1. Introduction
Forward flight represents a critical operating condition for helicopters blades, and it is well known that the most crucial aspect in helicopters rotors aerodynamics concerns the composition of velocities between the advancement motion of the system and the revolution of the rotor, which results in different conditions of the incident flow. The issue involves a differential distribution of the pressure load, which increases over the advancing blade, while simultaneously decreases for its retreating counterpart. As a consequence, an imbalance of the overall moment is induced [
1]. Therefore, the need to stabilise the attitude of the system entails the imposition of active control, accounting for the variation of the blades’ angle of attack. In these concerns, the most popular strategy consists in forcing the blade with a pitching motion, the latter described by a periodic law harmonised with the same period as the rotor revolution. In particular, the control system aims at increasing the angle of attack all through the advancing phase (
blade azimuth angle ) and at decreasing it throughout the retreating part (
) [
2]. However, the incidence variations can reach a certain angle of attack above the static stall condition, which may promote the aerodynamic instability of the blade; thus, a dynamic stall occurs by producing sudden excursions of the pressure load. In such conditions, forces and moments may reach values considerably higher than those loads typical of the steady case, and the possible coupling with the structure dynamics may lead to an earlier failure of the system. As a result, this event represents a propulsive limit in helicopters engineering and explains such attention paid by researchers over the years to reach a more in-depth knowledge of its occurrence [
3]. Overcoming this constraint is necessary and more and more required by the increased demand for higher performance in the helicopters field, especially in the military context [
4]. Although many control techniques were proposed to reduce the detrimental outcomes of dynamic stall, there is still a lack of proper awareness of the phenomenon which may drive the design process of oscillating wings, and the related subsystems [
5].
Concerning the experimental side, although mentioned in previous works, the reference measurements performed for a better comprehension of the dynamic stall date back to the late 1970s and, in particular, to the pioneering works of Martin et al. [
6], Carr et al. [
7] and McAlister et al. [
8] over a NACA 0012 aerofoil: The latter still being as milestones for the topic. Later on, a detailed physical description of the dynamic stall occurrence and attempts made to predict it was drawn by McCroskey [
9]. Experiments conducted by Leishman [
10] on a NACA 23012 aerofoil section showed that the evolution of a secondary vortex is typical of dynamic stall occurring at low Mach numbers. Wernert et al. [
11] analysed the unsteady dynamics of an oscillating NACA 0012 aerofoil with both Particle Image Velocimetry (PIV) and laser-sheet visualisation, together with computations with Baldwin-Lomax algebraic turbulence model and stressed the strong non-reproducibility of the detached conditions during the downstroke phase. In this path, more recently, the work of Gerontakos [
12] has become the most popular experimental reference for the academic community involved in computational modelling.
On the other hand, numerical modelling still struggles the inherent complexity related to the dynamics stall phenomenon, which is dominated by unsteadiness and flow characters ranging from laminar to turbulent. As a consequence, satisfactory numerical investigations became possible only with the gradual evolution of computing resources and computational techniques applied to Computational Fluid Dynamics (CFD). Even if Large-Eddy Simulations (LES) are increasingly feasible in the field of aerodynamics problems, also accounting for moving objects [
13,
14,
15], at present the solution of Unsteady Reynolds-Averaged Navier-Stokes (URANS) equations, together with a turbulence closure model, consists of the most widely spread approach. Standard closure techniques for the URANS environment involve adopting one or two additional equations, from which the corresponding names are referred to as 1- or 2-equation models. As regards the first class, the literature shows promising results by the adoption of the Spalart-Allmaras model [
16], for both compressible [
17,
18] and incompressible cases [
19,
20]. A variant of the model was proposed by Edwards and Chandra [
21] the so-called Spalart Allmaras with Edwards modification (SAE). The different formulation of the turbulence production term has proved that a proper tuning of the model may improve the prediction abilities of the standard approach [
22,
23], even in the field of retreating blade stall [
24]. A detailed 2D-based comparison study concerning the role of the seven different turbulence models can be found in Singh and Páscoa [
25] while similar analysis are provided by Wang et al. [
26]. The latter authors discussed the difference between standard
and
SST for a low-Reynolds two-dimensional case experiencing dynamic stall evolution. The work was expanded by Wang et al. [
27] who accounted for the third dimension and compared standard URANS computations with a Detached Eddy Simulation (DES) approach: in particular, the former was coupled with either RNG
and Transition SST (also known as
Re
), while the latter was solved together with
SST. In Kai et al. [
28] available experimental data are reproduced through a 2D model. Simulations are conducted with overset mesh technique, by solving URANS equations together with
SST closure model. A further simulation is performed with the addition of coupled plunging and pitching conditions to assess the effects on the dynamic stall of the phase difference between the two motions. In recent times, Marchetto and Benini [
29] numerically addressed the pitching motion on a 2D RAE 2822 aerofoil. Attention was paid to the impact of height and chordwise location of Gurney flaps on the loads during flow evolution.
In some way, recognising the organisation underlying a phenomenon is the first step toward a better comprehension of the phenomenon itself. This lies in an enhanced ability to reproduce it even through simplified models which are, however, able to describe its features as reliably as with the demanding CFD approach. It is in this spirit that, starting from the 1970s, the identification of coherent structures in turbulent flows has become a central issue for the research community [
30]. To this end, a significant contribution was given by Lumley [
31] with the introduction of Proper Orthogonal Decomposition (POD) in the context of fluid mechanics. The method extracts the coherent structures through a modal decomposition, without the need to account for conditional criteria. The minimisation of the projection error guarantees the energetic optimality in the sense of the
norm [
32]. Thus, from either numerical or experimental data, the optimal projection basis is obtained by solving a classical eigenvalue problem. The energy modal ranking allows for a selection of highly correlated modes which are isolated from the high-order contributions, usually characterised by a significant signal-to-noise-ratio [
33]. However, since the turbulence complexity is still an open issue, the identification of organised characters represents the first step toward a more-in-depth comprehension of its nature and possible ways to control turbulent phenomena [
34]. To this purpose, the great appeal of POD lies in its being a linear procedure that, nevertheless, puts no linearity conditions on the problem where it is applied to [
35,
36]. Different formulations of the theory were devised over the years to enhance its suitability for turbulent issues and, currently, the most popular POD-based technique is the so-called
snapshot-POD by Sirovich [
37]. More recently, Sieber et al. [
38] proposed a spectral variation of the procedure; thus, named Spectral Proper Orthogonal Decomposition (SPOD). The method is grounded on the application of finite impulse response filter along with the diagonals of the POD correlation matrix. In particular, the methodology aims to redistribute the energy among the modes; thus, increasing the signal-to-noise-ratio of the higher-order structures. As a result, it is possible to recover new coherent structures that were previously hidden as the noise of the low-order modes, resulting in a chance to extract novel dynamics and flow properties that were once unknown. The method was already successfully adopted by Sieber et al. [
39] in the modal analysis of a swirl-stabilised combustor flow resulting from measurements with Particle Image Velocimetry (PIV). In particular, the study showed the SPOD’s ability to extract the coherent structures related to the swirl flames accurately. Similarly, Lückoff et al. [
40] addressed the impact of seven different actuators on the flow of a swirl-stabilised combustor showing the SPOD capability to draw the differences between the flow dynamics connected with two of the geometries. In Ribeiro and Wolf [
33], the modal analysis is conducted on a flow past a NACA 0012 aerofoil at a fixed incidence angle, solved with the LES approach. The investigation focused on an extensive comparison between snapshot POD, spectral POD and Fourier-POD regarding the extraction of organised flow features connected with the noise generated on the computational domain. Ricciardi et al. [
41] compared POD and SPOD concerning the identification of coherent structures inside the cavity of landing gear and accounting for a 3D-model DES database. The study proved that spectral POD variation showed superior ability in recovering enhanced harmonic correlations of the structures connected with the tonal noise in the cavity. Finally, a first attempt at studying dynamic stall through SPOD can be found in Wen and Gross [
1]. Here, Implicit Large-Eddy Simulations were carried out over a Sikorsky SCC-A09 aerofoil undergoing three different configurations of motion, respectively: pitching & surging, pitching & surging & rotation, and pitching & surging & yawing.
The copious production related to dynamic stall agrees upon the fact that the phenomenon is not yet entirely ascertained. This supports the intention to address the problem through novel approaches that were not previously considered. The definition of a new procedure that provides novel insights into the issue is expected by the research community, in that this would represent a solid backbone for the design of lifting rotors. In this regard the SPOD results in a promising strategy. The present work aims at introducing this enhanced tool as one of crucial importance in the investigation of dynamic stall, providing new insights into the flow dynamics and extracting previously hidden structures that may lead to a better knowledge of its nature. In this path, this novel technique emerges from the previous approaches as a further step towards an improved and reliable magnification on the innermost organisation of the turbulence dynamics. According to the expertise gained by the authors, this investigation represents the first attempt to provide a detailed framework for a proper tuning of the spectral decomposition in the context of oscillating aerofoils. The SPOD is applied to the numerical database obtained from simulations of a two-dimensional NACA 0012 aerofoil undergoing pitching motion, by decomposing the velocity and the pressure fields. The flow configuration is that of a case study of interest for the research group, in that it is currently part of an enhanced design preliminary study. Thus, the main intention is to assess the effects of the filter on the dynamics of a flow in deep dynamic stall conditions when this is applied to both vector- and scalar-valued fields. The analysis is performed by inspection of time and spatial behaviour of the modes and by reconstruction of the database with a low-order approach.
The paper is organised as follows: In
Section 2, the main properties of the POD and SPOD are outlined.
Section 3 focuses on the geometrical model and the computational domain. The solver is introduced as well, and simulations setups are drawn.
Section 4 presents the results. Numerical solutions from simulations are first discussed, then the modal analysis is thoroughly addressed. Finally,
Section 5 provides the conclusive remarks and the future developments.
2. Spectral Proper Orthogonal Decomposition
Turbulent flows are characterised by a broad spectrum of time and spatial scales whose evolution has always drawn the attention of the research community [
34]. Although addressed as an organised phenomenon already from the late 1950s, the first mention of coherent structures underlying turbulence should be referred to the work from Brown and Roshko in 1971 [
30]. At that time, Lumley [
31] had already introduced the Proper Orthogonal Decomposition (POD) as an unbiased tool to extract organised characters from turbulent flows, which became, henceforth, widespread.
In the present work, the spectral variation of the POD proposed by Sieber et al. [
38] is employed. A detailed mathematical assessment of the method can be found in Sieber et al. [
42] while here the authors follow a brief description. The technique stems from the snapshot POD [
37], where a signal
is divided into its ensemble average
and a fluctuating component
. Thus, the signal is expanded as a spatial modes series
where
denotes the modal components and
the time coefficients. Here
denotes the number of snapshots, i.e., the collection of time instants for which the flow variable is defined.
The determination of the optimal projection basis is then addressed as a classical constrained optimisation problem. From the calculus of variations theory, it is possible to prove that the problem can be reformulated in terms of an eigenvalue problem for the correlation matrix. In particular, the correlation between two snapshots should be intended in the sense of the
inner product,
, as:
where
D denotes the variable domain of definition. As a result, in a discrete path the snapshot correlation yields to the
correlation matrix as:
Here
is the
snapshots matrix, i.e., the matrix which collects all the
fluctuations as
columns while
is a diagonal matrix containing the cells volume weighting, transferred at all the
nodes. Such theoretical arrangement accounts for the computations of the time-coefficients
which are subjected to the following eigenvalues problem
thus, being the eigenvectors of the
matrix. Here the eigenvalues,
, are also known as modal energies, representing the fraction of the initial signal which belong to the corresponding reconstructed mode. Spatial modes (or coherent structures) are recovered from projection of the original signal onto the basis of time coefficients as
The results obtained from the decomposition may then be adopted for the definition of a reduced-order model, as suggested by Equation
1. Thus, the energetic ranking based on the eigenvalues provides a natural criterion to select the modes containing the most correlated and energetic characters of the signal. Accordingly, the normalised cumulative sum of the energy,
, is used as a leading parameter in getting a quantitative datum concerning the number of modes
which should be retained for the signal reconstruction; the latter holds as:
The procedure described so far for the snapshot POD is revisited by Sieber et al. [
38] in a spectral variation and, in this sense, is called Spectral Proper Orthogonal Decomposition (SPOD). This novel technique is based on the application of a low-pass filter, along with the diagonals of the correlation matrix. The filter,
g, is of size
. The problem is then re-assessed for the modified correlation matrix,
, the latter defined as:
Thus, an eigenvalues problem is solved as in POD and, for the sake of clarity, the nomenclature is re-defined for the energies
, the time coefficients
, and spatial modes
. The role of the filter consists of increasing the modal signal-to-noise-ratio by redistributing the energy among the modes ensuring the preservation of whole dynamic content. As a result, for the first modes, the values of
are lower than the corresponding values of
. Furthermore, the method provides a tool to continuously shift from the snapshot POD (
) to the Discrete Fourier Transform (DFT) (
). Sieber et al. [
38] also propose a method to couple modes connected with periodic structures, which can be interpreted as a single complex quantity,
, defined as:
The technique adopts the Dynamic Mode Decomposition (DMD) of the time coefficients to identify and rank the pairs according to the harmonic correlation. In addition, the pair energy and corresponding weighted averaged frequency are computed. Both the techniques stem from the correlation matrix’s computation, as the weighted inner product of the snapshot matrix. Although the weighting matrix,
, is diagonal, the dense nature of
makes it impossible to take advantage of the convenient compact storage framework for the computation of Equation (
3). This may raise scepticism about the applicability of the analysis in three-dimensional problems, where the rows of the snapshot matrix may reach dimensions that are prohibitive for common computational resources. Anyhow, it is essential to recall that in general one is interested in recovering the most energetic structures. This allows for decreasing the dimension of the problem either by selecting a proper smaller region of interest [
41] or by taking advantage of peculiar symmetries. The latter reducing the problem to an actual 2D decomposition of the spanwise averaged quantities [
1]. When the nature of the investigation is such that a massive grid region must be considered, a solution could be to implement a code through the Message Passing Interface (MPI) [
33].
3. Computational Setup
The numerical analysis reproduces the experimental setup of Lee and Gerontakos [
43]. Thus, a NACA 0012 aerofoil is chosen with a chord length
m as sketched in
Figure 1. The measurements were carried out on a finite wing with a span equal to 2.5 times the chord and adopting endplates to reduce the tip effects. In particular, hotwire probes showed a two-dimensional non-uniformity of the flow over the wing bounded to
. Based on this evidence, in agreement with the assumption of Wang et al. [
26], the present work considers it reasonable to adopt a 2D model.
The computational domain is built as a double-block O-Grid with an overlapping region and the mesh is created as three-dimensional with a single cell along the span. The inner block is generated with quadrilateral elements around the aerofoil and extends up to 1.7 times the chord, starting from the quarter chord location. This point coincides with the centre of the whole mesh and defines the pole of the pitching motion. As usual in pitching analysis, the rotating motion develops about the
y-axis. A first layer of cells is wall-normally extruded, starting from a nominal initial distance of
. Here
is the fluid density at the wall location,
is the friction velocity,
is the distance of the first grid point and
is the fluid laminar wall viscosity. The friction speed
is estimated based on aerofoil chord and the maximum of the considered free stream velocities (see
Table 1). This choice is conservative for the
experimental setup configuration which involves less critical boundary conditions. Nodes distribution is thickened in the neighbourhood of leading and trailing edges. Then the distribution of the elements’ sizes toward the edge of the block followed a root mean square principle.
The outer block is mainly composed of triangular elements, except for a refined structured region behind the aerofoil that extends up to 6 times the chord of the aerofoil. The far-field boundary is located at a distance of
, and the mesh counts about
elements with a maximum equiangular skewness lower than 0.69. An equiangular skewness value lower then 0.37 is reserved in the structured region
Figure 2 reports the overall mesh (
Figure 2a) an the magnification of the leading and trailing edges of the aerofoil (
Figure 2b,c).
The set of the simulations aim at tracing the unsteady behaviour of the flow during pitching, focusing in particular, on deep dynamic stall conditions. In particular, the aerofoil pitch angle is varied through a monochromatic sinusoidal law which holds as:
Here
is the time-depending pitch angle,
its offset and
denote its amplitude.
k is the reduced frequency of the system and
its angular frequency. The reference Reynolds number
is based on the flow free stream density
and viscosity
and the free stream speed
. The simulation data initially follow the experimental assessment of Lee and Gerontakos [
43]. Then a
case study setup is adopted to run simulations from which data are extracted for the spectral analysis. As stated in the introduction, such workflow falls in the context of a preliminary investigation aiming to improve the definition of the design parameters for an applicative configuration. For this reason, the numerical assessment has been primarily tested on the well-documented low-Mach conditions [
43]. However, being the latter very far from real applications, a
test case configuration is further analysed. A complete summary of the simulated conditions is listed in
Table 1.
Concerning the boundaries, the latter are prescribed as follow: A far-field condition is chosen for the outflow edges, while symmetry is enforced at the ends of the aerofoil span. To solve the flow between the overlapping zones, the free edges of the two blocks corresponding to the overset region are set as
Chimera boundary condition. The
Chimera boundaries technique is efficiently implemented in TAU [
44,
45,
46,
47] and in this case allowed for building a customised refinement without the need to impose motion on the whole grid. According to this technique, also known as simply overset grid, multiple blocks are built with overlapping common zones. A hole is cut on blocks where non-penetrable surfaces are identified, then the information is transferred between different blocks by interpolating over the overlapping cells at the boundary of the hole [
48]. It is essential to consider that such grid structure favours a higher resolution in the wake region and provides greater accuracy in heading the detached flow dynamics, which was not achievable with a simple O-grid. In particular, the latter would have required a much denser grid overall and, consequently, higher demand for computational resources. Instead, splitting the domain into blocks allowed for adopting an unstructured grid for a more significant part of the far-field domain.
The computations are performed using the TAU-solver by DLR. The solver has been extensively validated by previous publications [
44,
49,
50,
51] both performing 2D and 3D Navier-Stokes analysis. In the present work, the Unsteady Reynolds-Averaged Navier-Stokes (URANS) system of equations is solved using explicit dual time stepping. The convective fluxes are discretised through a central difference scheme with scalar dissipation, while turbulent scalars employ an upwind scheme. Time-integration takes advantage of a backward Euler relaxation method combined with Lower-Upper Symmetric Gauss-Seidel (LU-SGS) linear solver.
Concerning the turbulence model, both the 1-equation Spalart-Allmaras model with Edwards modification (SAE) and the 2-equation Shear Stress Transport (SST) model by Menter (2003) are used. Unsteady simulations are initialised from a steady-state condition; the latter solved up to 3000 iterations. Four complete pitching periods are computed, thus showing flow stabilisation starting from the third. Pitching cycles are discretised with 3600 time-steps, each solved for a maximum of 300 inner iterations. Stop criterion for iterations is set at a condition where root mean square of the residuals is equal to , and monitoring the relative fluctuation on the density, re-normalised at the beginning of each time step. This strategy for monitoring the residuals is chosen to promote the convergence of the single time step, especially for the ones related to the most critical flow conditions at the highest incidences. In this regard, the maximum number of inner iterations guarantees convergence for those steps, though being oversized for the remainder. The choice of the time discretisation is consistent with sampling at least 100 flow events over a single pitching cycle. The chosen physical time size evolves almost 155 times in the convective time scale of the flow, based on the most critical boundary conditions, i.e., .
Figure 3 shows the non-dimensional vorticity field distribution corresponding the instant where
. The configuration preludes the Leading Edge Vortex (LEV) sheds into the wake; thus it slightly advances the dynamic stall condition.
5. Conclusions
The unsteady flow evolution of a pitching 2D NACA 0012 aerofoil is numerically investigated through Unsteady Reynolds-Averaged Navier-Stokes (URANS) equations. Computations are performed to tune either the Spalart-Allmaras turbulence model with Edwards modification (SAE) and
SST model against the experimental data of Lee and Gerontakos [
43]. Both the two closures show a general difficulty to reproduce the pitching phenomenon, anticipating the detachment of the Leading Edge Vortex (LEV); a result that agrees with previous references [
25,
26,
28]. However, since the
SST turbulence model shows more stable behaviour, especially regarding the downstroke phase, the model is selected for a
case study simulations setup.
The CFD results are post-processed thought advanced stochastic tools and, in particular, both the Proper Orthogonal Decomposition (POD) by Sirovich [
37] and the Spectral Proper Orthogonal Decomposition (SPOD) by Sieber et al. [
38] are adopted. The SPOD method differs from the POD in filtering the diagonals of the correlation matrix and, in particular, the study demonstrates how a proper choice of the filter size allows for the extraction of new dynamic contents developing throughout the flow evolution. The different behaviour of the two techniques is observed concerning the post-processing of vector- and scalar-valued fields. In particular, the analysis shows how the SPOD method acts better for the velocity reconstruction while the POD better performs for pressure decomposition. The computation of the aerodynamic loads from a reduced-order reconstruction of the pressure field, compared to original CFD, benchmarks the ability of both the POD and the SPOD to recover the inherent characters of the pitching cycle.
The application of SPOD on the corresponding database shows that the filter allows for the extraction of the eleventh harmonic dynamic. In particular, for the velocity field, a wider window size enhances the response and correlation of the coherent structures around that frequency. On the contrary, the effects on the dynamics of the pressure field depict marked variations depending on the filter size. In conclusion, the aerodynamics loads’ reconstruction demonstrates that low-order models based on SPOD better perform in recovering the incidence of the oscillations occurring during the pitching cycle. However, this comes at the expense of the energetic optimality of the decomposition and, as a result, the POD-based low-order model is more reliable in reconstructing the actual values of the loads. Anyhow, forthcoming investigations are meant to dissolve the uncertainty about the behaviour of the analysis when applied to flows evolving with different Reynolds numbers. This broader study will draw a complete insight into how the results provided by this novel approach may represent an essential aid for the design process of rotorcrafts, facing a wide range of operating envelopes.
Future investigations on 3D models will also include the evolution of lower scales of turbulence. Thus, the consistency of the method will be analysed for flows where more advanced CFD techniques allow for the identification of a broader spectrum of unsteady dynamics’ characters. As a further step, the intention is to take advantage of the modes for building a reduced-order model and, thus, to prove how this can help in predicting the performance in the preliminary phase of the design process.