Vous êtes sur la page 1sur 13

International Journal of Heat and Fluid Flow 37 (2012) 123–135

Contents lists available at SciVerse ScienceDirect

International Journal of Heat and Fluid Flow


journal homepage: www.elsevier.com/locate/ijhff

Effects of 3-D channel blockage and turbulent wake mixing on the limit
of power extraction by tidal turbines
Takafumi Nishino ⇑, Richard H.J. Willden
Department of Engineering Science, University of Oxford, Oxford OX1 3PJ, UK

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

Article history: Three-dimensional incompressible Reynolds-averaged Navier–Stokes (RANS) computations are per-
Received 14 October 2011 formed for water flow past an actuator disk model (representing a tidal turbine) placed in a rectangular
Received in revised form 9 May 2012 channel of various blockages and aspect ratios. The study focuses on the effects of turbulent mixing
Accepted 11 May 2012
behind the disk, as well as on the effects of channel blockage and aspect ratio on the prediction of the
Available online 6 July 2012
hydrodynamic limit of power extraction. To qualitatively account for the effect of turbulence generated
by the turbine (rather than by the shear flow behind the turbine), we propose a new approach, called a
Keywords:
blade-induced turbulence model, which does not use any additional model coefficients other than those
Actuator disk
Blade-induced turbulence
used in the original RANS turbulence model. Results demonstrate that the power removed from the mean
Channel blockage flow by the disk increases as the strength of turbulent mixing behind the disk increases, being consistent
Tidal power with the turbulent shear stress on the interface between the bypass and core flow passages acting in such
a way as to decelerate the bypass flow and accelerate the core flow. The channel aspect ratio also affects
the flow downstream of the disk but has less influence upstream of the disk; hence its effect on the limit
of power extraction is relatively minor compared to that of the channel blockage, which is shown to be
significant but satisfactorily estimated using one-dimensional inviscid theory previously reported in the
literature.
Ó 2012 Elsevier Inc. All rights reserved.

1. Introduction the limit of power extraction from unconstrained flow (i.e., the
Betz limit) is derived using the conservation of mass and
The development of clean and sustainable energy technologies momentum of an inviscid incompressible flow. Recently, similar
is amongst the most pressing issues facing the world today. one-dimensional (1-D) theoretical analyses have been conducted
Tidal-stream power generation is—albeit its basic concept has been for tidal power applications (Garrett and Cummins, 2007; Houlsby
well known for many years—one of the less-developed technolo- et al., 2008; Whelan et al., 2009), where the constraint or blockage
gies that are receiving increasing interest. A key initiative towards of flow passage due to the proximity of the sea surface and seabed
the re-examination of tidal power generation in recent years is the has a major influence on the limit of power extraction. These
assessment of global and regional tidal energy resources: in the UK, extended theories can be used in estimating the maximum extract-
for example, such an assessment has been reported by the Depart- able power from a given constrained flow. However, it is still
ment of Trade and Industry (DTI, 2004; later updated by BERR unclear whether these theoretical 1-D results will produce a
(2008)). However, there is still a significant level of uncertainty physically accurate description of power extraction. Of particular
in the estimation of the maximum amount of power that can interest to the tidal energy community are the effects of aniso-
actually be extracted from those tidal energy resources. This makes tropic or three-dimensional (3-D) blockage of flow passage and
it difficult to benchmark, and subsequently to optimize, the also of viscous/turbulent wake mixing; these effects are difficult
performance of given tidal power generation systems. to study purely theoretically.
The (fluid-dynamic) limit of power extraction from fluid flow With the rapid progress of computer technology in recent years,
can be theoretically derived using the so-called Linear Momentum Computational Fluid Dynamics (CFD) is becoming a powerful
Actuator Disk Theory (LMADT), which was first introduced inde- investigation tool in wind/tidal engineering as well as in other
pendently by Lanchester, Betz and Joukowsky in the 1910s to engineering disciplines. The Reynolds-averaged Navier–Stokes
1920s (Burton et al., 2001; see also van Kuik, 2007) mainly for (RANS) approach has been used in many previous CFD studies on
wind power and rotorcraft applications. In the original LMADT, wind/tidal turbines since directly resolving high Reynolds number
turbulence requires huge computational cost; meanwhile, there
⇑ Corresponding author. Tel.: +44 (0)1865 273129; fax: +44 (0)1865 283301. exist several possible approaches on how to simulate the effect
E-mail address: takafumi.nishino@eng.ox.ac.uk (T. Nishino). of turbines on the wind/tidal flow. The principal approaches

0142-727X/$ - see front matter Ó 2012 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.ijheatfluidflow.2012.05.002
124 T. Nishino, R.H.J. Willden / International Journal of Heat and Fluid Flow 37 (2012) 123–135

include the use of actuator disk models, actuator line models, where the Reynolds stress tensor, ui uj , is modelled using the stan-
blade-element momentum (BEM) models and fully blade-resolved dard k–e model (Launder and Spalding, 1974):
computations; a recent review of these models/approaches has  
Dk @ mt @k
been given by Sørensen (2011). The selection of the most appropri- ¼ þ Pk  e ð2Þ
ate approach depends on the aim of each study; for example, a
Dt @xj rk @xj
 
BEM model is suitable for the design/optimization of turbine De @ mt @ e e e2
blades, whereas an actuator line model is suitable for investigating ¼ þ C e1 Pk  C e2 ð3Þ
Dt @xj re @xj k k
the time-dependent dynamics of turbine wakes and tip vortices.
The virtues of actuator disk models are their generality, simplic- @U i
Pk ¼ ui uj ð4Þ
ity and comparability to LMADT; these features are all desirable @xj
particularly when the limit of generic power extraction (rather than  
@U i @U j 2
the actual aero/hydrodynamic performance of a specific turbine) is ui uj ¼ mt þ  kdij ð5Þ
@xj @xi 3
of interest. Several CFD studies using a coupled RANS-actuator disk
model have been reported so far on tidal power generation: Sun 2
k
et al. (2008) investigated the effect of a nearby free surface on the mt ¼ C l ð6Þ
wake profile of an actuator disk placed in an open channel. Gant e
and Stallard (2008) performed unsteady simulations with various where k is the turbulent kinetic energy, e is the dissipation of k, Pk is
turbulence conditions specified at the inlet of their channel. the production of k and mt is the kinematic eddy viscosity. The mod-
Harrison et al. (2010) compared CFD results with their water el coefficients used are: Cl = 0.09, Ce1 = 1.44, Ce2 = 1.92, rk = 1.0 and
channel experiments, where disks of varying porosity were used re = 1.3 (following Launder and Spalding’s standard k–e model).
to change the level of thrust. They showed that the wake profile In order to take account of the effect of blade-induced turbu-
predicted by their RANS-actuator disk model agreed well with that lence, additional sources of k and e are added to the right-hand-
measured in their water channel except for the region just down- side of Eqs. (2) and (3), respectively, but only at the location of
stream of the disk (i.e., near wake region). More recently, Belloni the actuator disk. This will be described later in Section 2.3.
and Willden (2011) conducted simulations of bare, ducted and
open-center disks to compare their performance. They also com- 2.2. Actuator disk model
pared their results with LMADT to show that, for a bare disk placed
in a closed channel with relatively low blockage, the 1-D inviscid In the actuator disk model used in this study, the turbine is
theory agrees fairly well with 3-D RANS-actuator disk results. One modelled as a stationary permeable disk of zero thickness, which
important point to be noted here is that all these previous studies is placed perpendicular to the incoming flow. The effect of the tur-
have been performed for only one or a few specific channel geom- bine/disk on the (Reynolds-averaged, or mean) flow is considered
etries; a more systematic and thorough investigation is necessary to as a loss of momentum in the streamwise (x) direction at the disk
fully understand the effects of 3-D channel blockage. plane. The change in momentum flux (per unit disk-area) is locally
Another important issue that has not been considered in the calculated as
previous RANS-actuator disk studies for tidal turbines is the effect  
1
of turbulence generated by the turbine itself. The importance of Mx ¼ K qU 2d ð7Þ
2
this turbine-blade-induced turbulence on the flow characteristics
in the near wake region has recently been discussed for wind tur- where q is the fluid density, Ud is the local streamwise velocity at
bines (El Kasmi and Masson, 2008; Réthoré et al., 2009) and also for the disk plane and K is a momentum loss factor (which is a param-
tidal turbines (Turnock et al., 2011; Batten et al., 2011); however, eter to determine the load or thrust acting on the disk). For the pres-
its effect on the limit of power extraction is still unclear. As will ent study, a momentum source (over the disk, per unit area)
be described in Section 2.3, it is not a trivial issue to model the ef- SU = Mx/q is added to the incompressible RANS equation for the
fect of blade-induced turbulence in computations where turbine streamwise direction. The thrust acting on the disk is calculated as
blades (and turbulent boundary layers on them) are not resolved; Z  Z 
1
further studies are required on this issue. Thrust ¼ Mx dA ¼ K q U 2d dA ð8Þ
2
The objectives of the present study are to elucidate the effects of
(i) anisotropic or 3-D channel blockage and (ii) turbulent mixing in where the integration is taken over the disk area, A, and the thrust
the near wake region on the hydrodynamic limit of power extrac- coefficient is defined as
tion, by using a modified RANS-actuator disk model. A new ap-
proach is proposed and used to take account of the effect of Thrust hU 2d i
CT ¼ 1 ¼ K ð9Þ
blade-induced turbulence in a generic sense (rather than for a spe- 2
qU 2in A U 2in
cific turbine) in the framework of RANS-actuator disk computa-
where Uin is the inlet (far-upstream) velocity and h/i indicates the
tions. Results are presented for rectangular channels of 10
spatial average of a variable / over the disk. Similarly, the power re-
different cross-sections to show the effects of channel cross-sec-
moved from the mean flow at the disk plane is calculated as
tional aspect ratio and channel blockage ratio (which is the ratio  
Z Z
of disk area to channel cross-sectional area). 1
Power ¼ Mx U d dA ¼ K q U 3d dA ð10Þ
2
2. Mathematical models
and the power coefficient defined as

2.1. RANS model Power hU 3 i


CP ¼ 1 3
¼ K 3d : ð11Þ
2
qU in A U in
In this study we numerically solve the incompressible RANS
equations: Also, the axial induction factor is defined as
 
DU i 1 @P @ @U i U in  hU d i
¼ þ m  ui uj ð1Þ a¼ : ð12Þ
Dt q @xi @xj @xj U in
T. Nishino, R.H.J. Willden / International Journal of Heat and Fluid Flow 37 (2012) 123–135 125

The definitions of CT, CP and a described in Eqs. (9), (11), and (12) (This means that the dissipation rate assumed for the blade-induced
follow the definitions commonly used in LMADT. (Note, however, turbulence increases as the length scale for the blade-induced tur-
that in the present CFD study the value of Ud varies over the disk bulence decreases.) Then we further assume that, at the disk plane,
and hence the spatially-averaged values of Ud, U 2d and U 3d are used the blade-induced turbulence of kb and eb is instantly mixed with
in the definitions of a, CT and CP, respectively, whereas in LMADT the incoming ambient turbulence of ka and ea, resulting in the mixed
the value of Ud is assumed to be uniform over the disk.) turbulence that is described by its turbulent kinetic energy km and
its (effective) dissipation rate em. Here km is calculated as
2.3. Blade-induced turbulence model
km ¼ ka þ kb ð15Þ
The actuator disk model described above accounts for the effect whereas em may be modelled in several different ways. In this study
of a turbine on the mean flow and yields the power (or energy) that we model em as follows:
is removed from the mean flow at the turbine/disk. This is the abso-
km sm ¼ ka sa þ kb sb ð16Þ
lute upper limit of energy that may be extracted by the turbine; in
reality, some of the energy removed from the mean flow is not ex-
km
tracted but converted to the kinetic energy of turbulence, which is em ¼ ð17Þ
predominantly generated by the turbine blades. The proportion of sm
the energy that is converted to turbulence—we refer to this as where sa = ka/ea, sb = kb/eb and sm = km/em represent the time scale
blade-induced turbulence in this study—depends on actual turbine (or more specifically, initial eddy turnover time or lifetime) for the
design as well as on the operating conditions (such as the tip-speed ambient, blade-induced and mixed turbulence, respectively. The
ratio) of the turbine. Hence, to identify the proportion that is con- physical meaning of the model Eq. (16) is as follows: the time-inte-
verted to turbulence is beyond the scope of the present RANS-actu- gral of the turbulent kinetic energy for the mixed turbulence, which
ator disk study. The question to be answered is if, and how is approximated by 12 km sm when assuming that the turbulence just
significantly, this turbulence generated at the turbine location alters decays linearly at a constant (initial) dissipation rate em, is equiva-
the mean flow field and hence the upper limit of power extraction. lent to the sum of those time-integrals for the ambient and blade-
Several models to account for the effects of blade-induced tur- induced turbulence, which is approximated by 12 ka sa þ 12 kb sb when
bulence have been reported in previous RANS studies using k–e assuming that they also just decay linearly (and independently) at
type turbulence models: El Kasmi and Masson (2008) have pro- the rates of ea and eb, respectively (see Fig. 1). It should be noted that
posed to add an extra term to the equation for e to model the effect here we have selected the time-integral of linearly-decaying turbu-
of direct energy transfer from large-scale turbulence (correspond- lent kinetic energy as the quantity to be summed (or conserved)
ing to the incoming or ‘‘ambient’’ turbulence) to small-scale turbu- since it seems to better represent the dissipation characteristics of
lence (corresponding to that induced by the turbine), following the the mixed turbulence than the sum of the dissipation rates them-
concept of the extended k–e model of Chen and Kim (1987). selves (ea + eb). However it may also be appropriate to select other
Réthoré et al. (2009) used extra terms added to the equations for quantities to be conserved in the modelling of em.
k and e (two extra terms for each equation) to model the source One important issue arising when applying the above model to
and sink of k and e separately, following the idea of forest canopy RANS-actuator disk simulations is how to obtain the values of ka
turbulence modelling (Sanz, 2003). A drawback to these models is and ea that properly represent the characteristics of the ‘‘incoming
that they use one or more additional model coefficients that need (ambient) flow’’. From the modelling point of view, it is preferable
to be tuned empirically (similar to Ce1 and Ce2 in the original e to use local values of k and e just upstream of the disk (so that the
equation) and this tuning is not straightforward since those coeffi- model serves as a self-contained internal boundary condition, i.e.,
cients should depend on the actual design/conditions of a given the model does not require any flow information other than that
turbine. In other words, the effect of blade-induced turbulence is obtained along the disk surface). Our preliminary results have
case dependent and should therefore be incorporated using model shown, however, that the values of k and e just upstream of the
variables (that directly represent physically meaningful quantities) disk can be significantly affected by those just downstream of
rather than model coefficients; see Nishino and Willden (in press) the disk via the diffusion of k and e through the disk plane. In order
for further discussion on this issue. to avoid this problem, here we treat the actuator disk plane as a
In this study, we introduce an approach that does not use extra special internal boundary that does not allow the transport of k
model coefficients to be tuned empirically. Instead, two model or e via their diffusion (see Fig. 2). This treatment makes it possible
variables are introduced (as ‘‘input’’ for the simulations) to assess to define physically reasonable sources (over the actuator disk, per
the effect of blade-induced turbulence; these two variables directly unit area) of k and e to be added to the transport equations of k and
represent physically meaningful quantities, namely the ratio of the e, respectively, as
energy converted to turbulence to the energy removed from the
mean flow at the disk, b, and a representative length scale for the
blade-induced turbulence, lb. We first assume that the characteris-
tics of the turbulence induced by the turbine blades are described
by its turbulent kinetic energy, kb, and its dissipation rate, eb. Here
kb is calculated (locally over the actuator disk plane) as
τ τ τ
  ε τ
Mx 1 2
b lad e

kb ¼ b ¼ bK Ud ð13Þ
mi x

q 2
-ind u

ed

whereas eb is estimated (based on the high Reynolds number equi-


c ed

librium hypothesis used in most standard eddy-viscosity-based tur-


ambient
bulence models) as
3 3
C 4l k2b
τ τ τ τ
eb ¼ : ð14Þ
lb Fig. 1. A linear dissipation model to estimate the effective dissipation rate of
turbulence of two different scales instantly mixed at the disk plane.
126 T. Nishino, R.H.J. Willden / International Journal of Heat and Fluid Flow 37 (2012) 123–135

Fig. 2. Schematic of the blade-induced turbulence model applied to a RANS-


actuator disk computation.

Sk ¼ U d ðkm  ka Þ ¼ U d kb ð18Þ Fig. 3. Schematic of the computational domain (Case: 1.25d  1.25d) and the
coordinate system.
!
ðka þ kb Þ2
Se ¼ U d ðem  ea Þ ¼ U d 2 2
 ea : ð19Þ
ðka =ea Þ þ ðkb =eb Þ
Table 1
Channel cross-sectional aspect ratios and area blockage ratios.
In this study we first conducted an extensive investigation into the
effects of b and lb on the results; some of the key results are pre- Case Aspect ratio Blockage ratio
sented later in the Appendix. In the main body of the paper, how- 1.25d  1.25d 1 0.5027
ever, we present results only for lb = 0.1d, where d is the diameter 1.25d  1.5d 6/5 0.4189
of the disk, and for b = 0 and 0.1. The selected lb value roughly cor- 1.25d  2.5d 2 0.2513
responds to the chord length of the blades of common horizontal- 1.25d  5.0d 4 0.1257
1.5d  1.5d 1 0.3491
axis turbines and arguably represents the (effective) length scale 1.5d  2.5d 5/3 0.2094
for the blade-induced turbulence in most common circumstances.1 1.5d  5.0d 10/3 0.1047
It should be noted that the effect of blade-induced turbulence 2.5d  2.5d 1 0.1257
studied using the present model is somewhat hypothetical as the 2.5d  5.0d 2 0.0628
5.0d  5.0d 1 0.0314
structure of the wake of an actuator disk is different from that of
an actual turbine rotor. The interaction (to be simulated in this
study) between the modelled blade-induced turbulence and the
shear-generated turbulence behind the disk is a simplified repre- is expected to be even smaller than that represented by C P . This is
sentation of the rotor wake turbulence; see Nishino and Willden because the energy converted to turbulent kinetic energy is not
(in press) for further discussion on this issue. Nevertheless, results the whole but a part of the energy that is unextractable. Since our
obtained from the present model are still useful to examine the main interest in this paper is how significant, or insignificant, the
(qualitative) effect of turbulence on the limit of power extraction effect of the blade-induced turbulence is on the prediction of the
by tidal turbines. absolute limit of power extraction, we use the original CP defined
in Eq. (11) throughout the main body of this paper. Plots of C P are
2.4. Note on the definition of CP presented only in the Appendix.

It should be borne in mind that the power coefficient CP defined 3. Details of the computations
in Eq. (11) always represents the power that is removed from the
mean flow at the disk plane, regardless of whether we consider Computations were performed using a commercial CFD solver,
the blade-induced turbulence or not. This power can be interpreted FLUENT 12 (ANSYS Inc., 2009), together with its User-Defined
as the maximum extractable power (i.e., extractable as useful Function (UDF) module for modifications. The solver is based on
power by the turbine) when the blade-induced turbulence is not a finite volume method employing a procedure similar to that out-
considered (b = 0). When we consider the blade-induced turbu- lined by Rhie and Chow (1983) and is nominally second-order
lence (b > 0), it might be useful to define an alternative power coef- accurate in space. The SIMPLE algorithm (Patankar, 1980) was used
ficient C P that better represents the extractable power as for pressure–velocity coupling. All computations were performed
C P ¼ ð1  bÞC P ð20Þ as steady state.
Fig. 3 shows a schematic of the computational domain and the
since the mean-flow energy that is converted to turbulent kinetic coordinate system used in this study. An actuator disk of a diame-
energy (at the disk plane) is unextractable. It should be noted, how- ter d of 20 m is placed inside a rectangular channel and the disk
ever, that the power extracted as useful power by an actual turbine face is perpendicular to the flow through the channel. The disk cen-
ter is positioned at (x, y, z) = (0, 0, 0). Computations were performed
1
Actual turbine blades generate turbulence (or velocity fluctuations) of at least two for channels of 10 different cross-sections. The cross-sectional as-
different scales: the larger one corresponds to periodic passages of the rotating pect ratio and area blockage ratio of each channel are summarized
blades, which determine the spacing of the helical structure of the near wake, and the in Table 1. Here the blockage ratio, B, is defined by the ratio of the
smaller one corresponds to vortices within the shear layers separated from the
surface of the blades (see e.g., Vermeer, 2001). In the present study, however, we
disk area to the channel cross-sectional area. All computations
consider a single (effective) length scale for the blade-induced turbulence for were performed only for a quarter of the channel (i.e., for the
simplicity. region of y P 0 and z P 0 as shown by the thick broken lines in
T. Nishino, R.H.J. Willden / International Journal of Heat and Fluid Flow 37 (2012) 123–135 127

Table 2
Grid resolutions: number of cells (Case: 1.25d  1.25d).

Grid Cross-sectional Streamwise Total


Medium 1824 300 547,200
Fine 7296 600 4377,600

resolution in each direction (see Table 2) was also used to examine


the effects of grid resolution on the results, which will be presented
in Section 4.1. For lower blockage cases, the same mesh resolution
and topology were used in the vicinity of the disk (to preserve the
disk area resolution) with additional grid points used away from
the disk.

4. Results and discussion

Computational results are presented below. As already noted in


Section 2.3, results only for lb = 0.1d and for b = 0 and 0.1 are pre-
Fig. 4. Cross-sectional mesh (medium resolution, Case: 1.25d  1.25d). sented in the main body of this paper, whilst further descriptions
on the effects of lb and b can be found in the Appendix. Unless spec-
ified, all results are those for the Low FST level and computed on
Fig. 3) with symmetry boundary conditions applied to all four lon- the medium grid.
gitudinal faces of the computational domain. This means that we First we briefly describe the effect of grid resolution in Section
simulate flow through an infinite number of disks periodically ar- 4.1, followed by the effect of square cross-section channel blockage
rayed in both y and z directions. It should be noted that this flow in Section 4.2. The effects of turbulent wake mixing and channel
situation is related to the tidal fence concept proposed for future cross-sectional aspect ratio are then described in Sections 4.3 and
tidal power generation, where a number of turbines are arrayed 4.4, respectively. Finally further discussion is given in Section 4.5.
in the spanwise direction of a wide tidal basin but with vertical
constraints provided by the sea surface and seabed. 4.1. Grid sensitivity
The inlet of the computational domain is located 10d upstream
of the disk, whereas the outlet is located 15d downstream of the A grid sensitivity study was performed for the highest blockage
disk (preliminary computations with a longer domain have been case ‘‘1.25d  1.25d,’’ where the shear layer developed behind the
performed to confirm that the domain is long enough). Uniform disk was the thinnest and hence the grid sensitivity expected to be
velocity components of (U, V, W) = (Uin, 0, 0) are given at the inlet the most significant. Fig. 5 shows the disk performance character-
boundary,2 where the inlet (or freestream) velocity Uin = 2 m/s, istics (i.e., the power coefficient CP versus the axial induction factor
and zero-streamwise-gradient conditions are applied at the outlet a) computed on the medium and fine grids for b = 0 and 0.1. For
boundary. Two different levels of freestream turbulence (FST) are each b case the results are very nearly the same between the two
tested, although their effects are examined only qualitatively in this grids (the differences in CP are all less than 0.5%), demonstrating
study (since, for example, the effects of strong mean strain induced that the medium grid resolution is sufficient for the present study
by the turbine are not predicted correctly by the standard k–e mod- regardless of whether the newly-proposed blade-induced turbu-
el): For the ‘‘Low FST’’ level, k = 6  106 m2/s2 and e = 1.21  lence model is used or not.
109 m2/s3 (corresponding to a turbulent intensity of 0.1% and a
turbulent length scale of 0.1d when assuming the high Reynolds 4.2. Effect of square cross-section channel blockage
number equilibrium hypothesis) are given at the inlet. For the ‘‘High
FST’’ level, k = 0.06 m2/s2 and e = 0.00121 m2/s3 (corresponding to a Fig. 6 shows CP and CT curves computed for four different square
turbulent intensity of 10% and a turbulent length scale of 0.1d) are cross-section cases. Here we focus on the results for b = 0 (i.e.,
given at the inlet. This high FST gradually decays downstream; without using the blade-induced turbulence model). Also plotted
however the turbulent intensity at the disk location (x/d = 0) is still
reasonably high (about 5.7% in the case without the disk). The
density and viscosity of the working fluid are q = 1000 kg/m3 and 2.6
l = 0.001 kg/m s, resulting in a disk Reynolds number Re
(=qUind/l) of 4  107. 2.4
Fig. 4 shows a cross-sectional snapshot of the computational
mesh for the highest blockage case ‘‘1.25d  1.25d’’ (2Ly = 2Lz =
2.2
1.25d, where Ly and Lz denote the computational domain size in
the y and z directions). The same 2-D mesh was stacked in the
streamwise (x) direction to create 3-D multi-block structured grids 2
consisting of hexahedral cells. A finer grid with double the
1.8 Medium ( β = 0)
Fine ( β = 0)
2
As has been discussed by Garrett and Cummins (2005) and Vennell (2010), in a Medium ( β = 0.1)
practical tidal channel, the mass flow through the channel may change depending on 1.6 Fine ( β = 0.1)
the hydrodynamic load or drag acting on tidal devices (in addition to the change due
to the tidal cycle). The assumption of a constant mass flow for different device 0.3 0.4 0.5 0.6
operating conditions (at a given phase of the tidal cycle) is valid only when the drag
acting on the devices is negligibly small compared to the drag acting on the entire
channel. In practice this assumption may only hold for very long channels. Fig. 5. Effect of grid resolution on CP (Case: 1.25d  1.25d).
128 T. Nishino, R.H.J. Willden / International Journal of Heat and Fluid Flow 37 (2012) 123–135

2.5 6

β=0 β=0
1.25d
LMADT 1.25d x 5 LMADT
2 5 d
.2
x1
d
4 25
1.5d x 1.5d 1.
1.5
d
1.5
3 dx
1.5
1 2.5d x 2.5d
2
2.5d
2.5d x
0.5 5.0d x 5.0d
1
5.0d x 5.0d

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6

Fig. 6. Effects of blockage on CP and CT (for square cross-section channels).

As pressures must equilibrate between the ‘‘core’’ flow (i.e., the


flow going through the disk) and the bypass flow both far upstream
and far downstream of the disk, the increase in bypass pressure
drop results in an increase in pressure drop through the core flow
passage and hence across the disk as well, noting that there is little
change in the expansion of the core flow passage upstream of the
disk. It follows that: (i) if the axial induction factor is held constant
then the disk momentum loss factor increases (thereby the pres-
sure drop, CT and CP increase) as the blockage increases, and (ii)
if the disk momentum loss factor is held constant then the axial
induction factor decreases (i.e., velocity through the disk increases,
thereby the pressure drop, CT and CP increase) as the blockage in-
creases. Fig. 7 shows contours of static pressure coefficient for
two different blockage cases (on the y = 0 plane cutting through
the center of the disk). Also presented by the thick line is the
cross-section of the streamtube that divides the bypass and core
flow passages. Here we compare the two cases at the same axial
induction factor a = 0.412; hence the disk momentum loss factor
is higher for the higher blockage case (K = 6.7) than for the lower
blockage case (K = 4). It can be seen that the pressure drop through
the bypass flow passage (as well as through the core flow passage
and hence across the disk) significantly increases as the channel
blockage increases.

4.3. Effect of turbulent mixing


Fig. 7. Contours of static pressure coefficient and the streamtube dividing the
bypass and core flow passages (cross-sectional view on the y = 0 plane). Here we discuss the effect of turbulence in detail. Results are
presented for a square channel case ‘‘2.5d  2.5d,’’ which will be
compared with a rectangular channel case ‘‘1.25d  5.0d’’ later in
here for comparison are the solutions of 1-D LMADT for flow Section 4.4 to discuss the effect of channel cross-sectional aspect
bounded by a parallel-sided tube (Houlsby et al., 2008; see also ratio.
Garrett and Cummins, 2007). The blockage effect predicted by Fig. 8 shows CP and CT curves computed for three different tur-
the present 3-D RANS computations is similar to that predicted bulence conditions: b = 0 with low FST, b = 0.1 with low FST and
by the 1-D LMADT: as the blockage ratio increases, CP and CT signif- b = 0.1 with high FST. Note that b = 0.1 means that 10% of the en-
icantly increase and the axial induction factor required to achieve ergy removed from the mean flow at the turbine/disk location is
the maximum CP also increases. The computed values of CP and CT converted to turbulent kinetic energy (cf. Section 2.3) and this is
are only slightly higher than LMADT for lower blockage cases but a representative value to qualitatively account for the effect of
are slightly lower than LMADT for the highest blockage case; the blade-induced turbulence in common circumstances; results for
reasons for these slight differences will be discussed later in Sec- various b values are presented in the Appendix. Comparing the re-
tion 4.5. sults for b = 0 and 0.1, it can be seen that the thrust acting on the
As can be seen from the fairly good agreement between the 3-D disk and the power removed from the mean flow both increase
RANS computations and the 1-D LMADT, the effects of channel (albeit gently) when the effect of blade-induced turbulence is
blockage presented here are largely explained by an inviscid mech- taken into account. It can also be seen that the thrust and power
anism. Specifically, the significant increases in CT and CP are due to further increase as the freestream turbulence intensity increases.
increased acceleration of the ‘‘bypass’’ flow (i.e., the flow going In order to better understand these changes in CP and CT, we
through the gap between the disk and the channel wall) and an compare the flow field around the disk between the three cases.
attendant larger pressure drop through the bypass flow passage. The comparisons are made for a constant disk momentum loss
T. Nishino, R.H.J. Willden / International Journal of Heat and Fluid Flow 37 (2012) 123–135 129

1 2

1.8

0.8 1.6

1.4

0.6 1.2

0.4 0.8

0.6
β = 0 (Low FST) β = 0 (Low FST)
0.2 β = 0.1 (Low FST) 0.4 β = 0.1 (Low FST)
β = 0.1 (High FST) β = 0.1 (High FST)
LMADT 0.2 LMADT

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6

Fig. 8. Effects of blade-induced and freestream turbulence on CP and CT (Case: 2.5d  2.5d).

Fig. 9. Streamwise velocity and turbulent kinetic energy contours (on the y = 0 plane, Case: 2.5d  2.5d).

factor K = 4 rather than for a constant axial induction factor a or for intensity increases. The flow upstream of the disk, however, is
the exact maximum CP condition for each case; however the rela- not significantly affected by either the blade-induced or the free-
tionship between K and a is very similar for the three cases and stream turbulence.
K = 4 provides a CP value that is very close to the maximum CP va- The effect of turbulent mixing can be seen more clearly in
lue for each case. Fig. 9 shows contours of streamwise velocity and Fig. 10, which compares the profiles of streamwise velocity at six
turbulent kinetic energy for the three cases (on the y = 0 plane). different streamwise locations upstream and downstream of the
Comparing the cases for b = 0 and 0.1, it is apparent that the mean disk plane. Also plotted here for comparison are the results of
flow field downstream of the disk is significantly affected by the ‘‘quasi-inviscid’’ simulations, where the diffusion terms in the
blade-induced turbulence. When b = 0, turbulence is generated momentum equations were neglected.3 At the disk location (x/
only gradually due to the mean shear between the low-speed wake d = 0), the velocity profiles are only slightly different between the
flow and the high-speed bypass flow surrounding the wake, three turbulence conditions and the quasi-inviscid case. As the tur-
whereas when b = 0.1, turbulence is generated by the disk itself
as well as by the shear behind the disk, resulting in much stronger 3
We refer to this as ‘‘quasi-inviscid’’ since although the inviscid momentum
turbulent mixing in the near wake region. The mixing in the near equations are solved some numerical diffusion is present that renders the solution not
wake becomes even stronger as the freestream turbulence inviscid.
130 T. Nishino, R.H.J. Willden / International Journal of Heat and Fluid Flow 37 (2012) 123–135

1.25 1.25 1.25


Quasi-Inviscid
β = 0 (Low FST)
1 β = 0.1 (Low FST) 1 1
β = 0.1 (High FST)

0.75 0.75 0.75

0.5 0.5 0.5

0.25 0.25 0.25


x/d = - 0.1 x/d = 0 x/d = 0.1

0 0 0
0.4 0.6 0.8 1 1.2 0.4 0.6 0.8 1 1.2 0.4 0.6 0.8 1 1.2

1.25 1.25 1.25

1 1 1

0.75 0.75 0.75

0.5 0.5 0.5

0.25 0.25 0.25


x/d = 0.5 x/d = 1 x/d = 2

0 0 0
0 0.5 1 1.5 0 0.5 1 1.5 0 0.5 1 1.5

Fig. 10. Streamwise velocity profiles (on the y = 0 plane, Case: 2.5d  2.5d, K = 4).

bulence level increases and the mixing between the wake and the
bypass flow is enhanced, the streamwise velocity at the disk (mostly
near the disk edge) increases slightly, resulting in the slight changes
in CP and CT presented earlier in Fig. 8. The relationship between the
strength of turbulent mixing behind the disk and the streamwise
velocity at the disk plane may be explained as follows: Considering
that the incoming flow goes through either bypass or core flow pas-
sages, the drag acting on the bypass flow increases and that acting on
the core flow decreases as the mixing behind the disk (where pres-
sure has not yet equilibrated between the bypass and core flows) is
enhanced. This is because the turbulent shear stress on the interface
between the bypass and core flow passages (i.e., the surface of the Fig. 11. Schematic of turbulent shear stress acting on the interface between the
bypass and core (wake) flow passages.
streamtube dividing the two flow passages) acts in such a way that
it decelerates the higher-speed bypass flow and accelerates the low-
er-speed wake flow (see Fig. 11). Therefore a greater portion of the
incoming flow goes through the disk (i.e., the axial induction factor 1.4
decreases) as the mixing behind the disk is enhanced (when the disk
1.3 Quasi-inviscid
momentum loss factor is held constant; alternatively, the momen- β = 0 (Low FST)
tum loss factor increases when the axial induction factor is held con- 1.2 β = 0.1 (Low FST)
E(x)

stant, as already noted in Section 4.2 for the inviscid blockage effect). β = 0.1 (High FST)
1.1
Whilst the main focus of this study is on the limit of power
extraction by a single device (or devices periodically arrayed but 1
only in the direction perpendicular to the flow), another interesting -5 0 5 10 15
issue is the effect of turbulent mixing on the flow far downstream x/d
of the device. Figs. 12 and 13 show the streamwise variations of the
Fig. 12. Kinetic energy flux of the mean flow (Case: 2.5d  2.5d, K = 4).
kinetic and total energy fluxes of the mean flow through the chan-
nel, E(x) and G(x), respectively, for the three turbulence conditions
and the quasi-inviscid case. Here E(x) and G(x) are calculated as ~ 2 ¼ U 2 þ V 2 þ W 2 , P is static pressure and S is the cross-
where jUj
R 2 sectional area of the channel. Note that Fig. 13 plots G(x)  Gin,
1 ~ U dS
qjUj
2
EðxÞ ¼ ð21Þ where Gin is the value of G at the inlet of the channel (x/d = 10).
1
2
qU 3in S It can be seen from Fig. 12 that the mean-flow kinetic energy in-
R creases around the disk since the disk creates the velocity difference
~ 2 ÞU dS
ðP þ 12 qjUj between the wake and the bypass flow, as shown earlier in Figs. 9
GðxÞ ¼ ð22Þ
1
2
qU 3in S and 10. The mean-flow kinetic energy reaches a peak within a
diameter downstream of the disk plane, before gradually decreasing
T. Nishino, R.H.J. Willden / International Journal of Heat and Fluid Flow 37 (2012) 123–135 131

narrower gap just downstream of the disk (see Fig. 15c, at x/


Quasi-inviscid
0 β = 0 (Low FST) d = 0.2, around y/d = 0.6, z/d = 0). This very-high-speed bypass
β = 0.1 (Low FST) flow quickly diffuses and the circular wake profile gradually
G(x) - G

β = 0.1 (High FST)


changes to a more planar (or two-dimensional) profile further
-0.1 downstream.
Fig. 16 compares streamwise variations of E(x) for the two
cases. The results for b = 0.1 are also presented here for compari-
-0.2
-5 0 5 10 15 son. For each b value, the peak value of the mean-flow kinetic en-
x/d ergy flux behind the disk (around 0.5 < x/d < 1) is lower in the
Fig. 13. Total energy flux of the mean flow (Case: 2.5d  2.5d, K = 4).
rectangular channel than in the square channel. In the region fur-
ther downstream, however, the mean-flow kinetic energy flux de-
creases faster in the square channel than in the rectangular
further downstream (except for the quasi-inviscid case) as the channel, indicating that the mixing in this downstream region is
velocity difference diminishes due to turbulent mixing. The rate stronger in the square channel, where the cylindrical wake profile
of mixing depends on both the blade-induced and the freestream remains for a long distance, than in the rectangular channel, where
turbulence. Of interest is that it takes a rather long distance for the wake profile gradually changes to a more planar profile. Fig. 17
the mean-flow kinetic energy flux to decrease back to the original compares streamwise variations of the channel cross-sectional
value (E = 1) and thus for the total energy flux to reach a final down- average of turbulent kinetic energy. It can be seen that, regardless
stream level (which is slightly different among the three cases due of whether the blade-induced turbulence is taken into account or
to the slight difference in the disk performance), even for the high- not, the cross-sectional average of turbulent kinetic energy is high-
est mixing case (b = 0.1 with high FST). These results suggest that er in the square channel than in the rectangular channel.
the limit of power extraction by devices arrayed in the streamwise
direction may substantially depend on the streamwise spacing of
4.5. Further discussion
the devices as well as the strength of turbulent mixing behind each
device. Further investigations are needed to clarify this issue in fu-
Fig. 18 plots CPmax (the maximum or peak value of the ‘‘CP vs. a’’
ture studies.
curve) computed for all 10 different channel cases and for the four
different conditions discussed in Section 4.3. Also plotted here for
4.4. Effect of channel cross-sectional aspect ratio
comparison is the solution of 1-D LMADT for flow bounded by a
parallel-sided tube (Houlsby et al., 2008; see also Garrett and Cum-
Fig. 14 compares CP and CT curves computed for the cases
mins, 2007), for which it can be shown that
‘‘2.5d  2.5d’’ and ‘‘1.25d  5.0d.’’ Results for b = 0 and 0.1 are pre-
sented here for each case. Note that the channel cross-sectional  2
16 1
area is exactly the same between the two cases, making it possible C Pmax ¼ : ð23Þ
27 1  B
to examine the effect of channel cross-sectional aspect ratio for the
same blockage. Also note that the 1-D LMADT does not account for It can be seen that the effect of turbulence on CPmax is similar for all
the effect of channel cross-sectional aspect ratio. It can be seen that 10 cases: the value of CPmax increases as the strength of turbulent
both CP and CT are slightly higher in the narrow rectangular cross- mixing behind the disk increases, as discussed earlier in Section
section channel than in the square cross-section channel, for each b 4.3 for the case ‘‘2.5d  2.5d’’. (However it should be borne in mind
value. that, as noted earlier in Section 2.4, the power coefficient CP pre-
Fig. 15 shows contours of streamwise velocity and turbulent sented here represents the power removed from the mean flow at
kinetic energy at several streamwise locations for the two cases the disk plane, including the power that will be given up to
(for K = 4, which gives a CP value that is very close to the maxi- blade-induced turbulence.) The effect of turbulent mixing on CPmax
mum CP value for each case, and for b = 0). The channel aspect is even enhanced as the channel blockage increases, presumably be-
ratio significantly affects the flow pattern downstream of the cause the magnitude of shear stress between the bypass and wake
disk; however the flow at the disk location (x/d = 0) is less sensi- flows increases (since the velocity difference between them in-
tive to the channel aspect ratio, resulting in the relatively small creases) and hence the level of turbulent mixing has greater signif-
differences in CP and CT. In the rectangular channel, it is icance. It can also be seen that the effect of channel cross-sectional
remarkable that a very-high-speed bypass flow develops at the aspect ratio, which varies between 1 and 4 in the present study, is

1 2

1.8

0.8 1.6

1.4

0.6 1.2

0.4 0.8
2.5d x 2.5d ( β = 0) 2.5d x 2.5d ( β = 0)
1.25d x 5.0d ( β = 0) 0.6 1.25d x 5.0d ( β = 0)
0.2 2.5d x 2.5d ( β = 0.1) 0.4 2.5d x 2.5d ( β = 0.1)
1.25d x 5.0d ( β = 0.1) 1.25d x 5.0d ( β = 0.1)
LMADT 0.2 LMADT

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6

Fig. 14. Effects of channel aspect ratio on CP and CT.


132 T. Nishino, R.H.J. Willden / International Journal of Heat and Fluid Flow 37 (2012) 123–135

Fig. 15. Streamwise velocity and turbulent kinetic energy contours at several streamwise locations (K = 4, b = 0).
T. Nishino, R.H.J. Willden / International Journal of Heat and Fluid Flow 37 (2012) 123–135 133

1.3 relatively minor compared to that of the blockage ratio. Of impor-


2.5d x 2.5d ( β = 0) tance is that the CPmax plots for all 10 cases appear to be closely
1.25d x 5.0d (β = 0)
1.2 2.5d x 2.5d ( β = 0.1) aligned along the solution curve of 1-D LMADT, suggesting that 1-
1.25d x 5.0d (β = 0.1)
D LMADT would fairly reasonably predict the limit of power extrac-
E(x)

1.1 tion even in a generic 3-D channel where the blockage is


anisotropic.
1 Finally it should be noted that the ‘‘quasi-inviscid’’ computa-
tions in this study agree very well with 1-D LMADT for lower
-5 0 5 10 15
x/d blockage cases but yield somewhat lower CPmax than 1-D LMADT
for higher blockage cases. This discrepancy reflects that, in the
Fig. 16. Kinetic energy flux of the mean flow (K = 4). present computations, the non-uniformity of the flow over the disk
becomes more significant as the channel blockage increases. Fig. 19
shows contours of local CP (which is defined as KU 3d =U 3in ) on the disk
plane for four different square channel cases (for the maximum CP
0.03 condition for each case). Here the local CP values are divided by the
2.5d x 2.5d ( β = 0)
1.25d x 5.0d ( β = 0) theoretical CPmax value (obtained from 1-D LMADT for each block-
2.5d x 2.5d ( β = 0.1)
∫k A / U A

0.02 age) to explicitly show the deviation of the 3-D quasi-inviscid com-
1.25d x 5.0d ( β = 0.1)
putations from 1-D LMADT. For the lowest blockage case
0.01 ‘‘5.0d  5.0d’’, the local CP is slightly higher than the theoretical va-
lue at and around the center of the disk but is lower than the the-
0 oretical value near the disk edge, resulting in a good agreement in
-5 0 5 10 15 the (spatially-averaged) CP value between the computation and 1-
x/d D LMADT. As the blockage ratio increases, the deviation from the
theory becomes more significant near the disk edge and the aver-
Fig. 17. Streamwise variation of the channel cross-sectional average of turbulent
kinetic energy (K = 4). aged CP value falls below the theoretical value. This, together with
the effect of turbulent mixing (which increases CP and CT compared
to the inviscid case), also explains the reason for the slight differ-
ences in CP and CT between the 3-D RANS computations and 1-D
3
LMADT presented earlier in Section 4.2.
Quasi-inviscid
β = 0 (Low FST)
β = 0.1 (Low FST)
1.25d x 1.5d

2.5
β = 0.1 (High FST)
5. Conclusions
LMADT
1.25d x 1.25d

2 In this study we have performed 3-D incompressible RANS com-


putations of water flow past an actuator disk placed in a rectangu-
1.25d x 2.5d

lar channel of various blockages and aspect ratios. The main focal
1.5d x 2.5d
1.25d x 5.0d

1.5
points of the study were on the effect of turbulent mixing behind
1.5d x 5.0d
2.5d x 5.0d

the disk as well as on the effect of channel blockage and channel


1.5d x 1.5d

1 cross-sectional aspect ratio on the prediction of the hydrodynamic


limit of power extraction. In order to qualitatively account for the
2.5d x 2.5d

effect of turbine-blade-induced turbulence, we have proposed a


5.0d x 5.0d

0.5
new approach that does not require any additional model coeffi-
cients (other than those used in the original k–e turbulence model)
0 to be tuned empirically. Instead, two physical parameters are re-
0 0.1 0.2 0.3 0.4 0.5
quired to specify the characteristics of the disk in this approach,
namely the ratio of the energy converted to turbulence to that re-
Fig. 18. CPmax vs. channel blockage ratio. moved from the mean flow at the disk plane, b, and a representa-
tive length scale for the blade-induced turbulence, lb.

Fig. 19. Contours of local CP on the actuator disk plane, showing the deviation of 3-D quasi-inviscid CFD results from 1-D LMADT (maximum CP condition for each blockage
case).
134 T. Nishino, R.H.J. Willden / International Journal of Heat and Fluid Flow 37 (2012) 123–135

The study has shown that both blade-induced and freestream in the streamwise direction. Further investigations are needed to
turbulence significantly affect the strength of turbulent mixing clarify these issues in future studies.
and therefore the mean flow profile behind the disk; however
the effect on the flow through the disk plane is less significant. Acknowledgments
The thrust acting on the disk and the power removed by the disk
were both found to increase, albeit gently, as the strength of turbu- The authors gratefully acknowledge the support of the Oxford
lent mixing behind the disk increased. This effect may be explained Martin School, University of Oxford, who have funded this
by considering that the turbulent shear stress on the interface be- research.
tween the bypass and core flow passages acts in such a way that it
decelerates the higher-speed bypass flow and accelerates the low-
er-speed wake flow; hence, as the mixing behind the disk (where Appendix A. Effects of the scale and strength of blade-induced
pressure has not yet equilibrated between the bypass and core turbulence
flows) is enhanced, a greater portion of the incoming flow goes
through the disk whilst the disk momentum loss factor is held con- Fig. A.1 shows the effects of b and lb on the peak value of the ‘‘CP
stant (or alternatively, the disk momentum loss factor increases vs. a’’ curve computed for the case ‘‘2.5d  2.5d’’ (employing the
whilst the axial induction factor is held constant). low FST and high FST conditions). Here we show the results for five
It has also been shown that the channel cross-sectional aspect different values of lb ranging from 0.01d to 0.5d, although we rec-
ratio significantly affects the mean flow profile behind the disk; ognize that 0.5d (one-half of the disk diameter) seems too large to
however the effects on the flow through the disk and thus on the be considered as an effective length scale for the blade-induced
power removed by the disk are relatively minor (at least within turbulence for any type of tidal turbine. As can be seen from the
the limits of the aspect ratio and blockage considered in the pres- figure, both b and lb affect the prediction of CPmax (as they affect
ent study) compared to the effects of the channel blockage, which the strength of turbulent mixing behind the disk), although the
are satisfactorily predicted by 1-D inviscid theory. In a narrow dependency on b appears less significant for b > 0.1 (except for
(high-aspect-ratio) cross-section channel, the cylindrical wake the largest lb case). It is also evident that the effects of b and lb
profile behind the disk gradually changes to a more planar profile, are similar for both low and high FST cases.
whereas in a square cross-section channel, the cylindrical wake It should be noted that, when a rather small length scale is as-
profile remains for a longer distance downstream. sumed for the blade-induced turbulence, CPmax slightly decreases as
Last but not least, it should be borne in mind that in the present b increases. This is because the turbulent kinetic energy very
study we simulated flow past an infinite number of disks periodi- quickly dissipates just downstream of the disk (due to the rather
cally arrayed in both spanwise and vertical directions. This flow small length scale and hence high dissipation rate assumed for
situation is related to the tidal fence concept proposed for future the blade-induced turbulence) and eventually the level of turbu-
tidal power generation, where a number of turbines are arrayed lent mixing behind the disk decreases overall as b increases. How-
in the spanwise direction of a wide tidal basin and the array verti- ever, CPmax never falls below that for the quasi-inviscid case since
cally bounded by the sea surface and seabed. In such an actual tidal the mixing behind the disk never becomes weaker than that for
flow situation, the influence of the nearby sea surface and seabed the quasi-inviscid case.
(and also the shear of the mean flow induced by them) may not Fig. A.2 shows the effects of b and lb on the peak value of C P
be negligible. Nevertheless, the physical insight into the effects of defined in Eq. (20), which represents the maximum extractable
3-D blockage and turbulent wake mixing obtained in this study power that accounts for the fact that the mean-flow energy con-
should remain valuable. Another interesting issue is the limit of verted to turbulent kinetic energy is unextractable (cf. Section
power extraction by devices arrayed in the tidal streamwise direc- 2.4). Except for the largest lb case, C Pmax only slightly increases as
tion. The present RANS results have shown that both the blade-in- b increases for b up to about 0.01–0.02 (depending on lb and the
duced turbulence and the aspect ratio of the flow passage affect the FST level) but then monotonically decreases as b further increases.
strength of turbulent mixing over a rather long distance down- These results suggest that, although the blade-induced turbulence
stream of the disk, suggesting that these factors may be of greater may enhance the turbulent mixing behind the disk/turbine and
importance when considering the performance of devices arrayed thereby increase the absolute upper limit of power extraction
(which is described by CPmax), this increase would not be large

1.05 1.05

1 1

0.95 0.95

0.9 0.9

0.85 0.85

0.8 0.8

Quasi-inviscid (0.765) Quasi-inviscid (0.765)


0.75 0.75
0 0.1 0.2 0.3 0 0.1 0.2 0.3

Fig. A.1. Effects of the scale and strength of blade-induced turbulence on CPmax (Case: 2.5d  2.5d): low FST (left) and high FST (right) cases.
T. Nishino, R.H.J. Willden / International Journal of Heat and Fluid Flow 37 (2012) 123–135 135

0.9 0.9

0.85 0.85

0.8 0.8

0.75 0.75

0.7 0.7

0.65 0.65

0.6 0.6
0 0.1 0.2 0.3 0 0.1 0.2 0.3

Fig. A.2. Effects of the scale and strength of blade-induced turbulence on C Pmax (Case: 2.5d  2.5d): low FST (left) and high FST (right) cases.

enough to compensate the hydrodynamic loss resulting from the Harrison, M.E., Batten, W.M.J., Myers, L.E., Bahaj, A.S., 2010. Comparison between
CFD simulations and experiments for predicting the far wake of horizontal axis
blade-induced turbulence itself (i.e., the loss of mean-flow energy
tidal turbines. IET Renewable Power Generation 4 (6), 613–627.
as it is converted to turbulent kinetic energy) in most common Houlsby, G.T., Draper, S., Oldfield, M.L.G., 2008. Application of Linear Momentum
circumstances. Actuator Disc Theory to Open Channel Flow. Report No. OUEL 2296/08,
Department of Engineering Science, University of Oxford, UK.
Launder, B.E., Spalding, D.B., 1974. The numerical computation of turbulent flows.
Computer Methods in Applied Mechanics and Engineering 3, 269–289.
Nishino, T., Willden, R.H.J., in press. Low-order modelling of blade-induced
References turbulence for RANS actuator disk computations of wind and tidal turbines.
In: Proc. EUROMECH Colloquium 528: Wind Energy and the Impact of
ANSYS Inc., 2009. ANSYS FLUENT 12.0 User’s Guide. ANSYS Inc., Canonsburg, PA, Turbulence on the Conversion Process, 22–24 February, Oldenburg, Germany.
USA. Patankar, S.V., 1980. Numerical Heat Transfer and Fluid Flow. Hemisphere
Batten, W.M.J., Harrison, M., Bahaj, A.S., 2011. The accuracy of the actuator disc- Publishing Corporation, Washington, DC, USA.
RANS approach for predicting the performance and far wake of a horizontal axis Réthoré, P.-E., Sørensen, N.N., Bechmann, A., Zhale, F., 2009. Study of the
tidal stream turbine. In: Proc. 9th European Wave and Tidal Energy Conference atmospheric wake turbulence of a CFD actuator disc model. In: Proc. 2009
(EWTEC 2011), 5–9 September, Southampton, UK. European Wind Energy Conference and Exhibition (EWEC 2009), 16–19 March,
Belloni, C.S.K., Willden, R.H.J., 2011. Flow field and performance analysis of Marseille, France.
bidirectional and open-centre ducted tidal turbines. In: Proc. 9th European Rhie, C.M., Chow, W.L., 1983. Numerical study of the turbulent flow past an airfoil
Wave and Tidal Energy Conference (EWTEC 2011), 5–9 September, with trailing edge separation. AIAA Journal 21, 1525–1532.
Southampton, UK. Sanz, C., 2003. A note on k–e modeling of vegetation canopy air-flows. Boundary-
BERR, 2008. Atlas of UK Marine Renewable Energy Resources. Department for Layer Meteorology 108, 191–197.
Business Enterprise & Regulatory Reform (BERR), UK. Sørensen, J.N., 2011. Aerodynamic aspects of wind energy conversion. Annual
Burton, T., Sharpe, D., Jenkins, N., Bossanyi, E., 2001. Wind Energy Handbook. John Review of Fluid Mechanics 43, 427–448.
Wiley & Sons Ltd., West Sussex, UK. Sun, X., Chick, J.P., Bryden, I.G., 2008. Laboratory-scale simulation of energy
Chen, Y.-S., Kim, S.-W., 1987. Computation of Turbulent Flows using an Extended k– extraction from tidal currents. Renewable Energy 33, 1267–1274.
e Turbulence Closure Model. NASA Contractor Report, NASA CR-179204. Turnock, S.R., Phillips, A.B., Banks, J., Nicholls-Lee, R., 2011. Modelling tidal current
DTI, 2004. Atlas of UK Marine Renewable Energy Resources. Department of Trade turbine wakes using a coupled RANS-BEMT approach as a tool for analysing
and Industry (DTI), UK. power capture of arrays of turbines. Ocean Engineering 38, 1300–1307.
El Kasmi, A., Masson, C., 2008. An extended k–e model for turbulent flow through van Kuik, G.A.M., 2007. The Lanchester–Betz–Joukowsky limit. Wind Energy 10,
horizontal-axis wind turbines. Journal of Wind Engineering and Industrial 289–291.
Aerodynamics 96, 103–122. Vennell, R., 2010. Tuning turbines in a tidal channel. Journal of Fluid Mechanics 663,
Gant, S., Stallard, T., 2008. Modelling a tidal turbine in unsteady flow. In: Proc. 2008 253–267.
International Offshore and Polar Engineering Conference, 6–11 July, Vancouver, Vermeer, L.J., 2001. A review of wind turbine wake research at TU Delft. In: Proc.
Canada. 39th AIAA Aerospace Sciences Meeting and Exhibit, 11–14 January, Reno, NV,
Garrett, C., Cummins, P., 2005. The power potential of tidal currents in channels. USA. AIAA Paper 2001-0030.
Proceedings of the Royal Society A 461, 2563–2572. Whelan, J.I., Graham, J.M.R., Peiró, J., 2009. A free-surface and blockage correction
Garrett, C., Cummins, P., 2007. The efficiency of a turbine in a tidal channel. Journal for tidal turbines. Journal of Fluid Mechanics 624, 281–291.
of Fluid Mechanics 588, 243–251.

Vous aimerez peut-être aussi