1. Introduction
The interaction of turbulent shear flows with solid surfaces is clearly of great engineering interest [
1,
2,
3] and too many papers have been written on the subject to be listed here. So, we mention only the well-known paper by Tufts, Wang and Wang [
2] who used numerical methods to analyze the acoustic radiation produced by the interaction of an aerofoil with a turbulent shear layer. However, these types of interactions can also be studied analytically by using Rapid Distortion Theory (RDT).
RDT was developed to analyze relatively fast changes in turbulent flows such as those that occur in turbulence/solid surface interactions. It applies when the turbulence intensity is small and the interactions take place over length (or time) scales that are short compared to the decay time (or length) of the turbulent eddies [
4,
5,
6,
7,
8]. These requirements make it possible to identify a distance that is very large (in fact, infinitely large when interpreted asymptotically) on the interaction scale, but still small on the length scale over which the turbulent eddies decay. The resulting flow will then be inviscid and non-heat conducting and will therefore be governed by the Euler equations linearized about an arbitrary nonlinear solution (often referred to as the base flow) to those equations.
1.1. The Kovasznay Result
The basic ideas are best understood by considering a uniform base flow. This case was first analyzed in the seminal paper by Kovasznay [
9] which decomposed the unsteady isentropic motion on this flow into a vortical component that carries no pressure fluctuations and a solenoidal component that accounts for the pressure fluctuations. Möhring [
10] pointed out that the latter, which is determined by a second-order wave equation in compressible flows, accounts for the acoustic component of the motion of these flows [
10]. The former, which is a purely convected quantity in the sense that it moves downstream at the mean flow velocity, can then be interpreted as the hydrodynamic part of the motion. Two of its velocity components can be arbitrarily specified as (usually time-stationary) upstream boundary conditions for the unsteady motion. The Kovasznay decomposition has turned out to be very useful for analyzing turbulence/solid surface interactions on uniform mean flows [
11,
12,
13], or on flows that become uniform far upstream [
4,
5,
6], since the hydrodynamic component of the solution can be used to represent the incident turbulence in these analyses.
1.2. The Orr Result
The analysis becomes much more interesting when the entire base flow is allowed to be non-uniform. The simplest case is arguably a two-dimensional and incompressible flow with uniform mean shear so that the mean velocity, say
, is of the form
with constant
and
denoting Cartesian coordinates, with
in the mean flow direction. The unsteady motion is then determined by the linearized incompressible vorticity equation (the compressible case was analyzed by Möhring [
10])
with
denoting the time and
denoting the spanwise vorticity perturbation. It was first pointed out by Orr [
14,
15] that Equation (2) or, equivalently, the two-dimensional Rayleigh equation
which determines the unsteady cross-gradient velocity perturbation,
, can be integrated to show that the spanwise vorticity perturbation
can be an arbitrary function, say
, of the indicated arguments and that
is determined by
Orr [
14] calculated the velocity and pressure fluctuations evolving from an initial state by solving an initial value problem associated with this equation. However, these solutions (or their long time limits) are not all that relevant to the time-stationary turbulent flows being considered here since the corresponding solutions to the full nonlinear equations can develop internal shear layers that can no longer be treated inviscidly and can support Kelvin–Helmholtz instabilities [
16,
17,
18]. It is, however, not unreasonable to use the time-stationary or steady-state solutions given by Equation (4) to represent the turbulence in these flows. The relevant solution can then be written as
where
and
are two-dimensional Cartesian coordinates,
represents a large time interval and
denotes the two-dimensional Green’s function that is determined by the equation
together with appropriate boundary conditions. The vorticity
, which is equal to the convected quantity
, can now be specified as a boundary condition since Equation (5) will satisfy Equation (4) for any choice of this quantity.
The inner integral in Equation (5) is assumed to be carried out over an unbounded or semi-bounded region of space, with the Green’s function
required to satisfy appropriate transverse boundary conditions in the latter case and taken to be the free space Green’s function
in the former case. The transverse velocity perturbation
will then be given by [
19]
with
when the convected vorticity
is taken to be the generic time-harmonic function
Figure 1, which is a plot of some typical values of
calculated from Equations (7)–(9) with
taken to be
shows that this quantity differs from the corresponding Kovasznay result on a uniform mean flow in that it now decays as
.
This behavior will also occur in surface interaction problems, which might arise when a flat plate with a leading edge at
is placed in the flow (see
Figure 2). These considerations show that it is not possible to impose upstream boundary conditions by specifying
at the upstream infinity for this type of problem. However, it follows from Equation (4) that the Laplacian
is equal to the streamwise derivative of the convected quantity
and, therefore, does not decay. This means that the former quantity can be specified far upstream (infinitely far in an asymptotic sense) on the interaction length scale, which can still be small (infinitely small in an asymptotic sense) compared to the scale on which the turbulence evolves.
The key result is that the arbitrary function can be determined by specifying an actual physical variable in a region where the flow is unaffected by the solid surface. The present paper shows that the situation is analogous but much more interesting for more complicated transversely sheared flows.
1.3. Scope of the Paper
A large number of papers [
7,
8,
21,
22] have used locally homogeneous RDT, which is a kind of local high-frequency approximation first introduced by Moffatt [
23], to study the turbulent motion on planar shear flows. The assumption of local homogeneity eliminates the requirement for upstream boundary conditions, but the present paper is only concerned with non-homogenous RDT, which usually provides a much more realistic description of the flow. There are also a large number of papers on this subject that have appeared in the literature, but the focus of the present paper is on the much narrower topic of non-homogenous RDT on transversely sheared mean flows. Its primary purpose is to bring together and describe in an integrated fashion a general methodology that has been developed in a number of different papers [
5,
6,
20,
21] to deal with this phenomenon. Some significant new results are also presented.
The basic equations are rewritten in terms of a gauge function in
Section 3 and a formal solution to the complete inhomogeneous RDT problem is given in
Section 4.
Section 5 discusses the role of causality and instability waves and shows that (as in the Kovasznay decomposition) the resulting formulas for the unsteady pressure and velocity fluctuations (which decouple from the entropy fluctuations) involve two convected quantities. However, unlike the Kovasznay result, they are not directly related to the physical flow variables. However, it is shown that they can be linked to these variables by using two very general conservation laws (derived in
Appendix A and discussed in
Section 7) that relate the convected quantities to the physical variables and a gradient-wise particle displacement (defined in
Section 5).
Appendix B shows that the latter quantity vanishes at upstream infinity and, therefore (as shown in
Section 8), that conservation laws can be used to obtain a set of upstream boundary conditions that relate the convected quantities to the physical variables.
Spatially growing instability waves- which are usually associated with coherent structures in turbulent flows-may also appear in the solutions. However, their amplitudes cannot be determined by imposing upstream boundary conditions since they decay exponentially fast at upstream infinity. They can, however, be determined as part of the solution when causality is imposed on the flow, which appears to be particularly appropriate in the present context since Creighton [
24,
25] has shown that the imposition of causality is equivalent to imposing a Kutta condition at a trailing edge.
Section 6 shows that the entropy fluctuations can be determined from the particle displacement once the solution for the pressure and velocity fluctuations is known. The result brings in a third arbitrary convected function, which is equal to the entropy fluctuations at upstream infinity and can, therefore, be specified as a third upstream boundary condition.
Section 9 shows how these results can be used to relate turbulent pressure spectrum (which is of principal interest in acoustic and structural vibration problems) to the upstream turbulent velocity spectrum (usually specified as an upstream boundary condition in turbulent surface interaction problems) and an appropriate model for this spectrum is introduced. Some applications of the theory that have appeared in the literature are described in
Section 11 and some brief conclusions are given in
Section 12.
2. The Basic Equations
We consider the flow of an inviscid and non-heat conducting fluid which is assumed to be an ideal gas with squared sound speed , and entropy where and denote the pressure and density, respectively, and denotes the specific heat ratio where are the specific heats at constant pressure and volume, respectively.
The pressure
and mass flux
perturbations (where
is the actual velocity perturbation) on a transversely sheared mean flow whose velocity
is in a single direction, whose pressure
is equal to a constant, and whose mean sound speed squared
depends only on the transverse coordinate
, decouple from the entropy fluctuations and are governed by the linearized momentum and continuity equations
and
where
,
and
denote the convective derivative based on the source point (while
denotes the convective derivative based on the observation point).
The entropy fluctuation
depends on the momentum fluctuations and can be determined from the energy equation (Equation (11) of [
26])
once
is known.
3. The Gauge Function Representation
The momentum Equation (12) will be identically satisfied for any arbitrary function
and any purely convected function
when the pressure fluctuation
and the momentum fluctuation.
are related to
and
by [
27]
and
where
plays the role of a gauge function,
is a kind “pseudo-particle displacement”,
denotes the Kronecker delta and
denotes the Levi–Cevita permutation tensor.
Since the Gauge function
is undetermined at this stage of the analysis it can be adjusted to ensure that the continuity Equation (13) is also satisfied by substituting Equations (16)–(18) into Equation (13) to obtain
where
denotes the linear operator
As in the Orr analysis discussed in
Section 1.2, this result can be integrated to show that the gauge function
is determined by
where
denotes a second arbitrary convected quantity.
4. Green’s Function Solution of Gauge Function Equation
Eliminating the mass flux perturbation
between Equations (12) and (13) shows that the pressure fluctuation
satisfies the well-known Rayleigh’s equation [
28]
where
denotes the usual Rayleigh operator, which turns out to be adjoint to the operator
defined by Equation (20) since
for any functions
[
29].
Since, to our knowledge, all applications of non-homogeneous RDT have been to steady-state turbulent flows the focus here will be on the time-stationary solutions of Equation (21), we, therefore, suppose that
is time stationary [
30] and that initial conditions imposed in the distant past do not affect the solution at the finite time
.
Let
denote the Green’s function for the Rayleigh operator
that satisfies
and behaves like an incoming wave as
. As usual we let the first two arguments of
represent the dependent variables and let the second two represent the source variables. Since
denotes value of the solution to Equation (25) at
due to a point sink at
, it should be related to its adjoint
by the reciprocity relation [
29,
31]
Then since satisfies Equation (21) with the source function replaced by and, therefore, corresponds to a direct Green’s function in the present context, it reasonable to require that it satisfy the causality condition , which implies that the Green’s function, , will vanish for all finite as , since it vanishes for all by definition.
We can require that the solution of the gauge function defined by Equation (21) along with its derivatives (and, therefore, the pressure fluctuations) vanish (for all finite times) as goes to plus or minus infinity when the source function is reasonably compact—even for globally unstable base flows since the signal generated at cannot reach these locations when is finite. This means that gauge function goes to zero as for all finite . We also assume that the initial conditions are such that and its derivatives vanish for all finite as .
Setting
equal to
in Equation (24), letting
denote a solution to Equation (21) and using the divergence theorem shows that Equation (21) possess the formal steady-state solution [
27]
where
is a large but finite time, the volume
is assumed to be bounded by cylindrical surface(s)
that can be of finite, semi-infinite or infinite length in the streamwise direction,
denotes the unit outward-drawn normal to
and
is defined by Equation (18). Since
for
we can replace the upper limit
by the time
.
Equation (27) expresses the solution to Equation (21) in terms of the volume source distribution and the gauge function distribution over one or more cylindrical surfaces . The formulation is unconventional since the direct Green’s function now plays the role of an adjoint Green’s function in the solution of Equation (27) for the Gauge function . The surface integral will not appear in Equation (27) and the Green’s function will be completely determined by Equation (25) together with the causality requirements given above when the integration volume is all of space and will be incompletely determined by these requirements when it is not.
7. Conservation Laws
Like the Kovasznay decomposition discussed in
Section 1.1 the present result involves two arbitrary convected quantities which we denoted by
and
. Equations (A2), (A6), (A8), (A13) and (39) show that these quantities are related to the physical variables and the gradient-wise particle displacement defined in Equation (39) by the conservation laws
where
is defined by Equation (39),
denotes a scaled velocity gradient and
denotes the density-weighted vorticity based on the
independent density-weighted velocity fluctuation
(defined by Equation (37)).
Equation (63) relates the arbitrary convected quantities and to the pressure , the gradient-wise particle displacement given by Equation(39) and the cross-gradient (density-weighted) vorticity components , where is defied by Equation (A2). Equation (64) relates the arbitrary convected quantity to the gradient-wise vorticity component and the gradient-wise particle displacement given by Equation (39) while Equation (62) relates the arbitrary convected quantity to the entropy fluctuation and the gradient-wise particle displacement given by Equation (39) in the important case where the level surfaces of mean pressure and velocity coincide.
8. Upstream Boundary Conditions
The conservation law Equations (63) and (64) cannot (by themselves) be used to relate the unknown convected quantities and to the physical variables because they involve the cross-gradient particle displacement that is not actually a physical quantity, but we shall now show that (actually the hydrodynamic component of ) at upstream infinity and, therefore, that these equations can be used to obtain upstream boundary conditions that relate these quantities to those variables.
As in the classical Kovasznay decomposition discussed in
Section 1.1, we suppose that it is the hydrodynamic component of the motion that should be related to physically measured variables and not the remaining scattered component.
Appendix B shows that transverse components of the hydrodynamic portion of the pseudo-particle displacement behave like
when causality is imposed on these flows. The purely convected quantity
is the Fourier transform of the function
introduced in
Appendix B.
Inserting Equation (67) into Equation (37) shows that the hydrodynamic component
of the
-independent component density-weighted velocity
behaves like
where
is given by Equation (37) and
has the obvious meaning.
Inserting Equation (68) into Equation (17) and using the result in the momentum Equation (12) shows that
It, therefore, follows from Equations (67) and (68) that the conservation law given by Equation (63) becomes
where
the continuity Equation (13) shows that
It, therefore, follows Equations (68) and (69) that
where
Equations (45), (46), (68), (70) and (74) then imply that
which (upon neglecting unobserved tones due to the neutral instability wave) expresses
in terms of the transverse component of the hydrodynamic portion
of the physical velocity
at upstream infinity and, therefore, provides a suitable upstream boundary condition for determining the unknown convected quantity
. Equation (76) can also be written as
where
when there exists a function that forms an orthogonal coordinate system with the level surfaces
of the mean velocity
.
11. Application of the Theory
The general theory is applicable to a wide range of surface geometries and boundary conditions, such as lined surfaces which would be of interest in trailing edge noise reduction studies.
The acoustic radiation resulting from the interaction of a round jet with the trailing edge of a flat plate was measured by Olsen and Boldman [
43] who compared their results with the RDT solution given in [
5,
21]. Their results showed that the shape of the radiation pattern and its change with jet velocity was accurately predicted by the theory which was not the case for alternative theories [
12], that did not account for mean velocity gradients (e.g., [
12]).
An earlier version of the theory given in [
5] was used by Ayton and Peake [
44] to calculate the high-frequency acoustic radiation resulting from the interaction of a periodic upstream disturbance with an airfoil embedded in a transversely sheared mean flow. Their results also showed that the radiated sound field was significantly affected by the mean shear. A more highly developed version of the theory given in [
27] was used by Baker and Peake [
45] to analyze the effect of boundary layer shear on trailing edge noise.
The authors of [
20,
46,
47] used this latter version of the analysis to calculated he noise produced by the interaction of a large aspect ratio rectangular jet with the trailing edge of a flat plate (see
Figure 3). They obtained an analytical formula relating the acoustic spectrum to the turbulence correlation function within the jet and represented this quantity by Equation (A35). The authors of [
20] used experimental data to determine the parameters in this formula and compared the predicted sound field with experimental data taken over a broader range of polar angles and three different Mach numbers at Glenn research center [
36,
37,
38,
39,
40]. As shown in
Figure 4, the computed spectra turned out to be in good agreement with the experimental data. The predicted sound pressure levels would change by as much as 10 Db at the highest Mach number considered if the mean shear and the resulting proper treatment of the convection velocity were not accounted for.
The low-frequency Green’s theory was used to predict the sound field produced by the interaction of circular jet with a trailing edge in [
42] (see
Figure 5). Their formulas turn out to be quite general and are expected to apply to many different flow configurations, such as the multiple jet configuration shown in
Figure 6. They carried out computations over a range of azimuthal angles and three different Mach numbers and again obtained good agreement with experiment. Some typical comparisons are presented in
Figure 7.
12. Concluding Remarks
This paper was written to bring together and present in a consistent fashion a general theory for the unsteady motion on a transversely sheared mean flow that has been developed over the years in a number of papers published in the Journal of Fluid Mechanics. The relevant equations are reformulated in terms of a gauge function in order to obtain expressions of the unsteady velocity and pressure fluctuations (which decouple from the entropy fluctuations) that involve two arbitrarily convected quantities. A pair of very general conservation laws are then used to derive upstream boundary conditions that relate two of these quantities to experimentally measurable flow variables.
Inflectional base flows are able to support spatially growing instability waves, which are often associated with the coherent structures in turbulent flows. Their amplitudes cannot be determined by imposing upstream boundary conditions, but can be uniquely determined as part of the solution when causality is imposed on the flow-which appears to be particularly appropriate in the present context since Creighton [
24,
25] has shown that the imposition of causality is equivalent to imposing a Kutta condition at a trailing edge.
The entropy fluctuations can be determined after the fact once the velocity and pressure fluctuations are known by specifying a third arbitrary quantity. The results, which can be used to analyze the unsteady motion resulting from the interaction of turbulent shear flows with solid surfaces, are applicable to a wide range of flow–surface interaction problems.