Anda di halaman 1dari 12

Renewable Energy 133 (2019) 1066e1077

Contents lists available at ScienceDirect

Renewable Energy
journal homepage: www.elsevier.com/locate/renene

Three-dimensionality of the wake recovery behind a vertical axis


turbine
Pablo Ouro*, Stefan Runge, Qianyu Luo, Thorsten Stoesser
Hydro-environmental Research Centre, School of Engineering, Cardiff University, The Parade CF24 3AA, Cardiff, UK

a r t i c l e i n f o a b s t r a c t

Article history: The wake recovery downstream of a vertical axis turbine operating in a turbulent channel flow is
Received 1 July 2018 investigated via detailed velocity measurements using an Acoustic Doppler Velocimeter. Three distinct
Received in revised form wake regions are identified: (i) a near-wake region which extends until two rotor diameters (2D)
7 October 2018
downstream and characterised by a low-momentum area isolated from the ambient flow and the
Accepted 29 October 2018
presence of energetic dynamic stall vortices; (ii) a transition region (2D-5D), characterised by a fast
Available online 1 November 2018
momentum recovery, high levels of turbulence and vertical expansion of the wake; and (iii) a far-wake
region beyond 5D where the velocity recovers to approximately 95% of the free-stream velocity. Albeit
Keywords:
Vertical axis turbine
the wake deficit recovery is mostly accomplished at 5D behind the turbine, rotor-induced effects are still
Wake recovery present beyond 10D as indicated by high-order flow statistics, such as high velocity fluctuations and flow
Array skewness. The analysis of the streamwise momentum budget reveals that advection is the main
Acoustic Doppler velocimeter mechanism for momentum replenishment through most of the wake and turbulent transport terms play
Turbulence only a minor role. This study evidences the anisotropic nature of the turbulence and asymmetry of the
Renewable energy flow in horizontal, vertical and cross-sectional planes downstream of the vertical axis turbine.
© 2018 Elsevier Ltd. All rights reserved.

1. Introduction The flow regime experienced by VATs depends on the operating


Reynolds number attained by the rotating blades, which will
With a growing demand for renewable and predictable energy, determine the relevance of dynamic stall of the blades or the
hydrokinetic energy from fast-flowing water bodies, such as ocean mixing of the turbine's wake with the ambient flow. Bachant and
currents and river or canal streams, could play a significant role in Wosnik [2] tested a high-solidity VAT in a towing tank for different
future energy strategies [1]. The design of hydrokinetic turbines can operational Reynolds numbers and found that drag forces and
be differentiated between horizontal axis turbines (HATs) and power coefficients tend asymptotically to a constant value, i.e. in-
vertical axis turbines (VATs). To date, HATs are the most common dependent of the Reynolds number, for ReD ¼ UD=n > 1,106 or
choice in commercial tidal energy projects where they are believed Rec > 2,105 . The acting forces on the blades are a consequence of
to have reached a certain maturity in their development in terms of the flow separation on their suction side which is fully turbulent
efficient rotor design. This is to some extend due to their similar- once the former flow condition threshold is achieved. Using 2D-
ities to wind turbines and the lessons learned from over 30 years of Particle Image Velocimetry (PIV), Sima~o-Ferreira [3] observed how
research and commercial application. On the other hand, invest- flow separation on the blades varies depending on its rotational
ment in R&D of VATs and their consequent application in tidal speed. Similarly, Ouro and Stoesser [4] used high-fidelity numerical
energy projects is limited despite their unique advantages such as simulations to evidence the different nature of dynamic stall
good performance in shallow waters (due to their squared/rect- vortices, the effect of which becomes more prominent at low tip
angular projected area), omni-directionality, low rotational speed speed ratios.
and limited technical design requirements, which make them an The performance and hydrodynamic blade loadings of VATs is
attractive technology to a wide range of flows. influenced greatly by these dynamic stall vortices, they are gener-
ated at the upstream side of the rotor's swept area and are subse-
quently convected through the rotor's inner swept area and hence
they eventually interact with or impinge on the blades that on the
* Corresponding author.
E-mail address: OuroBarbaP@cardiff.ac.uk (P. Ouro). downstream half of the rotation. Brochier et al. [5] identified from

https://doi.org/10.1016/j.renene.2018.10.111
0960-1481/© 2018 Elsevier Ltd. All rights reserved.
P. Ouro et al. / Renewable Energy 133 (2019) 1066e1077 1067

water tank experiments how such blade-vortex interaction is characteristics appeared to vary depending on the dynamic solidity,
characterised by leading- and trailing-edge vortices whose shed- a parameter that correlates with the rotor's geometric solidity to
ding pattern varies with the rotor's rotational speed. Somoano et al. the tip speed ratio. Low dynamic solidity values indicated a strong
[6] observed a varying wake recovery pattern at different tip speed ability of the free-stream flow to penetrate through the rotor while
ratios as a consequence of the vortex dynamics happening within high values of this parameter caused a larger obstruction of the
the swept rotor area. With similar scope, Ouro and Stoesser [4] turbine rotor to the flow and thus enhancing bypass flows.
visualised in their large-eddy simulations that at low rotational Consequently, the larger amount of fluid flowing through the rotor
speeds the blades undergo deep dynamic stall causing a drop in the faster its wake recovers and vice versa.
performance, whilst at high rotational speed, above peak perfor- Apart from dynamic stall as a key characteristic in the wake of
mance, the blades lack sufficient lift-generation for the turbine to vertical axis turbines, other key features in the wake of VATs were
generate torque at the shaft. Therefore, an adequate balance be- found by Tescione et al. [7] who identified a skewed nature of the
tween maximum lift and the dynamic stall vortices is a key aspect wake based on the time-averaged velocity distribution far down-
in the design of efficient vertical axis turbines. stream of the device. Rolin et al. [13] employed a stereo-PIV tech-
Early work on the three-dimensional nature of the wake behind nique and observed two pairs of asymmetrical-counter-rotating
vertical axis turbines was undertaken by Tescione et al. [7] who vortices at the horizontal edges of the wake as a result of a variation
adopted PIV to measure different planes throughout the turbine's in cross-flow momentum. Analogously, Ryan et al. [14] observed
wake. Their phase-averaged results indicated the mechanism that the strength of this pair of vortices is directly related to the
responsible for the generation of tip vortices as a result of the flow turbine's rotational speed.
moving over the blade's tip and the low-pressure area on the blade The present research considers the hydrodynamics downstream
suction side. These large-scale flow structures kept coherent until a of a three-bladed Gorlov-type vertical axis turbine. This turbine has
distance of two turbine diameters downstream, over which a ver- undergone extensive laboratory testing to optimise its design and
tical expansion was also observed due to the action of tip vortices has further been deployed and tested in real-life environments at
moving towards the outside of the turbine swept area. Thereafter two different scales [15]. The turbine has gone through the various
these rotor-induced energetic structures merge with the low- stages of the so-called Technology Readiness Levels (TRLs) and is
momentum wake which starts to develop a vertical contraction. currently being tested in an array configuration in a water
In the mechanisms involved in the wake recovery in the near- conveyance canal in Colorado, USA. The main purpose of this study
wake, vertical advection of momentum appears to play a relevant is to gather insights into the flow and turbulence of the rotor wake
role as discussed in Bachant and Wosnik [8]. Kinzel et al. [9] per- and its recovery by analysing first-, second- and third-order flow
formed field measurements of different arrays of vertical axis wind statistics obtained through velocity measurements using an ADV at
turbines showing that vertical momentum transport is also of great various cross-sections between 0.75 and 14 diameters downstream
significance in the wake recovery. It was further observed that of the turbine's rotor.
rotor-induced turbulence aids in recovering wake momentum with
wake velocity values close to the free-stream velocity at six di-
ameters downstream, which suggests a closer cluster capability of 2. Experimental setup
VATs compared to HATs [10]. Furthermore, Tescione et al. [7] also
identified a circulation-induced phenomenon known as Magnus 2.1. Laboratory configuration
effect in the near-to-far wake transition which was also observed in
the LES of Ouro and Stoesser [11] being this more noticeable at The experiments were conducted in a recirculating hydraulic
higher tip speed ratios. flume at Cardiff University which is 17 m long, 1.2 m wide and 1 m
Large-scale coherent motion in the wake of VATs were analysed deep. The flow rate was controlled by an axial flow impeller with a
by Araya et al. [12] for different turbine rotors equipped with 2e5 constant value of Q ¼ 0:75m3 =s providing a flow depth of H ¼
blades using spectral analysis and proper orthogonal decomposi- 0:65m. A flow-straightener was installed at the inlet of the flume in
tion. Three regions in the wake were observed: (i) near-wake, order to smoothen the approaching flow. Fig. 1 depicts the flume's
dominated by periodic blade vortex shedding; (ii) transitional- dimensions with the deployed turbine (whose characteristics are
wake, where the shear-layer formed in between the turbine's explained in Section 2.3), and shows the position of the measured
low-velocity wake and ambient flow became unstable; and (iii) far- cross-sections each consisting of 11 vertical profiles with 10 points
wake, dominated by bluff body-like wake oscillations. The wake per profile, i.e. 110 locations, yielding a total of 1100 measurements.
A number of previous studies performed in this flume provided

Fig. 1. Location of the cross-sections behind the vertical axis turbine along the hydraulic flume and including velocity measurement points in the cross-section (top left insert).
1068 P. Ouro et al. / Renewable Energy 133 (2019) 1066e1077

insights into the hydrodynamics of Savonious [16] and Darrieus [17] numbers under these flow conditions were ReH ¼ 6:3,105 and Fr ¼
vertical axis turbines as well as horizontal axis turbines [18]. 0:38, respectively. Note that due to ADV measurement limitations
An Acoustic Doppler Velocimeter (ADV, Nortek Vectrino) was close to the channel bed, side-walls and water surface, flow ve-
used to measure velocities in x, y and z directions with an accuracy locities were measured between 0:60 < z=D < 0:65 in the vertical
of ±0:5% and were taken at a frequency of 200 Hz over a sampling direction and 1:55 < y=D < 1:55 in the spanwise direction.
period (tn ) of 120s. Results were post-processed and filtered by According to Fig. 2a, the mean streamwise velocity fluctuation
removing those samples whose correlation coefficient values were (su =U0 ) was approximately equal to 15% at the centre of the channel
inferior to 70% [19], and a signal-to-noise ratio (SNR) of > 15 dB was whilst transversal (sv =U0 ) and vertical (sw =U0 ) turbulence in-
considered. The time-averaged velocity Ui ¼ ðU; V; WÞ is calculated tensities were 12.3% and 9.2%, respectively. The resulting aniso-
from the measured velocity field ui ¼ ðu; v; wÞ as, tropic ratio su : sv : sw is 1.0:0.8:0.6 being slightly higher than
those found in other channel flows, e.g. 1.0:0.71:0.55 [20], which is
tðn
1 most likely caused by the high Reynolds number of the flow, the
Ui ¼ ui ðtÞdt (1) relatively narrow section of the channel and the flow asymmetry
tn  t0
t0 induced by the flume's impeller. Turbulence intensity values along
the whole cross-section (not shown here) were fairly consistent
Reynolds decomposition was applied to calculate the root-
with those found for the centreline shown in Fig. 2a.
mean-square (rms) of the velocity fluctuation u0i ¼ ðu0 ; v0 ; w0 Þ
resulting from subtracting to the mean velocity Ui the instanta-
neous velocity values ui , i.e. u0i ¼ Ui  ui . The turbulence intensities 2.3. Description of the VAT
sui ¼ ðsu ; sv ; sw Þ are the rms of the velocity fluctuation u0i , and read,
A 1:10 scale three-bladed Gorlov-type VAT was placed in the
 0:5 previously described flume at 10 m downstream of the inlet. Blades
sui ¼ u0i u0i (2)
and arms were made of laser-sintered PA 2200 material. Hub, shaft
The turbulent kinetic energy (k) was calculated from the normal and submerged bearings were made of stainless steel. The rotor
Reynolds stresses as, diameter (D) was 0.36 m with a height (Ht ) of 0.30 m which gave a
rotor aspect ratio (D=Ht ) equal to 0.85. The blades featured a chord

1 0 0
ðu u þ v0 v0 þ w0 w0 Þ (3) length (c) of 0.08 m which resulted in a rotor solidity of s ¼ 21%
2 which is calculated as,
Note that from hereinafter u0i refers to its time-average value,
Nb c
and all flow variables are normalised with the free-stream velocity s¼ (4)
pD
U0 and the rotor diameter D.
with Nb denoting number of blades. The Reynolds number based on
2.2. Free-stream flow conditions the diameter, ReD , was approx. 3:5,105 and that based on the chord
was Rec ¼ 7:8,104 . Due to commercial sensitivity of the rotor
Velocity measurements of the flow in the flume without turbine design, no further geometry details can be given. The rotor swept
(hereinafter free-stream flow conditions) were obtained at 336 cross-sectional area extended between 0:5D < y < 0:5D and
points at the proposed turbine position prior to its deployment. The 0:4D < z < 0:45D occupying 14% of the channel's cross-section.
approaching flow is depicted in Fig. 2 which presents vertical The rotor was attached to a 0.02 m diameter shaft which drove a
profiles of the mean streamwise velocity (U=U0 ) and turbulence permanent magnet generator that converts the shaft's rotational
intensities (sui =U0 ) in the centreline of the flume (y=D ¼ 0:0) and motion into electrical energy and Fig. 3a depicts the turbine as
contours of the streamwise velocity in a cross-section of the flume deployed in the flume. Adjustable resistors were used to apply a
without turbine. The mean cross-sectional U-velocity was found to resistance to the electrical circuit and dissipate energy in order to
be 0.97 m/s and this is considered the free-stream or bulk velocity operate the rotor at the desired rotational speed. A combined
U0 . The bulk Reynolds (ReH ¼ U0 H=n) and Froude (Fr ¼ U0 =ðgHÞ0:5 ) encoder and torque sensor (Futek TRS605) was placed between the

Fig. 2. Data of the free-stream flow: (a) vertical profiles of mean streamwise velocity (solid line and filled symbols) and root-mean-square of velocity fluctuations (dashed lines and
blank symbols); and (b) cross-section with contours of normalised time-averaged streamwise velocity.
P. Ouro et al. / Renewable Energy 133 (2019) 1066e1077 1069

Fig. 3. (a) Turbine rotor, generator and torque transducer assembled in the hydraulic flume at Cardiff University. (b) Measured power coefficient, Cp versus tip speed ratio, l, curve
with the point of maximum efficiency indicated by a red symbol. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this
article.)

generator and rotor in order to measure rotational speed and tor- 3.1. Horizontal plane at mid-height of the turbine rotor
que. The rotor was held in place by a bearing adhered to the flume's
bed, and another bearing that was attached to a bridge which Fig. 4 presents an overview of the downstream evolution of the
rested on the side-walls of the flume well above the water surface wake behind the turbine with the distribution of mean streamwise
(see Fig. 3a). velocity (U=U0 ), u-velocity fluctuation (su =U0 ), turbulent kinetic
The performance of the turbine was calculated from the applied energy (k=U 20 ) and primary Reynolds shear stress (  u’v’=U02 ) along
torque (T) and rotational speed (U) monitored over a period of 60 s, a horizontal plane at mid-height of the rotor. The wake exhibits a
starting from free-wheeling conditions (no torque applied/open notable asymmetry in the horizontal plane with a low-momentum
electrical circuit) and progressively incrementing the applied tor- region immediately downstream of the rotor featuring velocities
que (increased electrical resistance) until the rotor stalled, i.e. below 0.25 U0 which are generally found on the y=D < 0:0 side.
stopped rotating. The turbine efficiency was measured in terms of Considering the rotor spins counter-clockwise, this side corre-
power coefficient (Cp ) defined as, sponds to the half of the rotor's swept circumference in which the
blades travel in the same direction as the flow, i.e. the downstroke
TU half of the revolution [4]. During the blade's downstroke motion
Cp ¼ (5)
1 rAU 3 the flow separates at the blade's inner side and dynamic stall
2 0
vortices are shed and consequently convected downstream [4,5,21].
where r is the density of water and A ¼ Ht D is the rotor's swept These flow structures are responsible for the isolation of the low-
area. The performance curve is presented in Fig. 3b in which rota- momentum wake area from the unperturbed high-velocity flow.
tional speeds are normalised by the water velocity and represented In the vicinity of the rotor, contours of high su and k (Fig. 4b and
by the tip speed ratio l, defined as, c) are found along the iso-lines of U=U0 ¼ 0.50 and 0.75 from
behind the turbine's rotor until x=D ¼ 4  5 when the turbulence
UD levels in the wake start to decay. The energetic dynamic stall
l¼ (6) vortices shed from the turbine rotor's blades move downstream
2U0
following a path that overlaps those iso-lines [4,12] and therefore,
The highest efficiency was attained at l ¼ 1:9 with a peak Cp they significantly contribute to such highly-turbulent areas in the
value of 0.56. During the wake velocity measurements, the turbine outskirts of the turbine rotor's wake. This phenomenon generates a
operated at peak efficiency, which was achieved by keeping the shear-layer between the ambient flow and the low-velocity area as
rotational speed of the device constant by controlling the torque reflected in the large values of -u’v’ shown in Fig. 4d.
applied to the rotor. Flow streamlines drawn in Fig. 4d indicate that in the centreline
of the wake the approaching flow diverts due to its interaction with
the rotor and starts to converge again at x=D ¼ 2, which eventually
triggers turbulent fluctuations and enhances wake deficit recovery.
However, the latter bypass flow at y=D > 0:5 is stronger than on the
3. Wake measurements
other side of the wake, i.e. at y=D < 0:5, as on either side the relative
velocity between the ambient flow and the counter-rotating blades
The wake behind the operating turbine's rotor was measured at
is different. Note that at x=D ¼ 0:0, y=D ¼ 0:5 the blades are moving
several cross-sections located downstream of the rotor at distances
into the flow experiencing the maximum relative velocity, whereas
of 0:75D, 1:0D, 1:5D, 2:0D, 2:5D, 3:0D, 4:0D, 5:0D, 7:0D, 10:0D until
at x=D ¼ 0:0, y=D ¼ 0:5 the blades move with the flow, i.e.
14:0D. Based on these data, additionally to each of the measured
attaining the minimum relative velocity. The rotor's rotation is
cross-sections, two planes of interest are reconstructed: a hori-
responsible for the wake asymmetry in the horizontal plane.
zontal plane at z=D ¼ 0, i.e. at the rotors mid-height, and a longi-
Fig. 5 presents horizontal profiles of streamwise mean velocity
tudinal plane at y=D ¼ 0.
1070 P. Ouro et al. / Renewable Energy 133 (2019) 1066e1077

Fig. 4. Horizontal plane at z=D ¼ 0 with distribution of mean (a) streamwise velocity, (b) streamwise velocity fluctuation, (c) turbulent kinetic energy, and (d) Reynolds shear stress
-u’v’. Solid line indicates U ¼ 0:25U0 .

(a) and velocity fluctuation (b), Reynolds shear stress -u’v’ (c), and the near-wake (at x=D ¼ 1:5), a distinct low-turbulence region is
streamwise velocity gradient (d) at mid rotor height and at selected identified immediately downstream of the rotor which coincides
locations downstream of the turbine rotor. These plots allow to with a low-momentum region and is the result of the blades
quantify the asymmetry of the flow and its turbulence features. blocking the flow. The two peaks in the x=D ¼ 1:5 su - profile signify
Minimum values of U are found predominantly over the y=D < 0 the dynamic stall and/or trailing vortices, respectively, which are
side and the asymmetry is more visible closer to the turbine's rotor. being flushed out of the inner rotor area. In the mid-wake, two
Note that the bypass flow, i.e. at jy=Dj > 0:5, features streamwise distinct peaks located at y=D ¼ 0:0 and 0.5 coinciding with re-
velocities of approximately U=U0 ¼ 1:5 as a consequence of the gions of significant turbulence, as seen in Fig. 4, and these are the
turbine rotor's flow blockage. In accordance with the distribution of signatures of aforementioned vortices being advected downstream.
U in Fig. 4a, the velocity asymmetry reduces after x=D ¼ 7  10 The peaks diminish further away from the turbine's rotor sug-
with minima of U above 0:8U0 . gesting that the vortices weaken as the wake is being filled with
The profiles of su in Fig. 5b exhibit two distinct distributions. In ambient fluid. Between x=D ¼ 2 and 5, previously high levels of

Fig. 5. Horizonal profiles at rotor mid-height of (a) streamwise velocity, (b) streamwise velocity fluctuations, (c) shear stress -u’v’, and (d) horizontal gradient of mean velocity
vU=vy.
P. Ouro et al. / Renewable Energy 133 (2019) 1066e1077 1071

streamwise turbulence in the bypass flow reduce to those found in wake region is reached, i.e. x=D > 5.
free-stream conditions. From x=D ¼ 7 onwards, the profiles of su Fig. 7 presents profiles of U, su , -u’w’ and vU=vz at different
smooth-out across the channel width until a uniform value of cross-sections downstream of the rotor. The vertical profiles of U
su =U0 ¼ 0:12 is attained at x=D ¼ 14. exhibit a clear non-uniformity in the wake recovery behind the
The location of the largest momentum exchange between the turbine, especially in the locations closest to the turbine at x=D < 3.
turbine's wake and the free-stream flow is identified by maxima The different velocity magnitude between the flow overtopping the
and minima in the -u’v’ profiles presented in Fig. 5c. Positive peaks rotor and that coming from the bottom part of the channel induce
of the transversal shear stress are found at approximately y=D ¼ 0 an uneven vertical distribution of the streamwise velocity in the
while negative peaks are found at y=D ¼ 0:5 at most locations profiles along the mid-wake. Analogously to the horizontal profiles
except for x=D ¼ 2. Momentum exchange is a result of the in Fig. 5, streamwise momentum is almost recovered at x=D ¼ 10 as
streamwise velocity gradients and these are presented in Fig. 5d its distribution matches that at x=D ¼ 14 being both very close to
with spanwise profiles of vU=vy, whose distribution is similar to the free-stream profile from Fig. 2a.
-u’v’, except in the near-wake dominated by flow overtopping and The profile at x=D ¼ 1:5 intersects the low-momentum region
undergoing the turbine's rotor. Hence, at the location of the peak of and the vertical profile of su exhibits values in the range of
the velocity gradient at x=D ¼ 1:5 the spanwise shear is relatively 0.12e0.15, i.e. close to the free-stream turbulence intensity which
low. suggests that the low-momentum area also features very low tur-
bulence intensities. Downstream of the near-wake, the mixing
3.2. Longitudinal plane through the rotor axis between the turbine rotor's wake and ambient flow increases as
well as streamwise turbulence intensity reaches a peak of
Longitudinal planes through the rotor axis are presented in approximately 0.35 at x=D ¼ 2 and 3 in between  0:45 < z=D < 0:4.
Fig. 6 depicting contours of U=U0 , su =U0 , k=U 20 and  u0 w0 =U 20 . The Note that near the free-surface, at 1.5 and 2 diameters downstream
low-momentum area extends over the entire projected rotor swept of the turbine rotor, su is triggered as a result of the turbulent
area, i.e. 0:4 < z=D < 0:45 and vanishes after x=D z 1:8. A certain downwash, i.e. flow coming from the top of the rotor. After x=D ¼ 3
asymmetry of the wake is observed in the vertical direction, there is a progressive decay of su that features the highest values in
evident from the iso-line of U=U0 ¼ 0:25 which is skewed towards the region  0:2 < z=D < 0:4. In the far-wake region, su exhibits a
the bottom half of the channel and extending until x=Dz2. This is similar value to those found in free-stream flow conditions.
caused by the downwash of fluid due to the flow overtopping the The mid-wake exhibits a notably irregular distribution of
rotor depicted from the flow streamlines in Fig. 6b, which aids in -u’w’ along the water depth as a result of the complex interaction
recovering wake momentum mainly above z=D ¼ 0. An upwards between the turbine's wake with the ambient flow, whilst in the
motion arises from the gap between the rotor's bottom tips and the far-wake the distribution of this shear stress is found to be more
channel bed which enhances the wake recovery for z=D < 0, uniform with negligible values beyond x=D ¼ 10. In Fig. 7d, the
although less intense than the overtopping flow. The difference in vertical gradient of mean streamwise velocity is plotted and its
wake recovery rate above and beneath the turbine rotor is similar distribution is somewhat similar to that of -u’w’.
to the behaviour of the wake behind vertical axis wind turbines The Boussinesq approximation for isotropic turbulent flow
operating in a boundary layer flow [13]. states that Reynolds shear stresses are proportional to the velocity
In the transitional-wake region, i.e. 2 < x=D < 5, the interaction gradients, i.e. u0 v0 ¼ nt vU
vy and u w ¼ nt vz where nt denotes eddy
0 0 vU

between the fast-moving water above the rotor and the low- viscosity. In the present flow, such an approximation would not be
momentum wake area triggers high levels of turbulence as depic- valid in the near- and mid-wake regions as the profiles of shear
ted in Fig. 6b and c. Contours of the Reynolds shear stresses u’w’ stress and velocity gradients do not show peaks at the same loca-
=U02 are plotted in Fig. 6d showing an asymmetric distribution with tion suggesting that the eddy viscosity would have to vary quite
significantly over the depth.
respect to z=D ¼ 0. It can be seen that this shear stress converges to
a fairly uniform distribution over the water depth once the far-

Fig. 6. Vertical plane at y=D ¼ 0 with distribution of (a) streamwise velocity, (b) streamwise velocity fluctuation, (c) turbulent kinetic energy and (d) Reynolds shear stress  u0 w0 .
Solid line indicates U ¼ 0:25 U0 .
1072 P. Ouro et al. / Renewable Energy 133 (2019) 1066e1077

Fig. 7. Vertical profiles at flume centre of (a) streamwise velocity, (b) streamwise velocity fluctuations (c) Reynolds shear stress -u’w’, and (d) vertical gradient of streamwise
velocity, vU=vz.

3.3. Cross-sectional planes At the section closest to the turbine (x=D ¼ 1), the region of U <
0:25U0 (solid lines) is predominantly found for y=D < 0. Velocity
The three-dimensional nature of the turbine wake is further vectors illustrate that in all considered sections the ambient flow
investigated in Fig. 8 showing the distribution of streamwise ve- entrains into the wake region more pronouncedly for y=D > 0 with
locity, streamwise velocity fluctuation and shear stresses in four large negative vertical velocities as a result of the downwash. At x=
cross-sectional planes (i.e. at x=D ¼ 1, 2, 3 and 5) normal to the flow D ¼ 1, the regions of high turbulence levels coincide with the
direction. projected rotor's swept perimeter which is in agreement with the

Fig. 8. Distribution of normalised streamwise velocity (U=U0 ), turbulence intensity (su =U0 ) and Reynolds shear stresses (-u’v’ and -u’w’) along four cross-sections downstream of
the turbine at x=D ¼ 1, 2, 3 and 5. Solid line indicates U ¼ 0:25U0 .
P. Ouro et al. / Renewable Energy 133 (2019) 1066e1077 1073

path described by the dynamic stall vortices shown in Fig. 4. The significant energy peaks at a frequency of 3fb are discerned. This
largest values of shear stresses are found along y=D ¼ 0:6 and 0.4 indicates that dynamic stall vortices are no longer coherent at this
for -u’v’ and z=D ¼ 0:35 and 0.55 for -u’w’. The peaks of -u’w’ stage of the wake which allows a larger entrainment of the bypass
almost coincide with the projected location of the blade's tips as flow into the low-velocity area of the wake.
they are slightly shifted upwards.
At two diameters downstream of the turbine, the low-
momentum region expands over a larger vertical extension than 3.5. Streamwise momentum recovery
the turbine's height and velocity minima surpass values above
0:25U0 . Velocity vectors evidence the inwards interaction of the The contributors to the replenishment of the momentum deficit
bypass flow into the low-velocity area generating a C-shaped dis- behind the turbine as a result of its interaction with the main flow
tribution of the wake. The largest values of su are observed be- are accounted via the Reynolds-Averaged Navier-Stokes (RANS)
tween the free-surface and top of the rotor as a result of the equation in the x-direction, which reads,
overtopping flow, and also at y=D z  0:5 and 0.0 where the 0 1
change from low to high streamwise velocities is relatively fast, as B C
depicted in the vU=vy profile in Fig. 5d. The highest -u’v’ and -u’w’ B 2 C
vU 1B vU vU vu’ 2 vu’v’ vu’w’ 1 vP v U C
values are located within the rotor's swept area as a result of the ¼ BV W     þn 2 C
vx UB vy vz vx vy vz r vx
B|fflfflffl{zfflfflffl}|fflfflfflffl{zfflfflfflffl} |fflfflffl{zfflfflffl} |fflfflffl{zfflfflffl} |fflfflfflffl{zfflfflfflffl} |fflfflffl{zfflfflffl} vx i C
C
larger entrainment of high-momentum flow from the ambient flow @ |fflfflfflffl{zfflfflfflffl} A
II III V
into the low-momentum wake. At x=D ¼ 3 and 4 (the latter not I IV VI VII
shown here), a similar distribution of U, su and -u’v’ to that at x=D ¼
2 is found as they belong to the transitional-wake region in which (7)
the wake maintains a similar pattern. Further downstream, at In the present analysis, the flow is deemed stationary in its time-
x=D > 5, the dynamics of the wake changes as the momentum is averaged characteristic, the streamwise derivative of the Reynolds
recovered and the C-shaped low-velocity region is no longer stress u0u0 (term III) is discarded due to the relatively large spacing
observed. The magnitudes of the Reynolds stresses are also reduced between cross-sections, the pressure term V is omitted as this
whilst their overall distribution remains similar to that found in the variable was not measured using ADV, and the last term VII (rep-
cross-sections at x=D ¼ 2 and 3. resenting the viscous diffusion) is also omitted due to the high-
turbulent nature of the flow and thus viscous effects are deemed
3.4. Power spectra negligible [8]. In Eq. (7) the momentum transport terms I, II, IV and
V denote y-advection, z-advection, y-turbulent transport and z-
The transition from near-wake to transitional-wake is marked turbulent transport, respectively, and these are integrated over
by the role of the dynamic stall vortices in the momentum recovery. every measured cross-section and normalised by D=U0 as pre-
The presence of these large-scale turbulent structures are identified sented in Fig. 10. In most of the cross-sections, y- and z-advection
in Fig. 9 via Power Spectral Density (PSD) of u’-velocity time-series representing secondary flow action are the dominant terms in
presented. In the calculation of the PSDs, each signal was decom- replenishing momentum into the wake whilst the turbulent
posed into 10 overlapping segments spanning 60s each and a Hann transport terms IV and V attain values between one to two orders of
window was applied in order to obtain each of their trans- magnitudes lower in most locations.
formations. The resulting spectra are the average of these Comparing the relative contribution between the advection
segments. terms at x=D ¼ 1, vertical advection has a greater value over the
A total of 12 PSDs are computed from points vertically aligned transversal advection suggesting that in the near-wake overtopping
with the top, centre and bottom of the wake at different locations and underflow of the turbine rotor are mainly responsible for filling
within the near-wake indicated in Fig. 6a. At distances of the turbine wake, which is in-line with the findings of Bachant and
x=D ¼ 0:75 and 1.0, the spectra show a clear energetic peak at a Wosnik [8]. At x=D ¼ 2, the lateral flow entrainment (i.e. y-
frequency of 3fb , with fb standing for the blade passing frequency, advection) reaches a peak approximately three times greater than
revealing the presence of dynamic stall vortices shed by the rotor the z-advection term. Within the region between x=D ¼ 3  5, the
blades. At x=D ¼ 1:5, the energy peaks become more diffused as a momentum deficit recovers and this is evidenced with a decrease
result of the energetic vortices losing coherence during their con- in the magnitudes of spanwise advection reaching similar values to
vection downstream while interacting with the turbulent ambient vertical advection. In the far-wake, the values of the advection
flow. In the onset of the mid-wake region, at x=D ¼ 2:0, no terms diminish whilst the turbulent transport terms seem to

Fig. 9. Power spectral density, F, of u’ -velocity obtained at three vertical locations at four locations downstream of the turbine.
1074 P. Ouro et al. / Renewable Energy 133 (2019) 1066e1077

PN 0 3
n¼1 ui ðtn Þ
Skui ¼ (8)
N s3ui

Fig. 11 presents contours of the streamwise (Sku ) and spanwise


(Skv ) skewness together with iso-lines of the streamwise velocity at
four different locations in the transitional- and far-wakes, namely
x=D ¼ 3, 4, 7 and 10. At x=D ¼ 3, areas of large positive Sku correlate
with low streamwise velocities and vice versa, whilst areas of
approximately zero skewness are found where the spanwise shear
stress, -u’v’, is maximum. Velocity arrows in Fig. 8c indicate the
inwards motion of the free-stream flow into the low-momentum
region at y=D > 0 resulting in negative values of Skv . The distribu-
tion of the skewness of u- and v-velocities at x=D ¼ 4 is similar to
Fig. 10. Cross-sectional average values of the streamwise momentum equation con-
those observed at x=D ¼ 3 although Skv features larger values at the
tributors along the locations measured. latter location due to a more pronounced interaction between the
turbine's wake and the free-stream flow. The pattern of Sku and Skv
at x=D ¼ 7 and 10 are comparable although with smaller values to
increase at x=D ¼ 7 and 10. those found at x=D ¼ 3 or 4, showing that rotor-induced effects in
These results indicate that advection is the key ingredient to the flow field remain present until 10D downstream of the device.
recover momentum in the wake close to the rotor vertical, while in Skewness contours at 14D (not shown here for brevity) show a
the transition from near-to mid-wake spanwise advection domi- similar distribution to those found in free-stream conditions.
nates. Along the mid-wake and over the far-wake, both advection The velocity fluctuations at x=D ¼ 3 are further analysed at three
terms have a similar net contribution, which results from the large selected locations P1, P2 and P3 (indicated in Fig. 11a and e) which
mixing of the rotor-induced wake with the ambient flow. belong to regions with distinct velocity skewness. Probability
Density Functions (PDFs) of u’ and v’ at those locations are shown in
Fig. 12aec. The PDFs at P1 indicate a reversed skewed distribution
of u’ compared to v’ being the median of u’ positive while that of v’
negative, albeit both of similar magnitude. The PDF of u’ features a
3.6. Skewness of the turbine's wake longer tail along the negative values until u’=su <  4 whilst in the
positive region of Sku it reaches almost zero values for u’=su > 2.
Skewness, or third-order moment, indicates the asymmetry of Such a distribution indicate there are fewer turbulent events on the
the probability density function of velocities. Chamorro et al. [22] negative side but these have a larger turbulence intensity while
outlined the importance of flow skewness together with mean events triggering positive u’ happen more frequently and with
velocity and turbulence intensity to determine the available power lower intensity. At the centre of the wake, i.e. at P2, both PDFs
that can be harnessed by wind turbines. Therefore, it is of particular exhibit an almost Gaussian distribution. On the other side of the
interest to quantify it in the wake downstream of the vertical axis wake, at P3, the PDFs of u’ and v’ are biased to the positive side with
turbine as it can provide further insights into high-order wake ef- a similar positive median which is again in good correlation with
fects towards designing future arrays of turbines. The skewness of contours of skewness (Fig. 11a and e) as these are both of negative
the i component of velocity fluctuations Skui is defined as: and of similar magnitude at this point.

Fig. 11. Contours of (aed) Sku and (eeh) Skv at different cross-sections behind the turbine.
P. Ouro et al. / Renewable Energy 133 (2019) 1066e1077 1075

Fig. 12. Probability density functions (top row) and quadrant analysis (bottom row) of velocity fluctuations at locations P1, P2 and P3.

These observations on the asymmetric distribution of velocity 3.7. Wake characteristics and recovery
fluctuations are complemented with Fig. 12def which plot nor-
malised values of velocity fluctuations in streamwise and spanwise The recovery of the flow along the channel centreline is pre-
directions. The quadrant analysis at the three reference points show sented in Fig. 13, in terms of the three components of time-
that the majority of the velocity fluctuation values reside within the averaged velocity (Ui ) and velocity fluctuation (sui ), the velocity
quadrants I, i.e. u’>0, v’>0, and III, i.e. u’<0, v’<0. In the centre of the deficit in the streamwise direction (DU ¼ ðU  U0 Þ) and the three
wake (P2), the symmetry in both PDFs agree with the quadrant Reynolds shear stresses -u’v’, -u’w’ and -v’w’.
analysis whose centre of mass is located at the origin of co- Fig. 13 shows that the lowest streamwise velocities are found in
ordinates. The predominant distribution of velocity fluctuations at the near-wake (n-w) with the minimum velocity Umin approxi-
quadrants I and III suggests that outward and inward interactions mately equal to 0:2U0 (which equates to a velocity deficit of
are predominant which underpins the observed anisotropic nature DU=U0 ¼ 0:8) located in the transition from near-to mid-wake (m-
of turbulence in the wake of a VAT. w) at about 2D downstream of the rotor's shaft. Based on wake
measurements behind multiple VATs, Araya et al. [12] proposed
that the location of the near-wake to mid-wake transition, xt , can
be approximated by,

Fig. 13. Distribution along the wake centreline of (a) normalised time-averaged velocity components, (b) turbulence intensities, (c) measured velocity deficit and exponential fit,
and (d) Reynolds shear stresses. n-w, m-w and f  w denote near-, mid- and far-wake, respectively.
1076 P. Ouro et al. / Renewable Energy 133 (2019) 1066e1077

inwards motion of the ambient flow entraining into the wake


xt =D ¼ 4:78  4:93sD (9) generating an asymmetric distribution of the velocity field; and (iii)
a far-wake that begins at approximately 5 diameters downstream
where sD is the dynamic solidity, a parameter that links the design of the rotor axis with an almost full recovery of the momentum
solidity of the turbine rotor (s, Eq. (4)) with its actual tip speed ratio deficit as streamwise velocities reach values above 0.8 times the
l, and is calculated as, bulk velocity, whilst high-order effects, such as elevated levels of
flow skewness in comparison to those in free-stream conditions,
1
sD ¼ 1  (10) persist until a distance of 14 rotor diameters downstream of the
2psl
turbine rotor's axis location.
For the current study sD is equal to 0.60 and thus, applying Eq. The contours of flow skewness and the probability distribution
(9), the value of xt is equal to 1:82D which is in good agreement function at selected locations evidenced the effect of the rotating
with the present observations. Considering the Reynolds number of rotor in producing an asymmetric wake in terms of flow velocities
this flow is almost one order of magnitude above that of Araya et al. and turbulence. The quadrant analysis of the streamwise and
[12]'s experiments, this match suggest that Eq. (9) provides spanwise velocity fluctuations suggested strong inwards and out-
reasonable estimate of the transition from near-to mid-wake for wards interactions again the result of flow asymmetry. The inte-
VATs and not very sensitive to the Reynolds number of the flow. gration of the streamwise Reynolds-averaged momentum equation
Second-order statistics reach their minima in the near-wake as quantified that transversal and vertical advection terms are main
shown in Fig. 13b and d. In the onset of the mid-wake, the contributors to momentum replenishment. In the near-wake, ver-
entrainment of external turbulent flow into the wake increases tical advection is the most relevant due to significant flow over-
momentum but also triggers turbulence and thus velocity fluctu- topping and undergoing the rotor, while spanwise advection is the
ations attain their maxima. In other words, there is a fast transition main contributor to momentum replenishment in the transitional-
from a quasi-steady region shortly behind the turbine (n-w) to a wake. Despite the complex mechanisms involved in the wake re-
stage of fast-rate of momentum recovery and high-turbulence covery, the velocity deficit recovery can be approximated with
levels (m-w). Spanwise and vertical velocities feature their largest reasonable accuracy with an empirical exponential distribution.
values in the mid-wake reaching a maximum of V z  0:25U0 and
W z 0:10U0 . Acknowledgements
A more subtle transition of flow quantities is observed from
mid-to far-wake (f-w). Here, the streamwise velocity features a This research was jointly funded by EMRGY Inc, Atlanta, GA and
value of U ¼ 0:8U0 and achieves a magnitude close to the free- Cardiff University.
stream velocity at x=D ¼ 10. Vertical and transversal velocities as
well as Reynolds stresses exhibit a progressive decay along the far-
References
wake reaching their minima at x=D ¼ 14 and at x=D ¼ 10 for the
shear stresses. [1] M. Esteban, D. Leary, Current developments and future prospects of offshore
Albeit the complexity of the wake recovery pattern, the nor- wind and ocean energy, Appl. Energy 90 (1) (2012) 128e136, https://doi.org/
malised streamwise velocity deficit (DU=U0 ) along the wake cen- 10.1016/j.apenergy.2011.06.011.
[2] P. Bachant, M. Wosnik, Effects of Reynolds number on the energy conversion
treline is in agreement with the following distribution: and near-wake dynamics of a high solidity vertical-Axis cross-flow turbine,
Energies 9 (2) (2016) 73, https://doi.org/10.3390/en9020073.
DU  x b [3] C. Sima~o Ferreira, G. Van Kuik, G. van Bussel, F. Scarano, Visualization by PIV of
¼a  Umin (11) dynamic stall on a vertical axis wind turbine, Exp. Fluid. 46 (1) (2009) 97e108,
U0 D https://doi.org/10.1007/s00348-008-0543-z.
[4] P. Ouro, T. Stoesser, An immersed boundary-based large-eddy simulation
in which a and b are 1.168 and 0.684 according to Araya et al. [12] approach to predict the performance of vertical axis tidal turbines, Comput.
who proposed such exponential decay of the velocity deficit for Fluids 152 (2017) 74e87, https://doi.org/10.1016/j.compfluid.2017.04.003.
[5] G. Brochier, P. Fraunie, C. Beguier, I. Paraschivoiu, Water channel experiments
VAT wakes. Note that Eq. (11) enables to estimate the wake deficit of dynamic stall on Darrieus wind turbine blades, J. Propul. Power 2 (5) (1986)
recovery by just knowing Umin which is found in the near-wake. 445e449, https://doi.org/10.2514/3.22927.
Fig. 13c shows that the velocity deficit attains its maximum value [6] M. Somoano, F. Huera-Huarte, Flow dynamics inside the rotor of a three
straight bladed cross-flow turbine, Appl. Ocean Res. 69 (2017) 138e147,
before the transition point xt , and beyond this location a good https://doi.org/10.1016/j.apor.2017.10.007.
match between the measured velocity deficit and the exponential [7] G. Tescione, D. Ragni, C. He, C.J. Sim~ ao Ferreira, G. van Bussel, Near wake flow
fit is observed. analysis of a vertical axis wind turbine by stereoscopic particle image veloc-
imetry, Renew. Energy 70 (2014) 47e61, https://doi.org/10.1016/
j.renene.2014.02.042.
4. Conclusions [8] P. Bachant, M. Wosnik, Performance measurements of cylindrical- and
spherical-helical cross-flow marine hydrokinetic turbines, with estimates of
exergy efficiency, Renew. Energy 74 (2015) 318e325, https://doi.org/10.1016/
The wake generated behind a vertical axis turbine operating at j.renene.2014.07.049.
its peak efficiency has been characterised based on an extensive [9] M. Kinzel, Q. Mulligan, J.O. Dabiri, Energy exchange in an array of vertical-axis
wind turbines, J. Turbul. 14 (6) (2012) N38, https://doi.org/10.1080/
experimental carried out in Cardiff University's hydraulics labora- 14685248.2012.712698.
tory. Using the method of Acoustic Doppler Velocimetry, the three- [10] P. Ouro, T. Stoesser, Impact of environmental turbulence on the performance
dimensional nature of the flow and turbulence characteristics of and loadings of a tidal stream turbine, flow, Turbul. Combust. https://dx.doi.
org/10.1007/s10494-018-9975-6.
the turbine rotor's wake have been revealed and quantified. [11] P. Ouro, T. Stoesser, Wake generated downstream of a vertical Axis tidal
The wake can be divided into three distinct regions: (i) a near- turbine, in: 12th Eur. Wave Tidal Energy Conf., Cork, Ireland, 2017.
wake region featuring a low-momentum area just downstream of [12] D.B. Araya, T. Colonius, J.O. Dabiri, Transition to bluff-body dynamics in the
wake of vertical-axis wind turbines, J. Fluid Mech. 813 (2017) 346e381,
the turbine rotor in which dynamic stall vortices play a key role and
https://doi.org/10.1017/jfm.2016.862.
limit the entrainment of ambient flow into the low-velocity region; [13] V.F.C. Rolin, F. Porte-Agel, Experimental investigation of vertical-axis wind-
(ii) a transitional-wake region, extending from 2 to 5 diameters turbine wakes in boundary layer flow, Renew. Energy 118 (2018) 1e13,
downstream of the turbine rotor that is characterised by a fast-rate https://doi.org/10.1016/j.renene.2017.10.105.
[14] K.J. Ryan, F. Coletti, C.J. Elkins, J.O. Dabiri, J.K. Eaton, Three-dimensional flow
of momentum recovery and high intensity of turbulence and, in field around and downstream of a subscale model rotating vertical axis wind
addition, a C-shaped distribution of the low-velocities due to the turbine, Exp. Fluid. 57 (38) (2016) 1e15, https://doi.org/10.1007/s00348-016-
P. Ouro et al. / Renewable Energy 133 (2019) 1066e1077 1077

2122-z. https://doi.org/10.1016/j.jfluidstructs.2017.03.009.
[15] S. Runge, T. Stoesser, E. Morris, M. White, Technology readiness of a vertical- [19] L. Cea, J. Puertas, L. Pena, Velocity measurements on highly turbulent free
Axis hydro-kinetic turbine, J. Power Energy Eng. 6 (2018) 63e85, https:// surface flow using, ADV, Exp. Fluid. 42 (3) (2007) 333e348, https://doi.org/
doi.org/10.4236/jpee.2018.68004. 10.1007/s00348-006-0237-3.
[16] T. Harries, A. Kwan, J. Brammer, R. Falconer, Physical testing of performance [20] I. Nezu, H. Nakagawa, Turbulence in Open-channel Flows, A. A. Balkema,
characteristics of a novel drag-driven vertical axis tidal stream turbine; with Rotterdam, The Netherlands, 1993.
comparisons to a conventional Savonius, Int. J. Mar. Energy 14 (2016) [21] D.B. Araya, J.O. Dabiri, A comparison of wake measurements in motor-driven
215e228, https://doi.org/10.1016/j.ijome.2016.01.008. and flow-driven turbine experiments, Exp. Fluid. 56 (2015) 150, https://
[17] L. Priegue, T. Stoesser, The influence of blade roughness on the performance of doi.org/10.1007/s00348-015-2022-7.
a vertical axis tidal turbine, Int. J. Mar. Energy 17 (2017) 136e146, https:// [22] L. Chamorro, M. Guala, On the evolution of turbulent scales in the wake of a
doi.org/10.1016/j.ijome.2017.01.009. wind turbine model, J. Turbul. 13 (27) (2012) 1e13, https://doi.org/10.1080/
[18] P. Ouro, M. Harrold, T. Stoesser, P. Bromley, Hydrodynamic loadings on a 14685248.2012.697169.
horizontal axis tidal turbine prototype, J. Fluid Struct. 71 (2017) 78e95,

Anda mungkin juga menyukai