+ All Categories
Home > Documents > International Journal of Computational Fluid Dynamicsanalysis with complex derivatives',...

International Journal of Computational Fluid Dynamicsanalysis with complex derivatives',...

Date post: 21-Apr-2020
Category:
Upload: others
View: 4 times
Download: 0 times
Share this document with a friend
21
This article was downloaded by:[University of Central Florida] On: 2 November 2007 Access Details: [subscription number 769428830] Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK International Journal of Computational Fluid Dynamics Publication details, including instructions for authors and subscription information: http://www.informaworld.com/smpp/title~content=t713455064 2D unstructured mesh generation for oceanic and coastal tidal models from a localized truncation error analysis with complex derivatives D. M. Parrish a ; S. C. Hagen a a Department of Civil and Environmental Engineering, University of Central Florida, 4000 Central Florida Blvd., Orlando, FL 32816-2450, USA Online Publication Date: 01 August 2007 To cite this Article: Parrish, D. M. and Hagen, S. C. (2007) '2D unstructured mesh generation for oceanic and coastal tidal models from a localized truncation error analysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article: DOI: 10.1080/10618560701582500 URL: http://dx.doi.org/10.1080/10618560701582500 PLEASE SCROLL DOWN FOR ARTICLE Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf This article maybe used for research, teaching and private study purposes. Any substantial or systematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.
Transcript
Page 1: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

This article was downloaded by:[University of Central Florida]On: 2 November 2007Access Details: [subscription number 769428830]Publisher: Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

International Journal ofComputational Fluid DynamicsPublication details, including instructions for authors and subscription information:http://www.informaworld.com/smpp/title~content=t713455064

2D unstructured mesh generation for oceanic andcoastal tidal models from a localized truncation erroranalysis with complex derivativesD. M. Parrish a; S. C. Hagen aa Department of Civil and Environmental Engineering, University of Central Florida,4000 Central Florida Blvd., Orlando, FL 32816-2450, USA

Online Publication Date: 01 August 2007To cite this Article: Parrish, D. M. and Hagen, S. C. (2007) '2D unstructured meshgeneration for oceanic and coastal tidal models from a localized truncation erroranalysis with complex derivatives', International Journal of Computational Fluid

Dynamics, 21:7, 277 - 296To link to this article: DOI: 10.1080/10618560701582500URL: http://dx.doi.org/10.1080/10618560701582500

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf

This article maybe used for research, teaching and private study purposes. Any substantial or systematic reproduction,re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expresslyforbidden.

The publisher does not give any warranty express or implied or make any representation that the contents will becomplete or accurate or up to date. The accuracy of any instructions, formulae and drug doses should beindependently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings,demand or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with orarising out of the use of this material.

Page 2: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

2D unstructured mesh generation for oceanic and coastal tidalmodels from a localized truncation error analysis with complex

derivatives

D. M. PARRISH* and S. C. HAGEN

Department of Civil and Environmental Engineering, University of Central Florida, 4000 Central Florida Blvd., Orlando, FL 32816-2450, USA

(Received 14 March 2007; in final form 19 July 2007)

A method for computing target element size for tidal, shallow water flow is developed anddemonstrated. The method, Localized truncation error analysis with complex derivatives (LTEA-CD)utilizes localized truncation error estimates of the linearized shallow water momentum equationsconsisting of complex derivative terms. This application of complex derivatives is the chief way inwhich the method differs from a similar existing method, LTEA. It is shown that LTEA-CD producesresults that are essentially equivalent to those of LTEA (which in turn has been demonstrated to becapable of producing practicable target element sizes) with reduced computational cost. Moreover,LTEA-CD is capable of computing truncation error and corresponding target element sizes at locationsup to and including the boundary, whereas LTEA can be applied only on the interior of the modeldomain. We demonstrate the convergence of solutions over meshes generated with LTEA-CD using anidealized representation of the western North Atlantic Ocean, Caribbean Sea and Gulf of Mexico.

Keywords: Localized truncation error analysis; Unstructured mesh generation; Shallow waterequations; Tidal computations; Complex derivatives; Western North Atlantic tidal model domain

AMS Subject Classifications: 30E10; 39-02; 39A99; 76-05; 76Bxx; 76Mxx

1. Introduction

The research presented herein is representative of our

progress toward creating an algorithm for automatically

generating target element sizes for application to two-

dimensional (2D) unstructured mesh generation for

oceanic and, more specifically, coastal domains. A finite

element model of a physical system requires a geometric

description of the system in the form of a mesh of

interconnected nodes and elements. In general, the level of

detail of the mesh affects the accuracy and stability of the

model. In this paper, we present an alternative and, we

propose, improved method for generating meshing criteria

for 2D models of shallow, tidal flow.

Existing methods of computing target element sizes for

coastal areas leave much to be desired. The a posteriori

Localized truncation error analysis (LTEA; Hagen 1998,

2001, Hagen et al. 2000, 2001) has shown more promise

than other methods, such as the wavelength to grid size

ratio (e.g. Westerink et al. 1994, Luettich and Westerink

1995) and the topographic length scale, because of its

(LTEA’s) basis in the shallow water (momentum)

equations themselves. However, a major disadvantage of

LTEA is that in order to compute values of the localized

truncation error—upon which are based the target element

sizes—a 9 £ 9 finite difference (FD) molecule, centred at

mesh nodes, is applied in computing the derivative terms

(up to fifth-order) of the localized truncation error

estimate. Points in the FD molecule must each lie in

different elements of a linear triangular mesh of the model

from which localized truncation error is to be computed.

This requirement results in numerous cases where the FD

molecule violates the mesh boundaries. These cases

include all boundary nodes and all nodes in the vicinity of

the boundary. Therefore LTEA is suitable only for

applications where the mesh area is large in comparison to

its boundary, that is nn/nb q 1, where nn is the number of

nodes in the mesh and nb is the number of boundary

nodes. Boundary shape alone is not the determining factor,

rather it is how that shape is discretized.

Therefore we are developing an algorithm that all but

eliminates the limitations imposed by the FD molecule

International Journal of Computational Fluid Dynamics

ISSN 1061-8562 print/ISSN 1029-0257 online q 2007 Taylor & Francis

http://www.tandf.co.uk/journals

DOI: 10.1080/10618560701582500

*Corresponding author. Email: [email protected]

International Journal of Computational Fluid Dynamics, Vol. 21, Nos. 7–8, August–September 2007, 277–296

Page 3: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

while maintaining the desirable qualities of LTEA. We

achieve this by recasting the localized truncation error

estimate in terms of complex derivatives (›/›z instead of

›/›x and ›/›y). This allows the production of a truncated

Taylor series (the mathematical basis of LTEA), the zero

to sixth-order (›6/›z6) terms of which are calculable using

only seven discrete points in a difference molecule, all of

which may be located within the “valence shell” of

elements surrounding a typical interior node in a linear

triangular mesh. For cases where the node is on the

boundary, eight points may be applied to estimate the zero

to seventh-order (›7/›z7) terms. The extra point is needed

in order to provide O[(DM)2] accuracy, where DM is the

size of a mesh element (units of length).

After developing the theory of and explaining the

procedure for applying LTEA-CD, we provide two test

cases. In the first, we apply LTEA-CD iteratively, allowing

each successive mesh to be more refined than the last, but

nowhere to be coarsened. In the second test case, we apply

LTEA-CD iteratively again, but this time only calculate a

new distribution of target element size that is applied to

the remeshing, while the number of elements remains

fixed within a tolerance of less than 0.01 (1%).

In order to produce target element sizes, a linear tidal

simulation is executed with an initial mesh. For ease of

mesh generation, this could be a uniform mesh, but

uniformity is not required. The LTEA-CD algorithm

computes target element sizes from model output; the

target element sizes are linearly scaleable. It is up to the

user to select the scale factor, but we provide some

guidelines. Note that were the flow field sufficiently

known from field data, no simulation would be necessary

in order to compute the target element sizes. However, for

most cases, flow fields must be computed due the scarcity

and unavailability of accurate, measurement-based data.

Eventually, we intend to develop LTEA/LTEA-CD

further by incorporating near shore and estuarine discrete

physics (i.e. quadratic bottom friction and advection) into

the estimation of truncation error and target element sizes.

This is a natural extension, since now we are able to

calculate target element sizes at and near the boundary.

Prior to that, however, we develop the theory behind

LTEA-CD and test the practicality of its application before

tackling the more complicated problems of the nonlinear

terms and corresponding tidal constituent interactions.

1.1 Alternative methods for computing mesh resolution

Several criteria for element size have been developed for

finite element models of ocean circulation. Researchers

investigating this problem include Le Provost and Vincent

(1986), Kashiyama and Okada (1992) and Westerink et al.

(1992). (See also Hannah and Wright 1995).

Greenberg et al. (2006) review several issues pertaining

to mesh resolution and the accuracy of coastal and ocean

circulation models. Based upon their review, it seems

that there are only about three quantitative relations

(though many qualitative ones) that should influence mesh

resolution: the Courant number, (Dt)(gh)1/2(Dx)21, where

Dt is the timestep, g is the acceleration due to gravity, h is

bathymetric depth and Dx is the grid spacing (Foreman

1984, Le Provost et al. 1995); the topographic length

scale, h=k7hk (Loder 1980, Lynch et al. 1995, Hannah

and Wright 1995) and Hagen’s (1998) localized truncation

error estimator (also Hagen et al. 2001, 2002, Hagen 2001,

Hagen and Parrish 2004):

tME ¼D2

4

ivþ t

2

� �›2u0

›x2þ

›2v0

›x2þ

›2u0

›y2þ

›2v0

›y2

� ��

þg

2

›3h0

›x3þ

›3h0

›x2›yþ

›3h0

›x›y2þ

›3h0

›y3

� ��

þD4

16

ivþ t

8

� �›4u0

›x4þ

›4v0

›x4þ 2

›4u0

›x2›y2

��

þ2›4v0

›x2›y2þ

›4u0

›y4þ

›4v0

›y4

þg

24

22

10

›5h0

›x5þ

›5h0

›x4›yþ 2

›5h0

›x3›y2þ 6

›5h0

›x2›y3

þ3›5h0

›x›y4þ

9

5

›5h0

›y5

��ð1Þ

where D is the length of an element’s edge, i 2 ¼ 21, t is

the linearized friction coefficient, u and v are complex

velocities in the x- and y-directions, h and is complex

deviation of the sea surface from the geoid for the chosen

tidal constituent of frequency v, and the subscript (0)

refers to the node at which tME is to be computed.

1.2 Similarities and distinctions between LTEAand LTEA-CD

The most distinguishing characteristic of LTEA-CD is its

ability to compute target element sizes at and near the

boundary. This capability opens up new possibilities in the

field of meshing for coastal circulation problems.

Both LTEA and LTEA-CD are a posteriori methods,

that is, they rely upon the results of a simulation in order to

compute optimal meshing requirements for future

simulations. By “optimal” we mean that the mesh is

designed so as to distribute truncation error uniformly. In

practice, the distribution of truncation error does not

become absolutely uniform, but is made more uniform.

From a theoretical perspective, the chief way in which

LTEA-CD differs from LTEA is that the localized

truncation error estimate is computed using derivatives

with respect to the complex quantity z ¼ xþ iy instead of

x and y (where the lateral coordinates of the mesh lie in the

x/y plane). The main consequence of this approach is the

dramatic simplification of the localized truncation error

estimator, which translates into reduced computing time

and the introduction of the capability of computing the

estimate at and near the boundary.

D. M. Parrish and S. C. Hagen278

Page 4: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

1.3 Complex derivatives and complex Taylor series:examples

There are many examples of the application of complex

derivatives to 2D engineering problems. Those new to

complex derivatives may find useful the following

references, which provide introductory material and

applications: Reddick and Miller (1938), Timoshenko

and Goodier (1970), Saada (1974), Daugherty et al.

(1985), Greenberg (1998) and Sadd (2005).

1.4 Outline of this paper

In the next section, we present the theory behind LTEA-

CD, including the development of a localized truncation

error estimator, derivative estimators and a corresponding

expression for target element size. We then test LTEA-CD

by comparing its results to those of LTEA. The section

concludes with a discussion of the advantages of LTEA-

CD over LTEA.

In Section 3, we examine the convergence properties of

LTEA-CD by applying the method to an idealized Western

North Atlantic tidal (WNAT) model domain. Two series of

simulations are undertaken; in the first, the number of

elements in each successive mesh is allowed to increase,

while the distribution of nodes is determined by LTEA-CD

results; in the second, the number of elements in each

successive mesh is held fixed, with the distribution

determined by LTEA-CD. Results are discussed apart

fromamore detaileddescription of either of these two series.

In a discussion section, we provide our interpretation of

the results, having presented the results proper in the

previous section. We also give recommendations on how

to apply LTEA-CD.

In the last two sections, we present our conclusions

based upon the material presented herein, and provide a

sketch of the further research in this line.

2. Theory

We consider only localized truncation error of the harmonic,

linearized shallow water momentum equations (e.g. Hagen

2001):

ðivþ tÞuþ g›h

›x¼ 0 ð2aÞ

and

ðivþ tÞvþ g›h

›y¼ 0; ð2bÞ

where the variables have been defined as in equation (1).

Note that terms excluded from equations (2a) and (2b)

become important in very shallow waters (e.g. near shore),

particularly the advective terms. Additionally, the friction

coefficient, t, here set constant in space and time, actually

varies in proportion to current speed and inversely with

depth. However, we apply only (2a) and (2b) to the

truncation error series for two reasons: (1) this is a first

study and (2) LTEA does likewise and has been shown to

produce favourable results (at least for tidal elevations),

even when applied to meshes on which fully nonlinear

simulations are executed (Kojima 2005, Hagen et al.

2006). In addition, this simplified study lays a foundation

for the next phase of research, namely to incorporate both

advection and variable bottom friction (dependent upon

water depth and velocity). The present theory and

succeeding applications, albeit idealized, lay the ground-

work for including nonlinearities, since with the present

theory, localized truncation error and corresponding target

element size may be computed at and near the boundary.

2.1 A local truncation error estimate

The momentum equations (2a) and (2b), are discretized

spatially over a submesh using Galerkin, linear triangular

finite elements. We define a submesh to be a central node

surrounded by a valence shell of equilateral triangular

elements, each consisting of three nodes, one of which is the

central node. The submesh does not necessarily coincide

with the elements of a mesh on which the solution is

computed, hence we avoid the term “stencil”, however, the

central node is located on a node of the mesh fromwhich the

solution is derived. The discrete form of the momentum

equations (2a) and (2b) are, in the x-direction,

ivþ t

12

X6j¼1

uj þ 6u0

!

þg

6Dð2h1 þ h2 2 h3 2 2h4 2 h5 þ h6Þ ¼ 0; ð3aÞ

and in the y-direction,

ivþ t

12

X6j¼1

vjþ6v0

g

2ffiffiffi3

pDðh2 þ h3 2 h5 2 h6Þ ¼ 0

ð3bÞ

(equations 6.22 and 6.23 in Hagen (1998)), where the

subscripts are the local indices of the central node (0) and its

neighbours (1–6, counter clockwise from the þx-axis, a

different scheme than in Hagen (1998)) andD is the distance

from the central node to that of any of its neighbours (Hagen

1998 defined the distance between neighbouring nodes as

2D, convenient when working in x- and y-coordinates).

In further departing from Hagen (1998), we develop an

expression for truncation error that is based upon an

analysis in the complex plane. Let z ¼ xþ iy. We place

the origin of the complex plane at a central node.

The discrete momentum equations (3a) and (3b) may be

expressed in terms of f0 [ {h0; u0; v0} and its derivatives

f kð Þ0 , k [ :, by substituting the complex Taylor series for

2D unstructured mesh generation for oceanic and coastal tidal models 279

Page 5: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

the fj [ {hj; uj; vj}, i.e.

f j ¼ f 0 þDj

1!f ð1Þ0 þ

D2j

2!f ð2Þ0 þ · · · þ

D6j

6!f ð6Þ0 þ HOT : ð4Þ

where mod Dj ¼ D, Dj ¼ ðxj 2 x0Þ þ iðyj 2 y0Þ, j [ {1, 2,

. . . , 6}, and HOT are the higher order terms. With our

chosen configuration of nodes and elements, the discrete

momentum equations therefore reduce to

ivþ t

14401440u0 þ D6uð6Þ0

� �þ

g

120120hð1Þ

0 þ D4hð5Þ0

� ¼ 0

ð5aÞ

and

ivþ t

14401440v0 þ D6vð6Þ0

� �þ

ig

120120hð1Þ

0 2 D4hð5Þ0

� ¼ 0;

ð5bÞ

where we have dropped the HOT. We multiply equation

(5a) by i and add the result to equation (5b), which yields:

ðivþ tÞði u0 þ v0Þ þ 2ighð1Þ0

þ D6 ivþ t

1440i uð6Þ0 þ vð6Þ0

� �¼ 0: ð6Þ

A localized truncation error estimator is determined by

subtracting (2b) and i £ (2a) from (6):

tME ¼ D6 ivþ t

1440i uð6Þ0 þ vð6Þ0

� �: ð7Þ

The terms involving h cancel through application of the

chain rule and because ›x/›z ¼ 1/2 and ›y=›z ¼ 1=2i

(see, e.g. Weisstein 2006).

2.2 Effects of submesh orientation on localizedtruncation error

The orientation of the submesh does not affect mod tME

when the derivative terms are evaluated with difference

equations. This can be shown by applying the rotation

equations to the tidal ellipse and regrouping the terms so

that the original form is attained, multiplied by a complex

number of unit magnitude. Were we solely concerned with

the truncation error itself, and the comparisons of the

truncation errors at various nodes of the mesh, it would be

necessary to reconcile the distinct orientation of each

submesh with each other submesh. However, our chief

concern is in computing target element sizes, which are

dependent upon mod tME only.

2.3 Derivative approximation

Two difference formulae are derived with which the

derivative terms of equation (7) may be computed. The first

is applied for the case in which the central node is on the

interior (V); the second for that in which the central node

lies on the boundary (G). In either case, it is important that

when computing f ð6Þ0 from a linear triangular mesh, the

points of the difference molecule lie within different

elements, since all derivatives beyond the first are zero

within a single element of the mesh.

2.3.1 Interior (V) case. Let f be approximated at any

interior point i by a sixth-order polynomial f i < Vpi ¼

aTDi, where the elements ak, k [ {0,1, . . . ,6}, of a are

complex constants, and the kth element of Di is Dki .

We construct a difference equation by applying the regular

hexagonal geometry of the submesh and requiring that

Vpð6Þ0 ¼ aTf ¼ 6!a6, where f ¼ {f0, f1, . . . , f6} and the

elements of a are complex constants. This condition

implies a set of seven simultaneous equations,

Da ¼ {0; 0; 0; 0; 0; 0; 6!}, where D ¼ D0;D1; . . . ;D6

�and 00 ; 1, that, when solved, yield

V f ð6Þ0 < Vpð6Þ0

¼ 120ð f 1 þ f 2 þ f 3 þ f 4 þ f 5 þ f 6 2 6f 0Þ=D6:

ð8Þ

That equation (8) is O(D6) accurate may be shown by

substituting complex Taylor series (in terms of f0) for the

fjs and simplifying. The hexagonal configuration of the

submesh is the key to the high order of accuracy.

In order to compute V f ð6Þ0 we size the submesh such

that D is equal to half the distance from the central node

the nearest mesh node. We check whether each point lies

in a different element; if not, we rotate the difference

molecule by p/6 rad (308) and check again; if there are still

two or more points that lie within a single element, we do

not calculate V f ð6Þ0 for that central node. In general, V f ð6Þ0

can be computed at interior nodes of valence six or more,

provided the maximum angle at the central node & 5

p/12 rad (758) in our application. Allowing the difference

molecule to be oriented in any direction allows for

maximum submesh angles approaching 2p/3 rad (1208);

one could also enlarge the submesh in order to overcome

local geometries that are less than ideal.

2.3.2 Boundary case. The derivative terms of tME may,

with O(D2) accuracy, be estimated at the boundary by

considering a semi-circular difference molecule of eight

points, where the central node coincides with the midpoint

of the semicircle (i.e. the midpoint along the arc, not the

centre of curvature of the arc). This orientation is preferred

to that of placing the central node midway between the

vertices of the semi-circle, because the former may still be

used when the boundary is locally concave.

The derivation of the difference equation for the

boundary case is similar to the interior case, except that

both the estimating polynomial and the difference

equation have an additional term:

Gfð6Þ0 < Gp

ð6Þ0

¼ 60½c1f 1 þ c2f 2 þ c3f 3 þ 12ð5 2 sÞf 4

þ �c3f 5 þ �c2f 6 þ �c1f 7 2 6ð9 þ 4sÞf 0�=D6;

ð9Þ

D. M. Parrish and S. C. Hagen280

Page 6: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

where s ¼ 31/2, c1 ¼ 29 2 s2 3ið13 þ 7sÞ, c2 ¼ 9þ

15s2 ið9 þ sÞ, and c2 ¼ 23 þ 4s2 3ið1 2 2sÞ.

In order to compute Gfð6Þ0 , we choose points such that

the midpoint along the semi-circle’s diameter lies

0.5 £ 31/2 times further away than the distance between

the central node and the mesh node within the valence

shell that is furthest from the central node. We check to see

that each point lies in a different element; if not, we

increase the size of the difference molecule by a factor of

1.5 and check again; if two or more points still lie within

the same element, we do not calculate Gfð6Þ0 for that central

node. The orientation of the boundary submesh is such

that the boundary segments make equal angles with a line

tangent to the semicircle passing through the central node

of the submesh. Note that in order to compute Gfð6Þ0 there

need to be at least two elements between opposing

boundary segments. For coastal and ocean models, this is

typically the case. Another constraint on the computation

of Gfð6Þ0 by this method is that boundary edges may have

an interior angle no smaller than 5p/6 rad (1508). A

regular polygon having edges that meet at this angle has

12 sides; it would require three edges to turn the mesh

boundary p/2 rad (908).

2.4 Meshing requirements from localized truncationerror estimates

We now have the means by which to compute tME at

almost any node of a well-constructed, triangular finite

element mesh. There are a variety of ways in which

element size may be derived from tME. Equation (7)

may be rearranged to solve for target element size, D*

(a positive real number), the product of an arbitrary scale

factor, a and a deterministic factor, D: D* ¼ aD. Any

choice of a implies a given target value for mod tME.

There is only indirect dependence of D* on v and t when

only a single tidal constituent is considered, since ivþ t

can be lumped together with a, but u and v depend upon v

and t. The scale factor a may be selected such that (1) D*

is never less than a certain value, (2) boundary D* does not

exceed a tolerance, (3) D* at a particular location is

specified, (4) the number of elements, ne, is specified

(it easier to specify ne, proportional to domain area, than

to specify the nn, proportional to both area and perimeter).

One may also select different values of a in different

regions of the domain. There are other options; in fact, we

use a fifth approach, explained below.

In order to specify the number of elements, first

compute the elemental density assuming a ¼ 1:

re1 ¼ 1/(D 2 £ 30.5/4). Next, integrate re1 over the domain

in order to determine a hypothetical number of elements:

ne1 ¼ÐVre1dV. One may approximate ne1 by setting

the bathymetric depth of the mesh equal to re1 and

computing the volume. Finally, select the number of

elements desired, ne2 and compute a ¼ (ne1/ne2)1/2. We

recommend that a specific a also be selected so as to

target a mesh where the adjacent element area ratio

criterion (Æ-ARC, “eye arc”) is, for every pair of adjacent

elements, less than 2 (big:small), since extreme gradients

in element size are a source of increased localized

truncation error and reduced model stability (at least

for the discretization considered herein). We choose

a # kfD*klimit/maxkfDk, where kfD*klimit is the

maximum acceptable gradient in D*. For Æ-ARC # 2,

and when computing actual local element size, DM, by

taking the mean of the lengths of element edges that share

a node, kfD*klimit < 3/4.

It is also possible to smooth the target element sizes,

beginning at locations of high resolution and progressing

to locations of lower resolution, enforcing Æ-ARC as

smoothing progresses. This has the advantage of enabling

the production of a mesh with a target number of nodes

that still meets Æ-ARC, but at the expense of producing a

mesh that does not fit the distribution of D*. For this

reason, herein we prefer scaling the meshing requirements

so that D* is met everywhere, allowing us to evaluate

LTEA-CD without obscuring it with Æ-ARC smoothing.

The advantage of this approach is, of course, that the

corresponding mesh will fit both the distribution of D* and

Æ-ARC, however the distinct disadvantage is that a

potentially infeasible number of nodes may be required to

construct the mesh. A compromise may be to apply pure

smoothing to D before constructing the mesh.

Herein, we apply Gaussian smoothing to D before

determining target element sizes from those smoothed

values; the parameters of the Gaussian weight function are

set so that one standard deviation coincides with DM.

Values from nodes further than three elements away from

a central node are not used to compute a smoothed value

of D. The application of certain more advanced smoothing

techniques would be expected to produce better results

than simple Gaussian smoothing, particularly because

Gaussian smoothing ignores significant local, directional

variation in localized truncation error that are the result of,

for example the presence of a shipping channel (one

would want the smoothing algorithm to smooth only along

the channel, not across it). Note that the Gaussian

smoothing applied herein is distinct from that applied by

Hagen et al. (2006), who applied Æ-ARC smoothing.

2.5 Computer code for automatically generatingmeshing requirements

Our code consists of three main components. The first

component creates an inverse connectivity table, i.e. a

lookup table that gives the element numbers of each

element connected to a given node. The second component

generates difference molecules for each central node. Both

the inverse connectivity table and the difference molecules

may be generated without the use of searching. Searching

is not required since each component makes use of direct

access storage of the mesh connectivity and inverse

connectivity. (The data could be stored in RAM or direct

access files, depending on the number of elements and

available RAM). In order to interpolate values to the nodes

of the difference molecules, the element within which the

2D unstructured mesh generation for oceanic and coastal tidal models 281

Page 7: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

point lies must be known. This can be determined with

consideration for the angles that the corresponding edges

make with an axis through the central node. The third

component computes D using the connectivity table and

difference molecules generated by the other components.

Note that for a mesh with static lateral nodal coordinates

and static connectivity, the first two components need be

run only once, while the third component may be run for

multiple model outputs (e.g. multiple linear harmonics).

2.6 Comparison of LTEA and LTEA-CD

In order to test LTEA-CD, we use the tidal simulation

results (Hagen and Parrish 2004) of an ADCIRC-2DDI

model (Advanced Circulation model for Oceanic, Coastal

and Estuarine Waters—2D Depth-Integrated option,

henceforth, ADCIRC; Luettich et al. 1992, Luettich and

Westerink 2004). In the present application, ADCIRC

computes tides (current velocities and sea surface

deviations from the geoid) and generates tidal harmonics

for each node of a finite element mesh. The mesh consists

of 3,33,701 nodes and 6,48,661 elements over the WNAT

model domain. It encompasses the Gulf of Mexico,

Caribbean Sea, and the portion of the Atlantic Ocean that

lies west of the 608 W meridian (figure 1). The simulation

applied the linearized shallow water equations with M2

tidal forcing only (v ¼ 1.405 £ 1024 s21) and

t0 ¼ t ¼ 0.0004 s21, where t0 is the generalized wave-

continuity weighting factor. A no-flow boundary condition

is enforced at all land boundaries and the tidal forcing

(depth only) is applied to the open ocean boundary. Fifteen

days of real time are simulated, ensuring that a dynamic

steady-state is achieved. A time step of 20 s is used and a

hyperbolic ramping function is imposed during the first 2

days. We compute both tME and tME with the tidal

harmonics generated by ADCIRC. We select the minimum

spacing to be 1000 m when computing tME. LTEA-CD has

no such parameter for calculating localized truncation

error, since D in equation (7) cancels with the D in the

difference equations.

The magnitudes of the localized truncation error

estimates (mod tME and mod tME) are similarly

geographically distributed (figure 2; note that we compare

only those portions of the domain where LTEA is capable

of computing tME, hence the dark areas in the figure).

Although mod tME and mod tME differ by orders of

magnitude (not unexpected; see Hagen et al. 2000), the

essential point is that when converted to distributions of

D*, LTEA and LTEA-CD produce, in essence, the same

information (details below).

Comparison between the target element sizes corre-

sponding to the LTEA and LTEA-CD methods depends

upon how the minimum spacing is selected. Target

element sizes of either method can be forced to equal each

other at a single node by adjusting a. Enforcement of equal

spacing throughout the model domain between LTEA and

LTEA-CD methods is likely impossible for a spatially

constant D because of round-off error, inaccuracies and

imprecision in model results, and because of the different

methods of computing the derivative terms. In addition,

the two methods likely differ in their sensitivity to

distortions in the mesh. We would expect tME to be more

sensitive though, because its difference molecule extends

far beyond the valence shell of the node for which

localized truncation error is calculated.

Both methods produce localized truncation error

estimates and D* distribution with considerable variability

across several elements. Both are sensitive to water depth,

bathymetric gradient (e.g. ›h/›x; note the highlights over

the continental shelf in figure 2), and changes in

bathymetric gradient (e.g. › 2h/›x 2; note the highlights,

along the continental shelf break and Blake’s Escarp-

ment—among the steepest features in the domain—which

runs north from the Bahamas) as represented by the

corresponding local flow field. Therefore, it is important

that where these sensitivities are present, accurate

bathymetric data are provided. LTEA produces a narrower

Figure 1. WNAT Model Domain (a) as represented by World Vector Shoreline 1:2,50,000 map (40m resolution. Political boundaries are part of thedata set (DMA 2006); the label and grid are added), (b) as represented in the idealized model (25 km length, p/12 rad (158) bearing resolution).The resolution of these figures is coarser. The latitude and longitude grid lines are spaced at intervals of p/18 rad (108). The lattice point [p/18 rad (108)N, p/3 rad (608) W] lies near the southeast corner of the domain.

D. M. Parrish and S. C. Hagen282

Page 8: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

range of localized truncation error values than LTEA-CD

(in our example, 8.6 vs. 15.0 decades for LTEA-CD,

dependent upon D in equation (1)), perhaps because of the

application of a greater number of points in its FD

molecule, which span more elements. Compared to

LTEA-CD, LTEA has a degree of built-in smoothing of

the truncation error. Contrary to this, it would seem,

LTEA produces a wider range of target element sizes and

more variability in spacing (details follow).

Since there is considerable noise in both the LTEA

and LTEA-CD results, it is reasonable to apply a

smoothing function to each data set when making

comparisons between them. We apply the smoothing

function exp{-[(Ds)/15]2/2} where Ds is the distance in

steps along a transect (figure 3). Ds < 1000 m <min(DM). The smoothing function is forced to zero for

the ^11th step and beyond. The smoothing function is

applied to normalized values of D* (D*/min[D*], treating

LTEA and LTEA-CD data sets separately). Next, the ratio

between the normalized, smoothed target element sizes

(D*LTEA-CD:D*LTEA) is computed and averaged (mean).

The normalized, D*LTEA-CD are adjusted by dividing

by the ratio. The resulting data sets, smoothed, normalized

D* (adjusted D*LTEA-CD) provide essentially the same

information for points on the interior (figure 3; see figure 4

for a representation of the bathymetry). Although some

differences near the extreme values are evident, the values

of D* rise and fall consistently with distance along the

transect. We cannot compare performance at the boundary

because LTEA does not produce any data there.

Again, the main advantage of LTEA-CD over LTEA is

that it provides more information: it is able to compute

Figure 2. Comparison of localized truncation error estimate magnitude as computed with LTEA (left) and as with LTEA-CD (right). The calculation oflocalized truncation error for LTEA assumes D ¼ 1000m in equation (1). The same geographic areas are highlighted in by either method. LTEA tends toproduce localized truncation error estimates that are greater than those produced by LTEA-CD.

Figure 3. Normalized target element sizes along a transect, based upon LTEA (solid line) and LTEA-CD (dashed line). The two curves follow the sametrend. Peaks and valleys are generally well matched, with no clear pattern in the differences between the results of the two methods. The transect is shownat right, along with domain boundaries and the subdomain over which LTEA is capable of computing localized truncation error.

2D unstructured mesh generation for oceanic and coastal tidal models 283

Page 9: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

localized truncation error and D* at and near the boundary.

The enhancement of availability of target element size

information for the selected domain and mesh is most

appreciated in the eastern Caribbean Sea, where LTEA is

unable to compute target element sizes within up to

170 km of the coast (figure 5). The ability to compute

target element sizes throughout the domain is critical to

enabling the construction of an efficient mesh (one that is

not over-resolved), especially since the DMs of finite

element meshes for tidal modelling in large domains

frequently span three orders of magnitude (note the legend

of figure 5).

2.7 Advantages of LTEA-CD over LTEA

The most significant advantage of LTEA-CD over LTEA

has is its ability to compute localized truncation error up to

and at the boundary. Additionally, LTEA-CD requires

fewer computations because the difference molecules are

smaller than those of LTEA (in spatial extent and in the

number of nodes) and because the localized truncation

error formula itself has fewer terms. Note that having

fewer computations not only increases the speed of the

calculation, but also decreases the degree of round-off

error. LTEA-CD uses information that is as topologically

Figure 4. Contours of bathymetry for (a) the DEM mesh, (b) mesh 25 km. The isobaths shown are separated in a quasi-logarithmic scheme, beingplaced at depths of 1, 5, 10, 20, 50, 100, 500, 1000, 2000 and 5000m. The deepest isobath shown in the Gulf of Mexico is 2000 m (G), as is the longcontour coincident with the model boundary (faint/dashed line) of Hispaniola’s (H’s) SouthWestern coast. The deepest isobath shown for the AtlanticOcean is 5000m (O).

Figure 5. Normalized, smoothed target element size for LTEA and LTEA-CD (detail for the eastern Caribbean Sea). LTEA-CD provides information inplaces that LTEA cannot (note the dark areas at left).

D. M. Parrish and S. C. Hagen284

Page 10: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

close as possible to the central node, while LTEA uses

information at points up to four elements away, that lie

within a rectangular FD molecule used to compute spatial

derivatives. LTEA ignores some information located

within the FD molecule in order to ensure that each point

of the FD molecule lies within a different element. LTEA

is incapable of estimating derivatives and therefore

truncation error at or near the boundary. LTEA-CD can

estimate derivatives for almost any interior node of

valence six or greater, and for most boundary nodes.

3. Applications

In this part of the paper, we demonstrate the utility of

LTEA-CD with two idealized cases. The two cases

correspond to two approaches to mesh development that

our projects have taken in the past. In the first approach, one

begins with an initial mesh and provides refinement where

the parameters of the project and experience dictate. In the

second approach, one selects a desired mesh size (in terms

of nn or ne) and alters target element sizes to achieve the

desired result. Each of these two approaches individually

represents the trade-off between model accuracy and speed.

Because LTEA-CD is capable of computing D* at the

boundary, there is the probability that, if not designed

properly, the boundary shape will be different between an

initial mesh and corresponding LTEA-CD mesh.

When applying LTEA-CD iteratively, nodes may

accumulate near the boundary because of feedback: as

more nodes are added in coastal areas, the model is able to

simulate greater variability in flow, which produces greater

localized truncation error. Conversely, as the other areas of

the mesh are coarsened, bathymetry is also coarsened and

flow is seen to vary less. As a result, convergence of

successive meshes may not occur. However, it is important

to note that although the bathymetry is coarsened, we

always interpolate bathymetric depths from the same

digital elevation model (DEM).

As one moves from the deep ocean into shallow, coastal

waters, bottom friction becomes increasingly important in

determining the tidal signal. Therefore, if an algorithm

that computes D* is to be fully consistent with the physical

processes there, a variable bottom friction should be

brought into the localized truncation error estimator.

Similarly, advection becomes increasingly important in

shallower and shallower waters. The incorporation of

advection and variable bottom friction is the aim of

continuing research, but is not addressed herein. Rather

than complicating the problem by adding this feature, we

investigate the performance of LTEA-CD using constant

bottom friction, and without advective terms.

3.1 Model domain and bathymetry

We run the ADCIRC model with an idealized boundary of

the WNAT model domain (figure 1(b)). The reason being

that this enables one to change freely the resolution at the

boundarywithout significantly affecting its shape. An initial

boundary is defined by cubic splines passing through points

which, when connected by line segments, are at least 25 km

in length; also, consecutive segments do not differ by more

than p/12 rad (158) bearing. The boundary is reshaped by

successively applying cubic splines and evenly distributing

boundary map vertices at 25 km separation until there is no

perceptible change (based upon visual inspection) in shape

from one iteration to the next. Each time a mesh is generated,

the same reshaped map defines the boundary.

The source of bathymetric data is ETOPO2 (NGDC

2007); a resolution of p/5400 rad (20) was selected. The

untriangulated grid of bathymetric data is interpolated

along meridians onto the points of a uniform triangular

grid (3883 £ 3883 m in the Carte parallelogrammatique

projection (CPP) used by ADCIRC). One edge of each

element of the uniform grid coincides with a meridian.

This equilateral interpolation and triangulation becomes

the DEM for the idealized models. Based upon visual

inspection, the initial mesh for the refinement series

(details below) represents the major bathymetric features

of the domain (figure 4).

3.2 Basis of comparison

We define the DEM mesh for the purposes of providing a

basis of comparison for our simulations. The DEM mesh

has 6,92,263 nodes and 13,79,315 elements. The DEM

mesh is identical to the DEM, with the following

exceptions: (1) The DEM has no boundary, but we

impose the idealized boundary (figure 1(b)) discussed

above, (2) near the coast, within about six elements, the

mesh is slightly distorted from equilateral so as to fit the

boundary as closely as possible and (3) depths shallower

than 1 m (including “depths” above the geoid) are set to

1 m. In this transition zone between equilateral and non-

equilateral elements, the elements range in size from 2600

to 4800 m. Solutions on the DEM mesh become our basis

of comparison for two series of simulations, presented

herein. The DEM mesh has over four times as many

elements and everywhere, and has nominal resolution at

least as high as any other mesh to which it is compared.

3.3 General model parameters and procedure

All simulations in section 3 apply the following parameters.

The friction factor t ¼ 0.0004 s21. The duration is a

simulated period of 15 days. Boundary conditions (M2 tidal

elevations at the open boundary, defined by Le Provost et al.

(1998)) are ramped up hyperbolically over a 5 days period.

The minimum depth is 1.0 m, sufficient to ensure that all

elements remain wetted throughout the simulation (a

condition of stability for a linear run). The solution is

harmonically analysed for steady and M2 constituents at

intervals of 360 s (6 min or 0.1 h); all time steps can be

divided evenly into 360 s.

Initial D* ¼ aD are generated from LTEA-CD results

for all but the last simulation in each series. The arbitrary

2D unstructured mesh generation for oceanic and coastal tidal models 285

Page 11: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

scale parameter a is selected such that nowhere in the

corresponding mesh would our recommended limit of Æ-

ARC (two) be exceeded. These initial meshing require-

ments are then adjusted according to the particular

purposes of each series (discussed below).

Meshing is accomplished via Surface Water Modeling

System (Zundel 2005), SMS.

3.4 Preliminary simulations: semi-uniform meshes

Again, for all practical purposes, LTEA-CD is an a

posteriori method. One must have model results before

applying the method. In order to generate flow fields for

LTEA-CD, we run ADCIRC on semi-uniform meshes.

Since we are applying LTEA-CD with two different

approaches, we provide two different semi-uniform meshes

for starting points.

Two distinct series of simulations are conducted in

order to experiment with different ways of applying

LTEA-CD. Both series of simulations are performed with

practicality being the most important consideration. In the

first series we begin with a coarse mesh and use LTEA-CD

to provide, after a start-up phase, a refined mesh at each

iteration. In the second series we select a fixed value of ne

and apply LTEA-CD iteratively.

3.4.1 Introduction to the refinement series. The basis

for the refinement series is that the modeller may wish to

investigate the accuracy (in terms of convergence to an

acceptable standard) of a series of meshes so that one may

be selected which enables modelling with acceptable

accuracy but has the fewest possible nodes and elements, so

that simulations may be performed repeatedly and quickly

for different scenarios. Therefore the refinement series

begins with a very coarse, semi-uniform mesh. The initial

mesh is semi-uniform because it is easy to generate—noD*

are required. A second question the modeller may have is

that of how much bathymetric information is needed to

provide acceptable accuracy. By beginning with a very

coarse mesh, we allow the method to determine how much

bathymetric information is needed rather than presume this.

We start with a mesh built of nearly equilateral triangles,

DM < 25 km, where, as for all the meshes of both the

refinement and constant ne series, bathymetry is

interpolated from the DEM. We refer to this mesh as

“mesh 25 km”.

We present, for the refinement series, maps of differences

between solutions over the mesh 25 km and the DEM

mesh and between solutions over R9 and the DEM

mesh (figure 6; the top row corresponds to mesh 25 km

while the bottom row corresponds to R9). Differences are

computed for each of six measures: (1) relative elevation

amplitude, (2) elevation phase, (3) relative major semi-

axis, (4) eccentricity, (5) major semi-axis direction and

(6) major semi-axis phase. These indicate that the final

(R9) solution is a vast improvement over the initial

solution (mesh 25 km).

As for the solution over mesh 25 km, we note the

following. Tidal elevation amplitudes are within 0.1 (10%)

of the DEM mesh solution for over half the domain; for

almost the entire Gulf of Mexico and for most of the

Caribbean Sea they are within 0.2 (20%) although there are

distinct regions where the solution is not converged. In the

South Eastern Caribbean Sea there is a large region where

the difference exceeds 0.2 (20%). Elevation phases are

withinp/18 rad (108) for all but a few places where there are

some spikes, and for a portion of the south eastern

Caribbean Sea where phase differences are as great as

p/9 rad (208). The velocity phase and eccentricity

Figure 6. Filled contours of differences between 25 km and DEM mesh simulations (top row) and between R9 and DEM mesh simulations (bottomrow) for the measures indicated above the graphics. Although some differences remain with the R9 simulation, it is obvious that LTEA-CD improved thesolution by selecting levels of refinement for the entire domain. Note the quasi-logarithmic scales.

D. M. Parrish and S. C. Hagen286

Page 12: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

differences are very small, eccentricity differences being

some millionths, and phase being some thousandths of

radians throughout the domain. Velocity direction and

relative major semi-axis differences are generally greater

than the phase and relative elevation amplitude differences.

Velocity major semi-axis directions are generally within

p/18 (108), but there are significant patches where

differences exceed this (additional details below). Note

that the velocity phase errors and tidal ellipse inclination

errors have been computed by minimizing the phase error,

i.e. adding or subtracting 2p (3608) to or from either error

measure, and adding or subtracting p(1808) to or from both

error measures (we measure phase as lag from the semi-

major axis). Further analysis and discussion of the

refinement series follows, in Section 3.5.1 (Results:

refinement series).

3.4.2 Introduction to the constant ne series. The

motivation for our second series is that the modeller may

desire to have a mesh of a certain size (ne or nn),

determined by available resources. For this application, we

set the target ne to 300,000. The initial mesh, called 08 km,

is semi-uniform, having nearly equilateral elements,

DM < 8750 m. Again, bathymetry for all meshes in this

series is interpolated from the DEM.

3.5 Details of refinement series

The purpose of executing the refinement series is to

demonstrate that LTEA-CD is capable of producing D*

distributions that, when implemented, produce sea surface

deviations and velocity fields that converge. We

emphasize that LTEA-CD is used to determine a

distribution of D*, not absolute requirements for D*.

In the refinement series, D* larger than the current mesh

are reset to DM, while D* smaller than the DEM resolution,

DDEM, are set to DDEM (3883m) for the practical purpose

of preventing the elements from becoming too small, and

because the DEM mesh is our basis of comparison.

We start with mesh 25 km. ADCIRC is executed using

mesh 25 km and D is computed from the results. The data

points where values of D are available are triangulated and

interpolated (linearly) or extrapolated (nearest neighbour)

onto the mesh from which they were derived. Triangu-

lation and interpolation/extrapolation are performed by

SMS. After smoothing D and maximizing a, constrained

by Æ-ARC (using our own codes), the resulting D* are

applied to the generation of the next mesh, C, by allowing

only coarsening, that is where D* , 25 km, we reassign:

D* ; 25 km. (The triangulation, interpolation, extrapol-

ation, smoothing and maximizing steps apply to all

meshes presented herein that rely upon D* computed from

LTEA-CD). The principal reason to apply a coarsening

step is that we know from previous research (Kojima

2005) that we are able to have elements as coarse as

140 km where the bathymetry is both deep (deeper than

about 5000 m) and gradually varying.

The model is run again on mesh C and D* is computed

from this solution. Therefore the semi-uniform mesh, its

solution, the coarse mesh, and its solution represent the

components of a start-up phase for the refinement series.

The meshing requirements from the solution on mesh C

are generated and applied as refinement only. That is,

where D* . DM, D* we reassign: D* ; DM. Also, where

D* , DDEM, D* we reassign D* ; DDEM. The adjusted D*

are applied to the generation of the next mesh, the first

refined mesh, called R1. The second through ninth refined

meshes, called R2, R3, . . . , R9 are generated with the

same process that R1 was generated from: each mesh is

generated from refinement only criteria that are based on

model results over the previous mesh. Mesh properties are

presented in table 1.

Bathymetric data from the DEM (not the DEM mesh)

are interpolated linearly onto each mesh of the refinement

series and of the constant ne series.

3.5.1 Results: refinement series. Meshes R1–R9 are

reflections of their predecessors. We present R9 in figure 7.

Iterations of LTEA-CD and corresponding meshing has

resulted in reduced differences between the solutions over

mesh 25 km and R9 by strategic refinement. However, the

relationship between magnitude of differences and the

level of refinement is not completely intuitive. Given a

location, near-field refinement can result in both near-field

and far-field improvements (figure 8); Hagen et al. (2000)

reported the same phenomenon in an application of

LTEA. In the South Eastern Caribbean Sea, there is, for

example, a spike in relative major semi-axis difference for

the mesh 25 km solution (indicated by the arrow in figure 8).

This spike is not present in the solution over R9.

Elimination of this spike is accounted for by far-field

refinement (cf. bottom-left and bottom-right quadrants of

figure 8). Note also that the refinements in the South

Eastern Caribbean Sea result in reduced differences in the

near-field. Near-field reduction in differences with

increased resolution is predominant, but not universal:

resolution enhancement is not always accompanied by

significant, local reductions in differences.

Table 1. Properties of the meshes of the refinement series.

Element sizes

ID ne nn Time step (s) Small Large

25 km 33,208 17,024 60 21,593 33,010C 16,624 8663 90 20,025 90,028R1 29,161 15,065 60 8824 86,310R2 60,887 31,146 36 3520 83,128R3 1,07,411 54,600 24 3314 79,015R4 1,45,698 73,834 24 3196 80,590R5 1,60,453 81,237 20 3316 75,496R6 1,85,963 94,047 15 3200 75,599R7 2,07,449 1,04,848 18 2891 74,977R8 2,30,985 1,16,694 15 3160 74,699R9 2,42,287 1,22,388 12 2029 76,357DEM 1,379,315 6,92,263 12 2594 4790

2D unstructured mesh generation for oceanic and coastal tidal models 287

Page 13: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

Cumulative area fraction error (CAFE) curves (Luettich

and Westerink 1995; see also Hagen et al. 2001) are used

to compare results of the 11 simulations in the refinement

series. A set of CAFE curve is produced for each of six

error measures of the tidal elevation (1–2) and tidal

ellipse (3–6): differences in (1) relative elevation

amplitude, (2) elevation phase (rad), (3) relative major

semi-axis velocity, (4) eccentricity, (5) major semi-axis

direction and (6) major semi-axis phase (figure 9(a)–(e)).

CAFE plots present cumulative area (fraction of total

domain) vs. positive and negative differences exceeded.

Note that a perfect solution would result in a single

vertical line plotted at zero on the horizontal axis. Each

plot is truncated along the vertical axis at a value equal to

0.01 (1%) of the cumulative area. The coordinate where

the horizontal axis intersects the curve displays the error

levels exceeded within 0.01 (1%) of the area of the

domain. Note that the error levels associated with 0.99

(99%) of the total domain area are plotted above the

horizontal axis. Again, the bases of comparison are the

results from the DEM mesh simulation. With each

successive mesh, the CAFE curves become steeper and the

range of errors at a given level become narrower.

Mesh C actually appears to perform better than mesh

25 km (note that the elevation differences, both amplitude

and phase, of mesh C are less than those of mesh 25 km,

figures 9(a) and (b)), even though the latter has, nominally,

at least as high resolution everywhere. It is likely that this

is because C, having fewer nodes, possesses smoother

bathymetry, particularly in the deeper, portions of the

domain that have lower bathymetric gradient (Hagen et al.

2006). Since we are comparing to the DEM mesh’s

Figure 7. Mesh R9 (final mesh) of the refinement series. Meshes R1–R8 show a gradual progression to this point.

D. M. Parrish and S. C. Hagen288

Page 14: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

solution, we expect also that mesh C better represents the

bathymetry than mesh 25 km. Luettich and Westerink

(1995), in another study over the WNAT model domain,

demonstrated the domain’s sensitivity to bathymetric

resolution alone in a set of two nonlinear M2 simulations:

increasing the bathymetric resolution from 75 to 19 to

9 km Westerink et al. (1994) mention, (our emphasis)

“semi-diurnal response (in the Gulf of Mexico is) very

sensitive to the grid resolution, bathymetry, as well as the

bottom friction coefficient”.

While nn and ne in each successive mesh appear to

increase at a rate that does not diminish (therefore the

mesh does not appear to converge), the performance of the

meshes appears to be consistent for up to ten iterations

(figure 10; note that differences are plotted vs. the number

of elements in each mesh). Additionally, the rate of

convergence for elevation amplitude and phase is

O[(DM)2] (1 / ne 21 / D2), while that of the tidal

ellipses (i.e. semi-major axis, eccentricity, inclination and

phase) appears to be O(DM).

Figure 8. Detail of south eastern Caribbean Sea. Mesh refinement does not always occur in regions of high error. Error level in one location may beaffected by model performance in another location. Top-left is relative semi-major axis error for mesh 25 km, filled contours. At top-right, the relativesemi-major axis error for mesh R9. Bottom-left: mesh 25 km. Bottom-right: mesh R9.

2D unstructured mesh generation for oceanic and coastal tidal models 289

Page 15: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

Figure 9. CAFE (cumulative area fraction “error”—read “difference”), curves for each mesh of the refinement series: (a) relative elevation amplitude,(b) elevation phase, (c) relative major semi-axis, (d) eccentricity, (e) major semi-axis direction, (f) major semi-axis phase. The selection of domain, rangeand line widths preserves distinction among meshes, while allowing the representation of the essential information. It is clear that the velocity solution isconverging. The elevation solution converges non-monotonically.

D. M. Parrish and S. C. Hagen290

Page 16: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

3.6 Details constant ne series

For the constant ne series, we apply LTEA-CD iteratively,

allowing only about 300,000 elements into each mesh

(about 0.22 (22%) as many elements as the DEM mesh).

This is accomplished by selecting a, then, where D* ,

DDEM, D* we reassign: D* ; DDEM. Again, the solution on

the DEM mesh is the basis of comparison. As with the

previous series, we do not allow a to increase to such a

value that the resulting mesh would violate Æ-ARC # 2.

Figure 10. Mean differences of solutions of the refinement series, area-weighted over the domain (hollow dots,W), for: elevation (a) relative amplitudeand (b) phase and for: (c) relative major semi-axis velocity, (d) eccentricity, (e) major semi-axis direction and (f) major semi-axis phase. Differences foreach mesh are plotted against the number of nodes for each respective mesh. The elevation solution converges at O(D2

M / ne 21), solid line (——); thevelocity solution converges at approximately O(DM / ne 21/2), dashed line (– – –). For these calculations, eccentricity is calculable (due to file format)to 1026 round-off error; all other values presented are greater than round-off.

2D unstructured mesh generation for oceanic and coastal tidal models 291

Page 17: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

The Æ-ARC provides us with an indicator of a stopping

point for the series. The Æ-ARC allows us to compute D*

for three iterations, along with the corresponding meshes:

D1–D3 (table 2, figure 11). The mesh generated from the

results of the third simulation was in slight violation of Æ-

ARC; however, with some very minor manual editing, Æ-

ARC was decreased to 2.2 or less everywhere and ,2.0

for the vast majority of the nodes of the mesh.

3.6.1 Results: constant ne series. A second set of CAFE

curves are used to compare results of the four simulations

of the constant ne series (figure 12). The same parameters

are analysed as in the refinement series. Again, the bases

of comparison are the results from the DEM mesh

simulation. While there is notable symmetry among the

CAFE curves for the tidal ellipses, the curves for tidal

elevation (both amplitude and phase) are decidedly

asymmetric, however the asymmetry is consistent for all

four simulations (the curves are “heavier” on the same

side of zero), unlike that of the refinement series. There is

no clear distinction in performance among meshes D1–

D3. The CAFE curves for the various measures of

difference appear in different orders, e.g. for major semi-

axis phase difference (figure 12(f)), the ordering of the

curves is as one would expect: D1–D3, however for major

semi-axis direction they are ordered oppositely. It is

interesting that there is decidedly less intertwining of the

elevation curves for the constant ne series than for the

refinement series: one might expect the opposite to be the

case—keeping the same number of elements, one might

expect the curves to be less distinguishable, while adding

nodes, one might expect more distinction.

As compared to mesh 08 km, area-weighted mean

differences for the tidal ellipses improve by factors of three

(eccentricity and major semi-axis phase), two and one-half

(relative major semi-axis), and two (major semi-axis

direction).

4. Discussion

An idealized boundary has been applied in our examples so

that we may have full control over the resolution at the

boundary without significantly altering its shape. Again,

this is needed in order to be able to apply the results of

LTEA-CD, which can be used to calculate D* at the

boundary. Of course, our eventual goal is to be able to apply

LTEA-CD to a domain having a realistic, complicated

boundary. In a practical application, there would be some

trade-off between LTEA-CD’s D* and the boundary shape;

one may ameliorate the conflict by making a small enough

near the boundary. If the area of interest is much smaller

than the model domain, one may allow the far-field

boundary to change from iteration to iteration without

significantly affecting the near-field solution, insofar as

comparison to tidal elevations are concerned.

The applications of LTEA-CD presented herein are for

an idealized domain, so as to allow for flexibility and

consistency in the definition of the boundary throughout

the domain. While we recognize the practical need to

extend the method to domains having complicated

boundaries, we note that in order to properly address the

issue of complicated boundaries and resolution in near

shore and estuarine areas we will need to incorporate

variable bottom friction and advection into our LTEA-CD

scheme. These challenges are the aim of ongoing research.

Table 2. Properties of the meshes of the constant ne series.

Element sizes

ID ne nn Time step (s) Small Large

08 km 3,00,154 1,51,341 24 5507 9962D1 3,00,351 1,51,855 24 2806 26,557D2 2,99,184 1,51,352 18 3146 62,236D3 2,99,187 1,51,391 15 3175 1,05,474DEM 13,79,315 6,92,263 12 2594 4790

Figure 11. Actual element sizes for the initial and three successive meshes of the constant ne series. The meshes are too dense to present herein.

D. M. Parrish and S. C. Hagen292

Page 18: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

4.1 Refinement series

It is very clear from the CAFE curves of the tidal ellipse

differences that the tidal ellipses are converging with each

iteration of LTEA-CD since each successive curve lies

beneath its predecessor, with a few exceptions. In addition,

there is some intertwining of the major semi-axis direction

difference curves corresponding to the simulations on

meshes R3–R6.

In short, the convergence for the elevation amplitude

and phase is oscillatory. There is much intertwining of the

CAFE curves for the elevation differences, however there

Figure 12. CAFE curves for each mesh of the constant ne series: (a) relative elevation amplitude, (b) elevation phase, (c) relative major semi-axis, (d)eccentricity, (e) major semi-axis direction, (f) major semi-axis phase. The selection of domain, range and line widths preserves distinction amongmeshes, while allowing the representation of the essential information. There is marked improvement in the solution over mesh D1 as compared to thatover mesh 08 km. Repeated application of LTEA-CD for a fixed ne does not appear to be effective at further improving the solution.

2D unstructured mesh generation for oceanic and coastal tidal models 293

Page 19: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

is a trend toward narrowing, falling curves as one

progresses from iteration to iteration. The intertwining

results partly from an exchange between negative and

positive errors from one iteration to the next. If we lump

the positive and negative errors together, the separation in

differences among the simulations is clarified (figure 13).

Note that the R9 curve (corresponding to the final

simulation) is beneath every other curve for almost every

fraction of area. The curves for R4 and R8 are still

intertwined: their differences are about 0.003 (0.3%) down

to about the 0.03 (3%) area level, where separation begins.

Overall, mesh R9 performs best: it is associated with the

smallest differences for over 0.2 (20%) of the area of the

domain; over the remainder of the domain, the differences

associated with R9 are barely distinguishable from those

of any other mesh. In terms of absolute elevation phase

differences, R9 outperforms all other meshes for 0.69

(69%) of the area of the domain (R9 is best for area level 1

down to 0.4 (1 2 0.4 ¼ 0.6), and from 0.09 downward

(0.6 þ 0.09 ¼ 0.69)). The oscillatory nature of the

convergence may stem from the lack of an elevation

term in our equation (7) for computing tME. Extension of

the Taylor series would certainly provide elevation terms,

however this would also introduce computational

difficulties, for example, the choice of submesh shape is

unclear were eight nodes to be used for computation of

tME on the interior.

Elevation amplitude and phase are O[(DM)2] accurate

when nodes/elements are added according to the

refinement scheme presented herein. The velocity field,

however, is O(DM) accurate under the same scheme

(additional discussion below).

4.2 Constant ne series

There is a marked improvement in the performance of

each of the LTEA-CD meshes (D1–D3) as compared to

mesh 08 km. However, performing iterations with LTEA-

CD, while holding ne constant appears to be of little effect

in terms of solution convergence. Performing iterations

with LTEA-CD appears unprofitable for constant ne,

where elements smaller than a certain size are disallowed.

The fact that no gain in solution accuracy is to be had by

repeatedly remeshing with the same ne suggests that any

advantage gained in placing nodes in better positions to

theoretically better represent the flow field is offset by the

resulting increasing amount of distortion of the elements.

Distortion of elements is known to cause increases in

truncation error (Hagen 1998) and to cause the increase of

numerical artefacts in the form of a folding dispersion

relationship (response frequencies differing from those

of model forcing, and, in moderate to extreme cases, dual

wave number response for each forcing frequency;

Atkinson et al. (2004)).

4.3 Recommendations

Based upon observations represented thus far, we

recommend that applications of LTEA-CD be conducted

in one of two ways. The modeller may begin with a coarse

mesh (not necessarily uniform) and iterate with LTEA-CD

until the resultant mesh has about the number of

nodes/elements desired, or until Æ-ARC reaches its

limit. A second approach begins with a mesh that has the

number of nodes/elements desired, and a single iteration of

LTEA-CD is performed in order to produce a second mesh.

Note that if the second approach is taken, it is important

that the initial mesh enable an accurate representation (on a

per element basis) of the flow field, otherwise one cannot

expect LTEA-CD to produce from that flow field a better-

performing mesh in only one iteration.

It is also important to smooth the results of LTEA-CD,

just as it has been with LTEA. This is necessary because

meshes applying the unsmoothed results become imprac-

ticably large for typical selected values of Æ-ARC. We

cannot at this time recommend a particular smoothing

function; thus far however, we have used isotropic

Gaussian smoothing with LTEA-CD. We also recall that

Figure 13. CAFE curves for each mesh of the refinement series: (a) absolute value of relative elevation amplitude difference and (b) absolute value ofelevation phase difference. Convergence of the elevation solution is apparent.

D. M. Parrish and S. C. Hagen294

Page 20: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

LTEA appears to produce meshes that perform very well

when target element sizes are smoothed by imposing Æ-

ARC , 2 (Kojima 2005, Hagen et al. 2006). Ideally,

however, smoothing of D* should account for directional

features that are represented by only a few elements in the

transverse direction (e.g. a shipping channel whose cross

section is represented by two elements).

Reconsidering the constant ne series, we note that in our

study we have not chosen a specific area of interest. If

there were a particular region the modeller were interested

in, it may be worthwhile to iterate with LTEA-CD for a

few steps until the resolution of the mesh in the vicinity of

the area of interest becomes more refined (assuming that it

does). This would allow the modeller to further refine the

area of interest with greater ease as compared to refining

the uniform mesh, since the size of the elements near the

area of interest would, in comparison, already be

approaching that needed to resolve the area of interest.

5. Conclusion

LTEA-CD is capable of providing physically-based (i.e.

based upon discrete physics) target element size

distributions on the entire domain, both along the

boundary and on the interior. LTEA-CD is a substantial

improvement over LTEA (which can compute target

element size distribution only on the interior) for this

reason and because, on the interior of the domain, LTEA-

CD is capable of producing essentially the same

information as LTEA, which, in turn has been applied to

produce operational meshes—over which the fully non-

linear shallow water equations are solved—that produce

error statistics (model vs. historical, measurement based

data) that are within the errors of the historical data

themselves (e.g. Mukai et al. 2001).

LTEA-CD can be used to compute target element sizes

orders of magnitude faster and with greater ease than

LTEA (cf. equations (1) and (7), and consider the 9 £ 9

difference molecule of LTEA vs. the seven-node

difference molecule of LTEA-CD). This practicality of

LTEA-CD will lend itself well to adaptive mesh

refinement schemes.

Applied iteratively, LTEA-CD produces a series of

meshes that are O[(DM)2] accurate for the elevation field

and O(DM) accurate for the velocity field. To our

knowledge, LTEA has never been applied iteratively.

Any improvement in the solution by iteratively redis-

tributing the nodes using LTEA-CD appears to stem from

the addition of nodes, not from redistribution. None-

theless, an LTEA-CD-based mesh produces a more

converged solution than a uniform mesh having the

same number of elements. Conversely, a uniform mesh

requires more elements than an LTEA-CD-based mesh to

produce a solution converged to the same degree.

Although LTEA-CD is capable of computing target

element size at the boundary, it should be noted that

LTEA-CD is based upon the linearized shallow water

equations. Therefore, LTEA-CD does not directly account

for depth dependent bottom friction (but note that it does

account variation in velocities, a function of depth and

friction and, although not implemented here, geographi-

cally varying bottom friction coefficient) or advection,

both important processes in the real ocean.

In introducing LTEA-CD, we noted that simulation

results are actually not required in order to apply, LTEA-

CD, only a sufficient amount of data on the flow field. We

imagine a future in which depth-integrated velocities over

the globe could be measured via remote sensing

technology. A resultant data set could be used in an

application of LTEA-CD. It is already possible to measure

not only depth-integrated velocities, but velocity profiles

using acoustic Doppler current profilers; Visbeck (2002)

provides an example application. Mollo-Christensen et al.

(1981) estimated ocean current velocity using an infrared

image. Crocker et al. (2007) discuss some of the

difficulties related to estimating surface currents from

infrared and ocean colour satellite imagery.

6. Future work

During the next phase of research, we will develop an

upgraded localized truncation error and target element

size-computing algorithm that will account for, at a

minimum, nonlinearities associated with variable bottom

friction and advection. Only then can we expect the

algorithm to be representative of the physical processes in

shallow, coastal areas. The upgraded algorithm will be

tested on a complicated estuary or similar coastal system.

The key to finding a practicable expression for localized

truncation error and corresponding target element size lies

in accurately computing the derivative terms in the

truncation error series (i.e. more accurately than the model

computes them). Once the technique for computing the

derivative terms has been determined, the remainder of the

computations are comparatively straightforward.

With the upgraded algorithm, computational efficiency

will be a more important consideration than with LTEA-

CD, since it is expected that time-stepping will be

necessary to accommodate the nonlinearities. In addition,

one of our goals is to make the mesh generation as

automatic as possible, reducing the time span and human

time required to generate workable meshes.

Acknowledgements

We acknowledge Deidre A. Parrish for a brief but

significant suggestion. This research was in part

conducted under award NA04NWS4620013 from the

National Oceanic and Atmospheric Administration

(NOAA), US Department of Commerce. The statements,

findings, conclusions and recommendations are those of

the authors and do not necessarily reflect the views of

NOAA or the Department of Commerce. This research

2D unstructured mesh generation for oceanic and coastal tidal models 295

Page 21: International Journal of Computational Fluid Dynamicsanalysis with complex derivatives', International Journal of Computational Fluid Dynamics, 21:7, 277 - 296 To link to this article:

Dow

nloa

ded

By:

[Uni

vers

ity o

f Cen

tral F

lorid

a] A

t: 17

:59

2 N

ovem

ber 2

007

was in part conducted under Award N00014-02-1-0150

from the National Oceanographic Partnership Program

(NOPP) administered by the Office of Naval Research

(ONR). The statements, findings, conclusions, and

recommendations are those of the authors and do not

necessarily reflect the views of ONR or NOPP and its

affiliates. The authors gratefully acknowledge the support

of the College of Engineering & Computer Science and

the I2Lab at the University of Central Florida.

References

Atkinson, J.H., Westerink, J.J. and Luettich, Jr, R.A., Two-dimensionaldispersion analyses of finite element approximations to the shallowwater equations. Int. J. Numer. Methods Fluid, 2004, 45(20),715–749.

Crocker, R.I., Matthews, D.K., Emery, W.J. and Baldwin, D.G.,Computing coastal ocean surface currents from infrared and oceancolor satellite imagery. IEEE Trans. Geosci. Remote S., 2007, 45(2),435–447.

DMA (Defense Mapping Agency, now the National Geospatial-Intelligence Agency), World Vector Shoreline. Available online at:http://rimmer.ngdc.noaa.gov/coast/ (accessed 29 June 2006).

Daugherty, R.L., Franzini, J.B. and Finnemore, E.J., Fluid Mechanicswith Engineering Applications, 8th ed., pp. 142–143, 1985(McGraw-Hill: New York).

Foreman, M.G.G., A two-dimensional dispersion analysis of selectedmethods for solving the linearized shallow water equations.J. Comput. Phys., 1984, 56(2), 287–323.

Greenberg, M.D., Advanced Engineering Mathematics, 2nd ed.,pp. 1177–1178, 1998 (Prentice Hall: Upper Saddle River, NJ).

Greenberg, D.A., Dupont, F., Layard, F.H., Lynch, D.R. and Werner, F.E.,Resolution issues in numerical models of oceanic and coastalcirculation. Cont. Shelf Res., Submitted 2006.

Hagen, S.C., Finite element grids based on a localized truncation erroranalysis, PhD thesis, University of Notre Dame, 1998.

Hagen, S.C., Estimation of the truncation error for the linearized, shallowwater momentum equations. Eng. Comput., 2001, 17(4), 354–362.

Hagen, S.C. and Parrish, D.M., Unstructured mesh generation for thewestern North Atlantic tidal model domain. Eng. Comput., 2004,20(2), 136–146.

Hagen, S.C., Westerink, J.J. and Kolar, R.L., One-dimensional finiteelement grids based on a localized truncation error analysis. Int.J. Numer. Methods Fluid, 2000, 32(2), 241–261.

Hagen, S.C., Westerink, J.J., Kolar, R.L. and Horstman, O., Two-dimensional, unstructured mesh generation for tidal models. Int.J. Numer. Methods Fluid, 2001, 35(6), 669–686.

Hagen, S.C., Horstman, O. and Bennett, R.J., An unstructured meshgeneration algorithm for shallow water modeling. Int. J. Comput.Fluid Dyn., 2002, 16(2), 83–91.

Hagen, S.C., Zundel, A.K. and Kojima, S., Automatic, unstructured meshgeneration for tidal calculations in a large domain. Int. J. Comput.Fluid Dyn., 2006, 20(8), 593–608.

Hannah, C.G. and Wright, D.G., Depth dependent analytical andnumerical solutions for wind-driven flow in the coastal ocean.In Quantitative Skill Assessment for Coastal Ocean Models(Coastal and Estuarine Studies 47), edited by D.R. Lynch andA.M. Davies, pp. 125–152, 1995 (American Geophysical Union:Washington, DC).

Kashiyama, K. and Okada, T., Automatic mesh generation method forshallow water flow analysis. Int. J. Numer. Methods Fluids, 1992,15(9), 1037–1057.

Kojima, S., Optimization of an Unstructured Finite Element Mesh for

Tide and Storm Surge Modeling Applications in the Western North

Atlantic Ocean. Master’s thesis, University of Central Florida, 2005.

Le Provost, C. and Vincent, P., Some tests of precision for a finite element

model of ocean tides. J. Comput. Phys., 1986, 65(2), 273–291.Le Provost, C., Genco, M.-L. and Lyard, F., Modeling and predicting

tides over the world ocean. In Quantitative Skill Assessment for

Coastal Ocean Models (Coastal and Estuarine Studies 47), edited by

D.R. Lynch and A.M. Davies, pp. 175–201, 1995 (American

Geophysical Union: Washington, DC).

Le Provost, C., Lyard, F., Molines, J.M., Genco, M.L. and Rabilloud, F.,

A hydrodynamic ocean tide model improved by assimilating a

satellite altimeter-derived data set. J. Geophys. Res., 1998, 103(C3),5513–5529.

Loder, J.W., Topographic rectification of tidal currents on the sides of

Georges Bank. J. Phys. Oceanogr., 1980, 10(9), 1399–1416.Luettich, Jr. R.A. andWesterink, J.J., Continental shelf scale convergence

studies with a Barotropic tidal model. In Quantitative Skill

Assessment for Coastal Ocean Models (Coastal and Estuarine Studies

47), edited by D.R. Lynch and A.M. Davies, pp. 349–371, 1995

(American Geophysical Union: Washington, DC).

Luettich, Jr. R.A. and Westerink, J.J., Formulation and Numerical

Implementation of the 2D/3D ADCIRC Finite Element Model

Version 44.XX. Available online at: http://www.adcirc.org/

adcirc_theory_2004_12_08.pdf (accessed 10 February 2007).

Luettich, Jr. R.A., Westerink, J.J. and Scheffner, N.W., ADCIRC: an

Advanced Three-Dimensional Circulation Model for Shelves, Coasts

and Estuaries, Report 1: Theory and Methodology of ADCIRC-2DDI

and ADCIRC-3DL, (Dredging Research Program Technical Report

DRP-92-6) 1992 (US Army Corps of Engineers: Washington, DC).

Lynch, D., Ip, J., Naimie, C. and Werner, F., Convergence studies of

tidally rectified circulation on Georges Bank. In Quantitative Skill

Assessment for Coastal Ocean Models (Coastal and Estuarine Studies

47), edited by D.R. Lynch and A.M. Davies, pp. 153–174, 1995

(American Geophysical Union: Washington, DC).

Mollo-Christensen, E., Cornillon, P., Da, S. and Mascarenhas, A.,

Method for estimation of ocean current velocity from satellite

images. Science, New Series, 8 May 1981, 212(4495), 661–662.Mukai, A.Y., Westerink, J.J., Luettich, Jr., and Mark, D., Eastcoast 2001,

a Tidal Constituent Database for the Western North Atlantic, Gulf of

Mexico and Caribbean Sea, 2001 (US Army Corps of Engineers:

Washington, DC).

NGDC (National Geophysical Data Center), ETOPO2 2-Minute Global

Relief Ver 2. Available online at: http://www.ngdc.noaa.gov/mgg/

gdas/gd_designagrid.html (accessed 21 February 2007).

Reddick, H.W. and Miller, F.H., Advanced Mathematics for Engineers,

3rd ed., pp. 426–433, 1938 (Wiley: New York).

Saada, A.S., Elasticity: Theory and Applications, p. 210, 1974

(Pergamon: New York).

Sadd, M.H., Elasticity: Theory, Applications, and Numerics, p. 249, 2005

(Elsevier: Burlington, Mass).

Timoshenko, S.P. and Goodier, J.N., Theory of Elasticity, 3rd ed.,

pp. 18–58, 1970 (McGraw-Hill: New York).

Visbeck, M., Deep velocity profiling using lowered acoustic doppler

current profilers: bottom track and inverse solutions. J. Atmos. Ocean.

Tech., 2002, 19(5), 794–809.Weisstein, E.W., Cauchy–Riemann equations, Available online at: http://

mathworld.wolfram.com/Cauchy–RiemannEquations.html (accessed

9 June 2006).

Westerink, J.J., Luettich, Jr., R.A. and Muccino, J.C., Modeling tides in

the western North Atlantic using unstructured graded grids. Tellus,

1994, 46A, 178–199.Zundel, A.K., Users Manual for the Surface-Water Modeling System,

Version 9.0. 2005 (Brigham Young University: Provo, Utah).

D. M. Parrish and S. C. Hagen296


Recommended