-
PDF
- Split View
-
Views
-
Cite
Cite
A D G Hales, L J Ayton, Analytical insights into semi-infinite plate scattering: the wiener–hopf technique and two-sided linear boundary conditions, The Quarterly Journal of Mechanics and Applied Mathematics, Volume 77, Issue 1-2, February-May 2024, hbae003, https://doi.org/10.1093/qjmam/hbae003
- Share Icon Share
Summary
The Wiener–Hopf Technique is popular throughout applied mathematics, particularly for wave scattering problems. One such problem is the scattering of an incident wave impinging upon a semi-infinite surface. For scattering problems in which one applies separate boundary conditions on each side of the diffraction medium, we must adapt the standard Wiener–Hopf approach via a generalised technique to deal with the resulting matrix system. Such problems are present in the literature, where the Wiener–Hopf problem is reduced to solving Hilbert equations, leading to the so-called Wiener–Hopf–Hilbert method. However, both boundaries tend to have the same mathematical form, which is overly simplistic when describing realistic compliant boundaries or surface coatings. In some other literature, the matrix factorization problem remains the focus and is solved for arbitrary entries, which may prove challenging to use in complex scattering problems. We solve a more general problem while keeping an applied scattering framework, drawing attention to the additional analytical considerations needed to adapt previous methods. More specifically, we study the Wiener–Hopf matrix kernel, which shows that it holds the key to distinguishing these problems from those studied in previous literature. Finally, we present an example of a sound wave scattering off a plate that has a one-sided surface coating and is in flow. We treat this as a two-sided boundary and model the upper side with the Ingard–Myers impedance boundary condition and the lower side with the rigid Neumann boundary condition. Results are presented using impedance values from existing literature that reflect real materials.
1 Introduction
The Wiener–Hopf technique [1] is an analytical method that is commonplace in applied mathematics. Areas of its use vary from scattering problems in both aeroacoustics and electromagnetism to finance, hydrodynamics and beyond [2]. This article aims to generalise existing work on the Wiener–Hopf technique in the context of solving scattering problems with a view to applications in aeroacoustics. In particular, we will focus on constructing a clear framework to solve scattering problems wherein an incident wave diffracts off a semi-infinite flat plate. We assume this boundary has different material properties on its upper and lower sides that can be mathematically modelled with some generalised linear boundary conditions. By developing a general and precise mathematical method, we may be able to study new materials and surfaces that permit aerodynamic noise reduction. This noise reduction is of particular appeal to the scientific community due to the health hazards associated with aircraft noise, particularly for those living near airports. Moreover, noise is a drawback of implementing wind turbines that produce excessive low-frequency noise [3].
This article positions itself at the interface of two subtopics in applied mathematics: the theoretical matrix Wiener–Hopf factorisation problem and the direct application of the Wiener–Hopf technique to wave-scattering problems. Considering the former, considerable research has been conducted to understand what sort of matrices can be factorised as a standard matrix multiplicative product where one factor is analytic and non-zero in some upper region of the complex plane and the other is analytic and non-zero in the lower region. This procedure has been reviewed thoroughly and excellently in [2], where the most common approaches to this factorisation problem that arise in wave scattering problems are outlined. This approach usually revolves around considering a general class of matrices to which a specific factorisation technique may be applied, such as those of Daniele–Khrapkov type [4–7], Jones type [8–11] or Rawlins type [12?–15]. Regarding the second subtopic, problems involving a plane wave impinging upon a semi-infinite boundary are common in the literature due to their importance to aeroacoustics and beyond. The most common representation of a compliant boundary (a ‘non-rigid’ boundary in the no-slip Neumann sense) is using an impedance boundary condition, such as in [16–22]. In these examples, the boundary is taken as ‘two-sided’, meaning we prescribe distinct boundary conditions on each side of the plate, a matrix Wiener–Hopf kernel arises of a specific type, as described in [12, 13]. Reference [12] discusses the problem of multiplicatively factorising the particular class of matrices that arises in more general scattering problems; however, this study is purely theoretical and does not provide a constructive method to solve the diffraction problem.
Conversely, numerous papers [18–21] study two-sided boundary condition problems in which the same ‘type’ of boundary condition occurs on each side. The only mathematical difference between the two boundary conditions is the change in some parameter values, not the differential operator’s order or structure. Moreover, although theoretical work on generalised linear boundary conditions such as [23] exists, there are still caveats as to which boundary conditions this applies. Similarly, some conditions still exist for the entries to Rawlins generalised matrix in [12], such as the assumed non-zero nature of the functions. Amending all these omissions is an arduous yet worthwhile task; we aim to lay the groundwork for this and present a method that can tackle various problems in the future.
When dealing with two-sided conditions, the matrix Wiener–Hopf equation must be handled carefully. This is mainly due to the branch cuts that arise when solving the initial Helmholtz equation and their impact on the analyticity of factors in preselected regions of the complex plane. With this in mind, our analysis will act as a bridge between these two approaches. We still consider a general setup as in [12]; however, the underlying scattering problem directly motivates our approach. A specific example not covered by the previous literature is a semi-rigid plate; we define this as a plate in which one side is rigid, and the other has some coating modelled by an impedance boundary condition. It was shown in [24] that introducing velvety structures on the upper side of a plate can reduce aerodynamic noise, mimicking the surface of an owl’s wing. A theoretical model for this would be to solve a scattering problem in which we consider a boundary that is rigid below and has an impedance boundary condition above to represent the coating on top. However, it is unclear what condition is best. Impedance boundary conditions are prevalent in the literature and have many forms with varying intricacy [25–27].
A benefit of our generalised approach is that we can consider any linear differential operator within our boundary conditions. In our example, we apply the more commonly used Ingard–Myers boundary condition that accounts for the presence and influence of constant mean flow [28] on a traditional impedance surface instead of the standard Robin condition used in [19–21]. Although this boundary condition was addressed in [16], no results were presented, and only one condition was applied on the plate.
The article is structured as follows: Section 2 discusses the governing equations and how they may be used to construct a matrix Wiener–Hopf equation. We discuss the analytic requirements for boundary conditions due to their importance when modelling realistic behaviour and provide an overview of the tasks that will be needed to find the solution. This overview will include brief discussions on edge conditions, zeros in the scalar kernels and their influence on regions of analyticity and surface waves. Section 3 formally solves this problem by addressing these tasks and discusses how the theory developed in previous work can be extended to our general linear boundary conditions. It particularly focuses on solving the Hilbert equations that arise when using the Wiener–Hopf–Hilbert approach. Section 4 demonstrates this theory by solving an example related to acoustic diffraction; a sound wave scatters off a rigid edge with some theoretical upper coating, which we model with the Ingard–Myers impedance boundary condition [28]. We demonstrate the significant changes in the diffracted field when altering one side of a boundary while using realistic parameter values for impedance and free stream Mach number.
2 Constructing Wiener–Hopf equations for scattering problems in flow
We begin with a more abstract set of governing equations to form the type of Wiener–Hopf equations we seek to solve using a new factorisation process. The setup is primarily for scattering problems in which the scattering medium is a semi-infinite plane; however, the method can be altered for other setups thanks to the versatility of the Wiener–Hopf technique. We include constant background flow , which can be set to zero for examples without flow.
While constructing our governing equations for a more general wave problem, we must not only consider an incidental plane wave but also account for the possibility of a wake forming downstream from the plate. This is a complication of diffraction problems in flow and is modelled in [19, 29, 30] using a jump condition in x > 0 with some specific exponential forcing, we choose to implement this below with a scaling constant that can be set to zero if no wake is needed.
2.1 Governing equations and initial assumptions
Our problem will consist of an incident wave scattering off a two-sided semi-infinite boundary that occupies the negative x-axis, as seen in aerodynamic noise problems. We assume each wave has a (suppressed) term, thereby converting the linear 2D wave equation to a 2D Helmholtz equation. Our scattered variable will therefore satisfy the Helmholtz equation with some wavenumber k, which may or may not be the acoustic wavenumber (ω the angular frequency and c0 the speed of sound), depending on the context. Due to the influence of constant mean flow, a convective transform of the governing equations is required to shift this wavenumber from the Helmholtz number k0 to some convective wavenumber that we denote k for simplicity. We also include a source term on the positive x-axis in the form of an aerodynamic wake, as implemented in [18, 19]. To begin our analysis, we assume there exist constants such that our incident wave is of the form and our wake can be modelled by a source term of the form downstream from the plate. The purpose of this is to permit the method to apply to any incident waves with , where is the Mach number of the constant mean flow.
We may also include incidental surface waves in which could be arbitrary, possibly imaginary, constants. A final possible choice relevant to aeroacoustics is hydrodynamic incident waves- known as ‘gusts’. These waves represent turbulence and require real constants δ2 and .
We outline this setup in Fig. 1.

Illustration of the general scattering problem we will solve with the Wiener–Hopf technique.
We include the common assumption in wave scattering problems: a physical solution. This is usually accounted for by including the Sommerfeld radiation condition. Mathematically speaking, we avoid ab initio the possibility of surface waves that would blow up exponentially along the plate. If these functions did not decay, the common procedure is to give the wavenumber a small imaginary part ϵ, which is taken to zero at the end of the analysis. When this parameter is taken to zero, our problem transitions from a Wiener–Hopf to a Riemann–Hilbert problem [31].
2.2 Constructing the Wiener–Hopf equation
Notation and definitions
Before proceeding, we clarify the key notation and important terms for the rest of the article.
We rely on two defined regions of the complex plane DU and DL in which functions are analytic. These are the upper and lower regions, respectively, overlapping in some non-zero areas where our integration contour will lie. They are chosen under the assumption that
The additive factorisation of some scalar function or matrix of functions refers to an additive decomposition , where is analytic in DU and is analytic in DL. Additive factors will have lower case symbols to refer to which region they are analytic in.
The multiplicative factorisation of some scalar function or matrix of functions refers to a multiplicative decomposition , where f+ is analytic in DU and f– is analytic in DL. Multiplicative factors will have upper case symbols to refer to which region they are analytic in.
The following analysis is dependent upon branch cuts arising from the function . These are labelled and referring to which branch point of γ they emanate from. They extend to at an angle ϑ satisfying .
The discontinuity in γ across the branch cuts is important to exploit when solving our problem. When evaluating functions on the upper side of (for example) the branch cut, we use the notation . Similarly, when we evaluate a function on the lower side of this branch cut, we denote this .
We set the half Fourier transform of the jump in , including any possible junction condition that arises from this, equal to some unknown function l1 that we will solve for in our Wiener–Hopf equation. From (2.4), we can see that l1 will be analytic in DL and 1 will be analytic in DU.
It is worth noting that the unknowns can be written in terms of the half-range transforms (written, for example, as in [32]) and also terms from the series expansion of about 0, such as . In some instances, this could provide helpful information on our solution and its values at the origin or even help with ensuring the uniqueness of the solution. Indeed, it may relate closely to the aforementioned usage of junction conditions in specific examples.
Assuming we can multiplicatively factorise the matrix kernel, we may rearrange the equation so that the left-hand side is analytic in DU, and the right-hand side is analytic in DL. This permits the analytic continuation of the equation to the union of these regions- the entire complex plane. Hence, each side of the equation must be equal to some entire function that, via the generalised Liouville theorem, must be a polynomial of some order determined by the asymptotic behaviour of the Wiener–Hopf equation. This polynomial can be determined using junction conditions or other restrictions on regularity, such as the Kutta–Joukowski condition. Good examples in which extra conditions are required to resolve this entire function occur commonly for problems with higher order boundary conditions such as poroelastic plates [33–35].
A similar result can be shown for , but using or has no impact on the final result or the required analytical approach.
2.3 Additional analytical requirements
Although we calculated and in (2.13), several features of the Wiener–Hopf technique were omitted. We will review some neglected areas that may influence our solution and require extra care when modelling more complex real-world phenomena.
2.3.1 Choosing edge conditions
For clarity, we derive the large α scaling of each unknown term in (2.4), (2.7), (2.5) and (2.8) for a specific example at the end of this section.
The unknowns here are determined after inserting this general solution into the problem’s boundary conditions, from which a and b can be found and applied elsewhere.
Example: A ‘semi-rigid’ Ingard–Myers boundary condition
As a motivational example of this method, we define a ‘semi-rigid’ boundary condition as a boundary with one side considered rigid (having a Neumann boundary condition) while the other has some other condition. We choose the commonly used Ingard–Myers boundary condition [16, 38, 39] and revisit this in section 4.
2.3.2 Regions of analyticity
With our general Wiener–Hopf equation constructed, we formally define initial regions DU and DL for our chosen governing equations. First, we can determine the regions of analyticity of the additional known terms from their half-range integrals once we have chosen ϵ such that . Omitting the analysis, we find that is analytic in the region while are analytic in .
We assume all zeros lie off the real line. We may achieve this by introducing an imaginary part into the boundary conditions on the plate (in most aeroacoustics-related examples, this will follow from the ϵ imaginary part introduced into k). This step is for convenience; however, if computing factorisations numerically, it is possible to keep these zeros on the real line and deform the integration contour around them, as described in [6].
Figure 1 demonstrates the regions . As an arbitrary example, we assume there are four zeros in that are defined with the subscript U or L denoting whether they are a member of or .
2.3.3 Acknowledging surface waves
To finish this section, it is important to mention the possibility that surface waves may arise depending on the chosen boundary conditions. In particular, one source of such waves may lie in the zeros of the kernels S1 and S2 or otherwise. Referring back to (2.13), these zeros would become poles in A, B that must be accounted for during the integration process. Surface waves are discussed throughout the literature, and a thorough parametric study of how they can arise for the Ingard–Myers boundary condition is tackled in [22]. For most practical examples, absorbing boundaries will ensure these zeros do not lie in the cut α plane, as shown in [16] for the Ingard–Myers boundary condition and [20] for the convective Robin boundary condition. One exception to this rule is [18], in which surface waves are explored in detail when purely imaginary impedance values are examined. In this case, the surface waves are equivalent to zeros of the kernel lying on the real axis. Therefore, their residues must be accounted for when integrating over the real line.
Extending this to more complicated boundary conditions, or even more generalised conditions, is a non-trivial but worthwhile study. However, since this article primarily focuses on adapting the Wiener–Hopf technique to more general scattering problems, we omit an in-depth study for brevity. Still, in Appendix 5, we outline a way to find the zeros of a scalar kernel .
For many other applications, the far-field (large r) scattering is required. In this case, the steepest descent method is performed to ensure the integration occurs along the path of greatest decay. As we deform our integration contour from the real line to the path of steepest descent, we may well pick up poles that are equivalent to hydrodynamic or acoustic modes, as described in [22, 39, 41, 42].
3 Solving the Wiener–Hopf equation
Before rearranging for one must investigate the asymptotic behaviour of each side of the equation to deduce ; this is the aforementioned imposition of the Kutta condition and edge conditions at the tip. To do so, the first task would be to factorise the matrix kernel multiplicatively. Then, we require additive factorisations for and . More general source terms can be computed directly using the integral equations (1.17) from [32] or use the method of pole decomposition under the assumption that these functions are meromorphic. These tasks will be performed for our specific wave problem later in this section.
To solve the Wiener–Hopf equation (2.11), we carefully generalise the method presented in [21] to split our arbitrary kernel and then briefly discuss how we can calculate the other components of (2.7) to obtain the unknown vector .
The significant changes from the literature that we must consider are that are now of the form , where each are polynomials of α of any positive order that may or may not have zeros in the cut α-plane.
By adapting this process to two general linear conditions, we can investigate plates that are different above and below not only in impedance (via some impedance parameter) but are more fundamentally different in having two distinct boundary conditions where more than one parameter is altered.
3.1 Application of the Wiener–Hopf–Hilbert method
As in references [19–21], we perform a multiplicative splitting of our matrix by reducing the problem to solving a coupled pair of Hilbert equations, hence why this technique is sometimes referred to as the Wiener–Hopf–Hilbert method.
This equation provides an analytic continuation of from DU to In Fig. 2, we demonstrate our branch cut notation and the jumps in γ we expect to see on each side. This is crucial to the analysis that follows in this section.

Demonstration of branch cuts and the values takes on each side. We also include the notation for the evaluation of matrix above and below each branch cut.
These Hilbert problems can be solved using techniques in [12, 19, 21]; however, it is clear that to approach this more generally, care must be taken to avoid zeros and poles in our factors which can be handled on a case-by-case basis. Furthermore, we pay more attention to the imaginary part of k, causing the branch cuts to be at an angle and affecting the calculations needed for W. This is included since we will use a wavenumber with a small imaginary part in our example since this aids the numerical integration.
We can determine the value of n by ensuring (3.9) is satisfied and our matrices are non-singular at the branch point, just as in [21]. We postpone this until after we have solved (3.11b).
A key issue omitted in the literature is how the presence of zeros (and consequently poles) in the kernel can influence this process. For our more general setup, although we can choose to factorise a kernel with finite determinant as in [19], this matrix will have poles and zeros. Therefore, our splitting must reflect this but with poles and zeros placed in the correct domain, DL for and DU for . We must check that this will not affect our solution at the end of the procedure.
Another interesting consequence of (3.18b) is that if the branch point is a root of either or we will see that W can have poles or zeros at the branch point. This problem will likely arise if Neumann is the upper or lower boundary condition. However, we will show later that the structure of the solution will ensure this does not pose a problem to the analyticity of our factors. This additional jump will be reflected in from (3.9), so we can still choose the same n value for V to ensure the factors are non-singular.
3.2 Finding the particular solution for
.
and are non-singular in the entire cut plane.
is analytic in DU.
is analytic in DL.
We can choose and ensure our matrices are non-singular.
Regarding , our matrix consists of ‘plus functions’ and W. It can be observed that W is fully analytic in the complex plane, excluding the branch cut , so it is analytic in DU.
This demonstrates there is no discontinuity in across the lower branch cut.
For two-sided wave scattering problems in the literature, the boundaries are assumed to be absorbing (or supporting surface waves), which ensures either have either no zeros in the cut plane or manageable zeros on the real line. The presence of zeros can be analysed using other matrix factorisation methods, as described in section 1. Still, for this study, we are more interested in a specific subclass of matrices relating to our application. Therefore, we suggest an amendment to the factorisation that can account for potential zeros in DL arising from the terms in .
In [23], it is noted that zeros in the kernels can be accounted for with the introduction of eigensolutions to the governing equation. This effectively ensures that a zero in one of the kernels in DL (without loss of generality) can be transferred to and vice versa. This ensures the definitions of the analytic splitting can be adhered to while also satisfying the governing equations. Our goal is to account for zeros similarly within our matrix kernel.
We replace our solution of V in 3.11a by , which is analytic in DU and meromorphic in DL. Inserting this new solution into will not affect the analyticity in DU. In contrast, when we then calculate , a factor of will ensure there are no longer any poles or zeros in DL. We note that will also be singular at the zeros in DL, but this will not affect the formation or rearrangement of our Wiener–Hopf equation.
3.3 Solving for
Having factorised the matrix, all that remains is to find every component of (2.7). For this, we may complete two more straightforward tasks on a case-by-case basis. First, additive factorisations of the functions and can be found using the method of pole decomposition [32].
Second, we must deduce the entire function using asymptotic arguments. This can be done case-by-case using the correct edge conditions and the solution’s small r expansion.
4 Example: scattering a sound wave off a two-sided edge in flow
To demonstrate how we may use this general solution and methodology, we shall find the solution to a specific problem in which a sound wave scatters off the edge of a semi-infinite plate with an upper layer added made from some hypothetical material used to reduce noise. For simplicity, we model this plate by considering some two-sided boundary, one rigid and the other satisfies the Ingard–Myers impedance boundary condition for this pre-determined impedance [28].
To emphasise how our method generalises the results from existing literature such as [19–21], we change the fundamental mathematical structure of the boundary conditions above and below (not just the impedance parameter) and extend the method to account for a more mathematically intricate impedance boundary condition. The Ingard–Myers condition is known to be a better physical representation of the behaviour of a plate in flow; therefore, it is vital to create a more general method to handle more realistic conditions.
4.1 Governing equations

Picking DU and DL for a general system with and some positive real that satisfy and .
The complex α plane, including the poles, branch cuts, and regions of analyticity, is represented in Fig. 4, to be compared with Fig. 1. As shown in Fig. 4, DU is the region while DL is the region . As desired, the real line will lie in the region of overlap .

Geometrical setup for our diffraction problem. The two-dimensional model ignores the spanwise direction and focuses on z = 0.

α-plane representation of regions of analyticity for the wave scattering example.

Contour plots of for the rigid boundary problem alongside some semi-rigid examples inspired by real materials. For this example, we choose frequency f1, Mach number M1 and incident angle θ1

Contour plots of for the rigid boundary problem alongside some semi-rigid examples inspired by real materials. For this example, we choose frequency f1, Mach number M2 and incident angle θ2

Contour plots of for the rigid boundary problem alongside some semi-rigid examples inspired by real materials. For this example, we choose frequency f2, Mach number M1 and incident angle θ1
4.2 Solving the Wiener–Hopf equation
Following section 3, we must first split the matrix before performing additive splittings and deducing the form of the entire function .
To find W, we need to calculate the roots of and respectively.
Appendix 5 outlines the procedure for applying the edge conditions derived in section 2. Moreover, we also apply the Kutta condition at the tip of the plate to enforce finite velocity and pressure. With this, we deduce and also show .
4.3 Results
We finish this section with a demonstration of the effects of our boundary condition; we plot near-field results of the fully rigid plate alongside the semi-rigid plate with various parameter values for and . We keep , which ensures the boundary absorbs energy [19] and allows us to disregard potential zeros in the kernel, as shown in [16].
The first impedance values we choose to test are taken from [20], , resembling an absorbing fibrous sheet, , resembling a perforated steel sheet. We expect similar reductions in ψs due to these coatings, but the presence of the rigid surface at and the effects of are harder to predict.
We observe many interesting phenomena when comparing our scattered field of the semi-rigid sheets with the fully rigid boundary. First, we notice expected reductions in ψs both above and below the plate for 6b and 6c. More specifically, it appears that in 6b the reflected wave from the upper surface is nominal. Conversely in 7b the reflected wave below the plate seems to have been either cancelled due to destructive interference or otherwise, while the reflected wave above seems to be more significant. We believe this is due to the interact dependence of the parameter W on our model parameters. This phenomena may have interesting consequences when choosing a material with a specific impedance to reduce sound.
The best performing boundary is the perforated fibre with impedance , particularly in the second quadrant. In 6c, we see there are regions where the wave field is reduced, potentially from the destructive interference of reflected or diffracted components. This is most prevalent at the higher Mach number M2 for . A final note for these plots is that we find some impedance values break the symmetry about the negative x axis found in cases 6a and 7a.
At the higher frequency, f2, we notice similar features and trends regarding the symmetry about and the effect of each impedance on the reflected waves above and below. There seem to be larger reduction regions, particularly for 8b and 8c. It can be observed that the angles where destructive interference occurs remain the same.
The varied natures of impedance values demonstrate interesting and varied effects on the scattered field ψs. For every case, the upper surface of the plate has absorbed some energy and demonstrates a weakened reflected wave. The only significant increase tends to lie in the first quadrant at the wake. This can be attributed to the choice of the wake constant CK varying with our parameters (particularly ) due to implementing the Kutta condition. One final interesting comparison is that by changing the sign of the imaginary part of from positive in 6c to negative in 8c we notice there is less prominent reduction in the second quadrant but more generally less impact upon the reflected field below the plate. Figure 11 test at f1, M2 and θ2. For these parameters, only retains the symmetry in the third and fourth quadrants, unlike Fig. 10, suggesting a non-trivial dependence for this feature on . We see that has significant reflected waves above the plate, suggesting that at higher Mach numbers, a purely imaginary impedance may not perform as well as others, perhaps due to convective effects. A feature repeated from Fig. 10 is the varying strengths in the jump across the wake. A final note is that we have chosen constant values for each impedance Z. In reality, Z is actually a function, . This can be straightforwardly incorporated into the model, which may prove very useful for some applications of the effects of metamaterials in aeroacoustics and beyond.

Contour plots of for the rigid boundary problem alongside some semi-rigid examples inspired by real materials. For this example, we choose frequency f2, Mach number M2 and incident angle θ2

Contour plots of for the rigid boundary problem alongside some semi-rigid examples inspired by real materials. For this example, we choose frequency f1, Mach number M1 and incident angle θ1

Contour plots of for the rigid boundary problem alongside some semi-rigid examples inspired by real materials. For this example, we choose frequency f1, Mach number M2 and incident angle θ2
5 Conclusions
In this article, we presented a generalised framework for solving scattering problems with a semi-infinite boundary: our boundary is a semi-infinite plate on which two arbitrary linear boundary conditions are applied on the upper and lower sides of the plate, while we may consider the problem with any incident wave field by accounting for forcing terms in our boundary conditions on the negative x-axis. At the same time, we introduce a jump condition on the positive x-axis to represent the possible presence of a wake.
The primary goal was to extend the approach in papers such as [19–21] to solve similar scattering problems to apply them to popular industrial problems in aeroacoustics, such as leading- and trailing-edge noise. Although [12] provides an outline for solving the matrix factorisation that arises in similar problems, we have instead focused on the generalised scattering problem for which the matrix factorisation is an important component but not the sole issue that requires attention.
We reduced the matrix factorisation problem to Hilbert problems by considering jumps across square root branch cuts. The extension from a simple first-order boundary condition to a general linear boundary condition follows through manageable extensions to the functions V and W, which must be carefully constructed by paying attention to scalar kernel functions, effectively reducing the matrix problem to a scalar one. The two functions V and W that arise from the governing Hilbert equation are solved by investigating a second-order scalar kernel that can be dealt with numerically at little extra cost than a simpler impedance boundary condition (such as the standard convective Robin boundary condition). During this generalised process, attention was focused on important analytical features of the Wiener–Hopf–Hilbert method that must be carefully accounted for. These include the possible zeros of the scalar kernel that can lead to the presence of surface waves in the solution and singularities in the kernel, the importance of choosing the correct edge condition, the use of the generalised Liouville theorem, and the implementation of a Kutta condition.
Finally, an industrially focused example is presented and solved. A coated rigid plate is modelled with a Neumann boundary condition below and the Ingard–Myers boundary condition. Previous examples in the literature consider a first-order boundary condition, and the impedance parameter varies above and below the plate. The benefit of our method is that we can easily adapt the approach from a simpler mathematical boundary condition to a more complicated yet physically realistic boundary condition. Impedance values from [20] that represent realistic absorbing sheets of different materials were tested for this model. Results demonstrated that both the real and imaginary parts of the impedance can reduce the scattered near-field in both magnitude and direction.
Footnotes
Our boundary conditions have no source term for the jump in , however, the analysis would allow for a suitable term here that would give rise to an upper analytic function.
The result is independent of which x-axis we integrate along
It can be shown that if there is a discontinuity in across a branch cut then it will have to hold for too. Furthermore, by construction, is continuous in the region DL that includes the branch cut
This is assuming all y derivatives in the boundary condition are derivatives concerning the normal pointing out of the plate
Each of these impedance values lie within the range of sensible values for the impedance of these steel sheet in [20]
Acknowledgements
A. D. G. Hales would like to thank I. D. Abrahams for his guidance with the formulation of the article and the excellent feedback from anonymous reviewers. A. D. G. Hales acknowledges support from EPSRC studentship EP/T517847/1, L. J. Ayton acknowledges support from EPSRC Early Career Fellowship EP/P015980/1.
References
Appendix A: Finding the zeros of a scalar kernel
Appendix B: Additional analysis for the semi-rigid wave scattering example
I: Finding the entire function
We note that and as allowing us to ignore two terms from each side of the equation when finding .
For some .