Accepted Manuscript
A Coastal Ocean Model with Subgrid Approximation Roy A. Walters PII: DOI: Reference:
S1463-5003(16)30010-5 10.1016/j.ocemod.2016.04.005 OCEMOD 1093
To appear in:
Ocean Modelling
Received date: Revised date: Accepted date:
3 January 2016 11 April 2016 16 April 2016
Please cite this article as: Roy A. Walters, A Coastal Ocean Model with Subgrid Approximation, Ocean Modelling (2016), doi: 10.1016/j.ocemod.2016.04.005
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
Highlights • A semi-implicit coastal ocean model balances stability, accuracy, and efficiency.
• The methods are applied to tidal power generation.
CR IP T
• Volume averaging is used to approximate subgrid objects of various kinds.
• Thrust and power coefficients convert from free stream to volume average velocity.
AC
CE
PT
ED
M
AN US
• Three conversion methods are an iterative approach and two analytical mappings.
1
ACCEPTED MANUSCRIPT
1
A Coastal Ocean Model with Subgrid Approximation
2
Roy A. Waltersa,∗ a Ocean-River
4
Hydrodynamics, Victoria, BC, Canada
CR IP T
3
Abstract
A wide variety of coastal ocean models exist, each having attributes that reflect specific application areas. The model presented here is based on finite element methods with
AN US
unstructured grids containing triangular and quadrilateral elements. The model optimizes robustness, accuracy, and efficiency by using semi-implicit methods in time in
order to remove the most restrictive stability constraints, by using a semi-Lagrangian advection approximation to remove Courant number constraints, and by solving a wave equation at the discrete level for enhanced efficiency. An added feature is the approximation of the effects of subgrid objects. Here, the Reynolds-averaged Navier-Stokes
M
equations and the incompressibility constraint are volume averaged over one or more computational cells. This procedure gives rise to new terms which must be approx-
ED
imated as a closure problem. A study of tidal power generation is presented as an example of this method. A problem that arises is specifying appropriate thrust and power coefficients for the volume averaged velocity when they are usually referenced
PT
to free stream velocity. A new contribution here is the evaluation of three approaches to this problem: an iteration procedure and two mapping formulations. All three sets
CE
of results for thrust (form drag) and power are in reasonable agreement. Keywords: coastal ocean, numerical model, unstructured grid, subgrid, volume
6
averaging, tidal power generation, Canada, Nova Scotia, Minas Passage.
AC
5
7
1. Introduction
8
A wide range of physical phenomena can be found in the area between the continen-
9
tal slope and the landward extent of the coastal ocean which may include complicated
10
interconnecting waterways. Among other things, stratification, rotation, strong advec∗ Corresponding author Preprint submitted Elsevier Email address:
[email protected] (Roy A. Walters)
April 20, 2016
ACCEPTED MANUSCRIPT
11
tion, wetting and drying, wave dispersion, scalar transport, and other types of dynamics
12
may be globally or locally important. Hence, it is not surprising that a large number of
13
coastal ocean models exist and they tend to specialize in a specific set of dynamics. The basic set of equations to solve are the well known 3-dimensional version of the
15
shallow water equations (SWE), perhaps with added features for nonhydrostatic flows,
16
dispersive waves, and scalar transport. In the early efforts to approximate these equa-
17
tions, a serious problem was encountered with spurious modes generated by coupling
18
between the pressure gradient in the momentum equations and flux term in the free
19
surface (continuity) equation. For finite difference approximations, this problem was
20
resolved with a staggered grid (C-grid) with structured grids [15] and a more recent
21
staggered grid approximation for irregular orthogonal grids [5]. For finite element ap-
22
proximations with unstructured grids, the problem led to considerably more research
23
before it was adequately resolved.
AN US
CR IP T
14
For the finite element approximations, two main strategies were used: mathemati-
25
cally manipulate the SWE to form a wave equation to replace the free surface equation
26
and use simple linear or p-order elements, or find element approximations that do not
27
support spurious modes [24]. The wave equation approach that is spectral in time pro-
28
vides an accurate and extremely efficient method to simulate tidal flows, particularly
29
in large domains. The main problem that remains is to deal with the compound tides
30
and overtides generated by the nonlinear terms and their influence on the main con-
31
stituents. The wave equation approach that uses time stepping can also be accurate and
32
efficient. The main problem encountered is approximating the advection term because
33
it is now a higher-order derivative, and treating wetting and drying with a node-based
CE
PT
ED
M
24
34
approximation. The examination of various element approximations that do not support spurious
36
modes has evolved over a number of years (see for instance [14]). Initially, mixed
AC
35
37
approximations were used such as linear approximation for sea level and a quadratic
38
approximation for velocity. This strategy required that large, global element matrices
39
be solved with the result that these methods are very inefficient and have run times
40
perhaps a thousand times longer than more recent models.
41
If we fast-forward to the present, a very useful element is the Raviart-Thomas ele3
ACCEPTED MANUSCRIPT
ment of the lowest order (RT0 ) [20]. This element uses a piecewise constant approx-
43
imation for sea level and a linear vector approximation for velocity on triangular and
44
quadrilateral elements. Many studies have shown that this element is accurate, robust,
45
and can lead to efficiencies comparable to wave equation approaches [26, 25, 30, 24].
46
Moreover, this approach has distinct advantages when approximating the advection
47
and Coriolis terms, and greatly simplifies the treatment of wetting and drying. This
48
element approximation is used in the model described in this report, RiCOM – River
49
and Coastal Ocean Model.
CR IP T
42
The next issue to consider is approximating the effects of subgrid objects. The ba-
51
sic problem is that the real world is continuous but numerical models are discretized.
52
Hence there is always some spatial and temporal scale where subgrid features are not
53
resolved. Some examples of subgrid features include submerged and emergent vegeta-
54
tion, arrays of pilings, aquaculture drop lines, and freestanding rotors for tidal power
55
generation. A common characteristic of these features is that they all can be simulated
56
using highly refined CFD models, but they are impractical to fully resolve with a larger-
57
scale coastal ocean model. Hence, some form of subgrid approximation is required. In
58
the end, the resolution is determined by a balance between computer resources and the
59
details of the problem being considered. Nonetheless, there is always some aspect of
60
the problem that cannot be resolved adequately so the question that arises is what can
61
be done about the effects.
ED
M
AN US
50
In the time domain, the problem has been largely resolved through Reynolds aver-
63
aging. The basic Navier-Stokes equations are time-averaged to arrive at the Reynolds
64
equations where the nonlinear terms give rise to averages of fluctuating components
CE
PT
62
known as Reynolds stresses. This approach then leads to a closure problem for the new
66
terms. Considerable research has been devoted to developing closures based on mixing
67
length theory, turbulence submodels, and other higher order closures.
AC
65
68
The averaging concept can be extended to the spatial domain through the use of
69
volume averages [16]. The approach leads to new terms that involve averages of spatial
70
fluctuations and requires closures similar to the time-averaging case. Although much
71
research has gone into this area, much remains to be learned.
72
The new spatially-averaged terms generally lead to some type of form drag and a 4
ACCEPTED MANUSCRIPT
73
dispersive term. The methods that can be used to approximate these terms range from
74
empirical formulations, to combined analytical-empirical formulations, and to coupled
75
large-scale and small-scale models. An example of empirical formulations is the approximation of form drag from
77
emergent and submerged vegetation using relations derived from laboratory experi-
78
ments [31, 29]. Here the drag coefficient can be formulated in terms of vegetation
79
diameter, length, spacing, and density. Reasonably accurate results were obtained in
80
comparing model results with experiments. However, the real world is characterized
81
by high spatial variability in both topography and the distribution of subgrid objects so
82
idealized experimental results may just provide a first guess.
AN US
CR IP T
76
Aquaculture is an example where a combined empirical and analytical approach
84
has been successful. An interesting aspect of this application is that the canopy is
85
suspended from the surface rather than emergent from the bottom of the water column.
86
Taking into account the different physical layers in the vertical, the problem can be cast
87
in terms of a two-dimensional model [19] or in terms of a three-dimensional model
88
[17].
M
83
Coupling large and small scale models seems attractive but presents another set of
90
difficulties. Small-scale computational fluid dynamics (CFD) models and larger-scale
91
coastal ocean (CO) models operate at much different time and space scales. Because
92
of this, it is very difficult to couple them directly due to issues with interface boundary
93
conditions, sub-cycling the CFD model, computational overhead of the CFD model,
94
and other problems. In [21] we develop an alternative approach where volume aver-
95
aging is used to pass information from the CFD model to the coastal ocean model and
CE
PT
ED
89
96
bypass most of these problems. The objective of this paper is to describe the methodology used in developing this
98
coastal ocean model and the approximation for the effects of subgrid objects. The
AC
97
99
approach adopted here is to use a volume average over one or more finite elements
100
(computational cells) and formulate the resultant closures in terms of form drag and
101
spatial dispersion.
102
As an example of this approach, the estimation of power output from free-standing
103
tidal turbines is considered. Currently, turbines for this purpose are being developed 5
ACCEPTED MANUSCRIPT
and tested in several countries and can take many forms, including but not limited to
105
vertical shaft rotors, and ducted and non-ducted impellers. An important question that
106
has arisen, is how to estimate the power that can be extracted for a given site. The
107
answer depends on many factors–the geometry of the site, the tidal range, the number
108
and size of the turbines, and the efficiency of the turbines as a function of flow. The
109
field site that has been selected is Minas Passage in the Bay of Fundy which provides an
110
interesting practical perspective for this problem and represents a single tidal boundary
111
problem that has been examined in [13, 30].
CR IP T
104
In section 2, a description of the model is presented along with the treatment of
113
tidal turbines as form drag. Section 3 contains a description the field site and the model
114
setup. Following that is a presentation of the results and discussion, and conclusions.
115
2. Approach
116
2.1. Basic equations
AN US
112
The starting point in the derivation is the Navier-Stokes equations and continuity
118
equation (incompressibility condition). In the derivation, these equations are first av-
119
eraged in time using traditional Reynolds averaging, then volume-averaged in space
120
using double averaging methods (DAM) described in [16]. The continuity equation is
ED
M
117
∂ρ + ∇3 · (ρu) = 0 ∂t
PT
and the Navier-Stokes equations in a rotating frame of reference are
CE
121
(1)
Du 1 + Ω × u + ∇3 p − ν∇23 u − g = 0 Dt ρ
(2)
where D/Dt is a material derivative, u is velocity, t is time, Ω is the rotation vector
123
which is twice the Earth’s angular velocity, ρ is density which is assumed constant, p is
AC
122
124 125
pressure, ν is kinematic viscosity, g is the gravity vector, and ∇3 is the 3-dimensional gradient operator.
126
¯ + u0 , and Then decompose velocity using a Reynolds decomposition , u = u
127
¯ = h¯ a spatial decomposition, u ui + u00 where the overbar denotes a time average,
128
brackets denote a spatial average, u0 is the velocity deviation in time, and u00 is the 6
ACCEPTED MANUSCRIPT
129
deviation of the time-averaged velocity in space. In addition, let φ = Vf /V be the
130
fluid fraction, where Vf is the volume of fluid in the averaging volume V .
∂φ + ∇3 · (φh¯ ui) = 0 ∂t 132
and the momentum equation becomes
CR IP T
Averaging over time and over space (see [16]), the continuity equation becomes
131
AN US
Dh¯ ui 1 1 ¯ i) − g + Ω × h¯ ui + ∇3 (φh¯ pi) − ∇3 · (φνh∇3 u Dt ρφ φ 1 1 + ∇3 · (φhu0 u0 i) + ∇3 · (φhu00 u00 i) φ φ I I 1 1 ¯) · n + p¯n ˆ dS − (ν∇3 u ˆ dS = 0 ρVf Vf
(3)
(4)
Some new terms have been created by averaging the nonlinear terms and are similar
134
in concept to the derivation of Reynolds stresses through time averaging. In eq. 4,
135
the sixth term is the spatially averaged Reynolds stress, the seventh term is a spatial
136
dispersion term, and the last two terms are form drag and skin friction. Because form
137
drag generally dominates skin friction, the last two terms can be written as a drag force
ED
M
133
fF D = (1/2)ρAf Cd |ur |ur
(5)
where Af is frontal area of an object measured in the plane that is normal to the di-
139
rection of flow, Cd is the non-dimensional drag coefficient, and ur is the reference
140
velocity. The properties of the spatial dispersion term are not known in any detail but
141
appear to control the wake mixing behind the subgrid object. These issues are discussed
142
in the results section.
CE
PT
138
Next, the equations are cast in the form of the 3-dimensional shallow water equa-
AC
143 144
145
tions by using the hydrostatic assumption. Rewriting the continuity equation as ∂Vf + ∂t
I
h¯ uidS = 0
and using the hydrostatic approximation in the momentum equation
7
(6)
ACCEPTED MANUSCRIPT
+ − +
146
1 g∇(φh¯ η i) φ ¯ 1 ∂ ∂u 1 (φhAv i) − ∇ · (φhAh ∇¯ ui) φ ∂z ∂z φ 1 ∇3 · (φhu00 u00 i) − FD = 0 φ
ˆ × h¯ fz ui +
CR IP T
Dh¯ ui Dt
(7)
where ∇ is the horizontal gradient operator, Reynolds stresses are written using the
147
eddy viscosity concept with coefficients Av and Ah , and FD contains the form drag
148
and friction terms in the last line of eq. 4.
In the computational domain where there are no subgrid objects, φ = 1 and the
150
equations reduce to the standard SWE with additional terms from the volume average.
151
In addition, for the types of coastal ocean problems considered here, φ ≈ 1 and is
AN US
149
152
constant so the equations can again be cast into a more traditional form where the
153
continuity equation is
∂w =0 ∂z
(8)
and the momentum equation expressed in nonconservative form is
ED
154
M
∇·u+
∂ Du ∂u ˆ × u + g∇η − + fz (Av ) − ∇ · (Ah ∇u) + F = 0 Dt ∂z ∂z
(9)
where all variables are time and space averaged so the overbar and brackets can be
156
dropped, the coordinate directions (x, y, z) are aligned in the east, north, and ver-
157
tical directions; u(x, y, z, t) is now the horizontal velocity with components (u, v);
158
ˆ is the upward w(x, y, z, t) is the vertical velocity; f (x, y) is the Coriolis parameter; z
159
unit vector; η(x, y, t) is the distance from the reference elevation to the free surface;
160
g is the gravitational acceleration; Av (x, y, z, t) and Ah (x, y) are the kinematic verti-
AC
CE
PT
155
161
cal and horizontal viscosity, respectively; ∇ is the horizontal gradient operator (∂/∂x,
162
∂/∂y); and F contains body forces that include form drag, skin friction, and spatial
163
dispersion terms derived above.
164
The Reynolds stress terms have been written in terms of an eddy viscosity coeffi-
165
cient and are incorporated into Av and Ah . These terms must be approximated through 8
ACCEPTED MANUSCRIPT
166
a turbulence closure submodel or by empirical methods. Note that because of volume
167
averaging, the stress terms should also be volume averaged. This is an active research
168
subject and more information may be found in [7]. For the most part, then, the time and volume averaged equations reduce to the
170
standard SWE with additional terms that account for the effects of the subgrid object
171
on the flow. Locally where the fluid volume fraction, φ, varies significantly, the more
172
exact averaged equations can be used.
The free surface equation is derived by vertically-integrating the continuity equa-
173 174
CR IP T
169
tion and using the kinematic free surface and bottom boundary conditions [18].
AN US
Z η ∂η +∇·( udz) = 0 ∂t h
(10)
Boundary conditions for Eqs. 8 to 10 generally fall into two categories: conditions
176
at open (sea) boundaries and conditions at solid (land) boundaries. At open bound-
177
aries sea level η, radiation conditions, or a combination of these are generally set. In
178
addition, discharge may be specified for river or other inflow. At land boundaries, the
179 180
M
175
b ) = 0 where n b is the unit normal. normal component of velocity vanishes so that (u · n If the flow is viscous, then stress conditions also need to be specified. Either a stress
condition is specified by Eq. 11 or u|h = 0 and the bottom boundary layer needs to be
182
resolved by the vertical grid placement.
ED
181
The 2-dimensional vertically-averaged form of Eq. 9 is similar with u replaced by
184
the depth-averaged value and the vertical component of stress (fourth term) replaced
185
by (τ s − τ b )/ρH where τ b is the bottom stress, τ s is the surface stress which is
neglected here, ρ is a reference density, and H(x, y, t) is the total water depth given by
CE
186
PT
183
187
H = η − h where h(x, y) is the bottom elevation measured from a reference elevation.
For the derivation, see the development of the depth-averaged equations in [18] or
189
many other publications.
AC
188
190
191
The bottom stress τ b is given by τb = Cb |u|u = γu ρ
(z = h),
where Cb is a bottom friction coefficient and γ is defined by this equation. 9
(11)
ACCEPTED MANUSCRIPT
192
2.2. Form drag Form drag that is modeled as a subscale process (eq. 5) can be included into the
194
momentum equation in a general sense with the body force term F in Eq. 9. Then F
195
includes Fd which is the form drag per unit volume which is given by
Fd = (1/2)λf Cd |u|u
CR IP T
193
(12)
where λf is the frontal area Af of the object divided by the averaging volume. Note that
197
both Af and the averaging volume are important for scaling from the discrete turbine
198
scale to the larger-scale model. Moreover, some care must be taken as the volume-
199
averaged model velocity in eq. 12 will not be the same as the reference velocity in eq.
200
5.
AN US
196
From a model perspective, the obvious averaging volume is the volume of a compu-
202
tational cell- an element volume when using finite element methods. In the following,
203
A and V will refer to the horizontal area and volume of the computational cell.
204
2.2.1. 3D case
M
201
Expand Af = W ∆z and V = A∆z in the expression for λf so that
205
206
where W (z) is width of the object and A is element area at depth z.
207
2.2.2. 2D case
The 2-dimensional equations are derived by depth integrating the 3-dimensional
PT
208
¯ d is the depth-averaged form drag approximated as equations with the result that F
CE
209
211
212 213
¯ d = (1/2)λf Cd |U|U = (cf d /H)|U|U F
is non-dimensional. Hence, the form drag coefficient and bottom friction coefficient in a 2-dimensional
formulation can be merged to give τ /(ρH) = [(Cb + cf d )/H]|U|U
214
(14)
where U is the depth-averaged velocity and cf d = (1/2)(Af /A)Cd by definition and
AC
210
(13)
ED
Fd = (1/2)(W/A)Cd |u|u
where Cb is the non-dimensional bottom friction coefficient. 10
(15)
ACCEPTED MANUSCRIPT
215
2.3. Discretization and Solution Methods The objective of the time and space discretization methods is to optimize the model
217
for accuracy, stability, and efficiency. The momentum and free surface equations are
218
discretized in time using semi-implicit methods in order to remove the most restrictive
219
stability constraints on gravity waves and vertical stress [2, 4]. The material derivative
220
in the momentum equation is discretized using semi-Lagrangian methods to remove
221
stability constraints on advection [2, 27]. The Coriolis term is discretized with a 3rd-
222
order Adams-Bashforth scheme to provide stability [28]. In addition, the spatial dis-
223
cretization uses an unstructured grid with either triangular or quadrilateral elements for
224
flexibility in representing irregular multiply-connected domains [25]. More details are
225
provided in the subsections that follow.
226
2.3.1. Time Discretization
Using the θ-method, the discretized equations for the free surface and momentum
227
and
Z η Z η η n+1 − η n n+1 n + ∇ · θ( u dz) + (1 − θ)( u dz) = 0 ∆t h h
(16)
ED
229
can be written as [26]
M
228
AN US
CR IP T
216
PT
un+1 − u∗ ∆t
ˆ × un + g∇(θη n+1 + (1 − θ)η n ) + fz −
∂ ∂un+1 (Av ) − ∇ · (Ah ∇un ) + Fn+1 = 0 ∂z ∂z
(17)
where n denotes the time level tn = n∆t, u∗ is the velocity at the foot of the La-
231
grangian trajectory, the vertical component of stress and form drag are treated implic-
232
itly, and quadratic stress terms in bottom friction and form drag are approximated as
CE
230
(|u|u)n+1 = |un |un+1 .
AC 233
234
For θ > 1/2, the gravity wave terms are stable [4]. The remaining terms with
235
stability constraints are the horizontal stress and the Coriolis terms. In practice, the
236
horizontal stress is only used when resolved horizontal boundary layers are present.
237
In that case, the term is sub-stepped in time if necessary in order to satisfy a stability
11
ACCEPTED MANUSCRIPT
constraint [12]. As pointed out in [11], the stability of the Coriolis term has 2 aspects –
239
one relating to the time discretization and one to the spatial discretization (next section).
240
The standard explicit treatment of this term is not stable [28] and the instability is
241
masked when significant dissipation is present, but becomes evident once beyond the
242
shelf break. For our purposes, a stable 3rd-order Adams-Bashforth scheme is used
243
[28]. Other models [6] have had success with a 2nd-order Adams-Bashforth scheme
244
with modified coefficients.
245
2.3.2. Spatial Discretization
CR IP T
238
In each finite element, the dependent variables for sea level, η, and normal velocity
247
on an edge, un , are approximated as a constant and a linear function, respectively.
248
This particular element is known as the Raviart-Thomas element of lowest order [20].
249
Hence sea level is a piecewise constant globally and velocity is a piecewise linear
250
function globally. The approximation for velocity is somewhat unusual in that un is a
251
scalar whereas the approximation function is a vector.
AN US
246
This particular approximation method has a number of important advantages.
253
• With piecewise constant η, the free surface and continuity equations become
255
finite volume expressions that conserve mass exactly.
ED
254
M
252
• Normal velocity is continuous on each edge which simplifies the calculation of fluxes and Lagrangian trajectories. There is no need to evaluate a discontinuous
257
normal velocity on an edge such as with Discontinuous Galerkin Methods.
258
• All terms in the equations can be calculated directly from the velocity approximation. In particular, the Coriolis term is calculated directly from a vector prod-
CE
259
PT
256
260
uct between the unit vector in the vertical and the vector basis functions for
261
velocity (Fig. 2 in [26]). Thus, there is no need to reconstruct the full velocity field such as with irregular grid finite difference methods that use circumcenters
263
in an orthogonal grid.
AC 262
264
• There is no constraint on element shape except to satisfy accuracy requirements
265
(equilateral elements are most accurate). There is no limit on element corner
266
angles. 12
ACCEPTED MANUSCRIPT
267
On the other hand, there is a serious shortcoming to this approach. A direct ap-
268
plication of the Galerkin finite element methodology leads to element mass matrices
269
that are fully populated nv × nv matrices, where nv is the number of vertices in an element. After assembly, a global sparse matrix for velocity must be solved which re-
271
sults in a very inefficient model. Fortunately, there are at least 2 ways to diagonalize
272
the matrix without reducing the accuracy significantly. One of these ways, suggested
273
in [1], results in a diagonal matrix with the distance between adjacent circumcenters
274
on the diagonal; the other way results in a diagonal matrix with the normal distance
275
to an edge between adjacent centroids on the diagonal. The latter is similar in concept
276
to the Integrated Compartment Method proposed long ago [32] and reduces to the cir-
277
cumcenter formulation for equilateral and square elements. This centroid approach is
278
used in the model described here. A more detailed description of these approaches is
279
found in [26].
AN US
CR IP T
270
In the horizontal plane, the model is discretized as described above. In the ver-
281
tical, the element edges are defined by vertical lines extending from the nodes at the
282
free surface to the bottom and the elements are brick- or pie-shaped depending on
283
whether quadrilateral or triangular elements are used in the horizontal. The vertical
284
discretization uses terrain-following σ coordinates. Vertical velocity is approximated
285
as a piecewise constant on the upper and lower faces of each element. Horizontal ve-
286
locity can be discretized at the midpoint on the horizontal edges of each element or
287
the face-centers of the layers. The former uses linear basis functions so the momentum
288
equation reduces to a standard finite element tridiagonal system along each vertical line
289
connecting the midpoint of the edges. In the case of layers, velocity is constant on each
CE
PT
ED
M
280
290
face and a finite difference approach is used to form a tridiagonal system. In this study,
291
layers are chosen since they are more compatible with volume averaging. Using these methods, the discretized equation for the free surface can be written as
AC
292 293
294
[26]
Ae
Z Z Nes Nes X X ηen+1 − ηen = −θ li ( un+1 dz) − (1 − θ) l ( unj(i) dz) i j(i) n n ∆t H H i=1 i=1 j(i) j(i)
(18)
where subscript e = 1, . . . , Ne is the element index, Ne is the number of elements, 13
ACCEPTED MANUSCRIPT
Ae is the area of element e in the x-y plane, li is the edge length of side i, Nes is the
296
number of element sides, uj(i) is the normal velocity on edge j defined by side i of
297
element e, H is the water depth, and a suitable convention is used to determine normal
298
direction uniquely [5]. This equation states that the volume change at the surface is the
299
net volumetric flux through the sides.
CR IP T
295
For this element approximation, there is one degree of freedom for sea level η on
301
each element and one degree of freedom for the normal component of velocity uj on
302
each edge j. Hence it is more convenient to express the momentum equation in terms
303
of the normal velocity component. Since (17) is invariant under solid rotation in the
304
horizontal plane, the momentum equation for the normal velocity component can be
305
written
AN US
300
un+1 − u∗j ∂ ∂η n+θ ∂uj n+1 j − (Av ) + Fjn+1 = −g + Cjn + ∇ · (Ah ∇u)nj (19) ∆t ∂z ∂z ∂n where uj is the normal component of velocity positive outwards on side j, the body
307
force F n+1 is evaluated as γ|un |un+1 where γ are the coefficients used, Cjn is the
M
306
Coriolis term evaluated on side j, the superscript denotes the time level, η n+θ =
309
θη n+1 + (1 − θ)η n , and ∂/∂n is the gradient operator normal to the cell edge. Note
ED
308
that horizontal friction is generally neglected because it is usually small compared to
311
vertical friction. If it is included, the discretization follows standard methods used in
312
FEM. For the example presented later, horizontal friction is neglected.
PT
310
The Coriolis term is evaluated with the 3rd-order Adams-Bashforth scheme as
CE
313
314
316
317
318
f (23vjn − 16vjn−1 + 5vjn−2 ) 12
(20)
where vjn ≡ (ˆ z × u)nj is defined from the corresponding term in eq. 17 and can
be thought of as the tangential component of velocity positive in a counter-clockwise
AC
315
Cjn =
sense around each element. After assembly of all the element contributions, the discretized momentum equa-
tion for the normal velocity on a side can be written as [26] Aun n+1 = G − gθ∆t∆η n+1 Z 14
(21)
ACCEPTED MANUSCRIPT
where matrix A contains all the implicit contributions on the left hand side of (19), un
320
is a vector of the normal velocity components uj on the edges, G is a vector containing
321
the explicit terms in (19), ∆η n+1 is the difference between element values adjacent to
322
an edge, and Z is a vector of depth increments. These terms are defined in more detail
323
in [26]. In general, A is a large sparse matrix that is block tridiagonal at each velocity
324
node in the horizontal grid. This matrix is inverted by diagonalizing in the horizontal
325
as described earlier and solving the tridiagonal system in the vertical at each node [26].
326
2.3.3. Solution Method
CR IP T
319
In general, the system of equations given by (18) and (19) are solved by first solving
328
for un by solving the tridiagonal system in (19) and substituting this result into (18) to
329
form a discrete wave equation that involves only η at the n + 1 time level. This equa-
330
tion is positive definite and diagonally dominant, and is solved by standard conjugate
331
gradient methods.
AN US
327
However, the treatment of wetting and drying can be improved by writing the free
333
surface equation (Eq. 18) in finite volume form and adopting the method proposed by
334
Casulli [3]. Then the momentum equation is solved for un n+1 and substituted into the
335
finite volume free surface equation, so that the resulting discrete wave equation can be
336
written in compact vector form as [3]
V(ζ) + Tζ = b
(22)
PT
ED
M
332
337
where ζ is a vector of sea level with components ζi = ηin+1 , V is a vector of volumes
338
with components Vi (ηin+1 ), and T is the flux matrix arising from the second term in (18). As shown by Casulli [3], an efficient Newton-type algorithm for solving Eq. 22
340
is given by
AC
CE 339
341
342
343
[P(ζ m ) + T]∆ζ m+1 = [V(ζ m ) + Tζ m − b]
where m is the iteration index, P is the wet area, and ∆ζ m+1 = ζ m+1 − ζ m .
(23)
Once ζ is known, the values for η n+1 are backsubstituted into eq. 21 to solve for
un n+1 .
15
ACCEPTED MANUSCRIPT
344
3. Tidal Power Example From the development thus far, the tidal power problem has been reduced to finding
346
the appropriate values for the drag coefficient Cd defined by a known reference veloc-
347
ity, and a suitable closure for the spatial dispersion term in (4). From the perspective of
348
tidal turbines used for power generation, two coefficients are necessary: a thrust coef-
349
ficient CT (which is equivalent to a drag coefficient) and a power coefficient Cp , along
350
with a useable reference velocity. The double-averaged velocity produced by the model
351
is the obvious choice for reference velocity. What remains then is how to determine
352
the values for the coefficients based on this reference velocity. The spatial dispersion
353
term appears to be mainly significant in the wake region and is not considered in this
354
example.
In the context of turbines, the drag coefficient, Cd , is written as the thrust coeffi-
355 356
AN US
CR IP T
345
cient, CT . Using this notation, the drag force and power can be written (24)
357
where At is the frontal area of the turbine normal to the direction of flow, and ur is the
358
free stream reference velocity, and
359
ED
M
F = (1/2)ρAt CT |ur |ur
where ur is the corresponding free stream reference speed. Drag from the structure
360
supporting the turbine as well as low speed cut-in and high speed cut-off are neglected.
361
For an actuator disk in a free stream flow, values for the drag and power coeffi-
(25)
CE
PT
P = (1/2)ρAt Cp ur 3
cients can be derived using conservation of mass and momentum. The coefficients
363
at maximum energy extraction were derived independently by Lanchester (1915), by
364
Betz (1920), and by Zhukowsky (1920) and are generally known as the Lanchester-
AC
362
365
Betz limit. For this case, CT = 8/9 and Cp = 16/27 where the reference velocity
366
is the free stream velocity. Garrett and Cummins [9] also derive these results and in
367
addition derive a correction term for flows constrained by lateral boundaries. For that
368
case, the power can be larger due to the increased pressure drop across the turbines.
369
In practical terms, CT and Cp are defined for various turbines by their manufacturers. 16
ACCEPTED MANUSCRIPT
Typical values for CT range from 0.4 to 0.8 and for Cp range from 0.3 to 0.5 based
371
on free stream velocity as the reference velocity. For a simple straight channel with
372
constant depth and width and discharge, the free stream velocity is defined upstream.
373
However, for more complicated geometry and transient flows, defining a free stream
374
velocity is not possible in general. Additional complications arise with shear flows
375
such as the vertical shear in tidal flows and the determination of averaging volume for
376
multi-turbine deployments.
CR IP T
370
In most cases, the drag coefficient for various geometric objects and thrust coeffi-
378
cients for turbines use free stream velocity as the reference. Then the relation between
379
free-stream and volume-averaged velocity needs to be determined and the coefficients
380
adjusted accordingly. We can define a new set of coefficients CT∗ and Cp∗ as
AN US
377
F = (1/2)ρAt CT∗ |U|U 381
where U is the volume-averaged velocity, and
M
P = (1/2)ρAt Cp∗ U 3
382
(26)
(27)
where U is the volume-average speed.
Note that the model velocity is discretized on the edges so care must be exercised
384
to ensure that the volume average of the edge-based equations is equal to the volume-
385
averaged equations. For the linear terms this is a straightforward average. The ad-
386
vective fluxes through the edges are defined the same in both cases. The only term
387
needing special attention is the form drag term in eq. 26. Approximating the term as
388
(1/2)ρAt CT∗ |U |u on each edge results in an equivalent approximation for F.
CE
PT
ED
383
389
In a previous study [21], some of the issues with volume-averaged velocities were
390
addressed by using a detailed CFD model to calculate the flow around a turbine and then volume-averaging the resultant velocity to determine the parameters to be used in
392
a larger-scale coastal ocean model (COM). A few of the more important results that
393
relate to the present study are
AC 391
394
1. Both CT∗ and Cp∗ were robust parameters based on using volume-averaged veloc-
395
ity as the reference velocity. That is, they were essentially constant as a function 17
ACCEPTED MANUSCRIPT
396
of velocity when operating at maximum power generation. Hence, these param-
397
eters could be determined for the maximum velocity and used over the entire
398
range of tidal velocity. 2. The averaging volume must be at least twice the turbine diameter on each edge
400
to maintain accuracy between the small-scale CFD model and large-scale COM.
401
In a physical sense, the volume needs to be larger than the detailed flow varia-
402
tions generated around the turbine because the latter model cannot resolve them
403
nor does it contain the right physics at the turbine scale. One implication is
404
that the vertical variation in velocity for the COM must be averaged over the
405
total averaging-volume because the grid spacing in the vertical is usually much
406
smaller that the turbine diameter. Toward this end, local coefficients in each layer
407
are defined from the local velocity in each layer such that the thrust and power
408
sum to the correct total for the turbine.
AN US
CR IP T
399
3. Ultimately, the consistency between the CFD model and COM was dependent on
410
how accurately the COM could produce the same volume-averaged velocity as
411
the CFD model. For single turbines in a simple channel geometry, the agreement
412
was good. For more complicated field-scale geometry, differences in physics and
413
turbulence closures between the models led to greater errors.
ED
M
409
In the following section which describes a field-scale problem, other options are
415
considered for determining coefficients when detailed CFD calculations may or may
416
not be available. This example uses the same grid and sites as were used in [21]. The
417
location is Minas Passage in the upper Bay of Fundy.
PT
414
3.1. Minas Passage
CE 418
Minas Passage connects Minas Basin with the upper Bay of Fundy (Fig. 1) and is
420
well known for its large tidal currents and range. This area has considerable potential
AC
419
421
for tidal power generation and has been the subject of many studies from a theoretical
422
and practical perspective.
423
The model domain includes the Gulf of Maine, Bay of Fundy, Minas Passage,
424
and Minas Basin (Fig.1). This area is located on the east coast of North America
425
between approximately 44◦ and 46◦ North latitude. The resonant period of this system 18
M
AN US
CR IP T
ACCEPTED MANUSCRIPT
ED
Figure 1: Geometry and bathymetry of Bay of Fundy. AO is Atlantic Ocean, GB is Georges Bank, GoM is Gulf of Maine, BoF is Bay of Fundy, MP is Minas Passage, MB is Minas Basin, B is Boston, and CC is Cape Cod.
is slightly longer than the period of the dominant tidal constituent, M2 , [8, 10] and this
427
area is well known for the large tide range in the upper bay. The node point for the
428
oscillation is near the shelf break east of Georges Bank; however, this is complicated
CE
PT
426
429
by other resonances parallel to the shelf that extend down to Boston and Cape Cod. The dominant tidal constituent is the M2 and the amplitude varies from 0.5 m at
431
the open boundary to over 6 m in Minas Basin. In a previous study [30], 8 constituents
AC
430
432
are used in the open boundary conditions– M2 , S2 , N2 , K2 , K1 , O1 , P1 , and Q1 .
433
Most of these constituents are small compared to the M2 (particularly the diurnal con-
434
stituents) but were retained in order to assess the accuracy of the model against tide and
435
current observations. In general, the model reproduced the observed tidal harmonics
436
and ADCP current meter observations so that the model is sufficiently accurate for the 19
ACCEPTED MANUSCRIPT
437
evaluation of tidal power potential [30]. During flood tide, a flow separation develops at Cape Split at the entrance to Minas
439
Passage with a high velocity core in the northern part of the passage and a counter eddy
440
in the southern part (Fig. 2). Peak velocity approaches 4m/s in the deep part of the
441
passage outside the eddy. The reattachment length of the eddy and its strength increase
442
until peak flood.
PT
ED
M
AN US
CR IP T
438
CE
Figure 2: Minas Passage at flood tide. CS identifies Cape Split. denote the turbine deployment sites. Velocity vectors have constant length and are decimated by 5.
During ebb tide, the velocity is more uniform across the passage and an eddy devel-
AC
443
444
ops southwest of Cape Split. At the start of the ebb tide, the eddies that were created
445
during the flood are advected out Minas Passage. The passage of these eddies is re-
446
peatable over the tidal cycles and can be seen in the current meter data as well as the
447
model results.Overall, the flood/ebb behavior is consistent with other published results
448
[10, 22]. 20
ACCEPTED MANUSCRIPT
The original model grid was constructed from unstructured triangles that vary in
450
edge length from 12 km on the open boundary to 70 m in Minas Passage. This grid
451
was modified by inserting four 200 m by 200 m square grids at the simulated turbine
452
deployment sites. The square grids are composed of either a 3x3 or 5x5 array of square
453
elements (66.7 m or 40 m edge length). The turbine was placed at the center of the
454
square grids (Fig. 3). This particular grid configuration was created so that CFD cal-
455
culations could be made on the square grids. This same grid is used in the results
456
presented here.
CR IP T
449
A single turbine was placed at each test site. All the turbines were 16 m in di-
458
ameter and placed 15 m from the bottom of the water column. Note that each set of
459
square grids was rotated to align with the maximum flood velocity to simplify the CFD
460
calculations in a previous study.
AC
CE
PT
ED
M
AN US
457
Figure 3: Grid around the turbine sites in Minas Passage.
21
ACCEPTED MANUSCRIPT
For the present example, the small constituents are neglected and only the M2
462
is retained. In total, there are 69583 vertices and 135101 triangular and quadrilateral
463
elements. On a desktop computer using openmp with 5 cores, the model runs 270 times
464
faster (2D) and 60 times faster (3D) than simulated time. The calculations presented
465
for this study are 2D and a set of 3D calculations can be found in [21].
466
3.2. Estimation of coefficients
CR IP T
461
The estimation procedures adopted here use the velocity in the absence of a turbine
468
at the turbine site as an approximation for the free stream velocity. Then the coefficients
469
CT and Cp for the Lanchester-Betz limit are used to calculate thrust and power where
470
the reference velocity is the approximation for free stream velocity. From the model
471
perspective, the volume-averaged velocity is smaller than the free stream velocity when
472
form drag is applied so the coefficients must be adjusted to recover the same thrust and
473
power values. This adjustment is made in two ways: (1) the coefficients are adjusted
474
iteratively until the thrust and power from the initial estimate is reached, and (2) the
475
coefficients are mapped from a free stream reference to a cell averaged velocity using
476
two analytical approaches.
M
AN US
467
The general procedure is to spin up the model without form drag until it reaches a
478
periodic steady state. At a time near maximum flood currents at the turbine sites, a ref-
479
erence thrust and power are calculated based on the approximate free stream velocity
480
(case 0). In case 1, form drag from the turbines is introduced and the model is spun
481
up to the same time as before. Thrust and power are calculated based on the volume-
482
averaged velocity which gives values that are smaller than the free stream calculations.
483
The coefficients CT and Cp are then modified iteratively to estimate CT∗ = CT u2f s /u2m
484
and Cp∗ = CT u3f s /u3m where uf s is the free stream speed and um is the volume aver-
485
aged speed at iteration m. Iterations were continued until the errors in thrust and power
CE
PT
ED
477
were small (< 1%). Typically one or two iterations are sufficient. In case 2 and 3, the
487
coefficients CT and Cp are mapped into CT∗ and Cp∗ and thrust and power calculations
488
are compared to the reference calculations.
AC 486
489
In this study, the turbines are assumed to be omnidirectional; that is, the incident
490
flow is always normal to the face of the turbine. For a strongly bi-directional flow such
22
ACCEPTED MANUSCRIPT
491
as encountered here, the differences with fixed turbines are small and occur mainly
492
when the tide is turning and the currents are weak. In the first case using the Lanchester-Betz limit, CT = 8/9 and Cp = 16/27 are
494
referenced to the free stream velocity. The resulting values for thrust and power at the
495
4 sites is listed in Table 1. The values in the table are taken at 88.5 hrs from the start of
496
the simulation, a point near maximum flood currents.
CR IP T
493
Table 1: Coefficients, thrust, and power for case 1. The values for CT , Cp , T , and P are based on free stream velocity. CT∗ and Cp∗ are referenced to the volume-averaged velocity.
CT 0.8889 0.8889 0.8889 0.8889
Cp 0.5926 0.5926 0.5926 0.5926
CT∗ 1.0064 0.9222 0.9056 0.9904
Cp∗ 0.7139 0.6262 0.6094 0.6969
T (kN) 829 717 693 759
P (kW) 1678 1354 1287 1472
AN US
site 1 2 3 4
Generally a monotonic convergence would be expected because of negative feed-
498
back between changes in the coefficients and thrust and power. However, that was not
499
the case at site 1, indicating that there is a weak interaction between the turbines even
500
with the large spacing between them. The two upstream turbines (2,3) change the flow
501
slightly and affect the two downstream turbines (1,4). Here, the upstream direction is
502
taken at maximum flood velocity which is the time when maximum power is gener-
503
ated. In retrospect, a better estimate of free stream velocity at a particular site would
504
be when the other three turbines were in place. Recalculations result in larger thrust
505
and power at site 3 (≈ 3%) and 4 (≈ 4%), and smaller changes at sites 1 and 2 (Table
506
2). For the present case, this new estimate results in marginal changes compared to
507
the first estimate but is used as the reference thrust and power in this analysis. Note
508
that as the number of turbines increases, this procedure becomes more time consuming
CE
PT
ED
M
497
and tedious since a separate simulation is needed to define the approximate free stream
510
speed at each site.
AC 509
511
The values for CT∗ and Cp∗ follow a trend that has been defined previously in sim-
512
ulations with a single turbine in a simple channel flow. For a specific device with a
513
known thrust and power, the volume-averaged velocity becomes smaller as the averag-
23
ACCEPTED MANUSCRIPT
Table 2: Coefficients, thrust, and power for case 1. The values for CT , Cp , T , and P are based on free stream velocity with the other 3 turbines in place. CT∗ and Cp∗ are referenced to the volume-averaged velocity.
CT 0.8889 0.8889 0.8889 0.8889
CT∗ 1.0021 0.9221 0.9235 1.0266
Cp 0.5926 0.5926 0.5926 0.5926
Cp∗ 0.7094 0.6262 0.6309 0.7355
T (kN) 822 717 700 771
P (kW) 1659 1353 1307 1507
CR IP T
site 1 2 3 4
ing volume becomes smaller. Hence for the same thrust and power, the coefficients CT∗
515
and Cp∗ are larger with smaller averaging volume. As a rough measure, the coefficients
516
increase by approximately 4 % when a square element with edge length of 4 times the
517
turbine diameter is reduced to 2 times the diameter. There is a similar trend when depth
518
is reduced. In the present case, site 1 and 4 are contained in elements that are 40 m
519
on an edge and in water depths of 36 and 37 m. On the other hand, sites 2 and 3 are
520
contained in elements that are 66.7 m on an edge and in water depths of 55 and 48 m,
521
respectively. Using the rough scaling above, an increase of 8 % in the coefficients from
522
sites 2 and 3 to sites 1 and 4 is reasonable.
M
AN US
514
In the second case, the coefficients CT and Cp are mapped to CT∗ and Cp∗ using a
523
525
ED
CT = 4a(1 − a)
(28)
where a = (u0 − ut )/u0 is the axial induction factor, u0 is free stream speed and ut is the speed at the turbine disk. To relate u0 to ut , this expression can be written as
CE
526
method from [23]. Using actuator disk theory, CT is written as
PT
524
u0 =
(1 +
√
2 ut 1 − CT )
(29)
Waldman et al. [23] assume that this relation also holds for relating the free stream
AC
527
528
velocity to the area averaged velocity u in the grid cell in the plane of the turbine so
529
that
u0 =
(1 +
√
2 u 1 − CT )
24
(30)
ACCEPTED MANUSCRIPT
where = At /Ac corrects for the fact that the disk only occupies a part of the cross
531
section. Ac is the cross section area of the computational cell in the plane of the turbine.
532
Then eq. 30 can be substituted into eqs. 24 and 25 where u0 = ur . Comparing with
533
eqs. 26 and 27, the coefficients can be corrected to use the average velocity, resulting
534
in
Cp∗ = Cp
and
535
2 √ 1 + 1 − CT
2
2 1 − CT
3
√
AN US
= CT
CT∗
CR IP T
530
1+
(31)
(32)
536
These equations apply to the case where is small so that Ac is sufficiently large
537
that it contains all the flow variations induced by the presence of the turbine. In a
538
previous study [21], we found that accuracy deteriorated when a edge of the averaging
539
cube was less than twice the turbine diameter.
The calculated coefficients, thrust, and power for case 2 are listed in Table 3. Com-
541
paring this table with the results after numerical convergence in Table 2, the values for
542
the coefficients as well as thrust and power are all smaller. Sites 2 and 3 have the largest
543
elements that are 66.7 m on an edge or 4.1 times the turbine diameter and hence have
544
the largest Ac and smallest . At these sites the results are in reasonable agreement with
545
the other results with differences less than 1 or 2 %. Sites 1 and 4 have the smallest
546
elements that are 40 m on an edge or 2.5 times the turbine diameter and have smaller
547
values for Ac and larger . The differences at these sites are larger, approximately 5 %
548
for thrust and 8 % for power. Past experience would indicate that the differences would
549
increase significantly at even smaller averaging volumes.
CE
PT
ED
M
540
In the third case, another mapping is derived from continuity requirements between
AC
550 551
the upstream flow and the sum of the flow through and bypassing the turbine. From
552
continuity
Ac u = (Ac − At )ub + At ut
25
(33)
ACCEPTED MANUSCRIPT
Table 3: Coefficients, thrust, and power for case 2. The values for T and P are based on free stream velocity. CT∗ , Cp∗ , T ∗ , and P ∗ are referenced to the volume-averaged velocity.
CT∗ 0.9438 0.9108 0.9141 0.9458
Cp∗ 0.6483 0.6147 0.6180 0.6504
T ∗ (kN) 784 709 699 730
P ∗ (kW) 1543 1332 1303 1388
T (kN) 822 717 700 771
P (kW) 1659 1353 1307 1507
CR IP T
site 1 2 3 4
In the limit as becomes small, the bypass flow ub can be approximated as the free
554
stream flow. Then a relation between the cell averaged and free stream flow can be
555
derived as
AN US
553
−1 p u u0 = 1 − (1 − 1 − CT ) 2
(34)
−2 p CT∗ = CT 1 − (1 − 1 − CT ) 2
(35)
−3 p Cp∗ = Cp 1 − (1 − 1 − CT ) 2
(36)
and
ED
557
M
and the coefficients are corrected as
556
These corrections result in larger coefficients and hence larger thrust and power
559
(Table 4). In general, these results are in better agreement with the reference thrust and
560
power from case 1. The differences in thrust range from 3 % at sites 1 and 4 with 40 m
561
edge lengths to less than 1 % at sites 2 and 3 with 66.7 m edge length. Differences in
CE
PT
558
562
power range from 5 % to less than 1% at the same sites. In comparing the time series of power at the 4 sites, 2 of the sites (1 and 4) have
564
relatively large differences between the cases and 2 of the sites (2 and 3) have relatively
AC
563
565
small differences between the cases. For comparison, the reference calculation (case 0),
566
the results after numerical convergence (case 1) and the 2 results after mapping (case 2
567
and 3) are compared at sites 1 and 3 (Figure 4). Power at site 1 (the upper set of lines)
568
displays the largest differences in the 4 results, whereas power at site 3 (the lower set of
569
lines) displays the smallest differences. The results at site 4 are similar to those at site 26
ACCEPTED MANUSCRIPT
Table 4: Coefficients, thrust, and power for case 3. The values for T and P are based on free stream velocity. CT∗ , Cp∗ , T ∗ , and P ∗ are referenced to the volume-averaged velocity.
CT∗ 0.9703 0.9217 0.9264 0.9735
Cp∗ 0.6759 0.6257 0.6305 0.6792
T ∗ (kN) 801 716 704 739
P ∗ (kW) 1594 1349 1318 1415
T (kN) 822 717 700 771
P (kW) 1659 1353 1307 1507
CR IP T
site 1 2 3 4
1 but with slightly smaller amplitude. The results at site 2 are similar to those at site 3
571
but with larger values for power. The difference in results at the various sites probably
572
gives a good sense of the variability between different approximation methods. What
573
is really needed is a rigorous comparison between data from an installed turbine and
574
numerical model simulations using the different approximation methods. I am unaware
575
of any such data that is available.
576
4. Concluding Remarks
M
AN US
570
In the formulation of this coastal ocean model, an attempt is made to optimize the
578
model for accuracy, efficiency, and robustness. Toward this end, a semi-implicit time
579
integration removes all the major stability constraints, a semi-Lagrangian (SLM/ELM)
580
advection approximation removes Courant number restrictions, a finite element dis-
581
cretization gives a compact stencil, and solution of a wave equation at the discrete level
582
results in good efficiency. The advection approximation uses quadratic interpolation at
583
the foot of the trajectory which has negligible dispersion compared to linear interpola-
584
tion methods.
CE
PT
ED
577
585
The model also employs a volume averaging method to approximate subgrid ob-
586
jects as body forces. For a coastal scale model, the shallow water equations are modified to include a form drag and spatial dispersion term. The methods are applied to
588
tidal power generation from turbines that are subgrid in scale at a study site in Minas
589
Passage at the upper end of Bay of Fundy. Some options are presented for estimating
590
the thrust and power coefficients, including a numerical convergence procedure and 2
591
methods of mapping from a free stream velocity reference to a volume averaged veloc-
AC 587
27
M
AN US
CR IP T
ACCEPTED MANUSCRIPT
ED
Figure 4: Variation in power between different approximation methods. The thick line denotes site 1 and the thin line denotes site 3. A different color is used for each of the 4 cases. In the legend, C denotes the case (C0 is case 0, etc.) and S denotes the site (S1 is site 1, etc.).
ity reference. Although the two mappings are ostensibly for the same conditions where
593
is small and the flow bypassing the turbine is approximately the free stream veloc-
594
ity, the two sets of calculated coefficients differ. All sets of results are in reasonable
595
agreement, but the smallest difference is between the the coefficients from numerical
CE
PT
592
596
iteration (case 1) and the mapping based on continuity (case 3). In order to simplify the analysis, only the dominant tidal constituent (M2 ) was used
598
in the boundary forcing. Using the full set of constituents would introduce fortnightly
599
and monthly modulations in the tides and lead to greater variability in power output.
600
The results presented here would still apply under the same assumptions. For a more
601
complete analysis, the assumption that the turbines operate with coefficients given by
602
the Lanchester-Betz limit would have to be dropped and field-scale data used instead.
AC
597
28
ACCEPTED MANUSCRIPT
For this analysis, the locations of the turbines are at sites where tidal turbines will be
604
tested in the future. Choosing other sites depends on a number of factors including but
605
not limited to the maximum velocity, the water depth, navigation hazards, turbulence
606
intensity, accessibility for maintenance, access to the power grid, etc. This evaluation
607
is beyond the scope of this study.
CR IP T
603
In these examples, the spatial dispersion term was neglected, mainly because it is
609
not clear how to calculate this term except by small-scale CFD simulations. Intuitively,
610
this term should have a role in modifying the wake and the flow that bypasses the tur-
611
bine disk. When the averaging volume is of the order of the size of the turbine (a blade
612
diameter on an edge), then the flow in the laterally adjacent elements is accelerated
613
by the spatial dispersion in addition to the surface pressure gradient. Clearly, much
614
more work needs to be done to try and understand the magnitude and importance of
615
this term.
616
Acknowledgments
AN US
608
M
The author wishes to acknowledge helpful review comments by Michael Shives,
617
Phillip Gillibrand, and Vincenzo Casulli.
619
References
ED
618
[1] Baranger, J., Maitre, J.F., Oudin, F., 1996. Connection between finite volume and
621
mixed finite element methods. Mathematical Modelling and Numerical Analysis
622
30, 445-465.
PT
620
[2] Casulli, V., 1987. Eulerian-Lagrangian methods for hyperbolic and convection
624
dominated parabolic problems. In: Taylor, C., D.R Owen, Hinton, E., (Eds),
625
Computational Methods for Non-linear Problems, Pineridge Press, Swansea, pp.
AC
CE 623
626
239–268.
627
[3] Casulli, V., 2009. A high-resolution wetting and drying algorithm for free-surface
628
hydrodynamics. Int. J. Numer. Meth. Fluids 60, 391–408. DOI: 10.1002/fld.1896
29
ACCEPTED MANUSCRIPT
629
[4] Casulli V., Cattani E., 1994. Stability, accuracy, and efficiency of a semi-implicit
630
method for three-dimensional shallow water flow. Computers Math. Applications
631
27(4), 99-112. [5] Casulli, V., Walters, R. A., 2000. An unstructured grid, three-dimensional model
633
based on the shallow water equations. International Journal for Numerical Meth-
634
ods in Fluids 32 (3), 331-348.
CR IP T
632
[6] Fringer, O.B, Gerrirsen, M., Street, R.L., 2006. An unstructured grid, finite vol-
636
ume, nonhydrostatic, parallel coastal ocean simulator. Ocean Modelling 14, 139-
637
173.
AN US
635
638
[7] Finnigan, J.J., Shaw, R.H., 2008. Double-averaging methodology and its applica-
639
tion to turbulent flow in and above vegetation canopies. Acta Geophysica 56(3),
640
534–561.
[8] Garrett, C., 1972. Tidal resonance in the Bay of Fundy and Gulf of Maine. Nature 238, 441–443.
642
643
[9] Garrett, C., Cummins, P., 2007. The efficiency of a turbine in a tidal channel. J. Fluid Mechanics 588, 243-251. doi:10.1017/S0022112007007781
ED
644
645
M
641
[10] Greenberg, D.A., 1979. A numerical model investigation of tidal phenomena in the Bay of Fundy and Gulf of Maine. Marine Geodesy 2(2), 161–187.
PT
646
[11] Ham, D.A, Kramer, S. C., Stelling, G. S., Pietrzak, J. D., 2007. The symmetry and
648
stability of unstructured mesh C-grid shallow water models under the influence
CE
647
of Coriolis. Ocean Modelling 16, 47-60.
649
650
[12] Jankowski, J.A., (2009). Parallel implementation of a non-hydrostatic model for free surface flows with semi-Lagrangian advection treatment. Int. J. Numer. Meth.
652
Fluids 59, 1157?1179.
AC 651
653
[13] Karsten, R.H., McMillan J.M., Lickley, M.J., Haynes, R.D., 2008. Assessment of
654
tidal current energy in the Minas Passage, Bay of Fundy. Proc. of the Institution
655
of Mechanical Engineers, Part A: Journal of Power and Energy. 30
ACCEPTED MANUSCRIPT
656
[14] Le Roux D.Y., Rostand V., Pouliot B., 2007. Analysis of numerically induced os-
657
cillations in 2D finite-element shallow-water models Part I: Inertia-gravity waves.
658
SIAM Journal of Scientific Computing 29, 331-360. [15] Leendertse, J.J, 1967. Aspects of a computational model for long period water
660
wave propagation. Tech Report RM-5294-PR, Santa Monica, Rand Memoran-
661
dum.
CR IP T
659
[16] Nikora, V., McLean, S., Coleman, S.,Pokrajac, D., Walters, R., 2007. Double-
663
averaging concept for rough-bed open-channel and overland flows: Theoretical
664
background. J. Hydraul. Eng. ASCE 133, 873–883.
AN US
662
665
[17] O’Donncha,F., Hartnett, M., Plew, D.R., 2015. Parameterizing suspended canopy
666
effects in a three-dimensional hydrodynamic model. Journal of Hydraulic Re-
667
search, DOI: 10.1080/00221686.2015.1093036
668
[18] Pinder, G.F., Gray, W.G., 1977. Finite Elements in Surface and Subsurface Hydrology, Academic Press.
M
669
[19] Plew, D.R., 2011. Shellfish farm-induced changes to tidal circulation in an em-
671
bayment, and implications for seston depletion Aquaculture Environment Inter-
672
actions 1, 201–214. doi: 10.3354/aei00020
ED
670
[20] Raviart,P. A., Thomas, J. M., 1977. A mixed Finite Element Method for 2nd Order
674
Elliptic Problems. Mathematical Aspects of the Finite Element Method, Lecture
675
Notes in Math. Springer Berlin.
CE
PT
673
[21] Shives, M., Crawford, C., Hiles, C., Walters, R., 2013. Combining Numerical
677
Methods for Basin and Turbine Scales for Improved Modelling of in-situ Turbine
678
Arrays. Proc. 10th European Wave and Tidal Energy Conf., Aalborg, Denmark.
AC
676
679
[22] Sucsy, P.V., Pearce, B.R., Panchang, V.G., 2006. Comparison of two-and three-
680
dimensional model simulation of the effect of a tidal barrier on the Gulf of Maine
681
tides. Journal of Physical Oceanography 23(6), 1231–1248.
31
ACCEPTED MANUSCRIPT
[23] Waldman, S., Genet, G., Bast, S., Side, J., 2015. Correcting for mesh size de-
683
pendency in a regional models representation of tidal turbines. Proceedings of
684
the 11th European Wave and Tidal Energy Conference 6-11th Sept 2015, Nantes,
685
France.
686
CR IP T
682
[24] Walters, R.A., 2005, Coastal ocean models: two useful finite element methods, Continental Shelf Revsearch 25, 775-793.
687
[25] Walters, R. A., Gillibrand, P., Bell, R., Lane, E.M., 2010. A study of tides and
689
currents in Cook Strait, New Zealand. Ocean Dynamics 60, 1559–1580. DOI
690
10.1007/s10236-010-0353-8
AN US
688
691
[26] Walters, R.A., Hanert, E., Pietrzak, J.D., Le Roux, D.Y., 2009. Comparison of un-
692
structured, staggered grid methods for the shallow water equations. Ocean Mod-
693
elling 28, 106–117.
694
[27] Walters, R.A., Lane, E.M., Henry, R.F., 2008. Semi-Lagrangian methods for a finite element coastal ocean model. Ocean Modelling 19, 112–124.
M
695
[28] Walters, R.A., Lane, E.M., Hanert, E., 2009. Useful time-stepping methods
697
for the Coriolis term in a shallow water model. Ocean Modelling 28, 66–74.
698
doi:10.1016/j.ocemod.2008.10.004
699
ED
696
[29] Walters, R.A., Plew, D.R., 2008. Numerical modelling of environmental flows using DAM: Some preliminary results. Acta Geophysica 56(3), 918–934.
701
PT
700
[30] Walters, R.A., Tarbotton, M.R, Hiles, C.E., 2013. Estimation of tidal power potential. Renewable Energy 51, 255–262.
CE
702
[31] Wu, W., Shields Jr., F. F., Bennett, S. J,. and Wang, S. S. Y., 2005. A depth-
704
averaged two-dimensional model for flow, sediment transport, and bed topogra-
AC
703
705
phy in curved channels with riparian vegetation. Water Resources Research 41,
706
W03015.
707
[32] Yeh, G-T, 1981. Numerical solutions of Navier-Stokes equations with an inte-
708
grated compartment method (ICM). International Journal for Numerical Methods
709
in Fluids 1(3), 207-223. 32
ACCEPTED MANUSCRIPT
Figure Captions
710 711
Figure 1: Geometry and bathymetry of Bay of Fundy. AO is Atlantic Ocean, GoM
713
is Gulf of Maine, BoF is Bay of Fundy, MP is Minas Passage, MB is Minas Basin, B
714
is Boston, and CC is Cape Cod.
715
Figure 2: Minas Passage at flood tide. CS identifies Cape Split. denote the tur-
716 717
CR IP T
712
bine deployment sites. Velocity vectors have constant length and are decimated by 5.
718
AN US
Figure 3: Grid around the turbine sites in Minas Passage.
719 720
Figure 4: Variation in power between different approximation methods. The thick
722
line denotes site 1 and the thin line denotes site 3. A different color is used for each of
723
the 4 cases. In the legend, C denotes the case (C0 is case 0, etc.) and S denotes the site
724
(S1 is site 1, etc.).
AC
CE
PT
ED
M
721
33