(B+1)+∫B′1η2|(∇′)jg||(∇′)ju|dx′). |
Lastly, we use the bounds for η to obtain
‖xb2n+1∇(∇′)ju‖L2(B+1/2)≤C(j∑k=1‖xb2n+1(∇′)ku‖L2(B+1)+‖xb2n+1(∇′)jf‖L2(B+1)+(∫B′1|(∇′)jg||(∇′)ju|dx′)12). |
The estimate for j=0 is straightforward. Rescaling and summing over j∈{0,…,J} with the corresponding factors, yields the desired estimate.
In order to deduce the Carleman estimates from Propositions 1.7 and 1.8, we first prove Carleman estimates for second order (degenerate elliptic) equations. Here we proceed in two steps: First, we discuss the situation for constant coefficient metrics but in the presence of divergence form right hand side contributions, and then, in a second step, we deduce the estimates for variable coefficient metrics.
As a main ingredient in our argument we make use of the following (constant coefficient) Carleman estimate:
Proposition 3.1. Let b∈(−1,1) and let u∈H1(B+4,xbn+1) with supp(u)⊂B+4∖{0} be a solution to
∇⋅xbn+1∇u=f+n∑j=1∂jxbn+1Fj in B+4,limxn+1→0xbn+1∂n+1u=g on B′4, |
where f∈L2(B+4,x−bn+1), g∈L2(B′4) and F=(F1,…,Fn)∈L2(B+4,xbn+1)n with supp(f), supp(F)⊂B+4∖{0} and supp(g)⊂B′4∖{0}. Then, for each τ>τ0≫1 there is a weight function h(−ln(|x|)) such that there exists a constant C>0 which is independent of τ such that it holds
τ‖eh(−ln(|x|))xb2n+1(1+¯h)12|x|−1u‖L2(Rn+1+)+‖eh(−ln(|x|))xb2n+1(1+¯h)12∇u‖L2(Rn+1+)+τ1−b2‖eh(−ln(|x|))(1+¯h)12|x|b−12u‖L2(Rn×{0})≤C(‖eh(−ln(|x|))|x|x−b2n+1f‖L2(Rn+1+)+τ‖eh(−ln(|x|))xb2n+1F‖L2(Rn+1+)+τ1+b2‖eh(−ln(|x|))|x|1−b2g‖L2(Rn×{0})). | (3.1) |
Here ¯h(x):=h″(t)|t=−ln(|x|).
The proof of Proposition 3.1 relies on a splitting strategy, in which all inhomogeneities (be they bulk or boundary contributions) are dealt with in an elliptic estimate. As a consequence, the subelliptic part, i.e., the actual Carleman estimate, becomes rather clean. In particular, as shown in the following section, the estimates are strong enough to deal with only Lipschitz regular metrics in a perturbative way.
Proof. We proceed in three main steps: First, we construct an appropriate Carleman weight. Then, we deduce the desired Carleman estimate by a splitting argument in conformal polar coordinates. In a final step, we concatenate the obtained information.
Step 1: Construction of the weight. We begin by constructing a family of Carleman weights h(t):R→R. Anticipating the use of polar conformal coordinates, we require it to satisfy
h′∈(C−1τ,Cτ) for all τ≫1,14≤h″+dist(h′,spec(∇Sn⋅θbn∇Sn)),|h‴|,|h(4)|≤εh″≤Cτ, | (3.2) |
where ε>0 is a constant which is to be determined later (see Step 1 in the proof of Proposition 1.7), and C>1 is a constant independent of τ. We recall that by the results of [25] the operator ∇Sn⋅θbn∇Sn with θn:=xn+1|x| has a spectral gap if it is considered with Neumann (or Dirichlet) data (see Section 8.3 in the Appendix A in [25]). We follow the argument from [13] and [12] to obtain the desired properties for the Carleman weight. To this end, we consider a sequence {cj}j∈N∈ℓ1, ‖cj‖ℓ1<δ of non-negative numbers and define the sequence {aj}j∈N as the convolution of cj with 2−νj for some ν>0 small. Then, the sequence aj is slowly varying (i.e., 2−νaj+1≤aj≤2νaj+1) and obeys the bound cj≤aj. With this preparation, we define h(0)=0, h′(−∞)=⌊τ⌋+c1, where c1>0 is a suitable constant independent of τ, and
h″=∑jbjχ[j,j+1], | (3.3) |
with
bj{=0 if aj≤ντ−1,∈[τaj,2τaj] if aj≥ντ−1. |
In order to obtain the desired higher order regularity properties for h, we regularise this by convolution (on the unit scale). Thus, the last estimate in (3.2) is fulfilled for any given, small ε>0 provided that the difference of consecutive values of aj is small enough, i.e., if δ and ν are taken sufficiently small.
Step 2: Conformal polar coordinates and splitting argument. We proceed by a splitting argument. In order to obtain more transparent expressions, we pass to conformal coordinates by setting t=−ln(|x|), θ=x|x|. We further pass from the function u to the function ˜u(t,θ)=e1−b−n2tu(e−tθ). In these coordinates we consider (the weak form) of the equation
(θbn∂2t+∇Sn⋅θbn∇Sn+cn,bθbn)˜u=˜f−θbn∂t˜Ft+divSn+θbn˜F′ in Sn+×R,limθn→0θbnν⋅∇Sn˜u=˜g on ∂Sn+×R, | (3.4) |
where
˜f(t,θ)=e−n+3−b2tf(e−tθ)+n+b−12θbn˜Ft(t,θ),˜g(t,θ)=e(−1+b)te1−b−n2tg(e−tθ),˜Ft(t,θ)=e−n+1+b2t(n∑i=2θi−1Fi(e−tθ)±√(1−n∑i=1θ2i)F1(e−tθ)),˜Fj(t,θ)=e−n+1+b2tn∑i=1δi,j+1Fi(e−tθ)−θj˜Ft(t,θ), j∈{1,…,n},˜F′(t,θ)=(˜F1(t,θ),…,˜Fn(t,θ)),cn,b=−(n+b−12)2, θj=xj+1|x|, j∈{1,…,n}. |
Here divSn+ denotes the divergence with respect to the standard metric on the sphere and the choice of the sign in the expression for ˜Ft(t,θ) depends on the specific chart.
We split the problem into two parts ˜u=u1+u2, where u1 is a solution to the following elliptic problem
(θbn∂2t+∇Sn⋅θbn∇Sn+θbncn,b−θbnτ2K2)u1=˜f−θbn∂t˜Ft+divSn+θbn˜F′ in Sn+×R,limθn→0θbnν⋅∇Snu1=˜g on ∂Sn×R. | (3.5) |
Here K≫1 is a sufficiently large parameter (to be specified later). The function u2=˜u−u1 thus solves a corresponding problem. In order to derive the desired estimate, we discuss the bounds for u1 and u2 separately.
Step 2a: Bounds for u1. By virtue of the positivity of K≫1, the estimates for u1 are elliptic energy estimates. Indeed, by the Lax-Milgram theorem, we obtain that a solution to (3.5) exists in the energy space H1(Sn+×R,θbn). We test the weak form of (3.5) with the test function τ2e2hM,δ(t)u1, where
hM,δ(t):=min{M,max{h(t),−M}}∗ηδ(t), |
for M∈N and with ηδ denoting a standard mollifier.
This leads to the following identity
τ2(θbn∂tu1,e2hM,δ(t)∂tu1)+2τ2(θbn∂tu1,h′M,δe2hM,δ(t)u1)+τ2(θbne2hM,δ(t)∇Snu1,∇Snu1)−(cn,bτ2−K2τ4)(θbne2hM,δ(t)u1,u1)=−τ2(˜f,e2hM,δ(t)u1)−τ2(divθbn˜F′,e2hM,δ(t)u1)+τ2(θbn∂t˜Ft,e2hM,δ(t)u1)+τ2(˜g,e2hM,δ(t)u1)0. | (3.6) |
Here (⋅,⋅):=(⋅,⋅)L2(Sn+×R) and (⋅,⋅)0:=(⋅,⋅)L2(∂Sn+×R). Recalling that by definition (see (3.2))
|h′M,δ(t)|≤Cτ, | (3.7) |
an application of Young's inequality leads to
2τ2|(θbn∂tu1,h′M,δe2hM,δ(t)u1)|(3.7)≤14τ2|(θbn∂tu1,e2hM,δ(t)∂tu1)|+16C2τ4|(θbnu1,e2hM,δ(t)u1)|,τ2|(˜f,e2hM,δ(t)u1)|≤14τ4|(θbnu1,e2hM,δ(t)u1)|+C|(θ−bn˜f,e2hM,δ(t)˜f)|,τ2|(divθbn˜F′,e2hM,δ(t)u1)|≤14τ2|(e2hM,δ(t)θbn∇Sn+u1,∇Sn+u1)|+Cτ2|(θbn˜Fi,e2hM,δ(t)˜Fi)|,τ2|(θbn∂t˜Ft,e2hM,δ(t)u1)|(3.7)≤τ2|(θbn˜Ft,e2hM,δ(t)∂tu1)|+2Cτ3|(θbn˜Ft,e2hM,δ(t)u1)|≤14τ2|(e2hM,δ(t)θbn∂tu1,∂tu1)|+Cτ2|(θbn˜Ft,e2hM,δ(t)˜Ft)|+τ4|(e2hM,δ(t)u1,θbnu1)|+4C2τ2|(θbn˜Ft,e2hM,δ(t)˜Ft)|. |
Absorbing the contributions involving u1 as well as the other non-positive terms from (3.6) into either the positive derivative contributions or (for K2≫1 sufficiently large) into the coercive term involving K2 in (3.6), then yields
τ‖θb2nehM,δ(t)∂tu1‖+τ‖θb2nehM,δ(t)∇Snu1‖+K2τ2‖θb2nehM,δ(t)u1‖≤C(‖θ−b2nehM,δ(t)˜f‖+τ‖θb2nehM,δ(t)˜F‖+ϵτ3−b2‖ehM,δ(t)u1‖0+Cϵτ1+b2‖ehM,δ(t)˜g‖0). |
As above, here and in the sequel, we use the notation ‖⋅‖:=‖⋅‖L2(Sn+×R) and ‖⋅‖0:=‖⋅‖L2(∂Sn+×R). Applying the boundary-bulk interpolation estimate from Lemma 2.2, allows us to absorb the first boundary contribution into the left hand side, which then results in
τ‖θb2nehM,δ(t)∂tu1‖+τ‖θb2nehM,δ(t)∇Snu1‖+K2τ2‖θb2nehM,δ(t)u1‖≤C(‖θ−b2nehM,δ(t)˜f‖+τ‖θb2nehM,δ(t)˜F‖+Cϵτ1+b2‖ehM,δ(t)˜g‖0). | (3.8) |
Using the compact supports of ˜f, ˜F and ˜g, by dominated convergence, we may pass to the limits M→∞ and δ→0 which leads to
τ‖θb2neh∂tu1‖+τ‖θb2neh∇Snu1‖+K2τ2‖θb2nehu1‖≤C(‖θ−b2neh˜f‖+τ‖θb2neh˜F‖+Cϵτ1+b2‖eh˜g‖0). | (3.9) |
We remark that this estimate not only contains the right weighted bounds for u1 but also implies that u1 has (quantitative) fast decay as |t|→∞ (which corresponds to |x|→0 and |x|→∞). By the compact support assumption on ˜u a similarly fast decay then also holds for u2.
Step 2b: Bounds for u2. The estimates for u2 will be sub-elliptic (in τ) Carleman estimates. We recall that by construction, u2 is a weak solution of
(θbn∂2t+∇Sn⋅θbn∇Sn+θbncn,b)u2=−K2τ2θbnu1 in Sn×R,limθn→0θbnν⋅∇Snu2=0 on ∂Sn×R, | (3.10) |
i.e., it satisfies
−(θbn∂tu2,∂tφ)−(θbn∇Snu2,∇Snφ)+cn,b(θbnu2,φ)=−K2τ2(θbnu1,φ). | (3.11) |
We test this with a Neumann eigenfunction to the spherical operator, i.e., with a function ψλ which satisfies
∇Sn⋅θbn∇Snψλ=−λ2θbnψλ in Sn+,limθn→0θbnν⋅∇Snψλ=0 on ∂Sn+×R. |
We recall that the set {ψλ}λ forms an orthogonal, complete system in L2(θbn,Sn+). More precisely, we insert φ(t,θ)=ψλ(θ)γ(t) with γ(t) a compactly supported smooth function into (3.11). Since the set {ψλ} forms an orthonormal set in both H1(Sn+,θbn) and L2(Sn+,θbn), we infer that αλ(t):=(u2,θbnψλ) is a distributional (but then by uniqueness also a classical) solution to the ODE
α″λ−λ2αλ+cn,bαλ=−K2τ2βλ, | (3.12) |
where βλ(t)=(u1,θbnψλ).
Conjugating this modewise equation with the weight eh(t) yields the equation
˜α″λ−λ2˜αλ+|h′|2˜αλ+cn,b˜αλ−2h′˜α′λ−h″˜αλ=K2τ2˜βλ, |
where ˜αλ=eh(t)αλ and ˜βλ=eh(t)βλ. Noting that the symmetric and antisymmetric parts of the conjugated operator turn into
S˜αλ=˜α″λ−λ2˜αλ+|h′|2˜αλ+cn,b˜αλ,A˜αλ=−2h′˜α′λ−h″˜αλ, |
we expand the conjugated operator to infer
K4τ4‖˜βλ‖2L2(R)=‖S˜αλ‖2L2(R)+‖A˜αλ‖2L2(R)+([S,A]˜αλ,˜αλ)L2(R), | (3.13) |
where
([S,A]˜αλ,˜αλ)=4∫R(h′h″h′˜α2λ+h″(˜α′λ)2)dt−∫Rh(4)˜α2λdt. |
We observe that the first two contributions in the expansion of the commutator are non-negative. The last term which does not necessarily carry a sign can be absorbed into these positive contributions (recall the last condition in (3.2) for that) and can hence be neglected for τ≥τ0>1 sufficiently large, whence
([S,A]˜αλ,˜αλ)≥3∫R(h′h″h′˜α2λ+h″(˜α′λ)2)dt, | (3.14) |
if τ≥τ0>1 is chosen sufficiently large.
Next we exploit the spectral gap of the Neumann data version of the operator ∇Sn⋅θbn∇Sn (see [25,Appendix A, Section 8.3]) in connection with the symmetric and antisymmetric parts of the operator. To this end, we consider two cases: If λ2∈[0,2C2τ2) with C>1 as in (3.2), we have that |h′|≥C−2√2λ. Using the antisymmetric part of the operator, we estimate
12‖A˜αλ‖2L2(R)≥‖h′˜α′λ‖2L2(R)−12‖h″˜αλ‖2L2(R)≥C−42‖max{|h′|,λ}˜α′λ‖2L2(R)−12‖h″˜αλ‖2L2(R). | (3.15) |
If λ2≥2C2τ2, we estimate the symmetric part of the operator and integrate by parts
‖S˜αλ‖2L2(R)≥‖˜α″λ‖2L2(R)+2(˜α″λ,(|h′|2−λ2)˜αλ)L2(R)+‖(λ2−|h′|2)˜αλ‖2L2(R)−c2n,b‖˜αλ‖2L2(R)=‖˜α″λ‖2L2(R)+2(˜α′λ,(λ2−|h′|2)˜α′λ)L2(R)−4(˜α′λ,h″h′˜αλ)L2(R)+‖(λ2−|h′|2)˜αλ‖2L2(R)−c2n,b‖˜αλ‖2L2(R). | (3.16) |
We note that non-positive or not necessarily positive contributions are given by the third and the fifth term on the right hand side of (3.16), which we hence seek to bound. Since λ2≥2C2τ2≥|h′|22, taking τ>τ0 large enough, we have
c2n,b‖˜αλ‖2L2(R)≤12‖(λ2−|h′|2)˜αλ‖2L2(R). |
Thus the fifth term on the right hand side of (3.16) can be absorbed into the fourth term on the right hand side of (3.16). We estimate the remaining possibly non-positive contribution which is the third term (as λ2≥2C2τ2): To this end, we again integrate by parts and obtain that for sufficiently large τ≥τ0>1
4|(˜α′λ,h″h′˜αλ)L2(R)|=2|(˜αλ,(h″h″+h‴h′)˜αλ)|≤∫Rh′h″h′˜α2λdt. |
Thus, if λ2≥2C2τ2, we infer
12‖S˜αλ‖2L2(R)≥12‖˜α″λ‖2L2(R)+(˜α′λ,(λ2−|h′|2)˜α′λ)L2(R)−2(˜α′λ,h″h′˜αλ)L2(R)+14‖(λ2−|h′|2)˜αλ‖2L2(R)≥14‖(λ2−|h′|2)˜αλ‖2L2(R)−12∫Rh′h″h′˜α2λdt≥14dist(h′,spec(∇Sn⋅θbn∇Sn))‖max{|h′|,|λ|}˜αλ‖2L2(R)−12∫Rh′h″h′˜α2λdt. | (3.17) |
Hence, combining (3.15) and (3.17), we deduce that for some constant c0>0 which is independent of τ>1
12(‖A˜aλ‖2L2(R)+‖S˜aλ‖2L2(R))≥c0dist(h′,spec(∇Sn⋅θbn∇Sn))‖max{|h′|,|λ|}˜αλ‖2L2(R)−12∫Rh′h″h′˜α2λdt−12‖h″˜αλ‖2L2(R). | (3.18) |
Further using the antisymmetric part to bound (for τ≥τ0>1 sufficiently large)
12‖A˜αλ‖2L2(R)≥τ−2‖h′˜α′λ‖2L2(R)−τ−2‖h″˜αλ‖2L2(R), |
and combining this with (3.14) and (3.18) we arrive at
K4τ4‖˜βλ‖2L2(R)=‖S˜αλ‖2L2(R)+‖A˜αλ‖2L2(R)+([S,A]˜αλ,˜αλ)L2(R)≥c0‖max{|h′|,|λ|}˜αλ‖2L2(R)+‖h′(h″)1/2˜αλ‖2L2(R)+‖(h″)1/2˜α′λ‖2L2(R)+τ−2‖h′˜α′λ‖2L2(R). | (3.19) |
We remark that we have given up a factor τ2 in the antisymmetric estimate. This is due to the fact, that in undoing the conjugation with the weight eh(t), we obtain a term originating from the t derivative falling onto the weight. Without the loss of the factor τ2 this would carry a weight τ4. We would not be able to absorb this into the L2 contributions on the left hand side of the estimates.
We further complement the estimate (3.19) by a bound on the spherical part of the gradient. To this end, we make use of the symmetric part of the operator. Indeed, we have
(S˜αλ,h″˜αλ)=−(˜α′λ,h″˜α′λ)+12(˜αλ,h‴˜αλ)−λ2(˜αλ,h″˜αλ)+(|h′|2h″˜αλ,˜αλ)+cn,b(˜αλ,h″˜αλ). |
As a consequence, if c0>0 is sufficiently small,
c0λ2(˜αλ,h″˜αλ)≤c0|(S˜αλ,h″˜αλ)|+c0|(˜α′λ,h″˜α′λ)|+c02|(˜αλ,h‴˜αλ)|+c0|(|h′|2h″˜αλ,˜αλ)|+c0cn,b|(˜αλ,h″˜αλ)|≤c202‖Sαλ‖2+c202‖h″˜αλ‖2+c0|(˜α′λ,h″˜α′λ)|+c02|(˜αλ,h‴˜αλ)|+c0Cτ2‖(h″)12˜αλ‖2≤K4τ4‖˜βλ‖2. |
Here the last estimate follows from the previously deduced bounds from (3.19). Hence, we conclude
2K4τ4‖˜βλ‖2L2(R)=2(‖S˜αλ‖2L2(R)+‖A˜αλ‖2L2(R)+([S,A]˜αλ,˜αλ)L2(R))≥c0‖max{|h′|,|λ|}˜αλ‖2L2(R)+‖h′(h″)1/2˜αλ‖2L2(R)+‖(h″)1/2˜α′λ‖2L2(R)+τ−2‖h′˜α′λ‖2L2(R)+c0λ2‖(h″)1/2˜αλ‖2L2(R). | (3.20) |
By orthogonality, summing the estimate (3.20) over λ, integrating over Sn+, using the properties of h and undoing the conjugation, we thus obtain
τ‖eh(1+h″)1/2θb2nu2‖+‖eh(1+h″)1/2θb2n∇u2‖≤K2τ2‖θb2nehu1‖. | (3.21) |
Step 3: Conclusion. Last but not least, we combine the estimates from Steps 1 and 2 and deduce the Carleman estimate from Proposition 3.1 from this. We obtain
τ‖eh(1+h″)12θb2n˜u‖+‖eh(1+h″)12θb2n∇˜u‖≤τ‖eh(1+h″)12θb2nu1‖+‖eh(1+h″)12θb2n∇u1‖+τ‖eh(1+h″)12θb2nu2‖+‖eh(1+h″)12θb2n∇u2‖≤C(‖ehθ−b2n˜f‖+‖ehθb2n˜F‖+τ1+b2‖eh˜g‖0). |
Using the bulk-boundary interpolation estimate from Lemma 2.2, this can further be strengthened by a boundary contribution on the left hand side:
τ1−b2‖eh(1+h″)12˜u‖0+τ‖ehθb2n(1+h″)12˜u‖+‖ehθb2n(1+h″)12∇˜u‖≤C(‖ehθ−b2n˜f‖+‖ehθb2n˜F‖+τ1+b2‖eh˜g‖0). |
Transforming back into Cartesian coordinates yields the desired estimate.
Analogous arguments as in the proof of Proposition 3.1 allow us to derive a slight variation of the estimate from Proposition 3.1 which we will exploit in the sequel.
Corollary 3.2. Let K be a finite set of positive integers. Under the same assumptions as in Proposition 3.1, for each τ>τ0≫1 there is a weight function h(−ln(|x|)) such that there exists a constant C>0 which is independent of τ such that for any k∈K it holds
τ‖eh(−ln(|x|))xb2n+1(1+¯h)12|x|−k−1u‖L2(Rn+1+)+‖eh(−ln(|x|))xb2n+1(1+¯h)12|x|−k∇u‖L2(Rn+1+)+τ1−b2‖eh(−ln(|x|))(1+¯h)12|x|−k+b−12u‖L2(Rn×{0})≤C(‖eh(−ln(|x|))|x|−k+1x−b2n+1f‖L2(Rn+1+)+τ‖eh(−ln(|x|))xb2n+1|x|−kF‖L2(Rn+1+)+τ1+b2‖eh(−ln(|x|))|x|−k+1−b2g‖L2(Rn×{0})). |
Proof. Let h be the family of Carleman weights constructed in the proof of Proposition 3.1 and consider for k∈K the following new family
˜h(t)=h(t)−kt. |
It also satisfies (3.2) for τ>τ0 and τ0 sufficiently large: Notice that ˜h′ differs from h′ by a constant and hence all the higher order derivatives coincide. In particular, the first and the last estimates in (3.2) hold directly. Adding a suitable extra term of the form αt to the original h, we can ensure to keep a positive distance between ˜h′ and the spectrum of ∇Sn⋅θbn∇Sn for finitely many k.
Therefore, estimate (3.1) holds with ˜h playing the role of h. Finally, observe that
e˜h(−ln(|x|))=eh(−ln(|x|))|x|−k(1+¯˜h)=(1+¯h), |
so the claimed estimate is obtained.
Considering second order equations of the form
∇⋅xbn+1a∇u=f in B+4,limxn+1→0xbn+1∂n+1u=g on B′4, | (3.22) |
in the sequel, we seek to introduce variable coefficients in the Carleman estimate of Proposition 3.1. Throughout this section, the metric a is assumed to be of a block form as in (1.4) where the metric ˜a satisfies the conditions (A1)–(A3) from the introduction with μ=0.
We first note that the estimate in Proposition 3.1 remains valid for a constant coefficient metric in the block form (1.4). This follows immediately from a change of coordinates (only involving the tangential variables). In order to extend Proposition 3.1 to variable coefficient problems, we exploit the presence of the divergence contribution and in conformal coordinates localise the problem to scales of the size C(ajτ)−12 or of size one, respectively (depending on the size of the metric). Here {aj}j∈N denotes the sequence that was used in the definition of the Carleman weight h (see Step 1 in the proof of Proposition 3.1).
We follow the argument in [13] and argue in two steps: First, in the regime in which h is convex, we localise to very small scales (Lemma 3.3). In the regime in which no convexity is present anymore, we localise to scales of order one in conformal coordinates (Lemma 3.4). Finally, we patch these estimates together to derive the desired global bound of Proposition 1.7.
Lemma 3.3. Let τ≥1 and ϵ>0. Assume that h is convex and
h′∈[τ,2τ], h″∈[ϵτ,2ϵτ], |h‴|≤τ. |
Assume that
|x||∇aij(x)|≤δϵ in Iℓ:={x∈Rn+1+:|x|∈[e−ℓ−1,e−ℓ]} |
for some sufficiently small constant δ>0. Then, for all u with supp(u)⊂Iℓ and all τ≥τ0≥1 we have
τ‖eh(−ln(|x|))xb2n+1(1+ϵτ)12|x|−1u‖L2(Rn+1+)+‖eh(−ln(|x|))xb2n+1(1+ϵτ)12∇u‖L2(Rn+1+)+τ1−b2‖eh(−ln(|x|))(1+ϵτ)12|x|b−12u‖L2(Rn×{0})≤C(‖eh(−ln(|x|))|x|x−b2n+1∂ixbn+1aij∂ju‖L2(Rn+1+)+τ1+b2‖eh(−ln(|x|))|x|1−b2limxn+1→0xbn+1∂n+1u‖L2(Rn×{0})). |
Proof. Step 1: Restricted support. We first assume that u is supported on a ball BC0|x0|(ϵτ)−12(x0) for some x0∈Iℓ or in some half ball B+C0|x0|(ϵτ)−12(x0) for some x0∈Iℓ∩(Rn×{0}) and where the constant C0>0 is still to be determined (see Step 2). As the arguments are similar in both cases, we only discuss the case of the full ball in detail in the sequel. We note that for x∈BC0|x0|(ϵτ)−12(x0)∩Iℓ we have
|aij(x)−aij(x0)|≤supx∈BC0|x0|(ϵτ)−12(x0)|∇aij(x)||x−x0|≤CδϵC0|x0|(ϵτ)−12|x0|≤CC0δ(ϵτ−1)12. | (3.23) |
Next, we apply the Carleman estimate from Proposition 3.1 to the equation
∂ixbn+1aij(x0)∂ju=f+∂i(aij(x0)−aij(x))∂ju in BC0|x0|(ϵτ)−12(x0),limxn+1→0xbn+1∂n+1u=g, |
where f=∂ixbn+1aij∂ju and where we recall the block structure (1.4) of aij. By virtue of Proposition 3.1 we obtain
τ‖eh(−ln(|x|))xb2n+1(1+ϵτ)12|x|−1u‖L2(Rn+1+)+‖eh(−ln(|x|))xb2n+1(1+ϵτ)12∇u‖L2(Rn+1+)+τ1−b2‖eh(−ln(|x|))(1+ϵτ)12|x|b−12u‖L2(Rn×{0})≤C(‖eh(−ln(|x|))|x|x−b2n+1f‖L2(Rn+1+)+τ‖eh(−ln(|x|))xb2n+1(aij−aij(x0))∂ju‖L2(Rn+1+)+τ1+b2‖eh(−ln(|x|))|x|1−b2limxn+1→0xbn+1∂n+1u‖L2(Rn×{0})). | (3.24) |
In order to deduce the desired estimate under the support constraint, it suffices to bound the second bulk term on the right hand side of (3.24). To this end, we invoke (3.23) and estimate
‖eh(−ln(|x|))xb2n+1(aij−aij(x0))∂ju‖L2(Rn+1+)≤CC0δ(ϵτ−1)12‖eh(−ln(|x|))xb2n+1∇u‖L2(Rn+1+). |
For δ>0 sufficiently small (but independent of u), it is possible to absorb this contribution into the left hand side of (3.24).
Step 2: Localisation. We seek to apply the previous argument by localising a general solution u with supp(u)⊂Iℓ by a partition of unity. Here commutator estimates play a crucial role and provide a natural limitation to the possible localisation scale.
We consider a partition of unity {ψk}k∈{1,…,K} associated with the half annulus Iℓ and a finite collection {xk}k∈{1,…,K} of points in Iℓ such that supp(ψk)⊂BC0|xk|(ϵτ)−12(xk) or supp(ψm)⊂B+C0|xm|(ϵτ)−12(xm) (where in the latter case xm∈Iℓ∩(Rn×{0})) and such that the balls and half-balls BC0|xk|(ϵτ)−12(xk) and B+C0|xm|(ϵτ)−12(xm) cover the interval Iℓ (with controlled overlap). Without loss of generality, we choose the partition of unity and the points xk such that the following estimate hold:
|∇αψk|≤C−|α|1|xk|−|α|(ϵτ)|α|2 for |α|∈{0,1,2}, | (3.25) |
and C1=C1(C0)>0. Further, by for instance choosing ψk to have a dependence on |x−xk| only, we may assume that
limxn+1→0xbn+1∂n+1ψk(x)=0 |
in Iℓ. We then write u=K∑k=1ψku. With this in hand, we apply the triangle inequality as well as the Carleman estimate from Step 1:
τ‖eh(−ln(|x|))xb2n+1(1+ϵτ)12|x|−1u‖L2(Rn+1+)+‖eh(−ln(|x|))xb2n+1(1+ϵτ)12∇u‖L2(Rn+1+)+τ1−b2‖eh(−ln(|x|))(1+ϵτ)12|x|b−12u‖L2(Rn×{0})≤K∑k=1(τ‖eh(−ln(|x|))xb2n+1(1+ϵτ)12|x|−1ψku‖L2(Rn+1+)+‖eh(−ln(|x|))xb2n+1(1+ϵτ)12∇(ψku)‖L2(Rn+1+))+K∑k=1τ1−b2‖eh(−ln(|x|))(1+ϵτ)12|x|b−12ψku‖L2(Rn×{0})≤CK∑k=1(‖eh(−ln(|x|))x−b2n+1|x|fk‖L2(Rn+1+)+τ1+b2‖eh(−ln(|x|))|x|1−b2limxn+1→0xbn+1∂n+1(ψku)‖L2(Rn×{0})). | (3.26) |
We first consider the bulk contribution on the right hand side for which
fk=∂ixbn+1aij∂j(uψk)=ψkf+2aijxbn+1∂iψk∂ju+uaij∂i(xbn+1∂jψk)+uxbn+1(∂iaij)(∂jψk). |
Using the bounds for ψk from (3.25) as well as the estimate for |∇aij|, we obtain
‖eh(−ln(|x|))|x|x−b2n+1fk‖L2(Rn+1+)≤‖eh(−ln(|x|))|x|x−b2n+1fψk‖L2(Rn+1+)+C−11(1+(ϵτ)12)‖eh(−ln(|x|))xb2n+1ψk∇u‖L2(Rn+1+)+C−11(1+ϵτ)‖eh(−ln(|x|))xb2n+1|x|−1ψku‖L2(Rn+1+). | (3.27) |
Choosing C1=C1(C0)>1 sufficiently large (by choosing C0>0 appropriately), then allows us to absorb the second and third contribution from the right hand side of (3.27) into the left hand side of (3.26).
For the boundary term in (3.26), we use the fact that by construction of the partition of unity limxn+1→0xbn+1∂n+1(ψku)=ψklimxn+1→0xbn+1∂xn+1u.
Using this and the finite overlap of the supports of the functions ψk, allows us to turn (3.26) into
τ‖eh(−ln(|x|))xb2n+1(1+ϵτ)12|x|−1u‖L2(Rn+1+)+‖eh(−ln(|x|))xb2n+1(1+ϵτ)12∇u‖L2(Rn+1+)+τ1−b2‖eh(−ln(|x|))(1+ϵτ)12|x|b−12u‖L2(Rn×{0})≤CK∑k=1(‖eh(−ln(|x|))|x|x−b2n+1ψkf‖L2(Rn+1+)+τ1+b2‖eh(−ln(|x|))|x|1−b2ψklimxn+1→0xbn+1∂n+1u‖L2(Rn×{0}))≤C(‖eh(−ln(|x|))x−b2n+1|x|f‖L2(Rn+1+)+τ1+b2‖eh(−ln(|x|))|x|1−b2limxn+1→0xbn+1∂n+1u‖L2(Rn×{0})), | (3.28) |
which concludes the proof of the argument.
Similarly as in Lemma 3.3, it is also possible to deal with the situation of even smaller perturbations of the metric without invoking the convexity of the weight:
Lemma 3.4. Let τ≥1. Assume that
|x||∇aij|≤δτ−1 in Iℓ:={x∈Rn+1+:|x|∈[e−ℓ−1,e−ℓ]} |
for some sufficiently small constant δ>0. Then, for all u with supp(u)⊂Iℓ and all τ≥τ0≥1 we have
τ‖eh(−ln(|x|))xb2n+1|x|−1u‖L2(Rn+1+)+‖eh(−ln(|x|))xb2n+1∇u‖L2(Rn+1+)+τ1−b2‖eh(−ln(|x|))|x|b−12u‖L2(Rn×{0})≤C(‖eh(−ln(|x|))|x|x−b2n+1∂ixbn+1aij∂ju‖L2(Rn+1+)+τ1+b2‖eh(−ln(|x|))|x|1−b2limxn+1→0xbn+1∂n+1u‖L2(Rn×{0})). |
Proof. The argument follows along the same lines as the proof of Lemma 3.4, however now we directly localise to scales of the order C0|xk| around a finite number of points xk∈Iℓ. Using an associated partition of unity then yields the desired result.
Relying on the previous result, we obtain global Carleman estimates:
Proposition 3.5. Let the metric a:Rn+1+→R(n+1)×(n+1) be of a block form as in (1.4) where the metric ˜a is assumed to satisfy the conditions (A1)–(A3) with μ=0. Let u∈H1(B+4,xbn+1) with supp(u)⊂B+4∖{0}. Then, for each τ>τ0≥1 there exist a weight function h and a constant C>0 independent of τ such that it holds
τ‖eh(−ln(|x|))xb2n+1(1+¯h)12|x|−1u‖L2(Rn+1+)+‖eh(−ln(|x|))xb2n+1(1+¯h)12∇u‖L2(Rn+1+)+τ1−b2‖eh(−ln(|x|))(1+¯h)12|x|b−12u‖L2(Rn×{0})≤C(‖eh(−ln(|x|))|x|x−b2n+1∇⋅xbn+1a∇u‖L2(Rn+1+)+τ1+b2‖eh(−ln(|x|))|x|1−b2limxn+1→0xbn+1∂n+1u‖L2(Rn×{0})). |
Here ¯h(x):=h″(t)|t=−ln(|x|).
Proof. In order to deduce this, we use the properties of the weight h. By relying on a partition of unity, we localise the set-up to dyadic intervals. Then, with the constants aj as in Step 1 in the proof of Proposition 3.1, if ajτ>1, we apply Lemma 3.3, while if ajτ<1, we invoke Lemma 3.4.
As before, by the same arguments as in the proof of Proposition 3.5, it is possible to obtain a slight variation of this, which we will use in the sequel:
Corollary 3.6. Let K be a finite set of positive integers. Under the same assumptions as in Proposition 3.5, for each τ>τ0≫1 there is a weight function h(−ln(|x|)) such that there exists a constant C>0 which is independent of τ such that for any k∈K it holds
τ‖eh(−ln(|x|))xb2n+1(1+¯h)12|x|−k−1u‖L2(Rn+1+)+‖eh(−ln(|x|))xb2n+1(1+¯h)12|x|−k∇u‖L2(Rn+1+)+τ1−b2‖eh(−ln(|x|))(1+¯h)12|x|−k+b−12u‖L2(Rn×{0})≤C(‖eh(−ln(|x|))|x|−k+1x−b2n+1∇⋅xbn+1a∇u‖L2(Rn+1+)+τ1+b2‖eh(−ln(|x|))|x|−k+1−b2limxn+1→0xbn+1∂n+1u‖L2(Rn×{0})). |
Proof. The proof follows in the same way as Proposition 3.5 but instead of using Proposition 3.1 we use Corollary 3.2.
Proposition 1.7 arises as an iteration of the Carleman estimate from Proposition 3.5 (or directly from Proposition 3.1 if aij=δij). In the sequel, we present the variable and constant coefficient proofs simultaneously.
Proof of Proposition 1.7. We seek to iterate the second order Carleman estimates in order to obtain an estimate for the full system. To this end, we apply Corollary 3.6 with K={kj}mj=0, kj=2m−2j to ˜wj=(1+¯h)m−j2˜uj. We argue in two steps.
Step 1: Building block estimate. For j∈{0,…,m} we have
Lb˜wj(x)=(1+¯h)m−j2(Lb˜uj)(x)+2∇((1+¯h)m−j2)⋅a(x)∇˜uj(x)+˜uj(x)Lb((1+¯h)m−j2)=(1+¯h)m−j2(fj(x)+˜uj+1(x))+2∇((1+¯h)m−j2)⋅a(x)∇˜uj(x)+˜uj(x)Lb((1+¯h)m−j2). |
Hence, as consequence of Corollary 3.6, we deduce the estimate
τm+1−j‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−j2|x|−2m−1+2j˜uj‖L2(B+4)+τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)12|x|−2m+2j∇((1+¯h)m−j2˜uj)‖L2(B+4)≤C(τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)m−j2|x|−2m+1+2j(fj+˜uj+1)‖L2(B+4)+τm−j+1+b2‖eh(−ln(|x|))(1+¯h)m−j2|x|−2m+2j+1−b2gj‖L2(B′4)+τm−j‖eh(−ln(|x|))xb2n+1|x|−2m+1+2j∇((1+¯h)m−j2)⋅a∇˜uj‖L2(B+4)+τm−j‖eh(−ln(|x|))xb2n+1|x|−2m+1+2jLb((1+¯h)m−j2)˜uj‖L2(B+4)). | (3.29) |
By virtue of the triangle inequality, the second term on the left hand side of (3.29) can be estimated as
τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)12|x|−2m+2j∇((1+¯h)m−j2˜uj)‖L2(B+4)≥τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−j2|x|−2m+2j∇˜uj‖L2(B+4)−τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)12|x|−2m+2j˜uj∇((1+¯h)m−j2)‖L2(B+4). | (3.30) |
Using the last condition in (3.2), we obtain
|∇α((1+¯h)m−j2)|≤Cε|x|−|α|(1+¯h)m−j2, for |α|∈{1,2}, |
Hence, the negative term on the right hand side of (3.30) can be absorbed into the first left hand side contribution in (3.29) if τ≥τ0>1 is sufficiently large. Further by the regularity of the metric a, and by choosing ε sufficiently small, we may absorb the last two contributions on the right hand side of (3.29) into the left hand side of (3.29).
As a consequence, we obtain
τm+1−j‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−j2|x|−2m−1+2j˜uj‖L2(B+4)+τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−j2|x|−2m+2j∇˜uj‖L2(B+4)≤C(τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)m−j2|x|−2m+1+2j(fj+˜uj+1)‖L2(B+4)+τm−j+1+b2‖eh(−ln(|x|))(1+¯h)m−j2|x|−2m+2j+1−b2gj‖L2(B′4)). | (3.31) |
For later use, we remark that an analogous argument for ˜wj,ℓ:=(1+¯h)ℓ2˜uj for ℓ∈Z and k∈{kj}mj=0 yields
τm+1−j‖eh(−ln(|x|))xb2n+1(1+¯h)ℓ+12|x|−k−1˜uj‖L2(B+4)+τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)ℓ+12|x|−k∇˜uj‖L2(B+4)≤C(τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)ℓ2|x|−k+1(fj+˜uj+1)‖L2(B+4)+τm−j+1+b2‖eh(−ln(|x|))(1+¯h)ℓ2|x|−k+1−b2gj‖L2(B′4)). | (3.32) |
Step 2: Iteration. Using the C2m,1 coefficient regularity, we iterate the building block estimate:
τm+1−j‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−j2|x|−2m−1+2j˜uj‖L2(B+4)+τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−j2|x|−2m+2j∇˜uj‖L2(B+4)≤C(τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)m−j2|x|−2m+1+2jfj‖L2(B+4)+τm−j‖eh(−ln(|x|))xb2n+1(1+¯h)m−j2|x|−2m+1+2j˜uj+1‖L2(B+4))≤C(j+1∑k=jτm−k‖eh(−ln(|x|))xb2n+1(1+¯h)m−k2|x|−2m+1+2kfk‖L2(B+4)+τm−j−1‖eh(−ln(|x|))xb2n+1(1+¯h)m−j−12|x|−2m+1+2(j+1)˜uj+2‖L2(B+4))≤C(m∑k=jτm−k‖eh(−ln(|x|))xb2n+1(1+¯h)m−k2|x|−2m+1+2kfk‖L2(B+4)+τ1+b2‖eh(−ln(|x|))|x|1−b2g‖L2(B′4)). |
Here we used that gj=0 except for gm=g and applied the triangle inequality to separate the bulk terms on the right hand side. Summing these estimates from j=0 to m, then yields the desired bulk estimate in (1.7).
Step 3: Boundary-bulk interpolation. In order to deduce the boundary estimate, we apply Lemma 2.2 to the function
eh(−ln(|x|))(1+¯h)m+12|x|−2m+b−12˜u0(x) |
on each sphere |x|=r. {Recall that h and ¯h are independent of the spherical variables so the integration with respect to the radial directions implies
τm+1−b2‖eh(−ln(|x|))(1+¯h)m+12|x|−2m+b−12˜u0‖L2(B′4)≤C(τm+1‖eh(−ln(|x|))xb2n+1(1+¯h)m+12|x|−2m−1˜u0‖L2(B+4)+τm‖eh(−ln(|x|))xb2n+1(1+¯h)m+12|x|−2m∇˜u0‖L2(B+4)). |
Combining this with Step 2 then concludes the proof.
Proof of Proposition 1.8. We split the proof into two steps: First we deal with commutation relations arising in iterations of the second order estimates of Proposition 3.5 and then we iterate the resulting bounds.
Step 1: Commutators. We observe that
∂′kLb=Lb∂′k+(∂′k˜aij)∂′ij+(∂′ik˜aij)∂′j, | (3.33) |
where ∂′k denotes derivatives in tangential directions.
Due to the assumptions in condition (A2), the contributions in the Carleman estimate arising from ∇′Lbuj−1 for j∈{1,…,m} can be bounded from below by terms which are controlled by Lb∇′uj−1 if δ>0 is chosen sufficiently small: Indeed, by using the commutator (3.33), we first obtain
‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−j2|x|−2m+2j∇′Lb˜uj−1‖L2(B+4)(3.33)≥‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−j2|x|−2m+2jLb∇′˜uj−1‖L2(B+4)−‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−j2|x|−2m+2j|∇a||(∇′)2uj−1|‖L2(B+4)−‖eh(−ln(|x|))xb2n+1(1+¯h)m+1−j2|x|−2m+2j|∇2a||∇′˜uj−1|‖L2(B+4). |
Invoking the bounds from condition (A2) for the metric and interpolating, we obtain |x||∇a(x)|, |x|2|∇2a(x)|≤˜Cδ for x∈B+4 and thus
Applying the Carleman estimate from Step 1 in combination with equation (3.32) in the proof of Proposition 1.7 with , , it is possible to bound
whence, in combination with the previous estimates,
Choosing so small that , it is possible to absorb the last contribution into the left hand side, which leads to the estimate
This allows us to apply the estimate (3.32) from Step 1 in the proof of Proposition 1.7 with , and thus to deduce that for the following bounds hold
(3.34) |
Step 2: Iteration. With the additional bounds from (3.34) in hand, we can iterate the Carleman estimate. For and we infer
A similar estimate holds for the gradient term. Adding these estimates to the bounds from Proposition 1.7 and summing over all implies the desired bulk estimate.
Step 3: Boundary-bulk interpolation. The boundary estimates can be deduced as in the proof of Proposition 1.7. Applying Lemma 2.2 to the functions
on each sphere and integrating with respect to the radial directions we infer
Combining the previous steps concludes the proof.
In this section we study the strong unique continuation property for the system (2.3) and seek to reduce it to the weak unique continuation property. As in [8,10,19] we achieve this by careful compactness and blow-up arguments.
From a technical point of view, the main challenge is to control solutions to our system also in the normal direction in which we can only obtain information through the equation itself. Here two cases arise:
● If we had vanishing of infinite order in the tangential and normal directions, an immediate application of the Carleman estimate (1.8) would allow us to prove the strong unique continuation property.
● If we are however dealing with solutions which a priori do not vanish of infinite order in the normal {direction}, we have to argue more carefully, exploiting properties of our equations.
In this section, we consider solutions with of the system (1.5), (1.6) with satisfying the following conditions:
is of a block form (1.4) and such that satisfies (A1)–(A3) with and
with for and
Here is a sufficiently small constant (which is specified below), and are arbitrarily large, finite constants.
As the vanishing of infinite order in the normal directions is not a direct consequence of our assumptions on the infinite order vanishing in the tangential directions, we split the argument into two parts:
● In the case of infinite order vanishing in all directions, i.e., for all
we directly apply the Carleman estimate from Proposition 1.8 (see Section 4.2).
● If this is (a priori) not the case, i.e., there exist some , a subsequence and a constant such that
(4.1) |
we deduce doubling properties and then exploit these in a compactness argument to reduce the strong unique continuation property to the weak unique continuation property (see Section 4.1).
In the sequel, we seek to reduce the strong unique continuation property to a weak unique continuation result by a blow-up argument under the assumption that the solution vanishes to infinite order just in the tangential directions (but not in the normal directions, see (4.1)).
In order to deduce sufficient compactness for a blow-up argument, we first prove a doubling estimate for the functions . Here we exploit elliptic estimates and deal with the resulting boundary contributions by absorbing these into the bulk terms with finite order of vanishing (for sufficiently small radii).
Proposition 4.1 (Doubling). Let for be weak solutions of the system (1.5), (1.6) with satisfying the conditions from (C). Assume also that the tangential restrictions vanish of infinite order at and that there exist some , a subsequence and a constant such that (4.1) holds. Then, there exist a universal constant , a positive integer and a radius (the latter depending on ) such that for any and all we have
(4.2) |
Here denotes the radii from (4.1).
Remark 4.2. In the case of bounded potentials, the same result holds without assuming that the functions vanish of infinite order in the tangential directions. Moreover, in the setting of bounded potentials, the statement holds for all (there is no intersection with the interval around here), where is sufficiently small but independent of . We refer to the proof of Proposition 4.1 for further details on this.
Remark 4.3. Instead of restricting our doubling results to radii around , we could also have argued as in Section 3 in [36]. This would have allowed us to conclude that the vanishing order is defined not only through a subsequence of radii but is independent of such a sequence. As a consequence, we would have obtained the statement of Proposition 4.1 for any choice of radius less that . As our unique continuation argument does not rely on quantitative order of vanishing estimates, we do not further pursue this approach here.
Proof of Proposition 4.1. Consider , where is a radial cut-off function which is equal to one in , vanishes outside of and satisfies the following bounds
Here where is chosen sufficiently small (with a choice that is explained later). We note that the functions are solutions to the system from Proposition 1.7 with and . Hence the Carleman estimates (1.7) and (1.8) hold.
Step 1: Boundary contributions. Let us assume that for all it holds , where is a constant which will be specified below. We seek to absorb the boundary contributions from the right hand side of the estimate (1.8) in Proposition 1.8 into the ones on the left hand side. To this end, we compare the relevant contributions: The left hand side of the Carleman estimate controls contributions of the form
(4.3) |
while the boundary terms on the right hand side can be estimated from above by
(4.4) |
In order to carry out the absorption argument, we distinguish three cases:
● If , for any , it suffices to choose sufficiently large, in order to absorb the contributions from (4.3) into (4.4). Note that this always holds for and also for if .
● If and , it is still possible carry out this absorption argument in the case that provided the constant is small enough. More precisely, after plugging the estimate (4.3) into (1.8), the relevant boundary contribution will carry the prefactor . Requiring that then allows us to implement an absorption argument.
● Lastly, in the case of subcritical potentials, i.e., if , a wider range of values of is admissible by using the properties of . Indeed, by the construction of
Analogous estimates hold for the gradient contributions. Hence, by a sufficiently small choice of (in the construction of the Carleman weight in the proof of Proposition 3.1 which in particular is compatible with the other requirements there) it is possible to absorb all boundary contributions as long as for any finite constant by choosing sufficiently large. This enlarges the range of for to and for to .
Step 2: Bulk contributions. In discussing the bulk contributions, we first deal with the bulk terms of the right hand side of the Carleman estimate which are localised on the unit scale. We will absorb these into the left hand side of the Carleman estimate. Secondly, we treat the contributions on the small scale for which we deduce the desired doubling estimate.
In the sequel, we use the following abbreviations for the respective half annuli
For the convenience of the reader, we split the proof of the bulk estimates into two steps: In Step 2a, we deal with the case without gradient contributions. This allows us to introduce the ideas without resorting to too many technicalities. Then, in Step 2b, we deal with the full case including gradient terms.
Step 2a: Lowest order potentials. After having dealt with the boundary terms in Step 1, the Carleman estimate (1.7) turns into
(4.5) |
with .
As a first simplification step, we deal with the contributions on the unit scale: Using the monotonicity of , we infer
(4.6) |
where . Estimating the terms on the left hand side of (4.5) from below by
and relying on the monotonicity of , by choosing sufficiently large, we can absorb the contribution (4.6) into the left hand side of (4.5) (this yields a dependence of on , but only on the unit scale).
As a consequence, we are left with the estimate
(4.7) |
where has been fixed in the previous step. Using the monotonicity of , the bound of and the estimates on the derivatives of in , we obtain
(4.8) |
We observe that the difference is bounded independently of , since
where , and .
Thus, dividing (4.8) by and adding to both sides
we obtain
(4.9) |
Step 2b: Gradient potentials. The proof is similar to the one in Step 2a but instead of the Carleman estimate from Proposition 1.7, we here use the one from Proposition 1.8. After Step 1, estimate (1.8) becomes
Considering the bounds for derivatives of and the metric (i.e., ), and repeating the same arguments as in Step 2a, we arrive at
(4.10) |
Step 3: Caccioppoli's inequality. It remains to control the gradient terms of the right hand side in (4.10) and in (4.9). We can apply Lemma 2.3 to with , if and , if .
If we just consider the lowest order potentials (i.e., where in the bounds for only is needed), tangential derivatives are not necessary and after summing over with suitable factors we arrive at
(4.11) |
We seek to absorb the boundary terms which occur in (4.11) into the bulk term on the right hand side of (4.11). Before explaining the general case, we discuss the case . If , then the boundary term can be absorbed into the bulk terms for with some independent of : By scaling, it is enough to prove this for . If , we do not directly start from (4.11) but return to the proof of the Caccioppoli estimate which yields
(4.12) |
with as in the proof of Lemma 2.3. By Young's inequality we can then estimate the boundary term as
for some constant . Using the Poincaré type inequality for , we obtain the bound
In order to control the gradient term involving we note that it is possible to absorb it into the left hand side of (4.12) by choosing small enough. In order to bound the gradient contribution , we use again the Caccioppoli estimate (and observe that in this case no boundary terms appear). Therefore (4.2) is valid for and . For the case , we refer to Lemma 5.1 in [37].
Let us now return to the general case: If we also consider gradient potentials (i.e., where the full bound is needed), a similar estimates holds after considering in (2.5) tangential derivatives up to the order :
(4.13) |
Now the boundary terms in (4.11) and (4.13) (which in general are not in but might become singular) can be controlled as follows: We first notice that
and
with a suitable cut-off function and . Since for any , given any there is a radius such that if
Therefore
On the other hand, only vanishes of finite order, so choosing sufficiently small, the boundary term can be absorbed into the bulk terms {for suitable ranges of : Indeed, from (4.1) we know that for some and some there exist and such that if
and therefore if , with some constant (which we will fix later),
(where ). So the bulk term can absorb the boundary contribution in the right hand side of (4.11) and (4.13) if is small enough.
Hence, (4.11) becomes
and (4.13) turns into
(4.14) |
In order to deal with the remaining derivatives on the left hand side in (4.14), we notice that
(4.15) |
In order to estimate the remaining (lower order) gradient terms we seek to iterate the Caccioppoli estimate up to times. This leads to an estimate of the type (4.15) where the norms on the right hand side are evaluated in increasingly larger balls. In particular, in order to obtain radii of the size in the end, the beginning of the iteration should be carried out with radii of the size . Notice that in the Caccioppoli estimates the boundary terms only appear in the estimate for , so they will only appear in the first iteration, where . In order to carry out the iteration argument of the Caccioppoli estimate as outlined above, we thus have to ensure that for some . This is satisfied if with . For technical reasons appearing in the proof of Proposition 4.4, we choose .
With the doubling property in hand, we apply a blow-up argument reducing the strong unique continuation property to the weak unique continuation property. To this end, we introduce the following rescaled functions:
(4.16) |
We exploit the previous compactness arguments to pass to the blow-up limit which leads to a boundary weak unique continuation problem of the blown-up system:
Proposition 4.4. Let for be weak solutions of the system (1.5), (1.6) with satisfying the conditions from (C). Assume also that the tangential restrictions vanish of infinite order at and that there exist some , a subsequence and a constant such that
Let be the rescaled functions defined by (4.16) and let denote the sequence of radii from (4.1). Then, along a subsequence , for some , with we have strongly in , where the functions are weak solutions to the following elliptic system:
(4.17) |
with . Moreover, for all we have
and
(4.18) |
Proof. We first note that the functions are constructed in such a way that
Applying Proposition 4.1 twice with (so both and for any given ), we obtain
After rescaling, this turns into
which holds holds uniformly in for the whole family . By Rellich's compactness theorem, there exist a subsequence with and functions such that strongly in and weakly in and the normalisation (4.18) holds. In addition, since the embedding is continuous and is compactly embedded in , up to a redefinition of the subsequence, strongly in and weakly in .
The functions satisfy weakly the same system (1.5), (1.6) as the original functions (again with ) however with a rescaled metric and potentials and
Hence,
for any . As a result, in the limit
Here we have used that by the normalising condition (A3) the metric satisfies as and that the boundary integrals vanish of infinite order as we proved in Step 3 of the proof of Proposition 4.1. This shows that the functions are indeed weak solutions to the system in the statement.
Finally, we prove that the functions , , vanish on . Indeed,
and whereas the numerator vanishes of infinite order, by our assumption (4.1), the denominator vanishes of only finite order.
Here we deduce the strong unique continuation property for solutions which vanish of infinite order in tangential and normal directions. The proof relies on the associated Carleman estimates from Proposition 1.8.
Proposition 4.5. Let for be weak solutions of the system (1.5), (1.6) with satisfying the conditions from (C). Assume further that for all and all
Then, in .
Proof. Consider the functions , where is a smooth, radial cut-off function which satisfies the following bounds:
(4.19) |
where . The functions are solutions to the system from Proposition 1.7 with .
We insert the functions into the Carleman estimate (1.8). Notice that we can pass to the limit by virtue of the infinite rate of vanishing of . Arguing as in Step 1 of the proof of Proposition 4.1, we can drop the boundary contributions. Therefore (1.8) turns into
Finally, passing to the limit , we obtain in for all .
In this section we consider the weak unique continuation property for solutions to the system (4.17). In spite of weak unique continuation results for the fractional Laplacian already existing in the literature (see in particular [18]), both our argument and our result contain novel aspects: In contrast to the weak unique continuation results from Seo [18] our result is a localised unique continuation result (as we do not need the validity of the equation in ), and hence in particular it is formulated for a local equation (instead of working with the global fractional Laplacian).
Proposition 5.1. Let for be weak solutions of the system (1.5), (1.6) in with , and the metric of a block form (1.4) where satisfies the conditions (A1)–(A3). Assume also that for all the tangential restrictions vanish on . Then, in .
Proof. We bootstrap the system by applying the weak unique continuation property for scalar equations: Indeed, by the weak unique continuation property of solutions of the fractional Laplacian (see [10] and [19]) and regularity results from [25], we first infer that in . Iteratively, this then also entails that in since, once in , then satisfies the Caffarelli-Silvestre equation with zero Dirichlet and zero (weighted) Neumann data. We iterate this until we reach .
Remark 5.2. We remark that an argument for the WUCP had already been given by Riesz [39] (relying on certain regularity conditions, see the discussion in Remark 4.2 in [7]). Using a Kelvin transform he reduced it to the situation with data vanishing in the exterior of a domain. An argument of a related flavour for a much larger class of pseudodifferential operators was also used in [9] (see also [40]).
Remark 5.3. We remark that the (weak) unique continuation property requires the Lopatinskii condition to hold (see [41] and the references therein for a precise definition of the Lopatinskii condition). If this is violated even if "formally" there are sufficiently many boundary conditions prescribed, one will in general not be able to infer the vanishing of . This is for instance the case for problem
By simply invoking counting arguments these boundary conditions should yield an overdetermined system. They however do not (the function is a non-trivial solution), as the Lopatinskii condition is not satisfied.
As a consequence of the localised formulation of our weak unique continuation property, it for instance applies to settings which arise in inverse problems [7,8,9]. This allows us to prove the antilocality of the fractional Laplacian for any order with , i.e., it allows us to prove Proposition 1.9, which we postpone to the next section.
In this section, we rely on the connection between the systems representations for the higher order fractional Laplacian (see Proposition 2.1 as well as Propositions A.6 and A.7 in the Appendix) and–building on the previous compactness results-present the proofs of Theorems 1–4 and of Propositions 1.9 and 1.11.
We begin by proving Theorems 1–3:
Proofs of Theorems 1, 2 and 3. We seek to reduce the strong unique continuation properties for the fractional Laplacian to the previously deduced results on the systems case. We invoke Proposition 2.1 and rewrite the problem as a system of the form (2.3), where , and . We seek to apply a combination of Propositions 4.4, 4.5 and 5.1. To this end, we have to show that the functions with vanish of infinite order in the tangential directions on the boundary. By assumption, we have that the function vanishes of infinite order at in the tangential directions. In order to obtain the desired infinite order of vanishing of in the tangential directions on the boundary, we use an interpolation argument: Let be a smooth cut-off function which is equal to one on and which is supported in . Then,
Since for any and as , this implies that the same holds for and thus for with . Moreover, due to the previous identification of and in terms of and , the conditions from (C) are satisfied.
Hence, if we assume that (4.1) holds for the functions with , the blow-up argument from Proposition 4.4 is applicable and yields blow-up solutions with on for which simultaneously satisfy the normalisation condition (4.18). By the weak unique continuation result from Proposition 5.1, the tangential and normal vanishing of on however entails that in for , which contradicts (4.18). Therefore, (4.1) cannot hold and the functions for must vanish of infinite order in tangential and normal directions. Thus we invoke Proposition 4.5 to obtain in for . Using that the equation for the generalised Caffarelli-Silvestre extension holds globally, the vanishing of on propagates through the upper half plane : Indeed, by the weak unique continuation property for uniformly elliptic equation and by (2.3) we infer in . Plugging this into the equation for and again using the weak unique continuation property for solutions to uniformly elliptic equations in the upper half plane implies also in . Iterating this further leads to in , whence in . This concludes the argument.
In this section we prove Theorem 4 by reducing it to the weak unique continuation property for the generalised Caffarelli-Silvestre extension.
By the representations from Proposition 2.1 (see also Propositions A.6 and A.7) we rewrite the problem of Theorem 4 as a system of the form (2.3), where , and . Notice that by Proposition 2.1 and the assumptions in Theorem 4 we have that with and . By assumption, vanishes on a measurable set of density one at .
Under these assumptions and supposing that (4.1) holds, we prove an analogous blow-up result as in Proposition 4.4:
Proposition 6.1. Let with be the functions from above and let with be the associated rescaled functions defined in (4.16). Suppose further that (4.1) holds. Then, along a subsequence with we have strongly in , where is a weak solution to the following elliptic system
Moreover, for all we have
and
In order to obtain the properties of the blow-up limit, we deduce a smallness condition for the (not yet blown-up) function in tangential directions on the boundary. By virtue of an interpolation inequality, this will be inherited to all the (not yet blown-up) functions with on the boundary.
Lemma 6.2. Let with be as in Proposition 6.1. For any , there exists a radius such that if
Proof. Since is a point of density one in , given , there exists a radius such that if
On the other hand, using Hölder's inequality ()
(6.1) |
By Sobolev and trace inequalities
Now we use the estimate from Proposition 4.1, where according with Remark 4.2 no assumptions on the vanishing order of are necessary and it holds for with independent of :
Therefore, by combining this with (6.1) and recalling the definition of from above, we obtain
Choosing such that , the result holds.
Proof of Proposition 6.1. The proof of Proposition 6.1 follows along the same lines as the one of Proposition 4.4 until the moment of proving . Here we use Lemma 6.2 to obtain the same result: Indeed, after rescaling it implies
Therefore, in the limit ,
Since this holds for any , in particular for any sequence , we infer .
The proof of for relies on an interpolation result together with the previous bound: Considering a smooth cut-off function with in and , we obtain
By rescaling, we then also infer
Thus, the smallness of and also entails the smallness of and on the boundary. The remainder of the argument follows analogously as in the proof of Proposition 4.4.
Proof of Theorem 4. We again have a dichotomy: Rewriting the equation as a solution to a system of the form (2.3), we either have that all functions with vanish of infinite order at some point in or that (4.1) holds for the points in .
If (4.1) holds, we may apply Proposition 6.1 which leads to normalised blow-up solutions with which are non-trivial in . This however contradicts the weak unique continuation statement from Proposition 5.1 which implies {that in for . Hence, (4.1) cannot hold. Therefore, all the functions with must vanish of infinite order in tangential and normal directions at . This allows us to} directly apply Proposition 4.5, whence we conclude that in for and thus also in .
We turn to the proof of the antilocality result. As above we emphasise that in this case, we do not assume the validity of an equation on the whole space . Nevertheless the antilocality of the fractional Laplacian entails the claimed strong rigidity property.
Proof of Proposition 1.9. By Proposition 2.1 we can consider the extension and the functions for , which solve a system of the form (2.3). Thus, if and on , (2.3) reduces to the setting in Section 5, whence we conclude that on . Since the nonlocal equation is assumed to hold in , the vanishing of can be propagated to the whole upper half space , whence we conclude that in .
With the global unique continuation result of Proposition 1.9 in hand, the proof of Proposition 1.11 follows by a duality argument and the Fredholm property of the fractional Schrödinger equation (see [24]). In particular, this is of similar flavour as a number of approximation properties which had been used to treat inverse problems for nonlocal equation in [7,8,27].
We consider the fractional Schrödinger equation
(6.2) |
where is as in Proposition 1.11.
Considering the bilinear form
it is possible to show the well-posedness of the problem, provided zero is not a Dirichlet eigenvalue of the problem. This follows similarly as explained for instance in [7]. In this setting, we define the associated Poisson operator as
(6.3) |
where is a weak solution to (6.2). With this preparation, we address the proof of Proposition 1.11:
Proof of Proposition 1.11. It suffices to prove that the set
is dense in , where is any open subset. We can suppose without loss of generality that by assupmtion. As in [7] we rely on the Hahn-Banach theorem: Assuming that is such that for all , it suffices to show that . In order to infer this, we note that by the well-posedness results for the inhomogeneous fractional Schrödinger equation, we may define to be a solution to
Then, as in [7],
As a consequence, in . Thus, by Proposition 1.9, the function vanishes identically in , whence also . This concludes the argument.
The authors declare no conflict of interest.
In this section, in order to keep our presentation self-contained, we connect the previous discussion on systems with certain boundary conditions to the properties of the higher order fractional Laplacian. Here we mainly recall several known results from the literature and rely heavily on the observations from [1,15,21] but also refer to [42,43,44] and the references therein.
We split the section into two parts: First, we derive the representation of the constant coefficient higher order fractional Laplacian operators through a generalised Caffarelli-Silvestre extension. Next, we deduce analogous results for operators with non-constant coefficients.
The starting point of our discussion is the definition of the fractional Laplacian as a Fourier multiplier:
where denotes the Fourier transform. Since we seek to study the unique continuation property of the higher order fractional Laplacian by techniques which are available for local (possibly weighted) equations, we are particularly interested in Caffarelli-Silvestre type extension properties for the higher order fractional Laplacian. These exist in different generalities, we only recall two of these and refer to the literature for more general results. As we aim at applying these characterisations of the (higher order) fractional Laplacian for our study of the unique continuation property, we limit ourselves here to showing that starting from the fractional Laplacian of a function , it is possible to find a suitable and sufficiently regular extension of which obeys a corresponding equation/ a corresponding system of equations. We however do not address the full equivalence (in that we do not show that any solution to the system at hand is related to the fractional Laplacian of a suitable function). For this we refer to the literature cited above.
We begin by recalling that also the higher order fractional Laplacian can be realised as the solution to a degenerate elliptic, second order boundary value problem [1,15]:
For we consider the equation
(A.1) |
Here we are interested in solutions
(ⅰ) which (by elliptic regularity) are classical solutions in ,
(ⅱ) which attain the boundary data with in an sense as ,
(ⅲ) and which have some decay at infinity in the sense that , where for we define .
Solutions to degenerate elliptic equations of this form have been investigated in the literature, even in the context of fully nonlinear equations [45]. Working with extensions of a problem from into , in this section, we use the notational convention that
As we view the strong unique continuation property for the fractional Laplacian as a strong boundary unique continuation property of the associated degenerate extension problem, it is one of our main goals to identify the associated boundary values for the generalised Caffarelli-Silvestre extension. In particular, we aim at showing that, as in the original Caffarelli-Silvestre characterisation of the fractional Laplacian, the formulation (A.1) also allows one to compute the fractional Laplacian as an "iterated Neumann" map from the knowledge of its generalised Caffarelli-Silvestre extension .
Lemma A.1. Let , and assume that . Let denote the tangential Fourier transform. Then there exists an extension operator
such that is a solution to
and
(A.2) |
Here and denotes a modified Bessel function of the second kind.
Also, there exists a constant such that
(A.3) |
Proof. We first derive the desired representation of the extension operator: To this end, we solve (A.1) by means of a tangential Fourier transform. Fourier transforming in the tangential directions, one obtains for the partial Fourier transform the following ODE
We rewrite and deduce a similar ODE for this function, but where we can scale out the contribution:
Further, setting for some function , we are lead to a modified Bessel equation for (as a function of ):
with corresponding initial conditions. Since we are looking for a function with decay at infinity, by the asymptotics of modified Bessel functions (see [47]) we infer that , where denotes a modified Bessel function of the second kind.
Returning to our original variables and using the asymptotics of as , we thus obtain that , where and .
By the regularity of for the function is .
In order to observe the convergence from (A.2) we note that
Using that as , the boundedness of as and the fact that for , as , this implies the desired limiting behaviour.
Finally, in order to obtain (A.3), we now use the asymptotics and recurrence relations of the modified Bessel functions [47]. We have
(A.4) |
Recalling the expression for (or rather ), we thus obtain
Abbreviating , we thus infer
As a consequence, for ,
Due to the asymptotics of the modified Bessel functions (see (A.4)) and the regularity of in the limit this implies the desired result for a proper choice of .
Corollary A.2. Let and let be the extension from Lemma A.1. Then we also have the following bulk estimates:
(A.5) |
where denotes the tangential Fourier transform.
Proof. In order to deduce the bulk estimates from (A.5) we note that
where we used the change of coordinates . Using an analogous change of coordinates, we also obtain
and
Here, in the passage from to , we used the recurrence relations (A.4). This concludes the discussion of the mapping properties of and provides the estimates from (A.5).
While the formulation (A.3) already provides a convenient alternative local characterisation of the fractional Laplacian as an iterated and weighted Dirichlet-to-Neumann map for a second order equation in the upper half-plane, if it is not immediately associated in a natural way with a finite energy (the quantity diverges in general).
In order to remedy this, in the sequel, we recall that the fractional Laplacian is also related to a Dirichlet-to-Neumann map for a system (or, equivalently, a higher order equation) which can naturally be associated with a finite energy [15]. This provides the natural functional analytic framework for our discussion of the unique continuation properties of the higher order fractional Laplacian and explains our focus on unique continuation properties for systems with Muckenhoupt weights in the earlier sections.
In order to derive the desired higher order equation for , we begin by discussing the bulk equation:
Lemma A.3. (Lemma 4.2 in [15]). Let and let be a solution to the bulk equation in (A.1). Then, for the function with and satisfies
(A.6) |
In particular,
(A.7) |
The equation (A.7) provides us with the bulk equation which we are working with in the sequel. For self-containedness, we recall the argument for Lemma A.3 from [15].
Proof. We show that if a function solves
(A.8) |
with , then solves
To this end, we observe that
This concludes the proof.
Next we seek to complement (A.7) with suitable boundary conditions. To this end, we use the explicit form of which was deduced in the proof of Lemma A.1. It entails the validity of certain weighted Neumann conditions and provides tangential limits for the associated higher order Dirichlet data:
Lemma A.4. Let , , let and let be the solution to (A.1) from Lemma A.1. For and , define , where and . Then we have that for all it holds
(A.9) |
for some constant . The functions are in and for any they obey the bounds
(A.10) |
Moreover, for any
(A.11) |
and for some constant
(A.12) |
Proof. First, by induction, we show that
(A.13) |
For this is true by Lemma A.1. It thus suffices to prove the induction step. By Lemma A.3 and the equation for we have
Using the claimed representation for , i.e., and the asymptotic recurrence relations for modified Bessel functions (A.4), this directly implies
The representation from (A.13) together with the asymptotics from (A.4) then directly entails (A.9), (A.11) and (A.12). Here as in the proof of Lemma A.1 all limits should be understood in the corresponding Sobolev spaces. Indeed, for instance, if , by invoking (A.4), we have
Thus, similarly as in the proof of Lemma A.1, we infer
Relying on the asymptotics of the Bessel functions and passing to the limit then leads to (A.11). The remaining limits are obtained analogously.
Finally, the bounds from (A.10) are deduced as those from (A.5) in Corollary A.2.
In analogy to the notation from Lemma A.1, we introduce the notation .
We next show that it is possible to obtain localised regularity estimates for the functions from Lemma A.3. This is helpful in the discussion of the global unique continuation properties for the fractional Laplacian. These are particularly relevant in the analysis of associated fractional inverse problems.
Lemma A.5. Let and assume that for some and . Then for all we have that
The argument for this relies on the pseudolocality of the operator at hand.
Proof. We split
where is a cut-off function that is one on for some and vanishes outside of . For the claim is a direct consequence of Lemma A.4. It hence suffices to study the regularity of . To this end, we argue as in [9]. For convenience of notation we only prove the argument for ; the argument for is analogous. Let be a second smooth cut-off function which is equal to one on the support of and vanishes outside of . Then,
By an explicit computation, we obtain that (this exploits formula 9.6.25 in [46]). Using the explicit form of or heat kernel estimates (as outlined in the next section and in [14]), we obtain that for any
(A.14) |
As the convolution in the expression for is only active in regions in which for some suitable , by virtue of Schur's lemma and an integration by parts, we then deduce that
whence
Integrating in , then implies that
which is the desired statement.
For the estimate of the normal derivative we notice that a short computation shows that
Arguing as above then concludes the proof.
We summarise the results from this section for :
Proposition A.6. Let and let . Then the function is a solution to the scalar higher order problem
All limits are understood in an sense.
Setting and defining the functions for , this can also be rewritten as the following system of second order equations
(A.15) |
where . Again, all limits are understood in an sense.
In this section, we derive analogous results to Proposition A.6 in the presence of variable coefficients, i.e., we are now concerned with the operator , where and the coefficients satisfy the conditions stated in (A1)–(A3) with . In contrast to the previous argument in which the Fourier transform diagonalised the tangential operator, we here rely on a spectral decomposition. We argue analogously as in [21] and thus only present the arguments formally. For convenience of notation we set
where is as above and .
To this end, we recall that for the self-adjoint, positive operator we can carry out a spectral decomposition and obtain a unique associated resolution of the identity which is supported on the spectrum of with
Based on this we can define the action of the heat semigroup and of the fractional powers of :
where .
Our main result in the context of variable coefficients mirrors the statement of the constant coefficient case:
Proposition A.7. Let and let . Then the function
is a solution to the scalar higher order problem (2.2). The function can also be represented as
Setting and defining the functions for , allows one to rewrite (2.2) as the the system (2.3), where . All boundary conditions hold in an sense.
Remark A.8. With the systems representation being established, we also obtain that . Then, direct energy estimates also yield that for all , if .
Proof. The fact that solves the equation
follows as in [21,42]. By interior elliptic regularity estimates and the assumed coefficient regularity, this also implies the claimed regularity result.
We discuss the attainment of the Dirichlet boundary conditions for the function : To this end, we observe that
Here we used the change of coordinates . Thus, passing to the limit dominated convergence yields the claimed result.
Next, we seek to show that the function also satisfies the higher order equation (2.2) and the system (2.3). To this end, we set for . We claim that these functions solve the system (2.3), where all boundary conditions hold in an sense.
In order to observe this, analogously as in Lemma A.3, we infer that the functions solve the bulk equation
whence
We claim that
(A.16) |
The latter in particular also shows the equivalent representation for . We obtain the first representation for by induction and the following computation
Together with the representation for this yields the first identity in (A.16). Arguing along the same lines as above (where the argument is detailed for ), this also immediately implies the claim on the Dirichlet data.
We next deduce the alternative characterisation of : We have
Here we used the change of coordinates .
It remains to discuss the Neumann data:
If , dominated convergence allows us to infer
as the powers are still integrable in zero.
For , we argue similarly as in [42]:
Here we used the change of coordinates . By dominated convergence we may again pass to the limit and obtain
This concludes the argument.
[1] |
Okokpujie IP, Bolu CA, Ohunakin OS, et al. (2019) A review of recent application of machining techniques, based on the phenomena of CNC machining operations. Procedia Manuf 35: 1054–1060. https://doi.org/10.1016/j.promfg.2019.06.056 doi: 10.1016/j.promfg.2019.06.056
![]() |
[2] |
Natsu W, He J, Iwanaga Y (2020) Experimental study on electrochemical machining with electrolyte confined by absorption material. Procedia CIRP 87: 263–267. https://doi.org/10.1016/j.procir.2020.02.106 doi: 10.1016/j.procir.2020.02.106
![]() |
[3] |
Farooq MU, Anwar S, Bhatti HA, et al. (2023) Electric discharge machining of Ti6Al4V ELI in biomedical industry: Parametric analysis of surface functionalization and tribological characterization. Materials 16: 4458. https://doi.org/10.3390/ma16124458 doi: 10.3390/ma16124458
![]() |
[4] |
Happonen A, Stepanov A, Piili H, et al. (2015) Innovation study for laser cutting of complex geometries with paper materials. Phys Procedia 78: 128–137. https://doi.org/10.1016/j.phpro.2015.11.025 doi: 10.1016/j.phpro.2015.11.025
![]() |
[5] |
Gilmore R (1991) Ultrasonic machining: A case study. J Mater Process Tech 28: 139–148. https://doi.org/10.1016/0924-0136(91)90213-X doi: 10.1016/0924-0136(91)90213-X
![]() |
[6] |
Samsudeensadham S, Mohan A, Krishnaraj V (2021) A research on machining parameters during dry machining of Ti-6Al-4V alloy. Mater Today: Proc 46: 9354–9360. https://doi.org/10.1016/j.matpr.2020.02.821 doi: 10.1016/j.matpr.2020.02.821
![]() |
[7] |
Prasad AVSR, Ramji K, Datta GL (2014) An experimental study of wire EDM on Ti-6Al-4V alloy. Procedia Mater Sci 5: 2567–2576. https://doi.org/10.1016/j.mspro.2014.07.517 doi: 10.1016/j.mspro.2014.07.517
![]() |
[8] |
Rathod R, Kamble D, Ambhore N (2022) Performance evaluation of electric discharge machining of titanium alloy—A review. J Eng Appl Sci 69: 64. https://doi.org/10.1186/s44147-022-00118-z doi: 10.1186/s44147-022-00118-z
![]() |
[9] |
Nagadeepan A, Jayaprakash G, Senthilkumar V (2023) Advanced optimization of surface characteristics and material removal rate for biocompatible Ti6Al4V using WEDM process with BBD and NSGA Ⅱ. Materials 16: 4915. https://doi.org/10.3390/ma16144915 doi: 10.3390/ma16144915
![]() |
[10] |
Dewangan S, Kumar SD, Jha SK, et al. (2020) Optimization of micro-EDM drilling parameters of Ti-6Al-4V alloy. Mater Today: Proc 33: 5481–5485. https://doi.org/10.1016/j.matpr.2020.03.307 doi: 10.1016/j.matpr.2020.03.307
![]() |
[11] |
Garba E, Abdul-Rani AM, Yunus NA, et al. (2023) A review of electrode manufacturing methods for electrical discharge machining: Current status and future perspectives for surface alloying. Machines 11: 906. https://doi.org/10.3390/machines11090906 doi: 10.3390/machines11090906
![]() |
[12] |
Banu A, Ali MY (2016) Electrical discharge machining (EDM): A review. Int J Eng Mater Manuf 1: 3–10. https://doi.org/10.26776/ijemm.01.01.2016.02 doi: 10.26776/ijemm.01.01.2016.02
![]() |
[13] |
Galati M, Antonioni P, Calignano F, et al. (2023) Experimental investigation on the cutting of additively manufactured Ti6Al4V with wire-EDM and the analytical modelling of cutting speed and surface roughness. J Manuf Mater Process 7: 69. https://doi.org/10.3390/jmmp7020069 doi: 10.3390/jmmp7020069
![]() |
[14] |
Kamenskikh AA, Muratov KR, Shlykov ES, et al. (2023) Recent trends and developments in the electrical discharge machining industry: A review. J Manuf Mater Process 7: 204. https://doi.org/10.3390/jmmp7060204 doi: 10.3390/jmmp7060204
![]() |
[15] |
Ram JS, Jeavudeen S, Mouda PA, et al. (2023) The role of various dielectrics used in EDM process and their environmental, health, and safety issues. Mater Today: Proc. https://doi.org/10.1016/j.matpr.2023.05.137 doi: 10.1016/j.matpr.2023.05.137
![]() |
[16] |
Nadda R, Nirala CK, Singh PK, et al. (2024) An overview of techniques for monitoring and compensating tool wear in micro-electrical discharge machining. Heliyon 10: e26784. https://doi.org/10.1016/j.heliyon.2024.e26784 doi: 10.1016/j.heliyon.2024.e26784
![]() |
[17] |
Nguyen VQ, Duong TH, Kim HC (2015) Precision micro EDM based on real-time monitoring and electrode wear compensation. Int J Adv Manuf Tech 79: 1829–1838. https://doi.org/10.1007/s00170-015-6964-y doi: 10.1007/s00170-015-6964-y
![]() |
[18] |
Di Campli R, Maradia U, Boccadoro M, et al. (2020) Real-time wire EDM tool simulation enabled by discharge location tracker. Procedia CIRP 95: 308–312. https://doi.org/10.1016/j.procir.2020.01.176 doi: 10.1016/j.procir.2020.01.176
![]() |
[19] |
Mhatre MS, Sapkal SU, Pawade RS (2014) Electro discharge machining characteristics of Ti-6Al-4V alloy: A grey relational optimization. Procedia Mater Sci 5: 2014–2022. https://doi.org/10.1016/j.mspro.2014.07.534 doi: 10.1016/j.mspro.2014.07.534
![]() |
[20] |
Gu L, Li L, Zhao W, et al. (2012) Electrical discharge machining of Ti6Al4V with a bundled electrode. Int J Mach Tool Manu 53: 100–106. https://doi.org/10.1016/j.ijmachtools.2011.10.002 doi: 10.1016/j.ijmachtools.2011.10.002
![]() |
[21] |
Świercz R, Oniszczuk-Świercz D, Chmielewski T (2019) Multi-response optimization of electrical discharge machining using the desirability function. Micromachines 10: 72. https://doi.org/10.3390/mi10010072 doi: 10.3390/mi10010072
![]() |
[22] |
Maniyar KG, Ingole DS (2018) Multi response optimization of EDM process parameters for aluminium hybrid composites by GRA. Mater Today: Proc 5: 19836–19843. https://doi.org/10.1016/j.matpr.2018.06.347 doi: 10.1016/j.matpr.2018.06.347
![]() |
[23] |
Agarwal N, Irshad M, Singh M R, et al. (2022) Optimization of material removal rate of Ti-6Al-4V using Rao-1 algorithm. Mater Today: Proc 62: 6722–6726. https://doi.org/10.1016/j.matpr.2022.04.760 doi: 10.1016/j.matpr.2022.04.760
![]() |
[24] |
Klocke F, Schwade M, Klink A, et al. (2013) Analysis of material removal rate and electrode wear in sinking EDM roughing strategies using different graphite grades. Procedia CIRP 6: 163–167. https://doi.org/10.1016/j.procir.2013.03.079 doi: 10.1016/j.procir.2013.03.079
![]() |
[25] |
Kumar R, Roy S, Gunjan P, et al. (2018) Analysis of MRR and surface roughness in machining Ti-6Al-4V ELI titanium alloy using EDM process. Procedia Manuf 20: 358–364. https://doi.org/10.1016/j.promfg.2018.02.052 doi: 10.1016/j.promfg.2018.02.052
![]() |
[26] |
Shao F, Liu Z, Wan Y, et al. (2010) Finite element simulation of machining of Ti-6Al-4V alloy with thermodynamical constitutive equation. Int J Adv Manuf Tech 49: 431–439. https://doi.org/10.1007/s00170-009-2423-y doi: 10.1007/s00170-009-2423-y
![]() |
[27] | Caballero B, Finglas P, Toldrá F (2016) Encyclopedia of Food and Health, New York: Academic Press, 321–326. |
[28] |
Iqbal A, Khan AA (2010) Modeling and analysis of MRR, EWR and surface roughness in EDM milling through response surface methodology. Am J Eng Appl Sci 3: 611–619. https://doi.org/10.3923/jeasci.2010.154.162 doi: 10.3923/jeasci.2010.154.162
![]() |
[29] |
Liu Y, Fearn T, Strlič M (2021) Factorial experimentation on photodegradation of historical paper by polychromatic visible radiation. Herit Sci 9: 130. https://doi.org/10.1186/s40494-021-00602-4 doi: 10.1186/s40494-021-00602-4
![]() |
[30] |
Sultan T, Kumar A, Gupta RD (2014) Material removal rate, electrode wear rate, and surface roughness evaluation in die sinking EDM with hollow tool through response surface methodology. Int J Manuf Eng 2014: 259129. https://doi.org/10.1155/2014/259129 doi: 10.1155/2014/259129
![]() |
[31] |
Seshaiah S, Sampathkumar D, Mariappan M, et al. (2022) Optimization on material removal rate and surface roughness of stainless steel 304 wire cut EDM by response surface methodology. Adv Mater Sci Eng 2022: 6022550. https://doi.org/10.1155/2022/6022550 doi: 10.1155/2022/6022550
![]() |
[32] |
Jin HT, Wang F, Zhang W, et al. (2023) Linear regression analysis of sleep quality in people with insomnia in Wuhan city during the COVID-19 pandemic. Int J Clin Pract 2023: 6746045. https://doi.org/10.1155/2023/6746045 doi: 10.1155/2023/6746045
![]() |
[33] | Sarstedt M, Mooi E (2014) A Concise Guide to Market Research, 2 Eds., Berlin: Springer, 193–233. |
[34] |
Harane PP, Wojciechowski S, Unune DR (2022) Investigating the effect of different tool electrodes in electric discharge drilling of Waspaloy on process responses. J Mater Res Technol 20: 2542–2557. https://doi.org/10.1016/j.jmrt.2022.08.015 doi: 10.1016/j.jmrt.2022.08.015
![]() |
[35] |
Devarajaiah D, Muthumari C (2018) Evaluation of power consumption and MRR in WEDM of Ti-6Al-4V alloy and its simultaneous optimization for sustainable production. J Braz Soc Mech Sci 40: 400. https://doi.org/10.1007/s40430-018-1318-y doi: 10.1007/s40430-018-1318-y
![]() |
[36] |
Tang L, Du YT (2014) Experimental study on green electrical discharge machining in tap water of Ti-6Al-4V and parameters optimization. Int J Adv Manuf Tech 70: 469–475. https://doi.org/10.1007/s00170-013-5274-5 doi: 10.1007/s00170-013-5274-5
![]() |
![]() |
![]() |
1. | Lauri Oksanen, Mikko Salo, Inverse problems in imaging and engineering science, 2020, 2, 2640-3501, 287, 10.3934/mine.2020014 | |
2. | Ru-Yu Lai, Yi-Hsuan Lin, Angkana Rüland, The Calderón Problem for a Space-Time Fractional Parabolic Equation, 2020, 52, 0036-1410, 2655, 10.1137/19M1270288 | |
3. | María Ángeles García-Ferrero, Angkana Rüland, On two methods for quantitative unique continuation results for some nonlocal operators, 2020, 45, 0360-5302, 1512, 10.1080/03605302.2020.1776323 | |
4. | Joonas Ilmavirta, Keijo Mönkkönen, Unique continuation of the normal operator of the x-ray transform and applications in geophysics, 2020, 36, 0266-5611, 045014, 10.1088/1361-6420/ab6e75 | |
5. | Giovanni Covi, Angkana Rüland, On some partial data Calderón type problems with mixed boundary conditions, 2021, 288, 00220396, 141, 10.1016/j.jde.2021.04.004 | |
6. | Giovanni Covi, Keijo Mönkkönen, Jesse Railo, Unique continuation property and Poincaré inequality for higher order fractional Laplacians with applications in inverse problems, 2021, 0, 1930-8345, 0, 10.3934/ipi.2021009 | |
7. | Angkana Rüland, On Single Measurement Stability for the Fractional Calderón Problem, 2021, 53, 0036-1410, 5094, 10.1137/20M1381964 | |
8. | Gabriele Cora, Roberta Musina, The s-polyharmonic extension problem and higher-order fractional Laplacians, 2022, 283, 00221236, 109555, 10.1016/j.jfa.2022.109555 | |
9. | Giovanni Covi, María Ángeles García-Ferrero, Angkana Rüland, On the Calderón problem for nonlocal Schrödinger equations with homogeneous, directionally antilocal principal symbols, 2022, 341, 00220396, 79, 10.1016/j.jde.2022.09.009 | |
10. | Pu-Zhao Kow, Yi-Hsuan Lin, Jenn-Nan Wang, The Calderón Problem for the Fractional Wave Equation: Uniqueness and Optimal Stability, 2022, 54, 0036-1410, 3379, 10.1137/21M1444941 | |
11. | Tiffany Frugé Jones, Evdokiya Georgieva Kostadinova, Joshua Lee Padgett, Qin Sheng, A series representation of the discrete fractional Laplace operator of arbitrary order, 2021, 504, 0022247X, 125323, 10.1016/j.jmaa.2021.125323 | |
12. | Alessandra De Luca, Veronica Felli, Stefano Vita, Strong unique continuation and local asymptotics at the boundary for fractional elliptic equations, 2022, 400, 00018708, 108279, 10.1016/j.aim.2022.108279 | |
13. | Joonas Ilmavirta, Keijo Mönkkönen, Partial Data Problems and Unique Continuation in Scalar and Vector Field Tomography, 2022, 28, 1069-5869, 10.1007/s00041-022-09907-9 | |
14. | Manas Kar, Jesse Railo, Philipp Zimmermann, The fractional -biharmonic systems: optimal Poincaré constants, unique continuation and inverse problems, 2023, 62, 0944-2669, 10.1007/s00526-023-02468-9 | |
15. | Pu-Zhao Kow, Jenn-Nan Wang, Landis-type conjecture for the half-Laplacian, 2023, 0002-9939, 10.1090/proc/16093 | |
16. | Aingeru Fernández-Bertolin, Luz Roncal, Angkana Rüland, On (global) unique continuation properties of the fractional discrete Laplacian, 2024, 286, 00221236, 110375, 10.1016/j.jfa.2024.110375 | |
17. | Roberta Musina, Alexander I. Nazarov, Fractional operators as traces of operator-valued curves, 2024, 287, 00221236, 110443, 10.1016/j.jfa.2024.110443 | |
18. | Qi Wang, Feiyao Ma, Weifeng Wo, Unique continuation for fractional p-elliptic equations, 2024, 15, 1662-9981, 10.1007/s11868-023-00568-w | |
19. | Ching-Lung Lin, Hongyu Liu, Catharine W. K. Lo, Uniqueness principle for fractional (non)-coercive anisotropic polyharmonic operators and applications to inverse problems, 2024, 0, 1930-8337, 0, 10.3934/ipi.2024054 | |
20. | Angkana Rüland, Revisiting the Anisotropic Fractional Calderón Problem, 2025, 2025, 1073-7928, 10.1093/imrn/rnaf036 | |
21. | Joonas Ilmavirta, Pu-Zhao Kow, Suman Kumar Sahoo, Unique Continuation for the Momentum Ray Transform, 2025, 31, 1069-5869, 10.1007/s00041-025-10149-8 | |
22. | Hongwei Li, Sheng Zhang, Bo Xu, Integrable Riesz Fractional-Order Generalized NLS Equation with Variable Coefficients: Inverse Scattering Transform and Analytical Solutions, 2025, 9, 2504-3110, 228, 10.3390/fractalfract9040228 |