Sharp dispersive estimates for the wave equation on the 5-dimensional lattice graph
Abstract.
Schultz [24] proved dispersive estimates for the wave equation on lattice graphs for which was extended to in [3]. By Newton polyhedra and the algorithm introduced by Karpushkin [15], we further extend the result to the sharp decay rate of the fundamental solution of the wave equation on is Moreover, we prove Strichartz estimates and give applications to nonlinear equations.
1. introduction
Discrete dispersive equations in the form of difference equations have attracted much attention in the literature of mathematics and physics, since they constitute a natural way to approach numerically real physical laws. Indeed, spatial discretization would be the first step to implement finite difference schemes, transfering an equation on a continuum domain to that on a lattice graph. As for discrete wave equations, they appear in physical applications such as lattice dynamics and can be used to describe the vibrations of atoms inside crystals. A fundamental model is the monotonic chains, see [6, 7, 19]. On general graphs, wave equations have been studied in [8, 12, 20].
In this paper, we consider dispersive and Strichartz estimates for the discrete wave equation
(1.1) |
Here the discrete Laplacian is defined by
where is the standard basis of the lattice .
By the discrete Fourier transform, the fundamental solution of (1.1) is given by
(1.2) |
where and , see Section 2.1.
The pioneering work on sharp dispersive estimates for was initiated in Schultz [24], where he proves that decays like and when and respectively. On , the authors [3, Theorem 1.1] proved a sharp upper bound of order (or , in short) as . In all cases , the following oscillatory integral plays an important role,
(1.3) |
Note that is the imaginary part of .
Based on the analysis in [23, 24], we know the main obstacle is to describe long-time asymptotic behaviour for when is small, cf. Section 3. In this case, has degenerate critical points and the method of stationary phase breaks down. When , in the terminology of [2], only stable singularities () and appear. As increases, however, the singularity type becomes complicated. Instead of classifying all kinds of singularities, we seek a suitable way to obtain the stability of and find its optimal decay rate, uniformly in . Here uniformity in for the decay is the key issue.
In the sense of V. I. Arnold [1], for an oscillatory integral
(1.4) |
establishing the uniform estimate is to determine whether the decay estimate of could be extended to for with sufficiently small norm. If , this is Van der Corput lemma, cf. [25, Chapter 8]. If , the answer is also affirmative by [16]. However, Arnold’s conjecture is not always true when , even in the case that is linear. See the counterexamples in [11, 26]. For more results we refer to [10, 13, 22], all these work are closely related to Newton polyhedra.
In general, when and has degenerate critical points, it is difficult to establish the sharp uniform estimate of . Even for , it is still complicated to determine the oscillation index (cf. (2.3) below) of . Nevertheless, Karpushkin proposed an algorithm in [15] to determine the uniform upper bound of when is quasi-homogeneous. This algorithm reduces to a family of polynomial phases and can be iterated several times to get the results, see Section 4.
For odd , we proved in the previous paper [3, Theorem 1.5] the upper bound for with a fixed . This velocity governs the decay rates when . Due to the lack of uniformity in , the dispersive estimate for remains open when . In this paper, we prove the sharp decay estimate using Newton polyhedra and the algorithm of Karpushkin for .
Theorem 1.1.
There exists independent of such that
Remark 1.2.
(a) A related model is the discrete Klein-Gordon equation
where is the mass parameter. This model has been studied in Cuenin and Ikromov [5] for (see also [4]). Theorem 1.1 gives a partial answer to the conjecture in [5]. Indeed, let be the corresponding fundamental solution, then has additional singularity at the origin compared with . When , by the techniques in [23, 24] and slight modifications of our proof, one can show that shares the same decay estimate with . This yields sharp dispersive estimates for the discrete Klein-Gordon equation on .
For all , the main ingredient of the proof of the theorem is the uniform estimate of (1.4) with phase of the type
(1.6) |
where is a nondegenerate quadratic form. This polynomial appears naturally in the study of (1.3). Indeed, if has a degenerate critical point with corank (i.e. rankHess), then can be expressed as near . In particular, corresponds to the crucial phase , see [3, Lemma 3.6]. Note also that has type singularity, and can be reduced to (in this case has type singularity, cf. [5]).
In our context , the new case is the most complicated one, which has singularity of class , cf. [2, p. 253]. Karpushkin’s algorithm will be applied to this phase, see Section 4. More precisely, we consider the deformation of with rank . Using change of variables, we reduce to homogenous polynomials with three variables, involving many parameters. Then we repeat this process and reduce the problem to oscillatory integrals in , where the phases are expressed in adapted coordinate systems (cf. Section 2.2). Finally, we use results in [16, 26] to obtain the desired bound.
Note that the key difficulty is to deal with highly degenerate oscillatory integrals, and our approach is different from that in [5] for the discrete Klein-Gordon equation. To the best of the authors’ knowledge, it is the first time to adopt Karpushkin’s algorithm for estimating such oscillatory integrals with phase of corank . Our results are indeed valid for any analytic perturbation. However, our iteration approach will be tedious as increases. Furthermore, as a germ at the origin, is not finitely determined (cf. [21, Chapter 5]) for even since is not its isolated critical point. For these reasons, the dispersive estimates when remain open.
By Theorem 1.1 and a well-known result in [18], we obtain the following Strichartz estimate. Also, a standard argument can be used for the global existence of the solutions to nonlinear equation (1.1) with small initial data, see Theorem 5.1.
Theorem 1.3.
Let and be the solution to (1.1). If indices satisfy
(1.7) |
then there exists such that
where denotes the conjugate index of for any .
The paper is organized as follows.
We recall basic facts about the discrete setting and Newton polyhedra in Section 2. We also state some estimates concerning the stability of oscillatory integral in this section. In Section 3, we give the proof of Theorem 1.1. In Section 4, we prove the key result, Proposition 2.8, which is crucial for the proof of Theorem 1.1. In Section 5, we prove Strichartz estimates and give applications to the nonlinear equations. In Appendix we show the sharpness of Theorem 1.1.
Notation. We use and to denote the length and the inner product on Euclidean spaces, respectively, and the transpose of matrix . Let (resp. ) be the usual open ball in (resp. ) with center and radius , while (resp. ) denotes its closure.
The symbols will be used throughout to denote implicit positive constants independent of , which may vary from one line to the next. For non-negative functions and , we adopt the notation if there exists such that . For any , the translation . Also, we use to denote , (or ) the usual gradient, and Hess (or Hessξ) the Hessian matrix. Moreover, we use (resp. ) to denote the image (resp. preimage) of under map .
2. Preliminaries
2.1. The discrete setting
We denote by the standard -dimensional integer latticc graph in , that is, . For , is the -space of functions on with respect to the counting measure, which is a Banach space endowed with the norm
We shall also use to denote the norm of for notational convenience. Note that spaces are nested, that is, for . Moreover, for functions on , the convolution product is given by
The discrete Fourier transform of function is given by
while the inverse transform is defined as
In the notion of operator theory,
(2.1) |
Without loss of generality, we assume that unless otherwise stated. Then we get , where the is as in (1.2). Moreover, let , the relation yields , which gives
2.2. Newton polyhedra
Let be a smooth real-valued function on and real-analytic at 0 such that
(2.2) |
Let be as in (1.4), where with support near the origin. Then the following asymptotic expansion holds (cf. e.g. [2, p. 181]),
(2.3) |
where runs through finitely many arithmetic progressions not depending on , which consists of negative rational numbers. Let be the maximum over all pairs in (2.3) under the lexicographic ordering such that for any neighborhood of the origin, there exists for which . We call the oscillation index of at 0 and its multiplicity.
The pioneer work of Varchenko [26] connects (2.3) with the geometry of Newton polyhedra, which we shall recall in the following part. We will use basic notions from [26], see also [2, 9, 14].
The associated Taylor series of at 0 can be written as
(2.4) |
Without loss of generality, we assume that . The Newton polyhedron of , denoted by , is the convex hull of the set
The Newton distance is defined as
The principal face is the face on of minimal dimension containing . In particular, under certain nondegeneracy condition, it is proved in [26] that the oscillation index of at is if .
Since depends on the choice of coordinate systems, the height of is given by
(2.5) |
where the supremum is taken over all local analytic coordinate systems which preserve the origin, and is the Newton distance in coordinates . A given coordinate system is said to be adapted to if .
If , the following results, derived by [26, Proposition 0.7, 0.8] and [9, Lemma 7.0], can recognize whether a given coordinate system is adapted. Note also that adapted coordinates may not exist when by the counterexample in [26].
Proposition 2.1.
Let , be as in (2.2) and one of the following conditions holds:
-
(a)
, i.e. is a single point.
-
(b)
is unbounded.
-
(c)
is a compact edge. Moreover,
and does not have a real root of multiplicity larger than , where .
Then the coordinate system is adapted.
For instance, one can verify both and are expressed in adapted coordinates with Newton distance , while for the Newton distance is , cf. [3].
2.3. Results on uniform estimates
As we mentioned in Section 1, it is natural to consider the stability of (1.4). We need some notation initiated from [15].
Definition 2.2.
For any , the space is defined as
Definition 2.3.
Let and be real-analytic at 0, we write
if for sufficiently small, there exist , and a neighbourhood of the origin such that
for all and , where is as in (1.4), and
Theorem 2.4.
In the sequel, for , we write if . Also, we write , if implies that Moreover, if , then we write . And if , then we write .
For a given weight
(2.6) |
the one-parameter dilation is defined as
Definition 2.5.
A polynomial is called -homogeneous of degree , if
Definition 2.6.
Let be the set of -homogeneous polynomials of degree , be the linear space (over ) of -homogeneous polynomials of degree less than , and be the set of functions real-analytic at 0 with the associated Taylor’s series having the form , i.e. each monomial is -homogeneous of degree greater than 1.
Lemma 2.7.
Let be real analytic at .
-
(a)
If , then for any ,
-
(b)
If and , then
-
(c)
If with all , then
Proposition 2.8.
Let and be as in (1.6), it holds that
3. Proof of Theorem 1.1
The reader is recommended to have [3] at hand, since the following proof is similar to that of [3, Theorem 1.1], and we shall use the results in that paper without repeating all of the proofs here.
The strategy is as follows. For fixed , we first study the long-time asymptotic behaviour for (1.3) with . Then we prove the same decay estimate holds uniformly under (analytic) perturbation , that is, when is replaced by in the integrand of (1.3). In our context, note that
(3.1) |
Therefore, the same estimate holds uniformly for as long as belongs to some small neighborhood of . Then it suffices to apply a finite covering since (1.1) has finite speed of propagation.
Now we begin the proof, a direct computation gives that on , hence the critical points of only appear when . Thanks to the results [24, Proposition 2.1, Proposition 2.2, Proposition 3.10], there exists such that provided and .
Thus we restrict the attention to small . Since is periodic, there exists , such that the integral in (1.3) can be rewritten as
cf. e.g. [3, Section 3]. Choosing with support near the origin gives
By [24, Proposition 2.3], we know as . As for , its asymptotic is determined by the critical points of the phase .
Let be an integer. Note that HessHess, we set
Moreover, let . By [3, Lemma 3.1, Corollary 3.2], we have (only the first quadrant is considered by symmetry):
Lemma 3.1.
Let , then
-
(a)
consists of with exactly () components equal to for .
-
(b)
.
-
(c)
there exists such that .
-
(d)
and
Due to the compactness and Definition 2.3, in order to obtain the uniform estimate for , it suffices to establish local bounds in and then use partition of unity, where is the support of . Indeed, we have the following lemma, whose proof relies on (3.1) and can be found in [3, p. 13].
Lemma 3.2.
For any , suppose that
(3.2) |
Then for some .
In fact, we only need to handle finite pairs by a partition of unity, and is their maximum in lexicographic order. The “worst” index appears exactly when . More precisely, to establish (3.2), it suffices to prove:
Theorem 3.3.
Let and , then
Once proving Theorem 3.3, we finish the proof of Theorem 1.1. In summary, we have used Lemmas 2.7, 3.1, 3.2 and Theorem 3.3 (and hence Proposition 2.8).
Proof of Theorem 3.3.
We consider each case separately and use Taylor’s formula for at .
Case 1: . By [3, Lemma 3.6], there exists invertible linear transformation which preserves the origin such that
holds for near , where (recall Definition 2.6) with . Therefore, taking gives
Case 2: . By symmetries and Lemma 3.1 (a), we can assume that with . We use a change of variables
Then a direct computation yields
where with . Note that
(3.3) |
By Lemma 2.7 (b)-(c) and Proposition 2.8 , we have
4. Proof of Proposition 2.8
4.1. Preparation
We first give a little notation to state the algorithm (Theorem 4.3) in Karpushkin [15] and some simplification (Lemma 4.4), making it convenient to verify the conditions in his result.
Let , weight be as in (2.6) and . We set
Notice that . Let
where is the differential of at . Moreover, let be the Jacobi ideal of (cf. e.g. [21, p. 51]). We have the following definition, which is from [15].
Definition 4.1.
A subspace is said to be lower -versal, if
Recall that is defined in (2.4) and is the translation. Using [15, Proposition 4 on p. 1182, Lemma 21 on p. 1184 ], we have
Lemma 4.2.
Let and the first order partial derivatives of be linearly independent. Then for any lower -versal subspace , and any critical point of , there exists monomial such that .
Note that Definition 2.3 carries over to real analytic manifolds. In the sequel, we write .
Theorem 4.3 (Theorem 1 in [15]).
Let and the following two conditions hold.
-
(a)
There exists a lower -versal subspace , such that for any and any critical point of , it holds that
-
(b)
, and
Then , where
If holds and , then
Here the maximum is taken in the lexicographic order.
By Lemma 4.2, we know condition (a) can be simplified if the first order partial derivatives of are linearly independent, see Section 4.2.
To simplify condition (b), we define the projection
as well as the set
Then we have the following relations.
Lemma 4.4.
If and , we have
(4.1) |
Moreover, if all are equal, then
(4.2) |
Proof.
We first prove (4.1). By the -homogeneity of , we have
(4.3) |
where is the Euler vector field for . Thus, it follows easily that .
To see the reverse, taking , we assume that , and use the chart , where
Then and gives
(4.4) |
Combining (4.4) and the second equality in (4.3) with , we get (note that )
Since , we have . Then , which yields by (4.4) again. Therefore, and (4.1) is proved.
Now we prove (4.2). By the proof of (4.1), we know . It suffices to consider since . Taking , we can choose and then , the standard sphere in . Moreover, we assume , then the standard chart is
Since , a direct computation gives
Notice that , . Taking derivative in gives
Therefore, we simplify Hess and get that
Since is non-singular, we finish the proof of (4.2). ∎
With all the tools in hands, we are in a position to prove Proposition 2.8.
4.2. Proof of Proposition 2.8
Take . By (3.3), a change of coordinates and Lemma 2.7 (c), it suffices to prove
Since are linearly independent, Lemma 4.2 gives that for any lower -versal subspace and , the rank of at any critical point is nonzero. If the rank of at critical point is at least 2, the splitting lemma and Lemma 2.7 (c) yield
If the rank of at critical point is 1, it holds that (since )
(4.5) |
A change of variables , gives
By Lemma 2.7 (c) again, it suffices to prove that
We consider two cases and . It suffices to prove
These two estimates can be derived by, for instance, a combination of Theorem 2.4 and Proposition 2.1 (alternatively, one can use Theorem 4.3 again). In conclusion, we verify condition (a) in Theorem 4.3 with .
Next, by Lemma 4.4 and the formula of Hess , we know
By Lemma 4.4 again, we know each critical point of is non-degenerate. Therefore, .
As a result, Theorem 4.3 gives the desired estimate .
4.3. Proof of Proposition 2.8
Now we deal with , which has type singularity, see e.g. [2, p. 252]. To begin with, we give two elementary lemmas.
Lemma 4.5.
Suppose that
Then if and only if or has a real root with multiplicity four. Otherwise, we have .
Proof.
Assume that . We discuss the multiplicity for the roots of in by a similar argument in [21, p. 85]. Under proper linear transforms can be reduced to one of the following forms:
(4.6) |
Lemma 4.6.
Let and ,
If there exists a constant such that and has a real root with multiplicity four, then there exists such that .
Proof.
Without loss of generality, we assume
If , then and is the root of and , the conclusion is obvious.
If , since
we know both and have root with multiplicity 2. Then the proof is finished. ∎
Let for . It suffices to show by Lemma 2.7 (c). We apply Theorem 4.3. Since
the coefficient of must be zero. So , which yields for all as well. Therefore, the partial derivatives of first order of are linearly independent.
Moreover, . Indeed, the stationary equation gives
which has only zero solution. Then we use Lemma 4.4 to get the conclusion.
We are left with condition of Theorem 4.3.
Let weight , . By Lemma 4.2, for any lower -versal subspace and , the rank of at its critical point is at least 1. Since , it suffices to consider the cases that the rank equals to 1 or 2, by Lemma 2.7 (c) and the splitting lemma.
(1) The rank 1 case. Note that , by symmetry we can assume that at the critical point is the 5-parameter deformation (which is the counterpart of (4.5)):
where , and .
It suffices to show . Let . We use a change of variables and for . A direct calculation shows
where , and
(4.7) |
Using a change of coordinates and Lemma 2.7 (b)-(c), it suffices to prove
(4.8) |
We first focus on the case w.r.t. , in which for some . By symmetry we may assume that . Then a direct computation yields that this happens when
(4.9) |
In these cases , which has type singularity, and
Consider the weight , by Lemma 2.7 (b), in this case to prove (4.8) it suffices to establish the following estimate.
Lemma 4.7.
Proof.
Using a change of variables , the polynomial can be written as a unimodal germ which has type singularity in the terminology of Arnold, see [2, p. 273]. The uniform estimate associated to this phase has been studied by Karpushkin with index , which is matched with the oscillation index, see [17] and the reference therein. ∎
Thus, we can assume that for all .
Lemma 4.8.
Let be as in (4.7). If for , then .
Once proving Lemma 4.8, using Lemma 2.7 (b) again with weight , we know , and we shall get that in the rank 1 case.
Proof of Lemma 4.8.
We shall apply Theorem 4.3 on . We begin with the linear independence for the partial derivatives. Since
Let and display the coefficient of each monomial, we get:
where and A discussion on whether (or ) equals to 0 gives that for .
Now we verify condition (a) of Theorem 4.3 for . By Lemmas 2.7 (c), it suffices to consider the rank 1 case. Similar to the previous proof, it suffices to prove
By a change of variables , we have
where with , and
Note that both and should to be taken into account since, as we shall see, in some cases vanishes and the contribution would come from .
After a change of variables with , fixed, and using Lemma 2.7 (c), it remains to show
(4.10) |
Our strategy is to discuss whether vanishes. If , then we show has desired estimate. If , by the properties of binary cubic forms (see e.g. [21, p. 85]) and the Van der Corput lemma, we also get the desired result.
Through an elementary (but tedious) computation on the coefficients of the monomials, we see that if , it suffices to consider the following two cases (other possible cases are slightly but not essentially different):
(1) , and
(2) and is the real root of the equation . In this case,
In both cases, cannot process a root with multiplicity 4. Then we get (4.10) by Lemma 4.5. In conclusion, we verify condition (a) of Theorem 4.3 with .
Now we turn to condition (b) of Theorem 4.3 for .
By Lemma 4.4, it suffices to show that for any possible critical point of , we have rankHess. We argue by contradiction that is a critical point such that rankHess, then the three principal minors in Hess are singular. Combining this with the fact , by a direct computation we get the following six equations in :
where . Since , we claim that under the assumption of Lemma 4.8, this system has no solution. In fact, we solve from the first three equations and put them into the left three equations. Then we get a system of ,
To solve these equations, we compute the resultant of the first and the -th equation for and enumerate the solutions. Due to the symmetries, we take as an example and eliminate (or ) from the two equations, which gives , where
It is easy to check that or gives either or is of type (4.9). The latter case leads to for some , contradicting to the assumption in Lemma 4.8. For left possible solutions, should be two distinct real roots of in (the case implies above trivial solutions). However, a calculation on the discriminant of this equation gives that never has all roots real.
Finally, in the rank 1 case we get in Theorem 4.3.
(2) The rank 2 case. Like the rank 1 case, by symmetry we may assume that at the critical point is the 9-parameter deformation,
where , , and (since the rank is 2).
In the sequel, the strategy is the same as the proof of in the proof of Lemma 4.8 (under (4.10)). We first introduce a little notation, let
As before, in we set , and for . By a direct computation parallel to that of , it suffices to prove the following counterpart of (4.10),
(4.11) |
where ,
and
If , then we finish by Van der Corput lemma. If , using Lemma 4.6 we shall show that do not possess a root with multiplicity four. Thus (4.11) is valid by Lemma 4.5.
We classify all cases that . A direct calculation shows that, except for the trivial solutions, it suffices to consider the following two cases:
5. space-time estimates and nonlinear equations
To begin with, for , the mixed space-time Lebesgue spaces are Banach spaces endowed with the norms
with natural modifications for the case or . By an abuse of notations, for any function defined on , we write sometimes for simplicity. Recall that has been defined in (2.1).
In this section, the proofs follow the same line as those of [3, Theorems 1.4, 5.9].
Proof of Theorem 1.3.
By Duhamel’s formula,
(5.1) |
Considering the space , we set the operators , where is the characteristic function. From Theorem 1.1, we know that provided . Then as a direct consequence of Keel and Tao [18, Theorem 1.2], we have
for indices satisfying (1.7) and . This together with (5.1) gives
Finally, we utilize the boundedness of the operator , see e.g. [3, Lemma 5.5], to finish the proof. ∎
Theorem 5.1.
In (1.1), let with and . If with sufficiently small, then for any , the global solution exists in with
Proof.
For the convenience of notations we write . We only prove the theorem for , while the case for general is similar.
For the contraction mapping principle to work, we need to estimate the norm of the solution, i.e. , in terms of time and the norm of initial data . By interpolation between Theorem 1.1 and a trivial inequality , we deduce
Therefore, by Young’s inequality it holds that
(5.2) |
where .
Now we consider the metric space
with to be determined later and the map on ,
Given that , we see that
where we have used (5.2) and , verified by . Taking the supremum, choosing proper and supposing sufficiently small in order, we know . One can also check that
Therefore, is a contraction as long as is sufficiently small. Finally, admits a fixed point in , which is the global solution to (1.1). ∎
6. Appendix
Let and be as in (1.6), note that is finitely determined (cf. [21, Chapter 5]). We prove that the oscillation index (cf. Section 2.2) of at is with multiplicity . Then one can verify (1.5) by the proof of [3, Lemma 3.6] and the following proof with slight modifications.
Recall that for , the Newton distance . By the additivity of the oscillation index, it suffices to prove the following assertion.
Proposition 6.1.
There exist with , and such that
Proof.
Using a change of variables
it suffices to consider for proper (see (6.2) below), where
Let be a constant, and
We write
It is easy to see that , while on , the method of stationary phase gives
(6.1) |
Inserting this to yields an upper bound . The same estimate is valid for . Now it suffices to consider . Similar to (6.1), we obtain that
(6.2) |
provided large enough, where , and for some supported in .
Denote by the integrand in (6.2), by symmetries it suffices to estimate
where is the triangle with vertices and . We begin with , a change of variables , gives
(6.3) |
where the remainder comes from the integral on the triangle with vertices and in -plane.
A change of variable gives
Using the method of stationary phase (cf. [25, Chapter 8 §5]) to the inner integral yields the asymptotic as .
Similar argument can be applied to . Indeed, a simple computation gives that
In coordinate systems where and , we have
Writing gives
Noting that has zeros , we split
Obviously, and . For the left two terms we use stationary phase method as in the estimate of (for we use a change of variable in addition), and get that both of them have asymptotics as .
As a consequence, we check that there exists such that , which completes the proof of Proposition 6.1. ∎
Acknowledgement
C.Bi is grateful to Prof. James Montaldi and Titus Piezas III for helpful discussions and suggestions. B.Hua is supported by NSFC, No. 12371056, and by Shanghai Science and Technology Program [Project No. 22JC1400100].
References
- [1] V.. Arnold “Remarks on the method of stationary phase and on the Coxeter numbers” In Uspehi Mat. Nauk 28.5(173), 1973, pp. 17–44
- [2] V.. Arnold, S.. Gusein-Zade and A.. Varchenko “Singularities of differentiable maps. Volume 1” Classification of critical points, caustics and wave fronts, Translated from the Russian by Ian Porteous based on a previous translation by Mark Reynolds, Reprint of the 1985 edition, Modern Birkhäuser Classics Birkhäuser/Springer, New York, 2012, pp. xii+382
- [3] Cheng Bi, Jiawei Cheng and Bobo Hua “The Wave Equation on Lattices and Oscillatory Integrals”, 2023 DOI: 10.48550/arXiv.2312.04130
- [4] Vita Borovyk and Michael Goldberg “The Klein-Gordon equation on and the quantum harmonic lattice” In J. Math. Pures Appl. (9) 107.6, 2017, pp. 667–696 DOI: 10.1016/j.matpur.2016.10.002
- [5] Jean-Claude Cuenin and Isroil A. Ikromov “Sharp time decay estimates for the discrete Klein-Gordon equation” In Nonlinearity 34.11, 2021, pp. 7938–7962 DOI: 10.1088/1361-6544/ac2b86
- [6] Martin T. Dove “Introduction to lattice dynamics”, Cambridge topics in mineral physics and chemistry ; 4, 1993
- [7] Richard P. Feynman and Albert R. Hibbs “Quantum mechanics and path integrals” Emended and with a preface by Daniel F. Styer Dover Publications, Inc., Mineola, NY, 2010, pp. xii+371
- [8] Joel Friedman and Jean-Pierre Tillich “Wave equations for graphs and the edge-based Laplacian” In Pacific J. Math. 216.2, 2004, pp. 229–266 DOI: 10.2140/pjm.2004.216.229
- [9] Michael Greenblatt “The asymptotic behavior of degenerate oscillatory integrals in two dimensions” In J. Funct. Anal. 257.6, 2009, pp. 1759–1798 DOI: 10.1016/j.jfa.2009.06.015
- [10] Michael Greenblatt “Stability of oscillatory integral asymptotics in two dimensions” In J. Geom. Anal. 24.1, 2014, pp. 417–444 DOI: 10.1007/s12220-012-9341-1
- [11] Philip T. Gressman “Uniform estimates for cubic oscillatory integrals” In Indiana Univ. Math. J. 57.7, 2008, pp. 3419–3442 DOI: 10.1512/iumj.2008.57.3403
- [12] Fengwen Han and Bobo Hua “Uniqueness class of the wave equation on graphs”, 2020 DOI: 10.48550/arXiv.2009.12793
- [13] Isroil A. Ikromov and Detlef Müller “On adapted coordinate systems” In Trans. Amer. Math. Soc. 363.6, 2011, pp. 2821–2848 DOI: 10.1090/S0002-9947-2011-04951-2
- [14] Isroil A. Ikromov and Detlef Müller “Uniform estimates for the Fourier transform of surface carried measures in and an application to Fourier restriction” In J. Fourier Anal. Appl. 17.6, 2011, pp. 1292–1332 DOI: 10.1007/s00041-011-9191-4
- [15] V.. Karpushkin “Uniform estimates of oscillating integrals with a parabolic or a hyperbolic phase” In Trudy Sem. Petrovsk., 1983, pp. 3–39
- [16] V.. Karpushkin “A theorem on uniform estimates for oscillatory integrals with a phase depending on two variables” In Trudy Sem. Petrovsk., 1984, pp. 150–169\bibrangessep238
- [17] V.. Karpushkin “Dependence on amplitude of uniform estimates for oscillatory integrals” In Uspekhi Mat. Nauk 44.5(269), 1989, pp. 163–164 DOI: 10.1070/RM1989v044n05ABEH002277
- [18] Markus Keel and Terence Tao “Endpoint Strichartz estimates” In Amer. J. Math. 120.5, 1998, pp. 955–980 URL: http://muse.jhu.edu/journals/american_journal_of_mathematics/v120/120.5keel.pdf
- [19] P.. Kevrekidis “Non-linear waves in lattices: past, present, future” In IMA J. Appl. Math. 76.3, 2011, pp. 389–423 DOI: 10.1093/imamat/hxr015
- [20] Yong Lin and Yuanyuan Xie “Application of Rothe’s method to a nonlinear wave equation on graphs” In Bull. Korean Math. Soc. 59.3, 2022, pp. 745–756 DOI: 10.4134/BKMS.b210445
- [21] James Montaldi “Singularities, bifurcations and catastrophes” Cambridge University Press, Cambridge, 2021, pp. xvii+430 DOI: 10.1017/9781316585085
- [22] D.. Phong, E.. Stein and J.. Sturm “On the growth and stability of real-analytic functions” In Amer. J. Math. 121.3, 1999, pp. 519–554 URL: http://muse.jhu.edu/journals/american_journal_of_mathematics/v121/121.3phong.pdf
- [23] Pete Schultz “Nonlinear wave equations in multidimensional lattices” Thesis (Ph.D.)–New York University ProQuest LLC, Ann Arbor, MI, 1996, pp. 160 URL: http://gateway.proquest.com/openurl?url_ver=Z39.88-2004&rft_val_fmt=info:ofi/fmt:kev:mtx:dissertation&res_dat=xri:pqdiss&rft_dat=xri:pqdiss:9702055
- [24] Pete Schultz “The wave equation on the lattice in two and three dimensions” In Comm. Pure Appl. Math. 51.6, 1998, pp. 663–695 DOI: 10.1002/(SICI)1097-0312(199806)51:6¡663::AID-CPA4¿3.0.CO;2-5
- [25] Elias M. Stein “Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals” With the assistance of Timothy S. Murphy, Monographs in Harmonic Analysis, III 43, Princeton Mathematical Series Princeton University Press, Princeton, NJ, 1993, pp. xiv+695
- [26] A.. Varčenko “Newton polyhedra and estimates of oscillatory integrals” In Funkcional. Anal. i Priložen. 10.3, 1976, pp. 13–38