1. Introduction
The adjoint equations were introduced for design optimization in the field of computational aerodynamics by Jameson [
1,
2] and have since been extended to a variety of applications such as optimal shape design [
3] of aircraft [
4], ships [
5], and automobiles [
6], error estimation and goal-oriented mesh adaptation [
7], and stability analysis [
8], among many others.
Of particular interest is the application of adjoint methods to the shape design of supersonic aircraft. Early applications include the aerodynamic optimization of a supersonic transport configuration [
9] and the CAD-based shape optimization of a reentry capsule in hypersonic flow [
10]. More recently, and in an attempt to address environmental concerns of commercial supersonic flight, a program was initiated to apply adjoint methods to the reduction in the sonic boom. The adjoint method was used in [
11] to investigate the influence of geometry modifications on the pressure distribution at remote locations, while [
12,
13,
14] apply adjoint methods to compute the sensitivities of functionals based on the equivalent area distribution.
In all of these applications, the adjoint method is used to compute the linear sensitivity of a cost or objective function with respect to a number of independent variables defining perturbations of the flow. These can be design variables in shape optimization applications or numerical errors or tolerances in mesh adaptation or uncertainty quantification applications. From a mathematical perspective, the adjoint variables appear as Lagrange multipliers enforcing the flow equations in a variational analysis of the cost function. The vanishing of the variation requires the multipliers to obey a partial differential equation (the adjoint equation) and appropriate boundary conditions. Aside from their mathematical meaning, the adjoint variables also carry physical significance as giving the influence of point sources (Green’s functions) of mass, momentum, and energy, respectively, on the objective function [
15].
An important step in the application of adjoint methods is the development and verification of adjoint codes. In recent times, a program has been initiated by several authors to produce analytic predictions for the adjoint equations that can be used for those purposes (but also to gain understanding of the behavior of the adjoint equations). The connection of the adjoint variables to Green’s functions has made it possible to generate exact adjoint solutions for quasi-1D inviscid flows [
16,
17] and two-dimensional (2D) incompressible inviscid flows [
18,
19], and outline the solution for 2D compressible subcritical inviscid flows [
20]. For inviscid 2D/3D flows, entropy variables offer another exact solution for quantifying the net entropy flux across boundaries [
21,
22]. A closely related solution, corresponding to the near-field computation of aerodynamic drag, was recently discovered by the authors [
23]. For viscous flows, exact adjoint solutions for the Navier–Stokes equations have been proposed for both laminar [
24] and turbulent [
25] boundary layers. Lastly, two recent papers [
26,
27] derived ordinary differential equations that must be obeyed by the adjoint solutions along characteristic lines, which can be used as a verification tool for numerical adjoint fields.
This paper considers the characteristic structure of the adjoint Euler equations with particular emphasis on supersonic flow. For simplicity, the analysis is restricted to two-dimensional flows. Steady inviscid compressible flow in two dimensions obeys the Euler equations as follows:
where
In Equation (2),
are the Cartesian components of the velocity,
is the density,
p is the pressure, and
E and
H are the total energy and enthalpy, respectively. For a perfect gas,
and
where
is the modulus of the velocity and
is the ratio of specific heats. The flow Equation (1) can be written in quasi-linear form as follows:
where
are the inviscid flux Jacobian matrices. Equation (5) is the starting point for the characteristic analysis of the Euler equations, as will be shown momentarily.
The derivation of the adjoint Euler equations is quite standard [
28] and will not be repeated here. For definiteness, the focus will be set on an external aerodynamics problem consisting of an airfoil with profile
S immersed in a flow with incidence
. For this case, it is desired to compute the sensitivities of a cost function measuring aerodynamic drag given as the integral of the pressure along a wall boundary
S
where
is a normalization constant,
is a reference length—the chord length of the airfoil,
is the wall unit normal vector, and
lies along the inflow direction. The adjoint equations are as follows:
where
are the adjoint variables. The adjoint boundary conditions are chosen to eliminate the integral of
on the boundaries. This procedure results in dual characteristic boundary conditions (obtained from a locally one-dimensional characteristic decomposition) in the far-field, while at the wall the adjoint variables obey
The characteristic structure of Equation (5) is well known and depends on the local Mach number
M [
29]. For subsonic flow, there is only one family of characteristic lines (the streamlines), while for supersonic flows there are two additional families of characteristics, the Mach lines that are inclined at an angle
to the local flow direction. Along characteristics, the partial differential equations (PDEs) that describe the flow reduce to ordinary differential equations (ODEs) called compatibility equations. In certain circumstances, notably in the case of two-dimensional irrotational supersonic flow, the compatibility conditions can be integrated and are further reduced to algebraic equations that hold only along the characteristic lines. The information carried by the compatibility conditions can be used to simplify the computation of supersonic flows, one notable practical application being the design of supersonic nozzles [
29].
The adjoint Equation (7) shares the same characteristic structure as the flow equations but with the sign of each characteristic velocity reversed so that the characteristic information travels in the opposite direction [
1,
15]. As in the previous case, the adjoint variables obey compatibility conditions along the characteristics (see e.g., [
30] and references therein).
The adjoint equations are linear and, for supersonic flows, hyperbolic. As a result of this, the behavior of the analytic supersonic adjoint solutions is strongly constrained by the characteristic structure of the equations. Moreover, and barring special cases where cost functions are defined along certain lines, not necessarily characteristic, such as in [
11], the adjoint variables can only have discontinuities along characteristic lines. The jumps in the adjoint variables along these lines are constrained by jump conditions, as will be shown in
Section 2. Additionally, since in supersonic flow information travels along characteristic lines, supersonic adjoint solutions tend to follow characteristic lines emanating from significant features of the flow or the geometry, such as leading and trailing edges [
26,
31], shock feet [
32], expansion fan centers [
33], nozzle lips [
27], etc. (see also
Section 3), and these trends are clearly visible in adjoint-adapted meshes [
34]. This behavior is significant, as it is precisely in those regions, where the amplitude of the adjoint solution is usually highest, that the cost function is most sensitive to perturbations.
In this paper, the characteristic structure of the adjoint Euler equations is reviewed, and the corresponding compatibility and jump conditions along and across characteristic lines are obtained. Most of what will be said is well-known, but the aim is to present the material in a comprehensive and unified way. It is also shown how this information can be used to obtain quantitative predictions for the adjoint variables in supersonic flow. In this way, the results presented in this paper go beyond [
26,
27], where the compatibility conditions were derived and checked on several high-quality, numerical adjoint solutions.
The paper is organized as follows.
Section 2 is devoted to the review of the derivation of the characteristic structure of the flow and adjoint Euler equations in 2D, including the eigenvalues, characteristic directions, and compatibility and jump conditions associated to the characteristic lines. These conditions are applied in
Section 3 to the analysis of supersonic flow past a diamond airfoil, for which a closed-form analytic adjoint solution, valid in the near wall region, is presented and checked against a numerical adjoint solution on a very fine grid. In
Section 4, we summarize our findings and make some concluding remarks.
3. Application to a Supersonic Case: Flow Past a Diamond Airfoil
The method of characteristics can be used to solve for the flowfield in the case of steady, supersonic flow and can be applied to the design of supersonic nozzles for 2D shock-free, isentropic flow [
29], since in that case, the non-linear flow equations reduce to algebraic equations along the characteristic lines. Since adjoint compatibility conditions cannot be integrated in general, there is little hope that a similar approach can be used with the adjoint equations. However, it turns out that the characteristic structure of the equations can be used to obtain analytic results in particular cases. One such example is given by the supersonic flow past a symmetric double wedge (diamond shaped) airfoil. The airfoil has a chord length of one and a thickness of 0.06 (resulting in a half-angle of
τ = 6.85 deg). The free-stream Mach number is
M∞ = 2, and the angle of attack (AOA) is 0 degrees. The flow contains two sets of wedge-shaped oblique shocks attached to the leading and trailing edges and two expansion fans emanating at the mid-chord vertices (
Figure 2).
The flow solution can be computed exactly using shock-expansion theory [
29]. The flow is parallel to the free stream upstream of the leading edge shock and downstream of the trailing edge “fishtail” shock. Between the leading edge shock and the leading Mach line of the expansion fan, the flow is parallel to the front segment of the airfoil, with a Mach number
M1 <
M∞ that depends on the free stream Mach number and the wedge angle
τ, which also determine the shock inclination. The flow then turns across the expansion fan (with limit Mach numbers
M1 and
M2) so that behind the trailing Mach line of the fan, the flow is parallel to the rear segment of the airfoil until it reaches the fishtail shock. The Mach number along the rear section
M2 >
M∞ is determined by
M1 and the turning angle 2
τ. Finally, the wedge angle,
M2, and
M∞ determine the fishtail shock inclination. From shock-expansion theory,
M1 ≈ 1.755 and
M2 ≈ 2.254, which are in very good agreement with the numerical solution shown in
Figure 3, computed with DLR’s Tau solver [
40] on an unstructured mesh with over 8.2 million nodes and 5300 nodes on the airfoil profile, with an outer freestream boundary domain of around 50 chord lengths from the geometry.
The corresponding drag-based adjoint solution computed with Tau’s discrete adjoint solver is shown in
Figure 4. The adjoint variables are non-dimensionalized relative to Tau’s reference values
.
Figure 4 depicts the numerical adjoint solution along with the shocks and several notable characteristic lines (various streamlines, which are the nearly horizontal lines, as well as several Mach lines—the diagonal lines— including the limits of the expansion fan and two Mach lines emanating from the trailing edge and running diagonally towards the upcoming flow). It can be seen that the adjoint solution clearly follows the characteristic structure of the flow in a pattern that somehow mirrors that of the primal flow. The adjoint solution vanishes downstream of the two Mach lines impinging on the trailing edge since no perturbation past those lines can affect the flow about the airfoil. The solution is essentially concentrated along strips limited by Mach lines emanating from the trailing edge, the mid-chord vertices and the leading edge, and the adjoint solution is discontinuous along these characteristic lines. There is also a weak horizontal strip along the incoming stagnation streamline upstream of the forward shock, in agreement with the general structure described in [
41]. Finally, the solution along the wall is piecewise constant (see
Figure 5), and it is actually possible to use the jump conditions (20) and the wall boundary condition to predict the values of the adjoint variables.
In this particular case, the procedure starts at the trailing edge. On the upper side of the airfoil, the right-moving Mach line emanating from the trailing edge separates two regions where the flow and the adjoint solutions are constant. Hence, the Riemann functions are piecewise constant, and the only one that can jump is the one associated to that characteristic line. The other 3 are continuous across the line and, being constant on either side, maintain the zero value that they have downstream of the Mach line. Hence, the adjoint variables upstream of the Mach line obey the following equations:
as well as the wall boundary condition Equation (8), which on the rearmost upper segment is simply
. It is now easy to see that the above equations yield the following result for the adjoint variables in the rearmost segment of the wall:
where
and
c2 and
H2 are the speed of sound and total enthalpy at the rear part of the airfoil, respectively. Equation (41) also holds on the lower side with the opposite sign for
. This analytic solution is compared with the numerical solution in
Figure 5, showing excellent agreement.
As can be seen in
Figure 5, the adjoint solution along the wall is piecewise constant, with the values corresponding to the rear segment of the airfoil given by Equation (41). Away from the wall, the adjoint solution exhibits a fairly simple structure, at least in the immediate proximity of the wall, as can be seen in
Figure 6, where the adjoint solution along a streamline is depicted and compared with the solution along the wall. The solution is again piecewise constant, and 5 plateaus (that will be labelled with roman numerals I–V) can be clearly identified, which are separated by jumps across characteristic lines. The adjoint variable is zero (V) beyond the right-running Mach line emanating from the trailing edge (7). Then it stays constant (IV) until the fan (III), delimited by Mach lines (3) and (5), where it has a continuous variation. It achieves a second plateau (II), which does not appear in the wall solution, and then jumps again across the right-running Mach line (2) emanating from the vertex of the fan [
30,
33]. Finally, it jumps again across the right-running Mach line emanating from the leading edge (1). Notice that the plateau values I and IV agree fairly well with the wall values, which means that the adjoint solution is roughly constant throughout the corresponding regions.
It is also interesting to note that the Riemann function
(34), associated with left-running characteristics, is constant (and in fact vanishes) throughout the upper side of the airfoil. In fact, it is only non-zero in the region of the pressure side of the airfoil limited by the left-running characteristics emanating from the leading and trailing edges—see
Figure 7. The reason for this behavior can be found in the interpretation of the adjoint Riemann functions as adjoint variables to the flow compatibility conditions.
is the adjoint variable associated with left-running characteristics, and it is only non-zero in the region of the fluid domain where perturbations carried by left-running characteristics can reach the airfoil and, thus have an impact on the cost function.
This information can be used to extend the analytic solution to the forward part of the airfoil. Giles and Pierce’s Green’s function approach [
17] allows us to write the adjoint state (in fact, any 2D inviscid adjoint solution) in the following generic form:
where
pt is the total pressure (which is constant between the leading and trailing shocks) and
Ij, at any given point with coordinates
, are the linearized cost functions corresponding to 4 linearly independent point source perturbations (mass, force normal to the flow direction, enthalpy, and total pressure) [
15] inserted at
. For cost functions that only depend on pressure,
I3 = 0 (total enthalpy perturbation at constant pressure), while
requires that
. Using this in Equation (42) and rearranging, yields
This is the general analytic solution. It depends on two functions
and
, whose values, for the numerical solution, are depicted in
Figure 8.
In the extreme zones
I and
IV,
I2 is constant, and its value is related to the adjoint wall boundary condition.
where
is the unit wall normal vector pointing towards the fluid, yielding
and
respectively. Within the expansion fan (zone
III), the flow is homentropic but variable, so along right-running characteristics Equation (31) applies. Using also
,
, and
, where
, yields
where
is a constant. From Equation (47), the (constant) value in zone
II can be written as follows:
Lastly, in zone
IV,
, which fixes
as follows:
Focusing now on
, it is constant in zones
I,
II, and
IV, and actually
since
across the right-running Mach line emanating from the trailing edge and
downstream. Further information can be obtained by using the jump conditions across the right-running characteristic emanating from the midchord vertex to relate the solutions in zones
I and
II. In terms of the Riemann functions, the jump conditions are as follows:
The first and third equations in (50) are directly obeyed by the solutions on each side by construction, which leaves only the middle condition
which yields
and, thus
, in agreement with
Figure 8. Across the expansion fan,
changes between
and
. To obtain the value of
across the fan, recall that
is the effect on drag of a point perturbation to the stagnation pressure at constant static pressure [
15]. Its value at a point
is given by the integration along the local streamline of secondary sources of mass and force normal to the local flow direction [
18,
20],
In Equation (52),
is the total density, and the integral is taken from
(
s = 0) to the downstream farfield along the local streamline through
. Let
be a point inside the expansion fan. First, the contribution to the integral vanishes downstream of the trailing Mach line of the fan. Second, the flow is isentropic, so the following relations hold along any streamline:
Furthermore, inside the fan the flow variables are constant along left-running Mach lines, so the following holds
(here
μ is the Mach angle). Gathering Equations (52)–(54) and using
yields
where
is given by Equation (47), and the variable of integration has been changed to
, which is the angle that the local left-running Mach line makes with the
x-axis, and
is the corresponding value for the final Mach line of the fan. The final piece of information required to evaluate Equation (55) comes from the analytic solution for the expansion fan [
42]
where
Differentiating Equation (56) with respect to
yields
Gathering all the information, the analytic adjoint solutions can be written as follows:
in zone
I, where
gives the inclination of the fan’s leading Mach line relative to the
x axis,
in zone
II,
across the fan and
in zone
IV, in agreement with Equation (41). The analytic solution Equations (59)–(62) is in remarkable agreement with the numerical solution, as shown in
Figure 5 and
Figure 9.
3.1. Solution Across the Expansion Fan
Figure 10 sketches the fan region on the upper side of the airfoil. The figure shows the limiting left-running Mach lines of the fan (marked as 3 and 5), as well as the right-running Mach line emanating from the trailing edge (marked as 7), a streamline (6) impinging on the intersection of Mach lines 5 and 7, and an additional Mach line (4) picked at random within the fan.
Equation (61) gives the spatially varying solution throughout the fan. Notice that, as is clear from Equation (61), the adjoint solution, like the flow solution, remains constant along each Mach line of the fan, as can be confirmed with the numerical solution shown in
Figure 11, which depicts
I2 and
I4 and the adjoint variables along Mach lines 3, 4, and 5. The corresponding analytic values for
I2 and
I4 computed with Equations (47) and (55) are also shown for comparison, showing excellent agreement.
It can be seen that towards the center of the fan (located at
x = 0.5 in the plot),
I2 and
I4 and the adjoint variables attain a constant value along each line of the fan, the value changing from line to line.
I2 and
I4 behave differently in this regard, since the former jumps abruptly across the right-running Mach line (7), while the latter shows a smooth variation and reaches a plateau at the intersection with the limiting streamline (6). The reason for this different behavior can be explained as follows. Both
I2 and
I4 can be written in terms of the adjoint Riemann functions as follows:
Both and vanish everywhere on the upper side of the airfoil, so it turns out that in the fan and . Now, within the fan and are only different from zero in the shaded triangle bounded by Mach lines 3, 5, and 7. While (and, thus, I4) is continuous across Mach line 7, (and, thus, I2) is not, and its jump depends on the local value of the flow, which explains the difference between the plateau values along each line of the fan.
3.2. Summary of the Procedure
The overall procedure can be summarized as follows. Focus on the upper side, since the solution on the lower side can be obtained by symmetry considerations. The adjoint solution vanishes downstream of the characteristic line B emanating from the trailing edge (zone V in
Figure 12) since no perturbation downstream of this line can affect the flow on the airfoil. The adjoint solution has four components that we choose to parameterize with four functions,
I1,
I2,
I3, and
I4, corresponding to the linearized drag due to four linearly independent point source terms injecting mass, transverse momentum, enthalpy, and total pressure. Since the flowfield is piecewise constant (except across the fan), the functions
Ii are themselves piecewise constant, at least in the regions closest to the airfoil wall outside the fan. We divide the near-wall flowfield in 5 zones (labeled I to V in
Figure 12). These zones are separated by the forward shock, two right-running Mach lines (labeled A and B in
Figure 12), and the expansion fan delimited by left-running Mach lines C and D. A limiting streamline (labelled S and depicted in red) is also relevant since it sets a boundary for the adjoint solution within the fan. The solution procedure then amounts to finding the values of the
Ii on each zone using adjoint Riemann invariants and jump conditions. For pressure-dependent cost functions,
I3 is identically zero everywhere. Besides, the adjoint left-running Riemann invariant
, which yields
, and
I2 is fixed by the adjoint wall boundary condition in zones I and IV. Finally, jump conditions across right-running waves entail [[
I4]] = 0, so
I4 is constant (and equal to zero) in zone IV and throughout zones I and II, where its non-zero value is fixed by the leading edge of the fan. The missing information is obtained by patching up the adjoint solution across the fan (zone III), where the solution is constant along each Mach line between the fan center and the limiting streamline S. The value of
I2 is determined using the compatibility conditions along right-running waves that cross the fan, while that for
I4 its definition as an integral along streamlines is used.
4. Conclusions
The behavior of the steady solutions to the 2D adjoint Euler equations is severely constrained by the characteristic structure of the equations, particularly in supersonic flows, where the equations are hyperbolic and the solutions display distinctive traits along certain significant characteristic lines.
In this study, the characteristic structure of the adjoint Euler equations in two dimensions has been examined. The eigenvalues and characteristic lines are the same as for the Euler equations, and compatibility conditions can be derived that constrain the evolution of the adjoint variables along the characteristics. At each point, these compatibility conditions are equivalent to the adjoint Euler equations. Additionally, adjoint solutions can be discontinuous along characteristic lines, and the jumps of the adjoint variables along these lines are constrained by jump conditions that can be used to build analytic adjoint solutions in simple cases, such as the supersonic flow past a diamond airfoil considered in this study.
The adjoint solution obtained in this paper has a relatively simple structure, at least in the vicinity of the airfoil profile. Due to the supersonic nature of the flow, the adjoint variables vanish downstream of two Mach lines emanating from the trailing edge. The near-wall solution shows several constant patches delimited by the aforementioned trailing-edge Mach lines, the expansion fan, a Mach line of the opposite family emanating from the center of the fan, and two Mach lines emanating from the leading edge. The solution is parameterized in terms of the linearized objective functions corresponding to point sources of mass, transverse momentum, enthalpy, and total pressure, whose behavior in the different zones and across their boundaries has been established using characteristic information. The obtained solution agrees extraordinarily well with a numerical solution obtained on a very fine mesh.
The interest of the results described in this study is twofold. On the one hand, they provide benchmark solutions and constrains that can be used for verification of numerical adjoint solvers. On the other hand, by helping improve adjoint-based design tools, these insights may significantly impact methodologies used in the aerodynamic design processes and related applications in supersonic regimes, particularly as commercial supersonic flight becomes more relevant.