AIAA JOURNAL Vol. 54, No. 1, January 2016 Coriolis Effect on Dynamic Stall in a Vertical Axis Wind Turbine Hsieh-Chen Tsai∗ and Tim Colonius† California Institute of Technology, Pasadena, California 91125 Downloaded by CALIFORNIA INST OF TECHNOLOGY on June 22, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.J054199 DOI: 10.2514/1.J054199 The immersed boundary method is used to simulate the flow around a two-dimensional cross section of a rotating NACA 0018 airfoil in order to investigate the dynamic stall occurring on a vertical axis wind turbine. The influence of dynamic stall on the force is characterized as a function of tip-speed ratio and Rossby number. The influence of the Coriolis effect is isolated by comparing the rotating airfoil to one undergoing an equivalent planar motion that is composed of surging and pitching motions that produce an equivalent speed and angle-of-attack variation over the cycle. Planar motions consisting of sinusoidally varying pitch and surge are also examined. At lower tip-speed ratios, the Coriolis force leads to the capture of a vortex pair when the angle of attack of a rotating airfoil begins to decrease in the upwind half cycle. This wake-capturing phenomenon leads to a significant decrease in lift during the downstroke phase. The appearance of this feature depends subtly on the tip-speed ratio. On the one hand, it is strengthened due to the intensifying Coriolis force, but on the other hand, it is attenuated because of the comitant decrease in angle of attack. While the present results are restricted to two-dimensional flow at low Reynolds numbers, they compare favorably with experimental observations at much higher Reynolds numbers. Moreover, the wake-capturing is observed only when the combination of surging, pitching, and Coriolis force is present. max SPM SSPM sin surge VAWT Nomenclature CL c Em k p R Re Ro U∞ u^ u0 W x^ α α_ δ λ ν ρ θ ^ ω ω0 Δt Δx Ω l = = = = = = = = = = = = = = = = = = = = = = = = = = lift coefficient airfoil chord length complete elliptic integral of the second kind reduced frequency pressure radius of the turbine Reynolds Number Rossby Number freestream velocity velocity of the fluid in the rotating frame of reference velocity introduced by the change of variables incoming velocity position vector in the rotating frame of reference angle of attack pitch rate spatial distribution of the body-force actuation tip-speed ratio fluid kinematic viscosity fluid density azimuthal angle vorticity of the fluid in the rotating frame of reference vorticity introduced by the change of variables time step grid spacing angular velocity of the turbine ratio of the radius of the turbine to the chord length = = = maximum sinusoidal pitching motion sinusoidal surging–pitching motion sinusoidal variation surge velocity vertical axis wind turbine I. V Introduction ERTICAL axis wind turbines (VAWT) offer several advantages over horizontal axis wind turbines (HAWT), namely: their low sound emission (consequence of their operation at lower tip-speed ratios), their insensitivity to yaw wind direction (because they are omnidirectional), and their increased power output in skewed flow [1,2]. Dabiri et al. [3,4] showed that an array of counterrotating VAWTs can achieve higher power output per unit land area and smaller wind velocity recovery distance than existing wind farms consisting of HAWTs. The aerodynamics of VAWTs are complicated by inherently unsteady flow produced by the large variations in both angle of attack and incident velocity magnitude of the blades, which can be characterized as a function of tip-speed ratio. Typically, commercial VAWTs operate at a tip-speed ratio around 2–5, which produces an angle-of-attack variation with amplitude of 11.5–30° and an incident velocity variation with amplitude of 21.5–49% of its mean. Aerodynamics of wings at low Reynolds numbers have been well investigated due to the recent interest in the development of small unmanned aerial vehicles and micro air vehicles. Morris and Rusak [5] studied the onset of stall at low to moderately high Reynolds number flows numerically and provided a universal criterion to determine the static stall angle of thin airfoils. Taira and Colonius [6] simulated three-dimensional flows around low-aspect-ratio flat-plate wings at low Reynolds numbers, with a focus on the unsteady vortex dynamics at poststall angles of attack. Choi et al. [7] numerically investigated unsteady, separated flows around two-dimensional (2D) surging and plunging airfoils at low Reynolds numbers. Dynamic stall refers to the delay in the stall of airfoils that are rapidly pitched beyond the static stall angle, which is associated with a substantially higher lift than is obtained quasi-statically, and has been an active research topic in fluid dynamics for more than 60 years, largely because of the helicopter application [8]. Because of large variations in angle of attack, dynamic stall occurs on VAWT operating at low tip-speed ratios [9]. To study this phenomenon, Wang et al. [10] introduced an equivalent planar motion (EPM), which is composed of a surging and pitching motion that produces an Subscripts avg EPM inst = = = = = = average velocity equivalent planar motion instantaneous velocity Presented as Paper 2014-3140 at the 32nd AIAA Applied Aerodynamics Conference, Atlanta, GA, 16–20 June 2014; received 16 January 2015; revision received 13 June 2015; accepted for publication 19 July 2015; published online 15 September 2015. Copyright © 2015 by Hsieh-Chen Tsai. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission. Copies of this paper may be made for personal or internal use, on condition that the copier pay the $10.00 per-copy fee to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; include the code 1533-385X/15 and $10.00 in correspondence with the CCC. *Graduate Student, Mechanical Engineering. Student Member AIAA. † Professor, Mechanical Engineering. Associate Fellow AIAA. 216 217 Downloaded by CALIFORNIA INST OF TECHNOLOGY on June 22, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.J054199 TSAI AND COLONIUS equivalent speed and angle-of-attack variation over the cycle. They further simplified the EPM to a sinusoidal pitching motion (SPM) to investigate dynamic stall in a 2-D VAWT numerically. The results matched the experiments done by Lee and Gerontakos [11]. Several attempts have been made to model a VAWT at Re ∼ O105 , which is appropriate to the urban applications of VAWTs. Reynolds-averaged Navier–Stokes (RANS) with different turbulence models has been applied to a 2-D airfoil undergoing an effective planar motion [12–14] and to a multibladed 2-D VAWT [15,16]. Ferreira et al. [17] simulated dynamic stall in a section of a VAWT using detached-eddy simulation at Re 50; 000 and validated the results by comparing the vorticity in the rotor area with particle image velocimetry (PIV) data. Duraisamy and Lakshminarayan [18] numerically analyzed interactions of VAWTs with various configurations using RANS at Re 67; 000. Barsky et al. [19] investigated the fundamental wake structure of a single VAWT computationally by large-eddy simulation and experimentally by PIV. In this paper, a 2-D VAWT is investigated numerically at low Reynolds numbers in order to understand qualitative features of the flow field in a setting where a comparatively large region of parameter space can be explored than could be for full-scale, threedimensional computations. To explore the parameter space in relatively short computational time and have more understanding of the details of the vortex dynamics, flows are simulated at low Reynolds numbers, Re ∼ O103 . A major limitation of our present approach is the restriction of flow to a 2-D cross section of an otherwise planar turbine geometry. Comparisons with Ferreira et al. [17] show qualitative agreement, but a precise accounting for threedimensional effects awaits future simulations. We focus here on the Coriolis effect on dynamic stall by comparing the rotating airfoil to one undergoing an EPM. The influence of dynamic stall on forces is characterized as a function of tip-speed ratio and Rossby number. Moreover, inspired by Wang et al. [10], airfoils undergoing an SPM and a sinusoidal surging–pitching motion (SSPM) are also compared to see if these two motions can be an appropriate model for the VAWT. Furthermore, the coupling of the Coriolis effect with the angle-of-attack and incoming velocity variations is also examined. II. Methodology A. Simulation Setup Figure 1 shows a schematic of a VAWT with radius R rotating at an angular velocity Ω with a freestream velocity U∞, coming from the left. The chord length of the turbine blade is c. To systematically investigate the aerodynamics of a VAWT, four dimensionless parameters are introduced: Tip-speed ratio∶ λ ΩR U∞ (1) Radius-to-chord-length ratio∶ l Ro Rossby number∶ (2) U∞ c ν (3) U∞ l 1 2Ωc 2λ 4k (4) Re Renolds number∶ R c where ν is the kinematic viscosity of the fluid and k is the reduce frequency, k Ωc∕2U∞ λ∕2l. The instantaneous incoming velocity W inst and the angle of attack α can then be characterized as a function of the tip-speed ratio λ and the azimuthal angle θ: αλ; θ tan−1 sin θ λ cos θ (5) W inst λ; θ p2 1 2λ cos θ λ U∞ (6) Figure 2 shows the angle-of-attack variation and incoming velocity variation of the VAWT at λ 2. From Eq. (5), the maximum angle of attack αmax λ tan−1 1 p 2 λ −1 (7) occurs at θ cos−1 −1∕λ. To isolate the Coriolis effect on dynamic stall, a moving airfoil experiencing an equivalent incoming velocity and angle-of-attack variation over a cycle is proposed. This EPM is composed of a surging motion with a velocity W surge and a pitching motion around _ The airfoil is undergoing the EPM the leading edge with a pitch rate α. in a freestream velocity W avg. W avg , W surge , and α_ are shown to be W avg λ 1 2π U∞ Z 2π W 0 inst λ;θ U∞ 21λ E dθ π s! 4λ 1λ2 (8) p where the function Em ∫ π∕2 1 − m2 sin2 θ dθ is the complete 0 elliptic integral of the second kind, W surge λ; θ W inst λ; θ − W avg λ (9) 1 1 λ cos θ 2Ro 1 2λ cos θ λ2 (10) _ θ αλ; Moreover, due to the periodic oscillation of angle-of-attack and incoming velocity variation, Wang et al. [10] studied dynamic stall in a VAWT by investigating an airfoil undergoing a simplified SPM. Inspired by their work, sinusoidal variations in the angle of attack and incoming velocity, which are written as a function of the tip-speed ratio λ and the azimuthal angle θ in Eqs. (11) and (12) and shown in Fig. 2, are also considered: Fig. 1 Schematic of a VAWT and the computational domain. αsin λ; θ αmax λ sin θ (11) W inst;sin λ; θ λ cos θ U∞ (12) 218 TSAI AND COLONIUS 4 V.A.W.T. sinusoidal motion 30 V.A.W.T. sinusoidal motion 3.5 20 3 10 2.5 0 2 −10 1.5 1 −20 0.5 −30 0 Downloaded by CALIFORNIA INST OF TECHNOLOGY on June 22, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.J054199 Fig. 2 60 120 180 240 300 360 0 0 60 120 180 240 300 360 a) Angle of attack variation b) Incoming velocity variation Comparison of angle-of-attack variation and incoming velocity variation between the VAWT and the sinusoidal motion at λ 2. Although the sinusoidal motion shares the same amplitude, it overestimates the angle of attack in the upstroke phase, underestimates the incoming velocity in the downstroke phase, and slightly underestimates the instantaneous velocity over the entire half cycle. To search for the most appropriate model for a VAWT, we introduce two additional motions: SPM and SSPM. Airfoils undergoing both the SPM and SSPM pitch with the sinusoidal angleof-attack variation described in Eq. (11) in a freestream velocity W avg;sin λU∞ . Airfoils undergoing the SSPM also surge with a velocity W surge;sin U∞ cos θ. NACA 0018 airfoils are used as blades in the present study. The ratio of the radius of the turbine to the chord length l depends on the choice of the tip-speed ratio λ and the Rossby number Ro. In preliminary simulations of a three-bladed VAWT, as well as in previous studies [17], vorticity–blade interaction is only observed in the downwind half of a cycle. Because only the flow in the upwindhalf cycle is important to torque generation, we save computational time by modeling a single-bladed turbine. We compute about five periods of motion, which is equal to 5π∕Ro convective time units, to remove transients associated with the startup of periodic motion. For the largest Rossby number we examined, the starting vortex propagates far enough into the wake to have an insignificant effect on the forces on the blades after five periods. An additional five periods of nearly periodic stationary-state motion were then computed and analyzed below. B. Numerical Method The immersed boundary projection method (IBPM) developed by Colonius and Taira [20,21] is used to compute 2-D incompressible flows in an airfoil-fixed reference frame with appropriate forces added to the momentum equation to account for the noninertial reference frame. The equations are solved on multiple overlapping grids that become progressively coarser and larger (greater extent). The dimensionless momentum equation in the rotating frame of reference is ∂u^ 1 2 dΩ u^ · ∇u^ −∇p ∇ u^ − × x^ − 2Ω × u^ − Ω ∂t Re dt ^ × Ω × x (13) where u^ and x^ are the fluid velocity and the position vector in the rotating frame of reference and Ω 1∕2Ro is the dimensionless angular velocity of rotating frame of reference. We then introduce the change of variables ^ ∇ × u^ is the vorticity field in the rotating frame of where ω reference. By taking the divergence and the curl of Eq. (13), we have the following vorticity and pressure equation ∂ω 0 1 ∇ × ∇ × ω 0 ∇ × ^u × ω 0 − Re ∂t (16) 1 2 1 ^ − jΩ × xj ^ 2 ∇ · u^ × ω 0 ∇2 p juj 2 2 (17) Because the flow is incompressible and 2-D, the first term on the right-hand side of Eq. (16) is just the advection of vorticity with ^ Therefore, in the body-fixed frame of reference, the Coriolis velocity u. force does not generate vorticity except on the boundary so that it only changes the way vorticity propagates in free space. Moreover, because the magnitude of Coriolis force is proportional to the magnitude of velocity, fluid with high velocity will be deflected more rapidly. C. Verification and Validation The IBPM has been validated and verified by Colonius and Taira [20,21] and others for problems such as three-dimensional flows around low-aspect-ratio flat-plate wings [6,22], optimized control of vortex shedding from an inclined flat plate [23,24], 2-D flows around a NACA0018 airfoil with a cavity [25,26], and 2-D flows around surging and plunging flat-plate airfoils [7]. As noted in Sec. II.B, the Navier–Stokes equations are solved on multiple overlapping grids. Based on the analysis by Colonius and Taira [21], the IBPM is estimated to have an O4−Ng convergence rate, where N g is the number of the grid levels. Six grid levels are used for the present computations in order to have the leading-order error dominated by the truncation error arising from the discrete delta functions at the immersed boundary and the discretization of the Poisson equation. The coarsest grid extends to 96 chord lengths in both the transverse and streamwise directions of the blade. To show grid convergence, we examine a single-bladed VAWT with l 1.5 rotating at λ 2. The velocity field in the streamwise direction is compared with one on the finest grid with Δx 0.00125 at t 1. Figure 3 shows the spatial convergence in the L2 norm. The rate of decay for the spatial error is about 1.5, which agrees with Taira and Colonius [20]. All ensuing computations use a 600 × 600 grid, which corresponds to Δx 0.005. The time step Δt is chosen to make Courant–Friedrichs–Lewy number less than 0.4. III. Results A. Qualitative Flow Features in a VAWT u 0 u^ Ω × x^ (14) ^ 2Ω ω0 ω (15) We begin by examining flows at low tip-speed ratio, λ 2, Ro 1.5, and Re 1500, which gives a maximum amplitude of 30° in angle-of-attack variation and a reduced frequency k 1∕6. Figure 4 shows the vorticity field generated by the blade at different azimuthal angles over a cycle. Negative and positive vorticity are TSAI AND COLONIUS −1 10 −2 10 −3 Downloaded by CALIFORNIA INST OF TECHNOLOGY on June 22, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.J054199 10 −3 10 −2 10 Fig. 3 The L2 norm of the error of the velocity field in the streamwise direction in a single-bladed VAWT with l 1.5 rotating at λ 2 at t 1. Fig. 4 219 plotted in blue and red contour levels, respectively, and all vorticity contour plots use the same contour levels. At the beginning of a cycle (Fig. 4a), the airfoils are just returning to zero angle of attack, and there are still the remnants of earlier vortex shedding in the wake. The flow reattaches by the time airfoil reaches α 5∘ (Fig. 4b). When the angle of attack increases further, the wake behind the airfoil starts to oscillate and vortex shedding commences. Dynamic stall then takes place, and is marked by the growth, pinchoff, and advection of a leading edge vortex (LEV) on the suction side of the airfoil (Figs. 4c–4e). The vortices generated will propagate downstream into the wake of the VAWT or interact with the blades in the downwind half of a cycle. When the angle of attack starts to decrease, a trailing edge vortex (TEV) develops (Fig. 4f). Bloor instability [27] occurs in the trailingedge shear layer at this Reynolds number, which resembles the convectively unstable Kelvin–Helmholtz instability observed in plane mixing layers. This TEV couples with an LEV to form a vortex pair that travels downstream together with the airfoil (Figs. 4g–4i). Vorticity field for a (clockwise rotating) VAWT at various azimuthal angles at λ 2, Ro 1.5, and Re 1500. 220 TSAI AND COLONIUS B. Comparison of VAWT and EPM In this section, we compare flows around an airfoil undergoing the EPM and a single-bladed VAWT at λ 2, Ro 1.5, and Re 1500. We are interested in the tangential force response of the blade over a cycle because the power output is proportional to the tangential force acting on the blade when VAWTs operate at a constant tip-speed ratio. The tangential force acting on the blade can be written as a linear combination of lift and drag: Downloaded by CALIFORNIA INST OF TECHNOLOGY on June 22, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.J054199 This vortex pair interacts with the airfoil in the downwind half of a cycle (Figs. 4j–4l), which was also observed by Ferreira et al. [17]. When the blade rotates in the downwind half of a cycle, the angle of attack becomes negative. Vortices are now generated on the other side of the airfoil and shed into the wake of the VAWT (Figs. 4j–4p). For a multibladed VAWT, when a blade is traveling in the downwind half of a cycle, it interacts with vortices generated upstream from other blades or from the wake it generated at an earlier time (Fig. 4o). Fig. 5 Vorticity field for EPM and VAWT and the Coriolis force for VAWT at various azimuthal angles. 221 TSAI AND COLONIUS Downloaded by CALIFORNIA INST OF TECHNOLOGY on June 22, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.J054199 CT CL sin α − CD cos α CL sin α − 1 cos α CL ∕CD Figure 5 shows the comparison of VAWT and EPM motion in the surging–pitching configuration. Negative and positive vorticity are plotted in blue and red contour levels, respectively, and all vorticity contour plots are using the same contour levels. In the Coriolis force plots, black arrows show the direction of velocity, blue arrows point the direction of the Coriolis force, and the color contour plots the magnitude of the Coriolis force. Since the frame of reference is rotating clockwise, the Coriolis force deflects the fluid in the clockwise direction. Figure 6 shows a comparison of the lift coefficient against dimensionless time and angle of attack for a single rotation and for the average of both lift coefficients over five cycles. Although there are still the remnants of earlier vortex shedding in the wake when the airfoil just returns to zero angle of attack (Fig. 5a), the flow reattaches by α 5° (Fig. 5b), which leads to a smoothly increasing lift coefficient at low angle of attack. The differences in the lift coefficient between the EPM and VAWT are small (Fig. 6). As the angle of attack increases, dynamic stall commences (Figs. 5c–5e), which leads to rapidly increasing lift. EPM- and VAWT-induced (18) where α is the angle of attack of the blade. From preliminary simulations, a three-bladed VAWT with l 4 will be free-spinning with a time-averaged tip-speed ratio λ 0.95 at Re 1500, so that in the flow we are examining the average tangential force is expected to be negative. However, as the Reynolds number increases to the range where commercial VAWTs usually operate, Re ∼ O105 − 106 , drag coefficient drops dramatically while the change in the lift coefficient is small. This leads to a large increase in the lift-to-drag ratio CL ∕CD [28]. Therefore, the contribution of lift to the tangential force dominates at high Reynolds numbers. Moreover, the power of a VAWT is generated mostly in the upwind half cycle because large vorticity–blade interactions cancel out the driving torque in the downwind half cycle [17]. Therefore, in this study we will focus on the lift in the upwind half of a cycle. 3 3 VAWT EPM 3 VAWT EPM 2 2 2 1 1 1 0 0 0 −1 −1 −1 −2 0 −2 30 60 90 120 150 180 VAWT EPM −2 0 5 10 15 20 25 30 0 30 60 90 120 150 180 c)The average lift coefficients over five cycles. a) The lift coefficients over a cycle against b) The lift coefficients over a cycle against dimensionless time. angle of attack. Fig. 6 Comparing CL;VAWT and CL;EPM at λ 2, Ro 1.5, and Re 1500. 3 3 2 VAWT EPM SPM SSPM 2 1 1 0 0 −1 −1 −2 0 1 5 10 15 20 25 30 VAWT EPM SPM SSPM −2 0 1 0.5 0.5 0 0 −0.5 −0.5 0 2 4 6 8 10 12 14 VAWT EPM SPM SSPM 0 5 10 15 20 25 30 VAWT EPM SPM SSPM 2 4 6 8 10 12 14 Fig. 7 Lift coefficients of VAWT, EPM, SPM, and SSPM over a cycle at λ 2 and 4, Ro 0.75, and Re 1000. Downloaded by CALIFORNIA INST OF TECHNOLOGY on June 22, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.J054199 222 TSAI AND COLONIUS overestimated in the downstroke phase. Moreover, when this vortex pair travels downstream, it interacts with the airfoil in the downwind half of a cycle. This leads to a lift coefficient with large fluctuations and small mean, as was also observed by Ferreira et al. [17]. We can see from the second and third columns in Fig. 5 that the Coriolis force deflects the flow around the rotating airfoil in the clockwise direction. The magnitude of the Coriolis force acting on the background fluid decreases as the azimuthal angle increases. Therefore, the Coriolis force acting on the fluid around vortices becomes relatively important in the downstroke phase. A stronger Coriolis force is exerted on the fluid around the vortex pair, which deflects the fluid in such a way that the vortex pair travels with the airfoil (Figs. 5f–5h). flows are quite similar, with just a small phase difference when the airfoils pitch up. They result in comparable lift throughout the upstroke phase. When the downstroke phase starts (Figs. 5e and 5f), the development of a TEV leads to a decrease in lift. The aforementioned Bloor instability in the shear layer at the trailing edge produces highfrequency fluctuations in the lift coefficient. For EPM, the TEV sheds into the wake and a secondary vortex [29] appears as the angle of attack decreases further (Fig. 5g), which results in a sudden increase in the lift coefficient. On the other hand, for VAWT, as described in Sec. III.A, this TEV couples with the LEVand forms a vortex pair that travels together with the airfoil (Figs. 5g and 5h). This generates high pressure on the suction side and further decreases the lift. This vortex pair is “captured” by the rotating airfoil. By analogy with flow observed in insect flight by Dickinson et al. [30], we refer to this phenomenon as the wake-capturing of a vortex pair in VAWT. The wake-capturing occurs at a slightly different phase in each cycle and leads to a significant decrease in the average lift in the downstroke phase. In general, the lift of an airfoil undergoing the EPM is C. Comparison with an Airfoil Undergoing a Sinusoidal Motion Flows around an airfoil undergoing SPM and SSPM introduced in Sec. II.A are compared with one undergoing EPM and in a VAWT. A comparison of the lift response at λ 2 and 4, Ro 1.5, and Re 1000 is shown in Fig. 7. 3 3 3 2 2 2 1 1 1 0 0 0 −1 −1 −1 −2 −2 −2 −3 0 30 60 90 120 150 180 −3 0 30 60 90 120 150 180 −3 0 3 3 3 2 2 2 1 1 1 0 0 0 −1 −1 −1 −2 −2 −2 −3 0 30 60 90 120 150 180 −3 0 30 60 90 120 150 180 −3 0 3 3 3 2 2 2 1 1 1 0 0 0 −1 −1 −1 −2 −2 −2 −3 0 30 60 90 120 150 180 −3 0 30 60 90 120 150 180 −3 0 Ro=0.75, VAWT Ro=0.75, EPM Ro=1.00, VAWT Ro=1.00, EPM Ro=1.25, VAWT Ro=1.25, EPM 30 60 90 120 150 180 30 60 90 120 150 180 30 60 90 120 150 180 Fig. 8 Comparing lift coefficients of VAWT and EPM with Ro 0.75, 1.00, and 1.25 at λ 2, 3, and 4 and Re 500, 1000, and 1500. TSAI AND COLONIUS considered in this study. However, it overestimates the lift coefficients in the downstroke phase due to its inability to predict the wake-capturing phenomenon. D. Effect of Tip-Speed Ratio, Rossby Number, and Reynolds Number In this section, the effect of tip-speed ratio, Rossby number, and Reynolds number on the flow in a VAWT is investigated to understand when wake-capturing will occur. We compare the simulations of a rotating wing with a wing undergoing EPM. We examine the flows at tip-speed ratios λ 2, 3, and 4, and Reynolds numbers Re 500, 1000, and 1500. The corresponding lift coefficients with Rossby numbers Ro 0.75, 1.00, and 1.25 are shown in Fig. 8. As the tip-speed ratio increases, the amplitudes of angle-of-attack variation and the corresponding lift decrease. Because the maximum Downloaded by CALIFORNIA INST OF TECHNOLOGY on June 22, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.J054199 At lower tip-speed ratio, λ 2, in the upstroke phase, we can see that only CL;EPM is close to CL;VAWT at low angle of attack. CL;SPM and CL;SSPM overestimate the lift due to the overestimation of the pitch rate. In the downstroke phase, none of CL;EPM , CL;SPM , and CL;SSPM matches the behavior of CL;VAWT because of the strong effect on lift of the wake-capturing that occurs in the flows. At higher tipspeed ratio, λ 4, CL;SPM and CL;SSPM still overestimate the lift at the beginning of the upstroke phase. Nevertheless, as the angle of attack increases, and after vortex shedding starts, differences between the four lift coefficients are relatively small. In the downstroke phase, behaviors of CL;EPM , CL;SPM , and CL;SSPM are close to that of CL;VAWT due to the low angle of attack. We can see that, among all simplified motions, an airfoil undergoing the EPM is the best approximation to a rotating airfoil in a VAWT in the upstroke phase for the subscale Reynolds numbers 223 Fig. 9 The comparison of the vorticity fields of the LEV-filtered (a–f) and TEV-filtered (g–i) phase-averaged PIV data (gray scale), and of the corresponding simulations (color scale) at λ 2 and Ro 1. Downloaded by CALIFORNIA INST OF TECHNOLOGY on June 22, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.J054199 224 TSAI AND COLONIUS angle of attack is slightly above the static stall angle of a NACA 0018 airfoil predicted by Morris and Rusak [5], the lift coefficients of EPM are close to that of VAWT at λ 4 for all Rossby numbers and Reynolds numbers examined. Therefore, EPM motion is a good approximation of VAWT at larger tip-speed ratios due to the low angle of attack. However, at lower tip-speed ratios, CL;EPM remains close to CL;VAWT only in the upstroke phase. In the downstroke phase, the discrepancy in lift coefficients due to the wake-capturing effect becomes larger as Rossby number decreases and Reynolds number increases. As the VAWT rotates faster, on the one hand, the wake-capturing effect is strengthened due to the intensifying Coriolis force, which corresponds to decreasing Rossby number; on the other hand, it is attenuated because of the decreasing amplitude of the angle-of-attack variation due to the increasing tip-speed ratio. Therefore, the growth of the discrepancy depends subtly on the increase of the rotating speed of the VAWT. To probe the existence of wake-capturing at higher Reynolds numbers, the vorticity field in VAWT (Re 1500) for a single period is compared with phase-averaged PIV data from Ferreira et al. [17] (Re ≈ 105 ) at λ 2 and Ro 1 (l 4) in Fig. 9. The contours in gray are the phase-averaged vorticity field taken from the experiments. In Figs. 9a–9f, their phase-averaged field was filtered to plot only the LEV generated around θ 72° and the plot represents a composite of overlaid fields from various azimuthal angles. Similarly, in Figs. 9g–9i, the contours in the gray scale show the filtered, phase-averaged TEV evolution. To make qualitative comparisons, our vorticity fields at the corresponding azimuthal angles are overlaid in the color scale on top of the results from the experiments. Our blue contours correspond to negative vorticity, which should be compared with the LEV-filtered PIV data in Figs. 9a–9f, while the red contours correspond to positive vorticity that should be compared with the TEV-filtered PIV data in Figs. 9g–9i. The trajectories of the LEVand TEV from Ferreira et al. [17] seem to be reasonably captured by the simulation in the upwind half of a cycle. The disagreement in Figs. 9f and 9i may be due to strong vortex–blade and vortex–vortex interactions in the downwind half of a cycle. An LEV is generated around θ 72∘ and wake-capturing occurs around θ 90° , which forms a vortex pair traveling with the blade (Figs. 9a–9c). The vortex pair then detaches around θ 133° and propagates downstream (Figs. 9d–9f). The location of the vortex pair composed of the phase-averaged LEV and TEV agrees with the current simulation, especially at θ 158° (Figs. 9e and 9h). The qualitative agreement in the upwind half of a cycle suggests that wake-capturing may also be occurring in Ferreira et al. [17] experiment. E. Decoupling the Effect of Surging, Pitching, and Rotation The flow around a rotating airfoil in a VAWT is complicated not only by the Coriolis effect but also because the angle of attack and incoming velocity vary simultaneously. It is interesting to understand whether the Coriolis effect has strong coupling with the angle-of- 1.5 VAWT Equiv. motion 3 VAWT Equiv. motion 2.5 2 1.5 1 0.5 0 −0.5 −1 0 30 60 90 120 150 Fig. 11 Comparing lift responses of airfoils undergoing only pitching motion at λ 2, Ro 1.5, and Re 1500. attack or incoming velocity variations. Therefore, we independently examine airfoils undergoing the decoupled pitching and surging motion associated with the EPM. 1. Airfoil Undergoing Only Surging Motion We examine a surging motion with fixed angles of attack of 15° and 30° at λ 2, Ro 1.5 (k 1∕6), and Re 1500. A rotating airfoil undergoing only the surging motion of a VAWT is achieved by pitching the airfoil around the leading edge simultaneously as it rotates so that the angle of attack is fixed with respect to the incoming velocity. For an airfoil surging at an angle of attack of 15°, lift coefficients are shown in Fig. 10a. We can see that dynamic stall is relatively stable and no wake-capturing phenomenon is observed. Moreover, from the analysis by Choi et al. [7], when the reduced frequency is low enough, the flow can be approximated as quasisteady, which results in both lift coefficients for VAWT and EPM fluctuating about a slowly increasing mean value. For the case of α 30°, lift coefficients are shown in Fig. 10b. The flow is well separated so that there is no stationary vortex shedding. Moreover, no wake-capturing phenomenon is observed in the flow. 2. Airfoil Undergoing Only Pitching Motion We consider a pitching motion in a freestream velocity W avg λU∞ at λ 2, Ro 1.5, and Re 1500. A rotating airfoil undergoing only the pitching motion in a VAWT is achieved by rotating an airfoil in a VAWT without the external freestream and simultaneously pitching it around the leading edge with the exact angle-of-attack variation. The corresponding lift coefficients are shown in Fig. 11. We can see dynamic stall in both lift coefficients as angle of attack increases. However, there is no wake-capturing. The wake-capturing effect is therefore only present when pitching, surging, and the Coriolis force are all present. 3.5 VAWT Equiv. motion 3 2.5 1 2 1.5 0.5 1 0.5 0 0 30 60 90 120 150 180 180 0 0 30 60 90 120 150 180 b) α = 30° a) α = 15° Fig. 10 Comparing lift responses of airfoils undergoing only surging motion at λ 2, Ro 1.5, and Re 1500. TSAI AND COLONIUS Downloaded by CALIFORNIA INST OF TECHNOLOGY on June 22, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.J054199 IV. Conclusions In simulating the flow around a single-bladed vertical axis wind turbine (VAWT), an interesting wake-capturing phenomenon that occurs during the pitch-down portion of the upstream, lift-generating portion of the VAWT cycle was observed. This phenomenon leads to a substantial decrease in lift coefficient due to the presence of a vortex pair traveling together with the rotating airfoil. Our results show that this flow feature persists and grows stronger as tip-speed ratio and Rossby number are reduced and Reynolds number is increased. Therefore, the growth of this features depends subtly on the increase of the rotating speed of the VAWT, which, on the one hand, is strengthened due to the intensifying Coriolis force. On the other hand, it is attenuated because of the decreasing amplitude of the angle-of-attack variation. Moreover, although our study is restricted to 2-D flow at relatively low Reynolds numbers, the qualitative agreement of the leading edge vortex and trailing edge vortex evolutions with Ferreira et al. [17] experiment suggests that this feature may persist in real applications. The corresponding decrease in efficiency could be improved by implementing flow control (e.g., blowing) to remove this flow feature [31]. An equivalent planar surging–pitching motion was introduced in order to isolate the Coriolis effect on dynamic stall in a VAWT. Simplified planar motions consisting of sinusoidally varying pitch and surge were also examined. Except at the beginning of the pitch-up motion, all of the simplified motions are good approximations to VAWT motion at sufficiently high tip-speed ratios because the corresponding maximum angle of attack is close to or lower than the stall angle of the blade. However, at low tip-speed ratios, while the equivalent planar motion captures the pitch-up part of the cycle, all the motions show significant differences in forces during the pitchdown motion. The results show that the equivalent motion is a good approximation to a rotating airfoil in a VAWT in the upstroke phase where the Coriolis force has a relatively small effect on vortices. However, it overestimates the average lift coefficient in the downstroke phase by eliminating the aforementioned wakecapturing. The flow by decomposing the planar motion into surging- and pitching-only motions was further investigated. Wake-capturing was observed only when the combination of surging, pitching, and rotation are present, which suggests that this feature is associated with an unique combination of angle-of-attack variation, instantaneous velocity variation, and the Coriolis effect. Acknowledgments This project is sponsored by the Caltech Field Laboratory for Optimized Wind Energy with John Dabiri as principal investigator under the support of the Gordon and Betty Moore Foundation. We would like to thank John Dabiri, Beverley McKeon, Reeve Dunne, and Daniel Araya for their helpful comments on our work. The parametric study in this work used the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by National Science Foundation grant number ACI-1053575. [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] References [1] Mertens, S., van Kuik, G., and van Bussel, G., “Performance of an HDarrieus in the Skewed Flow on a Roof,” Journal of Solar Energy Engineering, Vol. 125, No. 4, 2003, pp. 433–440. doi:10.1115/1.1629309 [2] Ferreira, C. S., van Bussel, G. J., and van Kuik, G. A., “Wind Tunnel Hotwire Measurements, Flow Visualization and Thrust Measurement of a VAWT in Skew,” Journal of Solar Energy Engineering, Vol. 128, No. 4, 2006, pp. 487–497. doi:10.1115/1.2349550 [3] Dabiri, J. O., “Potential Order-of-Magnitude Enhancement of Wind Farm Power Density via Counter-Rotating Vertical-Axis Wind Turbine Arrays,” Journal of Renewable and Sustainable Energy, Vol. 3, No. 4, 2011, Paper 043104. doi:10.1063/1.3608170 [4] Kinzel, M., Mulligan, Q., and Dabiri, J. O., “Energy Exchange in an Array of Vertical-Axis Wind Turbines,” Journal of Turbulence, Vol. 13, [21] [22] [23] [24] 225 No. 38, 2012, pp. 1–13. doi:10.1080/14685248.2012.712698 Morris, W. J., and Rusak, Z., “Stall Onset on Aerofoils at Low to Moderately High Reynolds Number Flows,” Journal of Fluid Mechanics, Vol. 733, Oct. 2013, pp. 439–472. doi:10.1017/jfm.2013.440 Taira, K., and Colonius, T., “Three-Dimensional Flows Around LowAspect-Ratio Flat-Plate Wings at Low Reynolds Numbers,” Journal of Fluid Mechanics, Vol. 623, March 2009, pp. 187–207. doi:10.1017/S0022112008005314 Choi, J., Colonius, T., and Williams, D. R., “Surging and Plunging Oscillations of an Airfoil at Low Reynolds Number,” Journal of Fluid Mechanics, Vol. 763, Jan. 2015, pp. 237–253. doi:10.1017/jfm.2014.674 Corke, T. C., and Thomas, F. O., “Dynamic Stall in Pitching Airfoils: Aerodynamic Damping and Compressibility Effects,” Annual Review of Fluid Mechanics, Vol. 47, No. 1, 2015, pp. 479–505. doi:10.1146/annurev-fluid-010814-013632 Ferreira, C. S., van Kuik, G., van Bussel, G., and Scarano, F., “Visualization by PIV of Dynamic Stall on a Vertical Axis Wind Turbine,” Experiments in Fluids, Vol. 46, No. 1, 2009, pp. 97–108. doi:10.1007/s00348-008-0543-z Wang, S., Ingham, D. B., Ma, L., Pourkashanian, M., and Tao, Z., “Numerical Investigations on Dynamic Stall of Low Reynolds Number Flow Around Oscillating Airfoils,” Computers & Fluids, Vol. 39, No. 9, 2010, pp. 1529–1541. doi:10.1016/j.compfluid.2010.05.004 Lee, T., and Gerontakos, P., “Investigation of Flow over an Oscillating Airfoil,” Journal of Fluid Mechanics, Vol. 512, Aug. 2004, pp. 313–341. doi:10.1017/S0022112004009851 Paraschivoiu, I., and Allet, A., “Aerodynamic Analysis of the Darrieus Wind Turbines Including Dynamic-Stall Effects,” Journal of Propulsion and Power, Vol. 4, No. 5, 1988, pp. 472–477. doi:10.2514/3.23090 Allet, A., Hallé, S., and Paraschivoiu, I., “Numerical Simulation of Dynamic Stall Around an Airfoil in Darrieus Motion,” Journal of Solar Energy Engineering, Vol. 121, No. 1, 1999, pp. 69–76. doi:10.1115/1.2888145 Paraschivoiu, I., and Beguier, C., “Visualization, Measurements and Calculations of Dynamic Stall for a Similar Motion of VAWT,” Proceedings of the European Community Wind Energy Conference, Commission of the European Communities, Luxembourg, 1998, pp. 197–201. Hansen, M., and Søresen, D., “CFD Model for Vertical Axis Wind Turbine,” Proceedings of the 2001 European Wind Energy Conference and Exhibition, WIP–Renewable Energies, Munich, 2001. Nobile, R., Vahdati, M., Barlow, J., and Mewburn-Crook, A., Dynamic Stall for a Vertical Axis Wind Turbine in a Two-Dimensional Study,” World Renewable Energy Congress, Linköping, Sweden, 2011. Ferreira, C. S., van Bussel, G., and Van Kuik, G., “2-D CFD Simulation of Dynamic Stall on a Vertical Axis Wind Turbine: Verification and Validation with PIV Measurements,” 45th AIAA Aerospace Sciences Meeting and Exhibit, AIAA Paper 2007-1367, 2007. Duraisamy, K., and Lakshminarayan, V., “Flow Physics and Performance of Vertical Axis Wind Turbine Arrays,” 32nd AIAA Applied Aerodynamics Conference, AIAA Paper 2014-3139, 2014. Barsky, D., Posa, A., Rahromostaqim, M., Leftwich, M. C., and Balaras, E., “Experimental and Computational Wake Characterization of a Vertical Axis Wind Turbine,” 32nd AIAA Applied Aerodynamics Conference, AIAA Paper 2014-3141, 2014. Taira, K., and Colonius, T., “The Immersed Boundary Method: A Projection Approach,” Journal of Computational Physics, Vol. 225, No. 2, 2007, pp. 2118–2137. doi:10.1016/j.jcp.2007.03.005 Colonius, T., and Taira, K., “A Fast Immersed Boundary Method Using a Nullspace Approach and Multi-Domain Far-Field Boundary Conditions,” Computer Methods in Applied Mechanics and Engineering, Vol. 197, No. 25, 2008, pp. 2131–2146. doi:10.1016/j.cma.2007.08.014 Taira, K., and Colonius, T., “Effect of Tip Vortices in Low-ReynoldsNumber Poststall Flow Control,” AIAA Journal, Vol. 47, No. 3, 2009, pp. 749–756. doi:10.2514/1.40615 Joe, W. T., Colonius, T., and MacMynowski, D. G., “Optimized Control of Vortex Shedding from an Inclined Flat Plate,” AIAA Paper 20094027, 2009. Joe, W. T., Colonius, T., and MacMynowski, D. G., “Optimized Waveforms for Feedback Control of Vortex Shedding,” Active Flow Control II, edited by King, R., Vol. 108, Notes on Numerical Fluid 226 [25] [26] [27] Downloaded by CALIFORNIA INST OF TECHNOLOGY on June 22, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.J054199 [28] TSAI AND COLONIUS Mechanics and Multidisciplinary Design, Springer–Verlag, Berlin, 2010, pp. 391–404. Olsman, W. F. J., Willems, J., Hirschberg, A., Colonius, T., and Trieling, R., “Flow Around a NACA0018 Airfoil with a Cavity and Its Dynamical Response to Acoustic Forcing,” Experiments in Fluids, Vol. 51, No. 2, 2011, pp. 493–509. doi:10.1007/s00348-011-1065-7 Olsman, W. F. J., and Colonius, T., “Numerical Simulation of Flow over an Airfoil with a Cavity,” AIAA Journal, Vol. 49, No. 1, 2011, pp. 143–149. doi:10.2514/1.J050542 Bloor, M. S., “The Transition to Turbulence in the Wake of a Circular Cylinder,” Journal of Fluid Mechanics, Vol. 19, No. 02, 1964, pp. 290– 304. doi:10.1017/S0022112064000726 Lissaman, P., “Low-Reynolds-Number Airfoils,” Annual Review of Fluid Mechanics, Vol. 15, No. 1, 1983, pp. 223–239. doi:10.1146/annurev.fl.15.010183.001255 [29] McCroskey, W., Carr, L., and McAlister, K., “Dynamic Stall Experiments on Oscillating Airfoils,” AIAA Journal, Vol. 14, No. 1, 1976, pp. 57–63. doi:10.2514/3.61332 [30] Dickinson, M. H., Lehmann, F.-O., and Sane, S. P., “Wing Rotation and the Aerodynamic Basis of Insect Flight,” Science, Vol. 284, No. 5422, 1999, pp. 1954–1960. doi:10.1126/science.284.5422.1954 [31] Tsai, H.-C., and Colonius, T., “Coriolis Effect on Dynamic Stall in a Vertical Axis Wind Turbine at Moderate Reynolds Number,” 32nd AIAA Applied Aerodynamics Conference, AIAA Paper 2014-3140, 2014. Z. Rusak Associate Editor
© Copyright 2026 Paperzz