1. Introduction
Junction flows (JFs) are common, such as those on bridges, buildings, wings, submarines, turbines, compressors, and the radiators of electronic equipment. Such corner flow is due to the complex three-dimensional (3D) separated flow caused by the upstream boundary layer (BL) colliding with an obstacle, usually one that is protruding from an attached plane. Occurring with both blunt and streamlined obstacles, this phenomenon is due to the sudden change in pressure gradient caused by the obstacle and the 3D effects caused by the separation of horseshoe vortices [
1]. Except at very low Reynolds numbers, and both laminar and turbulent BLs, JFs are triggered easily in a very wide range of Reynolds numbers [
1].
For aircraft, JFs generally occur at the junctions of (i) wings and body, (ii) nacelles and wings, and (iii) horizontal and vertical tail parts, among other components. Once a JFs exists, the flow in the corner region oscillates and flaps the connecting components, which may threaten aircraft safety. JFs also produce additional interference drag that affects aerodynamic performance and even aircraft stability in the longitudinal and transverse directions [
1]. Therefore, the flow mechanism for a separated flow in a junction area must be well understood, with this being very important for designing junctions such as wing–body, nacelle–wing, and horizontal–vertical tail junctions. For laminar-flow aircraft in particular, the airfoils used for the wings and horizontal and vertical tail parts have a small leading-edge radius, and the position of maximum thickness is relatively far back, thereby making flow separation in junction areas more likely [
2].
The flow interference in a wing–body junction is actually caused by the fuselage and wing BLs intersecting at the junction, and because of the mutual fusion and interference of the BLs on both sides, the flow at the junction is highly anisotropic and very unsteady [
1,
2]. In addition, this physical phenomenon changes greatly with the flight state, wing leading-edge radius, and other factors, so the flow caused by the mutual BL interference is very complex [
3]. Therefore, a better understanding of the flow mechanism for this mutual BL interference would greatly help in aircraft design and improve aircraft aerodynamic performance.
To date, the research on flows in junction areas has mainly been focused on the spatial separation point, bimodal instability, horseshoe-vortex oscillation, and other flow characteristics near the leading edge [
2], and early JF research was mostly experimental in nature. Devenport and Simpson [
4] conducted wind-tunnel experiments on a NACA0020 wing with a 3:2 elliptical leading edge, and other experiments showed that the generation and shedding of horseshoe vortices at the leading edge of an aircraft wing at the bottom of the corner area led to low-frequency oscillation in the leading-edge area of the wing [
5]. By measuring the velocity in the leading-edge region and the wake, Fleming et al. [
6] studied the span-wise flow velocity pattern caused by the transverse reverse pressure gradient. To study the anisotropy of near-wall turbulence, various researchers measured the strong cross flow at the stagnation point [
7,
8,
9,
10,
11].
In comparison, there has been relatively little research on corner separated flows at the trailing edges of junction regions. Gand et al. [
12] carried out a series of experiments and found that with an increase of Reynold number, the corner separation increases as well. Thus, this phenomenon involved in the flow at the junction between a wing and an aircraft body or a nacelle pylon can very commonly trigger a separation flow and thus impair the aerodynamics of the aircraft through so-called interference drag, which is estimated to constitute 10% of the total drag of a civil aircraft [
2]. Baber [
13] carried out a series of experiments to study the impact of an incoming BL on the corner separation, and the research showed that a thinner incoming boundary may cause a larger corner separation at the tailing edge. Jing et al. [
14] experimented to invest the static directional stability (SDS) contribution of the vertical tail of a conventional aircraft, and they found a SDS loss at the vertical tail and body junction region because of corner separation. From the above research, it appears that junction corner separation might cause directional static instability of an aircraft when it occurs in the region of a T-tail. It is therefore important to understand the mechanisms involved in junction flow, especially for industrial configurations, where corner separation is among the most critical matters [
15]. However, an incoming BL is affected by the leading edge of the wing and merges with the disturbed BL of the wing itself in the corner region. This simultaneous presence of two BLs and their merger in the junction region makes the flow in this region very complex and highly anisotropic. To date, corner separation flow represents one of the few remaining challenging issues in applied aerodynamics, and the design of belly fairings mostly relies on the designer’s experience [
15].
The corner separated flow generated by the mutual interference of two BLs is the second kind of Prandtl secondary flow [
1], and it is strongly anisotropy. For general Reynolds-averaged Navier–Stokes (RANS) methods, especially the Spalart–Allmaras (SA) and shear-stress transport (SST) methods and other commonly used methods based on linear eddy viscosity, although the flow solution has good accuracy outside the interference region, it is almost impossible to obtain ideal results in the influence range of the horseshoe–vortex system, especially in the separation region.
Due to the lack of nonlinear terms, the obtained separation region differs greatly from that obtained experimentally; this is the case even for the SST model, which is relatively good at capturing other flow details. Indeed, the deviations in separation size and position can be as large as 100% [
2]. Apsley and Leschziner [
16] used 12 turbulence models to simulate a wing/body junction numerically; although the results given by the two-equation SST model and the one-equation SA model were better than those given by the other models, all 12 turbulence models used in the calculation failed to capture well the separation position and size, and some models even failed to capture well the leading-edge saddle point and horseshoe–vortex structure. Parneix et al. [
17] used the V2F model to simulate wing–body corner flow; this captured the 3D separation line and the leading-edge horseshoe vortex and its intensity, but it failed to successfully capture the separation flow in the junction area. Lien et al. [
18] developed the improved the V2F method to calculate wing–body corner separation, but although the separation was captured successfully, the correct separation position and size were not.
Owing to the current lack of a deep understanding of flow separation in junction areas and the insufficient capture accuracy of the commonly used turbulence models based on linear eddy viscosity, since 2016 NASA has been carrying out a series of studies on the corner separated flow at the trailing edge caused by wing–body interference [
19,
20,
21,
22,
23]. The aims are to (i) study the flow characteristics in the corner area, and (ii) form another standard calculation example for computational fluid dynamics (CFD), to assess the solution accuracy of CFD technology. In particular, experiments are carried out on flow details such as Reynolds stress, which allows greater evaluation of CFD calculation methods and offers a clearer direction for improving them.
To explore how separated flows in junction areas influence the SDS of a demonstration aircraft and improved design methods, this paper begins by introducing the complexity of flow in a junction area and the deficiency of the RANS methods, which are more mature in engineering applications at this stage. In
Section 2, the shear-layer-adapted improved delayed detached eddy simulation (SLA-IDDES) method is briefly introduced, and then the standard Rood model is calculated using SLA-IDDES and compared with experimental results, to verify the accuracy of the flow simulated in the junction area. In
Section 3, the SLA-IDDES method is used to perform calculations for a demonstration aircraft at a small angle of attack and different sideslip angles, and the results are compared with those from a wind-tunnel experiment, to verify the reliability of the method and speculate about the causes of the loss of SDS. In
Section 4, the flow in the flat–vertical tail junction area of the demonstration aircraft is controlled via a rectifier cone, to inhibit the flow separation in the junction area, so as to analyze the specific causes of the loss of SDS and provide improvement measures.
3. Simulation and Analysis of the Demonstration Aircraft
3.1. Experimental Results and Discussion
A demonstration aircraft with two airframes was designed for the purpose of performing better flying tests of newly design airfoils in the future, other than using a wing glove. To fully observe the test wing with these two airframes, the horizontal tail was designed to be thicker than usual, to place a special device. The wind tunnel experimental model is shown in
Figure 8. Low-speed tests were carried out in the FL-8 wind tunnel of the Aviation Industry Corporation of China (AVIC), in which a 1:3.25 scale model was used, the test Reynolds number was ca. 1.5 × 10
6 and the Mach number was 0.2. In these low-speed tests, an abdominally supported balance was used for force measurements, and the influence of its aerodynamic shape on the aircraft flow field and aerodynamic forces was subtracted by means of symmetrical balance measurement tests, as shown in
Figure 8a–c.
From the experimental data shown in
Figure 8d, it can be seen that the aircraft lost SDS under a small angle of attack of between −4° and 0° and a small sideslip angle of between −4° and 4°. By contrast, under a large sideslip angle or positive angle of attack, the aircraft had SDS; even at an angle of attack of 2°, it could basically maintain SDS, and it only lost SDS at a sideslip angle of between −2° and 2°, unlike the larger range of sideslip angles in the previous case.
The main components of the aircraft that provide SDS are the vertical parts of its tail, and based on the experimental data, we speculated that the loss of SDS may have been due to one or more flow separations in the junctions of the horizontal and vertical tail parts, resulting in the vertical tail parts failing to provide SDS. When the aircraft has a small sideslip angle, the yaw moment generated by the vertical tail parts is small, as is that generated by the separation flow in the junctions of the horizontal and vertical tail parts, so the demonstration aircraft loses SDS under a small sideslip angle. Under a large sideslip angle, the yaw moment generated by the vertical tail parts is large, and even if corner separation flows are generated between the horizontal and vertical tail parts, they do not substantially affect the SDS. In addition, for this T-tail aircraft, the horizontal tail part is much thicker than usual, so the junction regions may suffer strong separation under a negative angle of attack, whereas under a positive angle of attack the separation in the junctions may be eased or even eliminated by the incoming flow, thereby allowing the aircraft to maintain SDS. Therefore, the demonstration aircraft was characterized by SDS under a large sideslip angle, which may be why it lacked SDS under a small sideslip angle.
3.2. Computational Grid and Verification
To verify the above preliminary analysis, we subjected the experimental model to CFD analysis. We assessed whether the calculation method and the grid used in the calculations met the accuracy requirements for subsequently analyzing the loss of SDS of the demonstration aircraft, then we analyzed its flow details in the T-tail junction regions and obtained data to support the subsequently improved design scheme.
To analyze the level of grid dependence, we used three different O-grids, with ca. 30, 50, and 70 million nodes, i.e., coarse, medium, and fine grids, respectively, which satisfied the criterion of y+ ≤ 1.
Figure 9a shows the aircraft surface meshed with the coarse grid, and
Figure 9b shows the surfaces of the horizontal and vertical tail parts meshed with the fine grid.
Table 2 shows the grids of the tail parts in detail, the distribution details of the horizontal and vertical tail parts are denoted “spanwise × chordwise” in
Table 2.
The inlet and outlet boundary of the computation domain was 30 times the length of the span-wise in three directions (as the length of the span-wise is larger than the fuselage) and the aircraft was at the center of the domain, while the numerical set up was the same as mentioned above in
Section 2.2. In the calculations, the Mach number was 0.2 and the Reynolds number was 1.5 × 10
6, which is consistent with the experimental values. We calculated and verified the yaw moment of the whole aircraft at an angle of attack of between 0° and −4°.
Figure 10 compares the yaw moment coefficients between the experimental measurements and the SLA-IDDES calculations with the three different grids. As can be seen, the numerical results obtained with the medium and fine grids agree well with the experimental results; they capture well the loss of SDS of the aircraft under a small angle of attack and small sideslip angle, indicating that the calculation method with the medium or fine grid is suitable for capturing the overall macroscopic quantities of the demonstration aircraft. However, although the medium grid could capture the loss of SDS of the aircraft under a small sideslip angle, to analyze the flow separation in the T-tail junction regions in detail, the following results were all calculated using the fine grid.
3.3. Discussion
As discussed in
Section 3.2, the experimental model of the demonstration aircraft was subjected to CFD analysis, the results were compared with the experimental measurements, and SLA-IDDES was shown to capture the loss of SDS under a small sideslip angle. To verify whether the conjecture made in
Section 3.1 is correct, we must analyze the streamlines and
Cp on the vertical tail surfaces.
Figure 11 and
Figure 12 show the surface streamlines at the pressure and suction sides, respectively, of the vertical tail parts of the demonstration aircraft at an angle of attack of 0° and a sideslip angle of 0°, 2°, and 4°. The streamlines show that there is separation on both the pressure and suction sides of the junction region at a sideslip angle of 0° and 2°, whereas at a sideslip angle of 4° the separation exists only on the suction side. Comparing
Figure 11 and
Figure 12 shows that the separation region on the suction side grows with an increasing sideslip angle, whereas that on the pressure side shrinks with an increasing sideslip angle. When the sideslip angle exceeds 4°, the separation region disappears and the aircraft has SDS, so the loss of SDS may be due to the separation on the suction side.
To see clearly the flow separation in the junction regions, we use an Ω isosurface [
32] to analyze the flow detail.
Figure 13 and
Figure 14 show the Ω = 0.52 isosurface colored by the streamwise velocity normalized by the incoming velocity. Clearly, the separation regions on the suction side grow with an increasing sideslip angle. Separation begins after the point of maximum thickness of the local airfoil used for the vertical and horizontal tail parts; with an increasing sideslip angle, the separation start point moves upstream and the separation region grows sharply; and when the sideslip angle exceeds 2°, the separation shows a strong vortex–structure flow in the wake and even comes into contact with the horizontal tail part. Meanwhile, the separation vortices on the pressure side are limited in the junction regions by the incoming flow and have a limited effect on the vertical and horizontal tail parts, and with an increasing sideslip angle the separation is eliminated.
The question remains as to whether the main reason for the failure of the vertical tail parts is the separation on the pressure side, and this requires deep analysis. As the separated flow on the pressure side is smaller in extent and is restrained more easily than that on the suction side, flow-control measures to either weaken or eliminate the separated flow on the pressure side are taken, to control the flow in the horizontal-vertical tail junction areas, which is more conducive to analyzing the specific causes of the loss of SDS.
In addition, with the separation in the T-tail junction regions, not only will the aircraft suffer from extra drag (known as interference drag) and loss of SDS under a small sideslip angle, but the rudders at the trailing edges of the vertical tail parts may suffer from reduced efficiency at a small sideslip angle. Therefore, the separation in the junction regions must be weakened or better eliminated to improve the aircraft’s aerodynamic performance, and so flow control must be used in the junction regions.
5. Conclusions
In this study, numerical simulations of a Rood airfoil and the JFs of a specially designed T-tail aircraft under small sideslip and attack angles were conducted using SLA-IDDES. As a newly improved method, SLA-IDDES is able to provide richer a flow field details than can traditional RANS methods, especially for small separation flows. Therefore, IDDES can be used as a state-of-the-art CFD tool for investigating small-separation configurations with different geometric characteristics, to understand better the physics of the onset and effects of flow separation under a small sideslip angle.
To ascertain whether SLA-IDDES could capture the small separation caused by strongly anisotropic flow, a standard model of Rood-airfoil JF was first simulated. Then, because this captured the flow details well, we used this new improved method to simulate a specially designed T-tail aircraft, to find out why in wind tunnel tests it loses SDS only at small angles of attack and sideslip.
To verify the conjectured reason for the failure of the vertical tail parts, i.e., that the suction-side flow separation is the main reason why the aircraft loses SDS, we considered a small fairing cone (instead of a bullet one), to relieve or even eliminate the separation on the suction and pressure sides, the aim was to ascertain the reason for the loss of SDS and allow the aircraft to meet flight safety standards. From that research, we concluded that the onset and development of flow separation on the suction side of the vertical tail parts constitute the main reason for the loss of SDS, especially regarding the right (resp. left) vertical tail part under a positive (resp. negative) sideslip angle. Moreover, the qualitative analysis showed that fitting small fairing cones was sufficient for this T-tail aircraft to recover SDS, and doing so was effective for eliminating the loss of SDS and reducing the separation, so as to reduce the drag of the whole T-tail by at least 9%, including the interference drag and pressure drag. The specific conclusions of this study can be summarized as follows.
(1) SLA-IDDES captures macroscopic quantities fairly well, and although it is based on the k–ω SST RANS model, it avoids the inherent fault of the k-ω SST method (i.e., incorrectly predicting a four-vortex system) and captures sufficient flow details in the junction regions.
(2) Even with vertical tail parts that are quite large, the present T-tail aircraft loses SDS at a small sideslip angle when flow separation is triggered in the junction regions, this being because the vertical tail parts under that condition offer only very limited SDS from their other regions.
(3) Fitting small fairing cones relieves the flow separation in the junction regions; doing so completely eliminates the separation on the pressure side and limits that in the fairing-cone regions on the suction side. This avoids loss of SDS and rudder failure at a small sideslip angle, reduces the extra drag produced by flow separation, and allows the modified aircraft to retain SDS.
(4) The main reason for vertical-tail failure at a small sideslip angle is that the junction regions trigger separation flow on the suction side; this then reduces greatly the suction peak on the suction side, and a loss of SDS ensues when the vertical tail parts fail to provide sufficient SDS under this condition. Thus, the vertical tail parts fail at small sideslip angles.
Regarding future work, the present study considered only one type of fairing cone, and to extend the present results, it would be of interest to analyze and discuss the effects of fairing-cone radii, length, and location.