Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Annu. Rev. Fluid Mech. 1991.

23: 389-412
Copyright 1991 by Annual Reviews [nco All rights reserved

ANNUAL
REVIEWS

Further

Quick links to online content

COASTAL-TRAPPED WAVES
Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org
by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

AND WIND-DRIVEN
CURRENTS OVER THE
CONTINENTAL SHELF
K. H. Brink
Department of Physical Oceanography, Woods Hole Oceanographic
Institution, Woods Hole, Massachusetts 02543
KEY WORDS:

ocean-current variability, wind driving, coastal oceanography,


alongshore propagation

1.

INTRODUCTION

The 1970s saw a remarkable development in the understanding of wind


driven current variability over the continental shelf, allowing Allen (1980)
to summarize many of the keystones of our current understanding of the
subject. Since 1980, our understanding of such processes has become a
good deal more sophisticated, especially in terms of coastal-trapped-wave
theory. In fact, the point has been reached where models and observations
are now often compared with a reasonable expectation of quantitative
agreement. Along the way, some exciting new physical insights have also
been gained. It thus seems timely to revisit the problem of wind-driven
variability over the continental shelf, and therefore the present offering is
intended to be a continuation of Allen's (1980) original fine review.
The following material focuses mainly on variability having time scales
longer than a day but not more than a few weeks. In all cases, attempts
are made to emphasize the relation between theoretical concepts and
observed behavior in the coastal ocean. The approach is to address "water
column" models in which nonadiabatic effects, while often present, are
not of overwhelming importance. Within this classification, deterministic
problems expressed in terms of coastal-trapped-wave theory are treated
389
0066-4189/91/0115--0389$02.00

BRINK

390

first, followed by consideration of stochastic models that may or may not


be centered on wave models. A central focus on coastal-trapped-wave
theory is appropriate because it has proven to be a versatile approach,
yielding successful predictions as well as physical insight.

2.

COASTAL-TRAPPED-WAVE MODELS

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

2.1

Introduction

Many aspects of wind-driven variability over the continental margin can


be explained in terms of the properties of coastal-trapped waves. All
continental margins allow the existence of such waves, and the sense of
alongshore phase propagation is always such that the coast is on the right
(left) when looking in the direction of phase propagation in the Northern
(Southern) Hemisphere. Such waves generally have periods longer than
the inertial period, so that they often play a role in the response of the
ocean to atmospheric weather changes, which typically have time scales
of a few days. The waves are important because they are readily excited
and, through their propagation, spread the ocean's response in the along
shore direction. Many important wave properties have been summarized

by Huthnance et al (1986).
2.2

Formulation

Before proceeding to the properties of free coastal-trapped waves, it is


useful to consider the mathematical formulation of the problem. The
linearized equations of motion for a stratified, Boussinesq ocean are

u, -jv
v,+ju

= -

!,

( l a)

-Py+ -T,

( l b)

-Px+

Po

Po

Po

Po

( Ic)
( l d)
and
( Ie)
where u, v, and ware the velocity components in the x, y, and z directions,
respectively. The Coriolis parameter isj( = 2Q sin 1>, where Q is the Earth's
rotation rate and 1> is latitude), the acceleration due to gravity is g, the
pressure perturbation is p, and density is written as

WIND-DRIVEN CURRENTS OVER THE SHELF

391

Po+ p(z)+p'(x,y, z, t),

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

where Ip'l p Po. The horizontal component of turbulent stress is


(,X, ,V) and will be needed to represent wind driving and the bottom
frictional drag. Subscripts with respect to independent variables repre
sent partial differentiation. The offshore and upward directions are taken
to be x and z, respectively.
In the absence of turbulent stresses, it is straightforward to reduce the
system (1) to the vorticity equation

Pxxt+Pyyt+(J2 02)(pzt\h
+

ot2

N2

(2)

The buoyancy frequency squared is


N2

- !Lpz.

(3)

Po

The vorticity equation is to be solved subject to no normal flow at the


coast,

Pxt+Jpy
N2
pz+ P

at

x =

0,

(4a)

a free surface,
9

at

z =

0,

(4b)

no flow through the bottom

and boundedness far from the coast,

Px
p

-+

as

x -+

00.

(4d)

Free coastal-trapped waves are assumed to have the form


=

exp [i(wt+ly)]F(x,

(5)

with the depth here chosen to be a function only of distance offshore, i.e.
h h(x). The problem has now been reduced, for a given alongshore
wavenumber I, to an eigenvalue problem for F(x, z) and w. Even with
these simplifications, analytical solutions to (2) subject to (4) are very
difficult to find. Thus, F is usually found numerically via resonance iter
ation (Wang & Mooers 1976, Brink 1982) or inverse iteration (Huthnance
1978). Computer programs for obtaining modal structures and related
=

392

BRINK

properties (described by Brink & Chapman 1987) are freely available and
cover a fairly broad range of realistic situations.
System (2-4) can be simplified in the "long-wave" limit, where along
shore scales are large relative to cross-shelf scales (82/8y2 82/8x2) and
where frequencies are low (02/012 /2). In this case, the problem becomes
separable and is

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

Pxx+f 2

(N2)
Pzt

Pxt+fpy =
pz +

f2

N2

-p
g

()

0,

at

(6a)

x = 0,

at

+hx(Pxt+fPY)

(6b)

0,

(6c)

at

z =

-hex),

(6d)

and
Px-40

as

X-><X).

(6e)

The solutions can be written as


p

Y,ly, t)F,,(x, z),

(7a)

where
(7b)
The eigenvalue of the problem is now em and members of the complete set
(Clarke 1977) of eigenfunctions Fn are orthogonal subject to
fbnm

-h

FnFmdzlx=o+

roo FnFmdxlz=-h(x),
Jo hx

(8)

where bn m is the Kronecker delta. The particular normalization implied by


(8) must be used in order to conserve energy as the wave propagates
through a slowly varying medium (Brink 1989).
It is straightforward in the long-wave limit to solve for the wind-forced
problem (with no bottom stress). Further, it is frequently assumed that
forcing is only through an alongshore wind stress that does not vary in
the cross-shelf direction [rX 0, rJ' rJ'(y, t)]. In this case, the pressure can
still be expressed in terms of the eigenfunctions,
=

WIND-DRIVEN CURRENTS OVER THE SHELF

393

where now the amplitude functions obey


(9b)

Here , is the alongshore wind stress at the surface, and


Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org
by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

(9c)

is the coupling coefficient of mode n to the wind stress if the surface mixed
layer is infinitesimally thin (see, for example. Clarke 1977).
A scaling of (2) leads to a stratification parameter (Burger number)

where H, No, and L are typical values of depth, buoyancy frequency,


and cross-shelf horizontal scale, respectively. This dimensionless number
expresses the relative strength of stratification; alternatively, it can be
thought of as the ratio of the internal Rossby radius of deformation to the
horizontal length scale. The barotropic (unstratified) limit of (2), i.e. S --+ 0,
leads to the lowest order vorticity balance
( 10)
that is, that the vertical velocity does not vary with depth. In this barotropic
limit, the flow is then described by ( 10), (4b), and (4c) (e.g. Clarke & Brink
1985).

In the following, the properties of the free modes are first treated. The
excitation of coastal-trapped waves is then considered, and finally com
parisons of models with observations are discussed.
2.3

Free-Wave Properties

The general properties of free coastal-trapped waves are summarized nicely


by Huthnance (1975, 1978) and by Clarke (1977). For arbitrary topog
raphy and stratification, there is a free-wave mode similar to the baro
tropic Kelvin wave plus an infinite set of higher mode, more slowly
propagating waves. All of these wave modes have phase propagation
in the sense described above (Section 2.1), provided the depth increases
monotonically with distance from the coast. Slower (higher order) modes

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

394

BRINK

have progressively more complex spatial structures (Figure I). If the depth
increase is not monotonic, then an infinite number of waves will also exist,
supported by the reverse slopes associated, for example, with trenches
(Mysak et al 1979, Mysak & Willmott 1981, Brink 1983a) or by banks
(Brink 1983b). Increasing the stratification results in an increased fre
quency (hence phase speed) for a given wave mode. Further, making the
rigid lid assumption [neglecting the divergence due to the displacement of
the free surface, i.e. using pz = 0 in lieu of (4b)] leads to an increased phase
speed.
In the barotropic limit (Huthnance 1975), where stratification can be
neglected, all free-wave modes exist for all alongshore wavenumbers but,
except for the barotropic Kelvin wave, their frequencies are always less
than the inertial frequency (i.e. w <f). These subinertial waves in the
barotropic limit are usually called "shelf waves." Generally, long waves
have a group velocity in the same direction as the phase velocity, whereas
short waves propagate energy in the opposite direction, allowing for reflec
tion of wave energy. This behavior of barotropic waves is summarized by
the solid lines in Figure 2.
The effeet of density stratification is quantified in terms of the strati
fication parameter S. As the strength of stratification increases (larger S),
Distance Offshore (km)
300
200
100
Mode 1
1000

2000

Figure I

200

1000

2000

>0

300

Mode 2

Model structures of alongshore


computed for the northern
Oregon coast (after Battisti & Hickey
1984). Units are arbitrary.

velocity,

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

WIND-DRIVEN CURRENTS OVER THE SHELF

Alongshore Wavenumber, 1

Figure 2
oceanS

395

Schematic dispersion curves for an unstratified ocean

(solid lines) and for stratified

(dashed lines).

the wave frequency increases until it reaches the inertial frequency (dashed
lines in Figure 2). Huthnance (1978) has shown that the wave frequency
will r each ! if the maximum value of
A

17 :I,

(11)

evaluated at the bottom, is greater than or equal to unity. Once a dispersion


curve rises above f, strictly trapped waves apparently no longer exist.

Physic ally , this would

occur

because the coastal wave becomes

strongly

coupled, through the topography, to the continuum of freely radiating


oceanic internal waves. Thus, the coastal-trapped waves are apparently

the same sense as are edge waves (e.g. Chapman 1982). In the
S -> 00, the continental margin is much thinner than a typical

"leaky" in
limit as

internal Rossby radius of deformation, and the behavior of the wave


modes is like that of pure internal Kelvin waves

(e.g.

Gill

& Clarke 1974).

Once a dispersion curve reaches f, the wave's group velocity is generally


in the same direction as the phase velocity for all wavenumbers where the
wave exists. This substantially changes the nature of scattering off of

topographic obstacles, since reflection of energy is no longer possible


(Wilkin 1988).
A useful diagnostic for a given wave mode is the ratio (averaged over a
wave period) between kinetic energy K and potential energy P:
K
R=P

(Brink

1982). The potential energy is the sum of

(12)
two components: One is

396

BRINK

associated with the displacement of the free surface (Pr), and the other is
associated with the vertical displacement of water parcels in a stratified
ocean (PJ. In an unstratified ocean, Ps = 0 and Pr is small compared with
K (except for the barotropic Kelvin wave), so
Rl

(13)

for all shelf-wave modes. In fact, in the rigid-lid limit, we have


Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org
by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

D =

R -.

gH (fL)-2 -.00,
00.

(14a)
(14b)

In the opposite limit of Kelvin wave-type behavior (large S), we have


R -.1,
and the importance of gravity as a restoring force allows the equipartition
of wave energy. Once a specific wave mode is calculated, it is straight
forward to estimate f{, Pr, and Ps and thus estimate the importance of
baroclinic effects.
The presence of a mean alongshore flow over the continental margin
can alter the free-wave properties through Doppler shifting, modification
of the background vorticity field, and by allowing the growth of insta
bilities. The case of barotropic, stable, coastal-trapped waves has been
treated by Brooks & Mooers (1977), who were interested in the effect of
the energetic Florida current on coastal-trapped-wave propagation. Luther
& Bane (1985) considered the linearly unstable modes associated with the
Gulf Stream off the United States Carolina coast and found a strong
instability growing on the inshore edge of the mean current. They were
able to identify this mode convincingly with Gulf Stream "shingles," which
have been frequently observed (e.g. Bane et aI1981). When coastal-trapped
waves propagate in the opposite direction to that of the mean current, the
possibility of critical coastal-trapped waves arises (Gill & Schumann 1979).
This is a situation in which the direction of wave propagation is locally
reversed owing to the strength of the mean flow. Gill & Schumann (1979)
first raised this possibility to explain the observed absence of coastal
trapped-wave propagation off the Natal coast, where the Agulhas current
comes particularly close to shore (e.g. Schumann & Brink 1990). It now
seems more likely that the absence of propagation is related to enhanced
frictional damping associated with the distortion of the wave's modal
structure by the mean-flow vorticity field (Brink 1990).
It appears that the main mechanism for the decay of coastal-trapped
wave energy is bottom friction. It is straightforward to include its effects
in the long-wave limit, where the decay of wave modes is accompanied by

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

WIND-DRIVEN CURRENTS OVER THE SHELF

397

an onshore-offshore phase shift associated with the difference in frictional


efficiency versus depth (Brink & Allen 1978, 1983, Clarke & Van Gorder
1986). A perturbation approach to the damping of coastal-trapped waves,
valid even where the modes are not orthogonal but where the decay time
is long relative to the wave period, has been presented by Brink (1990). At
general frequencies and wavenumbers, the frictional problem has most
often been addressed by solving the complex eigenvalue problem associ
ated with the frictional version of (2)-(4) (Spillane 1980, Webster &
Holland 1987). This approach yields very general results at the cost of
substantially increased computational complexity. Power et al ( 1989)
used such an approach to show that, in addition to the onshore-offshore
phase shift, the free-wave modal structure also adjusts to increasing
friction by having the location of maximum fluctuating velocity shift
into progressively deeper water. The effect of this migration is to mitigate
somewhat the damping rate of the wave by putting more of its structure
into deeper water, where damping is relatively less effective.
2.4

The Excitation of Coastal-Trapped Waves

The study of tides over the continental shelf has led to an interesting
application of free coastal-trapped-wave theory. At latitudes poleward of
30, the diurnal tide occurs at a frequency below the inertial frequency, so
that it is possible for the diurnal tide to excite free coastal-trapped waves.
Fitting current and sea-level data to theoretical wave modes has shown,
in several locations, that the diurnal tide over the shelf includes a barotropic
Kelvin wave (which dominates sea-level variability but produces weak
currents) and a lowest mode coastal-trapped (shelf) wave, which is associ
ated with strong current fluctuations (e.g. Figure 3). The shelf wave is
apparently excited by topographic irregularities such as the Strait of Juan
de Fuca (Crawford & Thomson 1984, Flather 1988) or the Gulf of Maine
(Daifuku & Beardsley 1983). In addition, Crawford & Thomson (1984)
showed that including stratification in computing free-wave properties
improves the fit to data considerably and provides rationalization for the
neglect of the short shelf-wave mode (which appears in the barotropic case
but not when stratification is included). Such tidal studies provide some
of the most direct evidence for free coastal-trapped-wave propagation.
Most work on the excitation of coastal-trapped waves, however, has
concentrated on wind-forced motions in the long-wave limit (e.g. Gill
& Schumann 1974). Restriction to this limit is superficially quite reason
able because of the large scales of weather systems and because of the rela
tively short onshore-offshore scales of the continental margin. Given this
assumption, along with the absence of a mean alongshore flow, the free
coastal-trapped-wave modes have the convenient orthogonality property

398

BRINK

em/sec

10

Kelvin
Wave

Stratified
Shelf Wave

+E
Sum

Observed

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

15m r

45m r

81m r

Figure 3 Results of fitting eoastal-trapped-wave modes to diurnal tide data from the shelf
off British Columbia (after Crawford & Thomson 1984). The current contributions from
different modeled waves and their sum are shown, as are observed current ellipses.

(8) which reduces the problem, complete with bottom stress, to an infinite
set of coupled equations

c; I Ynt- Yny+ LanmYm


m

b;r:,

(ISa)

where Yn is the amplitude of wave mode n. The frictional coupling coeffi


cients are defined as
( ISb)
where a subscript B refers to a quantity evaluated at the bottom
[z = -h(x)], and por is a constant of proportionality between bottom
stress and bottom velocity, e.g. r porv. In the limit of weak friction, the
equations given by (15a) become uncoupled at lowest order, and ann can
be thought of as an inverse-decay distance for the frce wave.
The physical mechanism for the wave forcing is straightforward. An
alongshore wind stress causes an onshore-offshore transport within the
surface Ekman layer. In order to conserve mass, a compensating offshore
onshore flow must take place at greater depth. As this flow crosses isobaths,
vortex stretching leads to changes in the local relative vorticity. This
vorticity is expressed largely in terms of the alongshore velocity and,
=

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

WIND-DRIVEN CURRENTS OVER THE SHELF

399

in the presence of alongshore variability (%y "I 0), gives rise to wave
propagation.
A number of practical questions arise when solving coastal-trapped
wave problems (6 and ISa). First, when computing the free-wave modal
structures, there is some question about where to place an artificial vertical
coastal wall. Such an arbitrary boundary is used to exclude the highly
turbulent inner shelf so that the inviscid free-wave equations (6) remain
valid. This problem was addressed by Mitchum & Clarke (l 986a), who
showed that a reasonable choice is to place the "coast" at the isobath
where the water's depth is three times the scale thickness of the turbulent
(Ekman) boundary layer. In addition, they showed that the wind still
excites motions within the nearshore wedge, so that a correction (often
small) to the wave result, proportional to the local alongshore wind stress,
must be added to the wave solution's "coastal" sea level in order to achieve
accuracy. Their results are dependent on a highly idealized formulation of
the bottom stress but nonetheless appear to work well in practice (Mitchum
& Clarke 1986b).
A second question that often arises is, How many modes need to be
summed in order to achieve accurate results? [i.e. what is the range of n in
( IS)?] Early applications of coastal-trapped-wave theory yielded adequate
to very good results with only a single mode (Brink 1982, Battisti & Hickey
1984). Clarke & Van Gorder ( 1986), on the other hand, argued that seven
or more modes typically need to be used in order to obtain accurate results,
especially for alongshore velocity. Chapman (1987) addressed this question
in the context of hindcasting currents off northern California and con
cluded that no improvement in the hindcast occurred when more than
three modes were used. In realistic stratified situations, the accurate com
putation of higher modes tends to involve considerable expense, so this
issue is of some practical importance. Lopez & Clarke (1989) found a
solution to this difficulty by noticing that for higher modes (cn small), (ISa)
reduces to
C;; I Ynt+ LanmYm b;r.

(16)

In other words, alongshore variability (and hence wave propagation)


becomes unimportant. Lopez & Clarke (1989) then proposed the following
approach. First, solve for pressure (which converges faster than velocity)
for the first few wave modes by using (1Sa). This information is then used
to prescribe only Py in system (6), which is then solved for total pressure.
This method avoids the difficult calculation of higher modes and provides
a rational separation between "local" and "remotely forced" fluctuations.
The remotely forced information is that which would be obtained from

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

400

BRINK

solving (15) (i.e. directly, using winds from a large spatial domain), and
the locally forced information is that which could be obtained from (16)
or, equivalently, the residual of the Lopez & Clarke (1989) method.
Coastal-trapped-wave theory has often proved quite successful in prac
tice. For example, Battisti & Hickey ( 1984) and Chapman ( 1987) have
used it to demonstrate that a large portion of the sea-level and alongshore
current variability off the western United States is associated with wind
forced coastal-trapped waves. Typical results are shown in Figure 4. The
wind forcing occurs along the entire coast but is most effective where the
winds are strongest, off northern California. Other successful applications
of the theory have been carried out for the shelf east of Australia (Church
et a1 1986) and the west Florida shelf (Mitchum & Clarke 1986b). It is fair
to say that, in any coastal region where evidence of coastal-trapped-wave
phenomena has been seriously sought, it has been found. There is, however,
a consistent problem. Although the wave theory works quite well for
hindcasting pressure and the alongshore component of current, it has
little or no skill at hindcasting density or onshore-offshore currents, (e.g.
Chapman 1987). Some reasons for this partial failure are discussed in
Section 3. The Gulf of California might prove to be an exception, in that
preliminary results suggest that a substantial part of the temperature
variability there is related to wave propagation (Merrifield & Winant
1989).
Most studies of coastal-trapped-wave generation in nature involve the
use of a dynamical model. This approach imposes a bias in the study of
data, so it is desirable to seek more empirical approaches. One such effort
was made by Chapman et al (1988), who showed that for hindcasting sea
level given wind data over a spatial domain, a dynamical model based on
(I 5) works, within error, as well as an entirely statistical approach for the
region offshore of northern California. This, in effect, says that the first
order wave-equation model works about as well as any possible linear
model, given the inputs.
A more ambitious, purely empirical test of coastal-trapped-wave theory
was undertaken by Davis & Bogden (1989). They took the approach of

trying to explain oceanic variability over the shelf off northern California
by using a large suite of wind and sea-level information as inputs. They
found that a large part of the sea-level variation off California was coherent

and in phase, a result that they tentatively identified as a barotropic Kelvin


wave of unknown origin. The residual sea-level variation behaved largely
as would be expected of coastal-trapped waves. They found that currents
near 39N were influenced by this wave activity and by purely local wind
forcing. The local forcing contributed not only to alongshore currents, but
also to the onshore current component and to temperature variations.

b)

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

30

a)OlO t

C)

!\

20
10

-10
-20
-30
-40

(\,J
(,,)

0 .0

-0.10
0.10

-50

r \ /'Jill Al v. Nli \Il

50
40
3D

it)

-L.

--r-

20
10

-1..Q

::c

-20
-30

t1

C)

-40
0. 00

-0.10

::c

-50
50

N.' zAkNt\ J%
--L.

TIME (days)

40
3D

C)

(,.)

[;j

20

10

::c

-10

-20
-30
-40
-50

TIME (days)

Figure 4 Observed fluctuations (light lines) and modeled fluctuations (heavy lines) for the continental shelf off northern California (with permission
from Chapman 1987). (a) Bottom pressure (in decibars) from the inner shelf (C2) and the midshelf (C3); (b) alongshore velocity (in centimeters
per second) from the inner shelf at 20 m (C2), the midshelf at 55 m (C3), and the outer shelf at 90 m (C4).

S;;
;!5

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

402

BRINK

Overall, the results are consistent'with those that might be expected from
coastal-trapped-wave theory, but they have the added benefit of explaining
more variance than wave-type models have been able to rationalize.
Up to this point, the discussion has focused on the excitation of coastal
trapped waves by large-scale alongshore winds. Other possibilities cer
tainly exist. For example, Webster & Holland (1987) developed a fre
quency-domain analogue to (15) that can be used to solve for wind-forced
motions at any frequency. Gordon & Huthnance (1987) paid particular
attention to the forced behavior near the zero group velocity of the free
barotropic waves and showed how oscillating currents can be generated
in response to a wind impulse. Their results explain observed rotary current
variations in the North Sea and thus qualitatively account for some of the
variability in onshore-offshore flow, a very rare occurrence in terms of
coastal-trapped-wave modeling.
Other means of forcing coastal-trapped waves do not involve the wind
stress directly. For example, Kroll & Niiler (1976), and subsequently
others, have explored the possibility that fluctuations in the midlatitude
deep ocean drive shelf waves. No convincing observational evidence for
sueh excitation has yet been found. The situation at very low latitudes,
however, is different. The equator acts as a wave guide, which is interrupted
at the coasts. At the eastern edge of the ocean basin, incoming equatorial
wave energy can either be reflected or transmitted poleward along the
coast in the form of coastal-trapped waves (Moore 1968, Clarke 1983).
Enfield et al (1987) used observations to make a convincing case that
equatorial Yanai (mixed Rossby-gravity) waves excite the observed (Smith
1978, Cornejo-Rodriguez & Enfield 1987) free coastal-trapped-wave
activity off Peru.
Straits or estuaries represent another interesting interruption of the
coastal wave guide. A good example is provided by the effect of Hudson
Bay and the Hudson Strait on flow over the Labrador shelf. Wright et al
(1987) demonstrated that changes in air pressure over the bay lead to
fluxes out of the strait and onto the Labrador shelf. At the shelf-strait
junction, the coastal flux becomes analogous to a wind stress (Middleton
1988) in forcing waves over the shelf. Another interesting example is that
of Bass Strait, between mainland Australia and Tasmania. It appears that
along-strait winds generate waves within the strait, which in turn excite
coastal-trapped waves that propagate northward along the eastern coast
of Australia (Clarke 1987, Buchwald & Kachoyan 1987). This explanation
for the origin of the northward-propagating waves leaves one interesting
question, peculiar to this problem, unanswered: Where does the eastward
propagating energy along the Australian south coast go? Church & Free-

WIND-DRIVEN CURRENTS OVER THE SHELF

403

land ( 1987) made a convincing case that it does not pass around Tasmania,
and Clarke ( 1987) showed that it cannot pass through Bass Strait. It seems
possible that the waves are scattered into the open ocean as they pass the
sharp and topographically complex southern tip of Tasmania.

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

2.5

Summary

Coastal-trapped-wave theory provides a powerful conceptual mechanism


for understanding large-scale flow over the continental shelf. The under
lying idea of wind-driven variability propagating in the direction of the
coastal-trapped-wave phase velocity is conceptually simple and has been
verified repeatedly by observations. The theory works best for pressure
(sea level) and alongshore velocity, and only rarely shows any useful skill
for density or onshore velocity. This lack of skill is perhaps not surprising
in light of the tendency for observed onshorc currents to violate the long
wave assumption by their magnitude and by having short spatial scales
relative to the alongshore velocity (e.g. Kundu & Allen 1976).
3. STOCHASTIC MODELS OF WIND-DRIVEN
CURRENTS

The models discussed in the previous section all have the common thread
of determinism. That is to say, they could be used, given time series of
inputs, to produce time series of outputs that can then be compared
with observations. Another approach-the stochastic one-is to use the
statistics of input functions to hindcast the statistics of the observed varia
bility. This approach is, in some ways, more compact and offers some
physical insights that might not otherwise be so obvious.
3.1

Wave-Based Models

Allen & Denbo ( 1984) took advantage of the fact that the long-wave wind
forced problem could be simplified to ( 15) in order to investigate the
statistics of the wind-forced responsc. They further assumed that only a
single wave mode is needed. While this assumption may now seem a bit
naive, it appeared quite reasonable given information available at that
time (e.g. Battisti & Hickcy 1984). For the case of a single mode, or where
there are multiple, uncoupled modes (i.e. anm 0 if n "# m), the solution
of ( 15) is
=

YnCy,t)

cnbn

fO

exp(-cnannl1)'o[(y+cn1]),(t-1])]dl1

( 17)

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

404

BRINK

for an infinite extent of coast. Once this solution is known, it becomes


straightforward to express lagged covariances between Yn at different
locations, and between Yn and the wind stress. All of these covariances
depend only upon the modal coefficients (bm em and ann) and upon the
lagged autocovariance function of the wind-stress forcing, which is pre
sumed known.
Typical results of the Allen & Denbo (1984) calculations are shown in
Figure 5. The autocorrelation function of wind stress (R<T; not shown) is
used as input to the model. It is assumed to have its major and minor axes
along the axes of the plot, signifying that there is no preferred tendency
for wind fluctuations to propagate alongshore. The upper panel of Figure
1200
600
-...

Rpp

E:

0
-600
-1200
-4

-2

-2

1200

-...

Rrp

600
0
-600
-1200
-4

Figure 5

Theoretical correlation diagrams using single-mode theory (with permission from

1988). (Top) Correlation of sea level with sea level at other alongshore (YL)
(IL) lags. (Bottom) Same for correlation of sea level at (0, 0) with wind stress at

Chapman et al

and time

f (days)

other space or time lags.

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

WIND-DRIVEN CURRENTS OVER THE SHELF

405

5 shows the predicted autocorrelation function of sea level Rppo (For a


single mode, the function for alongshore velocity is identical.) The tilt of
the ellipses indicates alongshore propagation of the signal in the sense of
free-wave propagation. The slope of the major axis of the elliptical con
tours represents a value intermediate between the phase speed of the
forcing (infinite in this case) and of the free coastal-trapped wave. In
contrast, the correlation between wind and pressure RTp (bottom panel) is
not so symmetric. The maximum wind-pressure correlation occurs at a
location "upstream" with regard to wave propagation. The location of the
maximum correlation lies on a straight line passing through the origin and
having a slope exactly equal to the free-wave propagation speed. The Allen
& Denbo (1984) results are very powerful in that they can readily be
compared with observed statistics. Perhaps most interesting from an obser
vational standpoint is the finding that pressure (or alongshore velocity)
should be best correlated with nonlocal winds, an effect that is indeed
observed in nature (e.g. Halliwell & Allen 1984, Winant et al 1987).
The Allen & Denbo ( 1984) analysis has been extended to the case with
mUltiple modes by Chapman et al (1988). Two substantial changes occur
with this generalization. First, the correlation diagrams become functions
of location and of variable. This follows because
n

v =

j- 1 L YnFn.(X,

z ,

and because the modal functions Fn are different for each n. Also, the tilts
of the major axes of the R,p correlation diagrams vary so as to represent
a weighted (by modal structure) average of the phase speeds of the free
modes. Chapman et al ( 1988) used these results to show that three-mode
correlation diagrams yielded appreciably better theoretical results than
single-mode diagrams when compared with observations made off north
ern California.

3.2

More Complete Models

The Allen & Denbo (1984) stochastic approach, used with many modes,
in principle yields correct results for a broad range of problems where the
long-wave limit applies and forcing consists of only an alongshore wind
stress independent of cross-shelf distance. This approach, while very
enlightening, is thus somewhat limited in its applicability. A more general
approach to the linear problem is made by relaxing these assumptions.
In this case, the approach used is to Fourier transform the equations of

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

406

BRINK

motion (1) with respect to both time and alongshore direction. This leaves
a two-dimensional boundary-value problem to be solved for each fre
quency and alongshore wavenumber (Spillane 1980, Brink et al 1987),
yielding a frequency-wavenumber transfer function. The wind forcing
enters both at the coast and at the free surface, while bottom friction
provides a damping mechanism. The forcing of the model is prescribed
through the frequency-wavenumber spectrum and the cross-shelf structure
of the wind stress (Figure 6). One obtains statistics relating currents to
winds as a function of (x, z) location and frequency by integrating results
over alongshore wavenumber, using the wind-stress spectrum as a weight
ing function.
Some representative transfer functions at a fixed frequency are shown
in Figure 7. Pronounced spikes, associated with rapid phase shifts (not
shown), appear in the transfer functions for v and p. These represent the
contributions from coastal-trapped-wave resonances. Upon integration,
these spikes tend to dominate the amplitude of the response, especially
since the wind spectrum is "red" (that is, it has greatest amplitude at long
wavelengths). The phase shifts associated with these peaks, however, lead
to relatively small co-spectra, or relatively low coherence between wind

and v or p. This is the mathematical expression of the fact that propagating


energy arrives from locations sufficiently distant that remote, incoherent
forcing plays a substantial role.
The transfer functions for onshore current and density are not so domi
nated by resonance spikes. Indeed, the u function increases for short scales
and the p function is flat for short scales. This is in distinct contrast to
05 +--'----'--t---'---,---+

00 -l---r--,-----,--+--.--1
22
-1.1
0.0
-2.2
11

Figure 6

Wavenumber fey k";i'x f03)

Frequency-alongshore wavenumber spectrum of alongshore wind stress for


the West Coast of the United. States during summer 1982. Units are the base 10 log of
(dynes cm-2) (cycles km-1)-1 (cycles day-l)-l (with permission from Brink et aI1987).

WIND-DRIVEN CURRENTS OVER THE SHELF

407

25

20

15
',"" .

15

..

',.

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

....

,<.

<0

10

10

liil

-10

..

"
<!.
,-'"
.......

:;::.
',""
..

010

"-

.
:' 005
,

-5

JR(cm-' x fO}

10

liil
------------10

-5

0
5
J.R (cm'x fO)

10

Figure 7 Amplitudes of wind-stress transfer functions of midshelf oceanic variables for


w =

1.08

10- 5 s- I for northern California (with permission from Brink et al 1987).

the v or p functions, which decline for short scales (large alongshore


wavenumber I). This tendency for increased contribution to u at large /
only occurs in the realistic case of the wind stress varying with distance
offshore. Physically, the trends in the transfer functions imply that u and
p are most sensitive to very short alongshore scales in the wind stress
scales so short as to be unresolved by existing estimates of the wind
spectrum. Nonetheless, aircraft-based measurements of the wind field indi
cate that there is often substantial energy in scales as short as 10 km (e.g.
Stuart 198 1, Winant et aI1988).
The results of Brink et al (1987) show fairly good agreement with
observations for alongshore velocity and pressure. This is consistent with
the quality of the wind input, which resolved large scales well but small
scales poorly. Also, as with many forced-wave calculations (Figure 4b;
e.g. Chapman 1987), the amplitude of the alongshore velocity was also
somewhat underpredicted. The model performed poorly for u and p,
presumably at least partly because the wind spectrum did not resolve the
shortest scales. It is difficult to tell at this time to what extent the linear
model physics is also to blame for this inadequacy. The advantages with
stochastic models are that they relax some of the restrictive approximations
associated with most forced-wave models, and that they allow different
physical insights than those of the wave models. The two approaches are
complementary.

408

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

4.

BRINK

CRITIQUE

The above sections really address only a limited physical regime: fluc
tuating motions driven by fluctuating winds in a linearized, nearly inviscid
context. While this type of model is successful in some regards, it leaves
considerable room for improvement. It thus seems reasonable to consider
briefly a few topics that could not be discussed fully here.
In the ocean, turbulent surface and bottom boundary layers almost
always exist. Since these are typically of 0(10 m) thick, they occupy a
substantial part of the shelf water column. Although some of the effects
of these boundary layers are accounted for above, their considerable cur
rent and thermal perturbations are not considered. Turbulent boundary
layers over the shelf have dynamics that are at least two-dimensional
and hence substantially more complex than in simpler settings. The most
obvious effect occurs near the surface, where horizontal temperature gra
dients associated with upwelling can lead to the generation of variable
boundary-layer thickness and of near-surface fronts (de Szoeke & Richman
1984, Rudnick & Davis 1988). Likewise, flow in a bottom boundary layer
beneath a stratified interior and above a sloping bottom leads to buoyancy
induced anomalies in the boundary-layer structure (Weatherly & Martin
1978). In either case, it is clear that turbulent boundary-layer processes
over the shelf and slope will be complex and are important components
of a complete coastal-circulation model.
Fronts are a common phenomenon in the coastal ocean. Once a front
is formed its advection by wind-driven currents can complicate the thermal
response of the ocean (e.g. Ou 1984). Further, the frontal distortion of the
ambient vorticity distribution can modify coastal-trapped waves and allow
for the existence of new, frontal-trapped modes (e.g. Luther & Bane 1985).
Finally, various types of coastal fronts appear to be hydrodynamically
unstable (e.g. Flagg & Beardsley 1978, Narimousa & Maxworthy 1985,
Barth 1989a,b, Gawarkiewicz 1989). Instabilities lead to current fluc
tuations that are not directly related to the wind, and that may prove to
be a substantial mechanism for the exchange of water and materials across
the continental margin.
Addressing the problem of steady, wind-driven currents over the con
tinental margin at first seems to be as simple as taking the steady limit of
(1). This is not true, since the resulting suppression of the vertical velocity
(Ie) will then lead to singularities. Thus, it is necessary to add new physics
to any steady, stratified model. One approach has been to add horizontal
and vertical mixing, a procedure that leads to results that can be quali
tatively thought of as the steady limit of coastal-trapped-wave generation
(McCreary & Chao 1985). Although wind-driven variability, whether
remote or local, seems to dominate alongshore current fluctuations over

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

WIND-DRIVEN CURRENTS OVER THE SHELF

409

the shelf, it is not clear that steady winds play a similarly dominant
role. Various influences of the deep ocean on the shelf cannot be entirely
discounted (Huthnance 1984, McCreary et al 1986, Kelly & Chapman
1988), and estuarine outflows can also drive alongshore currents (e.g.
Beardsley & Hart 1978) and coastal-trapped waves (Whitehead & Chap
man 1986). Further, nonlinear processes can lead to the generation of a
mean flow driven by fluctuating currents (Loder 1980, Haidvogel & Brink
1986, Samelson & Allen 1987). It is, in fact, difficult to use observations
to test models of mean currents. This is so, at least partly, because the
absence of fluctuations (by definition of "mean") implies only one real
ization and hence too few degrees of freedom to make any meaningful
statistical test.

5.

CONCLUSIONS

It would certainly be fair to conclude that we have developed a sophis


ticated understanding (and predictive capability) for variations in sea
level and alongshore current over the continental shelf for typical periods
ranging from the inertial period out to about 20 days. Despite the impres
sive results of existing models (e.g. Figure 4), there are still some trouble
some problems. First, existing models tend to underpredict the amplitude
of observed fluctuations of alongshore current and sea level, typically by
about 10-50%. Second, such models tend to underestimate grossly the
observed strength of density and cross-shelf current changes. The latter
difficulty appears to be at least partly associated with the long-wave
assumption, which suppresses the small spatial scales that typify actual
cross-shelf currents. Any of these problems could also be related to frontal,
nonlinear, or turbulent processes that are not included explicitly in most
existing wind-driven models.
Our present understanding of currents over the continental shelf, while
satisfying in some regards, is clearly inadequate. This is so because many
problems of interest, such as oceanic frontogenesis, coastal upwelling, and
the dispersal of pollutants, involve the exchange of properties across the
continental margin. Understanding transport across the continental
margin requires an understanding of onshore-offshore currents. Such an
understanding is presently lacking and is likely to be the focus of con
siderable research in the coming years.
ACKNOWLEDGMENTS

David Chapman and Steven Lentz provided valuable comments. Support


for this effort was provided by the Office of Naval Research (ONR)
through the Coastal Sciences Program, code 1122CS.

410

BRINK

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

Literature Cited
Allen, J. S. 1980.Models of wind-driven cur
rents on the continental shelf. Annu. Rev.
Fluid Mech.12: 389-433
Allen, J. S., Denbo, D. W. 1984. Statistical
characteristics of the large-scale response
of coastal sea level to atmospheric forcing.
J. Phys. Oceanogr.14: 1079-94
Bane, J. M. Jr., Brooks, D. A., Lorenson,
K. R. 1981. Synoptic observations of the
three-dimensional structure and propa
gation of Gulf Stream meanders along the
continental margin. J. Geophys. Res. 86:
6411-25
Barth, J. I 989a. Stability of a coastal upwell
ing front, I, Model development and sta
bility theorem. J. Geophys.Res.94: 10,84456
Barth, J. 1989b. Stability of a coastal upwell
ing front, 2, Model results and comparison
with observations. J. Geophys. Res. 94;
10,857-83
Battisti, D. S., Hickey, B. M. 1984. Appli
cation of remote wind-forced coastal
trapped wave theory to the Oregon and
Washington coasts. 1. Phys.Oceanogr.14:
887-903
Beardsley, R. C., Hart, J. 1978. A simple
theoretical model for the flow of an estu
ary onto a continental shelf. J. Geophys.
Res.83: 873-83
Brink, K. H. 1982. A comparison of long
coastal trapped wave theory with obser
vations off Peru. J. Phys. Oceanogr. 12:
897-913
Brink, K. H. 1983a. Some effects of strati
fication on long trench waves. J. Phys.
Oceanogr.13:496-500
Brink, K. H. 1983b. Low-frequency free
wave and wind-driven motions over a sub
marine bank. J. Phys. Oceanogr.13: 10316
Brink, K. H . 1989. Energy conservation in
coastal-trapped wave calculations. J.
Phys. Oceanogr.19: 1011-16
Brink, K. H. 1990. On the damping of free
coastal-trapped waves. J. Phys.Oceanogr.
20: 1219-25
Brink, K. H., Allen, 1. S. 1978. On the effect
of bottom friction on barotropic motion
over the continental shelf. J. Phys.
Oceanogr.8:919-22
Brink, K. H., Allen, 1. S. 1983. Reply (to
T. J. Simons' comments on K. H. Brink
and J. S. Allen, J. Phys. Oceanogr. 8:
919-22,1978).J. Phys. Oceanogr.13: 14950
Brink, K. H., Chapman, D. C. 1987. Pro
grams for computing properties of
coastal-trapped waves and wind-driven
motions over the continental shelf and
slope (second edition). Rep. No. WHO[-

87-24,

Woods Hole Oceanogr. Inst.,


Woods Hole, Mass. 119 pp.
Brink, K. H., Chapman, D. c.. Halliwell. G.
R. Jr. 1987.A stochastic model for wind
driven currents over the continental shelf.
J. Geophys.Res. 92: 1783-97
Brooks, D. A., Mooers, C. N. K. 1977.Free,
stable continental shelf waves in a sheared,
barotropic boundary current. J. Phys.
Oceanogr.7: 380-88
Buchwald, V. T., Kachoyan, B. J. 1987. Shelf
waves generated by a coastal flux. Aust.J.
Mar. Freshwater Res.38:429-37
Chapman, D. C. 1982. Nearly trapped inter
nal edge waves in a geophysical ocean.
Deep-Sea Res.29(4A): 525--3 3
Chapman, D. C . 1987. Application of wind
forced, long, coastal-trapped wave theory
along the California coast. J. Geophys.
Res. 92:1798-1816
Chapman, D. c., Lentz, S. J., Brink, K. H .
1988. A comparison o f empirical and
dynamical hindcasts of low-frequency,
wind-driven motions over a continental
shelf. 1. Geophys.Res.93: 12,409-22
Church, J. A., Freeland, H. 1. 1987. The
energy source for the coastal-trapped
waves in the Australian Coastal Experi
ment region. J. Phys. Oceanogr. 17: 289300
Church, J. A., White, N. J., Clarke, A. J.,
Freeland, H . J., Smith, R . L. 1986. Coast
al-trapped waves on the east Australian
continental shelf. Part II: Model veri
fication. J. Phys. Oceanogr. 16: 1945-57
Clarke, A. J. 1977. Observational and
numerical evidence for wind-forced coast
al trapped long waves. J. Phys. Oceanogr.
7: 231-47
Clarke, A. J. 1983. The reflection of equa
torial waves from oceanic boundaries. J.
Phys. Oceanogr.13: 1193-1207
Clarke, A. J. 1987. Origin of coastally
trapped waves observed during the Aus

tralian Coastal Experiment. J.

Phys.

Oceanogr. 17:1847-59
Clarke, A. J., Brink, K. H. 1985. The
response of stratified, frictional flow of
shelf and slope waters to fluctuating large
scale low-frequency wind forcing. J. Phys.
Oceanogr.15:439-53
Clarke, A. J., Van Gorder, S. 1986. A
method for estimating wind-driven fric
tional, time-dependent, stratified shelf and
slope water flow. J. Phys. Oceanogr. 16:
1013-28
Cornejo-Rodriguez, M., Enfield, D. B. 1987.
Propagation and forcing of high-fre
quency sea level variability along the west
coast of South America. J. Geophys. Res.
92:14,323-34

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

WIND-DRIVEN CURRENTS OVER THE SHELF

Crawford, W. R., Thomson, R. E. 1 984.


Diurnal-period continental shelf waves
along Vancouver Island: a comparison of
observations with theoretical models. J.
Phys. Oceanogr. 14: 1 629-46
Daifuku, P. R., Beardsley, R. C. 1983. The
K I tide on the continental shelf from
Nova Scotia to Cape Hatteras. J. Phys.
Oceanogr. 1 3 : 3-1 7
Davis, R . E., Bogden, P . S . 1989. Variability
on the California shelf forced by local and
remote winds during the Coastal Ocean
Dynamics Experiment. J. Geophys. Res.
94: 4763-83
de Szoeke, R. A., Richman, J. G. 1 984. On
wind-driven mixed layers with strong hori
zontal gradients-a theory with appli
cation to coastal upwelling. J. Phys.
Oceanogr. 1 4: 364-77
Enfield, D. B., Cornego-Rodriguez, M . ,
Smith, R. L., Newberger, P. A. 1 9 8 7 . The
equatorial source of propagating varia
bility along the Peru coast during the
1982-1983 EI N ino. J. Geophys. Res. 92:
14,335-46
Flagg, C. N., Beardsley, R. C. 1 978. On the
stability of the winter shelf water/slope
water front south of New England. J. Geo
phys. Res. 83: 4623-3 1
Flather, R. A. 1 988. A numerical model
investigation of tides and diurnal-period
continental shelf waves along Vancouver
Island. J. Phys. Oceanogr. 1 8 : 1 1 5-39
Gawarkiewicz, G. 1989. Linear stability
models of shelfbreak fronts. PhD thesis.
Univ. Dei., Newark. 1 66 pp.
Gill, A. E., Clarke, A. J. 1 974. Wind-induced
upwelling, coastal currents and sea level
changes. Deep-Sea Res. 2 1 : 325-45
Gill, A. E., Schumann, E. H. 1 974. The gen
eration of long shelf waves by the wind. J.
Phys. Oceanogr. 4: 83-90
Gill, A. E., Schumann, E. H. 1 979. Topo
graphically induced changes in the struc
ture of an inertial coastal jet: application
to the Agulhas Current. J. Phys. Ocean
ogr. 9: 975-9 1
Gordon, R. L., Huthnance, J. M. 1 987.
Storm-driven continental shelf waves over
the Scottish continental shelf. Cont. Shelf
Res. 7: l O I S -48
Haidvogel, D. B., Brink, K. H. 1986. Mean
currents driven by topographic drag over
the continental shelf and slope. J. Phys.
Oceanogr. 16: 2 1 59-71
Halliwell, G. R., Allen, J. S. 1984. Large
scale sea level response to atmospheric
forcing along the west coast of North
America, summer 1973. J. Phys. Ocean
ogr. 14: 864-86
Huthnance, J. M. 1 975. On trapped waves
over a continental shelf. J. Fluid Mech. 69:
689-704

411

Huthnance, J. M. 1 978. On coastal trapped


waves: analysis and numerical calculation
by inverse iteration. J. Phys. Oceanogr. 8:
74-92
Huthnance, J. M. 1 984. Slope currents and
"JEBAR." J. Phys. Oceanogr. 14: 795-8 1 0
Huthnance, J . M . , Mysak, L . A . , Wang,
D.-P. 1 986. Coastal trapped waves. In

Baroclinic Processes on Continental


Shelves. Coast. Estuar. Sci. 3, ed. C. N.

K. Mooers, pp. 1 - 1 8 . Washington, DC:


Am. Geophys. Union
Kelly, K. A., Chapman, D. C. 1 988. The
response of stratified shelf and slope
waters to steady offshore forcing. J. Phys.
Oceanogr. 1 8 : 906-25
Kroll, J., Niiler, P. P. 1 976. The transmission
and decay of barotropic topographic
Rossby waves incident on a continental
shelf. J. Phys. Oceanogr. 6: 432-50
Kundu, P. K., Allen, J. S. 1976. Some three
dimensional characteristics of low-fre
quency current fluctuations near the
Oregon coast. J. Phys. Oceanogr. 6: 1 8 199
Loder, J. 1 980. Topographic rectification of
tidal currents on the side of Georges Bank.
J. Phys. Oceanogr. \ 0: 1 399- 1 4 1 6
Lopez, M . , Clarke, A . J. 1 989. The wind
driven shelf and slope water flow in terms
of a local and a remote response. J. Phys.
Oceanogr. 19: 1 09 1 - 1 1 0 1
Luther, M . E., Bane, J. M . Jr. 1 985. Mixed
instabilities in the Gulf Stream over the
continental slope. J. Phys. Oceanogr. 1 5:
3-23
McCreary, J. P. Jr., Chao, S.-Y. 1 985. Three
dimensional shelf circulation along an
eastern ocean boundary. J. Mar. Res. 43:
1 3-36
McCreary, J. P. Jr., Shetye, S. R., Kundu, P.
K. 1986. Thermohaline forcing of eastern
boundary currents: with application to the
circulation off the west coast of Australia.
J. Mar. Res. 44: 7 1-92
Merrifield, M . A., Winant, C. D. 1989. Shelf
circulation in the Gulf of California: a
description of the variability. J. Geophys.
Res. 94: 1 8, 1 33-60
Middleton, J. F. 1988. Long shc1fwaves gen
erated by a coastal flux. J. Geophys. Res.
93: 1 0,724-30
Mitchum, G. T., Clarke, A. J. 1 986a. The
frictional nearshore response to forcing by
synoptic scale winds. J. Phys. Oceanogr.
1 6: 934-46
Mitchum, G. T., Clarke, A. J. 1 986b. Evalu
ation of frictional, wind-forced long-wave
theory on the west Florida shelf. J. Phys.
Oceanogr. 1 6: 1 029-37
Moore, D. W. 1968. Planetary-gravity waves
in an equatorial ocean. PhD thesis. Har
vard Univ., Cambridge, Mass.

412

BRINK

Mysak, L. A., LeBlond, P. H . , Emery, W. 1.


1979. Trench waves. J. Phys. Oceanogr. 9:
100 1-1 3

Mysak, L. A., Willmott, A. J. 1981 . Forced


trench waves. J. Phys. Oceanogr. 1 1 : 148 1 1 502

Narimousa, S., Maxworthy, T. 1985. Two


layer model of shear-driven coastal up
welling in the presence of bottom topog
raphy. J. Fluid Mech. 159: 503-31
Ou, H. W. 1 984. Wind-driven motion near a
shelf-slope front. J. Phys. Oceanogr. 1 4 :

Annu. Rev. Fluid Mech. 1991.23:389-412. Downloaded from arjournals.annualreviews.org


by Pontificia Universidad Catolica de Valparaiso on 08/09/10. For personal use only.

985-93

Power, S. B., Middleton, J. H . , Grimshaw,


R. H. J. 1 989. Frictionally modified con
tinental shelf waves and the subinertial
response to wind and deep-ocean forcing.
J. Phys. Oceanogr. 1 9: 1486-1 506
Rudnick, D. L., Davis, R. E. 1988. Fronto
genesis in mixed layers. J. Phys. Oceanogr.
18:434-57

Samelson, R. M., Allen, J. S. 1987. Quasi


geostrophic topographically generated
mean flow over the continental margin. J.

. Schumann, E. H., Brink, K. H. 1990. Coast


al-trapped waves off the coast of South
Africa: generation. propagation and cur
rent structures. J. Phys. Oceanogr. In press
Smith, R. L. 1 978. Poleward propagating
perturbations in currents and sea level
along the Peru coast. J. Geophys. Res. 83:
Phys. Oceanogr. 1 7: 2043-64

6083-92

Spillane, M. C. 1980. A model of wind-forced


viscous circulation near coastal boundaries.

PhD thesis. Oreg. State Univ., Corvallis.


88 pp.
Stuart, D. W. 198 1 . Sea-surface tem
peratures and winds during Joint II. Part
II. :remporal fluctuations. In Coastal

Upwelling, ed. F. A. Richards, pp. 32-38.


Washington, DC: Am. Geophys. Union
Wang, D.-P., Mooers, C. N. K . 1976. Coast
al trapped waves in a continuously strati
fied ocean. J. Phys. Oceanogr. 6: 853-63
Weatherly, G. L., Martin, P. J. 1978. On
the structure and dynamics of the oceanic
bottom boundary layer. J. Phys. Ocean
ogr. 8: 557-70

Webster, 1., Holland, D. 1987. A numerical


method for solving the forced baroclinic
coastal-trapped wave problem of general
form. J. A lmos. Ocean. Technol. 4: 22026

Whitehead, J. A., Chapman, D. C. 1986.


Laboratory observations of a gravity cur
rent on a sloping bottom: the generation
of shelf waves. J. Fluid Mech. 172: 373-99
Wilkin, J. L. 1 988. Scattering of coastal
trapped waves by irregularities in coastline
and topography. PhD thesis. Mass. Inst.

Technol.jWoods Hole Oceanogr. Inst.,


Cambridge/Woods Hole, Mass. (available
as Rep. No. WHO/-88-47, Woods Hole
Oceanogr. Inst., Woods Hole, Mass.)
Winant, C. D., Beardsley, R. C., Davis, R .
E. 1987. Moored wind, temperature and
current observations made during CODE1 and CODE-2 over the northern Cali
fornia continental shelf and upper slope.
J. Geophys. Res. 92: 1 569-1604
Winant, C. D., Dorman, C. E., Friehe, C. A.,
Beardsley, R . C. 1 988. The marine layer off
northern California: an example of super
critical channel flow. 1. A lmos. Sci. 45:
Wright, D. G., Greenberg, D. A., Majaess,
F. G. 1 987. The influence of bays on
adjusted sea level over adjacent shelves
with application to the Labrador Shelf. J.
3588-3605

Geophys. Res. 92: 14,610-20

You might also like