influence of wind stress and ambient flow on a high...

49
1 Influence of Wind Stress and Ambient Flow on a High Discharge River Plume I. Garc a Berdeal, B.M. Hickey and M. Kawase School of Oceanography Box 357940, University of Washington, Seattle, WA 98195-7940 Abstract The response of a high discharge river plume to an alongshore ambient flow and wind forcing is studied with a three-dimensional numerical model. The study extends prior model studies of plumes by including (1) a very large volume discharge (14,000 m 3 s -1 , about twice the maximum used in other models), (2) ambient flow in a direction opposite to that of the propagation of coastally trapped waves and (3) a sequence of wind direction reversals. The magnitude of the ambient flow, wind stress, estuary width and river outflow are based on typical values for the Columbia River on the Washington coast. The model results challenge two longstanding notions about the Columbia plume: first, that the plume orientation is in a relatively stable southwest position in summer – with average discharge conditions (7000 m 3 s -1 ) a summertime downwelling event can erode the southwestward plume and advect it to the north of the river mouth over several days. Second, the plume is not always uni-directional; branches can occur both upstream and downstream of the river mouth simultaneously. The model also provides an explanation for the observation that the plume rarely tends southward during the winter season – in contrast to summer conditions, the rotational tendency of the plume and the ambient flow are in the same direction, so that wind stress must be be significant (> 1.4 dynes cm -2 for at least 2 days) to reverse the plume direction. Distinct anticyclonic freshwater pools form in modeled plumes both north and south of the river mouth under steady forcing conditions when ambient flow is present. The scale of modeled pools is consistent with features observed in the Columbia plume. 1. Introduction The plume from the Columbia River is a dominant feature in the hydrography of the U.S. West Coast. The Columbia River discharge varies between 3000 and 17,000 m 3 s -1 over a typical year (Hickey et al., 1998) and accounts for 77% of the coastal drainage on the U.S. West Coast. The largest outflows occur during spring due to snowmelt (May-June) and during winter storms

Upload: others

Post on 28-Jul-2020

11 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

1

Influence of Wind Stress and Ambient Flow on a High Discharge River Plume

I. Garc�a Berdeal, B.M. Hickey and M. Kawase

School of Oceanography Box 357940, University of Washington, Seattle, WA 98195-7940

Abstract

The response of a high discharge river plume to an alongshore ambient flow and wind

forcing is studied with a three-dimensional numerical model. The study extends prior model

studies of plumes by including (1) a very large volume discharge (14,000 m3 s-1, about twice the

maximum used in other models), (2) ambient flow in a direction opposite to that of the

propagation of coastally trapped waves and (3) a sequence of wind direction reversals. The

magnitude of the ambient flow, wind stress, estuary width and river outflow are based on typical

values for the Columbia River on the Washington coast. The model results challenge two

longstanding notions about the Columbia plume: first, that the plume orientation is in a relatively

stable southwest position in summer – with average discharge conditions (7000 m3 s-1) a

summertime downwelling event can erode the southwestward plume and advect it to the north of

the river mouth over several days. Second, the plume is not always uni-directional; branches can

occur both upstream and downstream of the river mouth simultaneously. The model also

provides an explanation for the observation that the plume rarely tends southward during the

winter season – in contrast to summer conditions, the rotational tendency of the plume and the

ambient flow are in the same direction, so that wind stress must be be significant (> 1.4 dynes

cm-2 for at least 2 days) to reverse the plume direction. Distinct anticyclonic freshwater pools

form in modeled plumes both north and south of the river mouth under steady forcing conditions

when ambient flow is present. The scale of modeled pools is consistent with features observed in

the Columbia plume.

1. Introduction

The plume from the Columbia River is a dominant feature in the hydrography of the U.S.

West Coast. The Columbia River discharge varies between 3000 and 17,000 m3 s-1 over a typical

year (Hickey et al., 1998) and accounts for 77% of the coastal drainage on the U.S. West Coast.

The largest outflows occur during spring due to snowmelt (May-June) and during winter storms

Page 2: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

2

due to rainfall. The plume can extend as far north as the Strait of Juan de Fuca (47°N) in winter

and as far south as 40°N in summer (Fig. 1). The seaward extent of the plume can be as large as

400 km (Barnes et al., 1972). The Columbia River plume therefore affects a vast area of the

coastal ocean, including the coastal estuaries in the Pacific Northwest (Hickey et al., 1999), not

only by reducing salinity, but also by changing the distribution of other water properties such as

nutrients. The plume plays an important role in the transport of dissolved and particulate matter,

phyto- and zooplankton, larvae, contaminants, etc. (Barnes et al., 1972; Grimes and Kingsford,

1996).

Winds and ambient coastal flow both play important roles in determining the

characteristics of buoyant plumes. The seasonal cycle of wind in this region is determined by the

alternation of atmospheric pressure systems over the North Pacific. During winter the Aleutian

Low results in northward winds along the coast; in summer, the North Pacific High results in

southward seasonal mean winds (Barnes et al., 1972; Hickey, 1989). Superimposed on this

seasonal mean are fluctuations with time scales of 2-10 days (Hickey, 1989). In the winter, these

episodes have a predominantly northward alongshore wind stress (“downwelling-favorable”)

with a magnitude typically ranging between 0.5 dyne cm-2 and 3 dynes cm-2, reaching 4-5 dynes

cm-2 during severe storms. Episodes of southward alongshore wind stress (“upwelling-

favorable”) during winter rarely exceed 1 dyne cm-2 and are usually about 0.5 dyne cm-2 (Hickey

et al., 1998). The latter are associated with good weather in the region. In contrast, in summer,

northward and southward wind stress events are comparable in magnitude (~ 0.5 dyne cm-2).

Mean coastal currents off Washington and Oregon also exhibit a strong seasonal cycle

(Hickey, 1989). In fall and winter the monthly mean flow is northward; during spring and

summer the surface (upper 50 m) coastal waters flow southward, although deeper layers may

flow northward. The seasonal mean currents have typical amplitudes of 5-20 cm s-1. Fluctuations

in the coastal currents are strongly wind-driven in all seasons, with forcing being more local in

winter, more remote in summer (Hickey, 1989). Typical fluctuations are ~10-50 cm s-1.

Both in situ hydrographic data (Hickey et al., 1998; Barnes et al., 1972) and satellite-

derived data (Fiedler and Laurs, 1990) show that the average position of the Columbia River

plume also varies with season. During the winter when the prevailing winds and coastal currents

are northward the plume from the Columbia River is usually observed north of the river mouth.

During periods of strong northward winds the plume “hugs” the coast. Under lighter northward

Page 3: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

3

or southward wind events the plume “relaxes” to a west to northwest position (mean winter

position) (Hickey et al., 1998) as seen in the sea surface temperature (SST) satellite image shown

in Figure 1a. When the weather systems change after the transition to spring conditions and the

prevailing winds and coastal currents turn southward, the plume generally adopts a

southwestward orientation (mean summer position) as seen in the coastal zone color scanner

(CZCS) image shown in Figure 1b.

In summer, episodes of inclement weather (when the winds turn northward) result in a

plume with a northward "winter" orientation. Fiedler and Laurs (1990) describe an event with

northward wind stress of 0.5 dynes cm-2 for about 5 days in July 1979. Following this episode

CZCS data show the plume hugging the coast north of the river mouth as in winter. These

summer plume reversals, although not often seen in satellite images due to the cloud cover that

invariably accompanies bad weather, may be relatively common. For example, in the summer of

1998 several reversals were detected in the salinity signature of Willapa Bay, an estuary 75 km

north of the Columbia River (Hickey et al., 1999).

In this paper we describe a series of numerical experiments designed to better understand

the response of a high discharge river plume to varying wind forcing and coastal currents, and in

particular, the processes involved in direction reversals of such a river plume. A number of

authors have previously addressed the dynamics of lower discharge buoyant plumes using

numerical models. Chao (1986) modeled an estuary and coastal ocean to describe the basic three-

dimensional structure of an unforced river plume as a function of prescribed vertical mixing and

bottom stress. He later explored the influence of bottom slope on the shelf and in the estuary on

plume structure. He also classified plumes as supercritical or subcritical according to an

empirical Froude number (Chao 1988a). Chao (1988b) investigated the effect of wind on pre-

existing estuarine plumes in a coupled shelf-estuary system and again the influence of bottom

slope. He found that a sloping bottom reduced the offshore extent of the plume. Oey and Mellor

(1993) introduced a turbulence closure scheme to calculate the vertical mixing coefficients in

their unforced, flat bottom estuarine plume model. Kourafalou et al. (1996a and 1996b) studied

the influence of wind on a pre-existing plume and made a simulation with realistic wind stress

for an individual source and for a “line source” of freshwater. Fennel and Mutzke (1997) used a

stratified non-tidal coastal ocean but flat bottom slope in their model to study the dynamics of a

river plume with wind forcing. They found that with a stratified ocean, a secondary bulge

Page 4: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

4

develops downstream. A classification of river plumes as surface or bottom-advected based on

the vertical structure was given by Yankovsky and Chapman (1997). Fong (1998) and Fong and

Geyer (2001) studied the effect of winds and downstream ambient flow on a river plume with an

average discharge of 1500 m3 s-1 and a sloping shelf. In particular, these papers addressed

advection and mixing of a surface-trapped plume during an upwelling wind event. Garvine

(1999) investigated the dependence of the alongshelf penetration of an unforced buoyant coastal

discharge on parameters such as bottom slope, background diffusivity, tidal amplitude and river

discharge. In a recent paper Xing and Davies (1999) explore the horizontal spreading and

vertical mixing of a buoyant plume with a discharge of 2000 m3 s-1 as a function of turbulence

closure scheme, wind direction and bottom slope.

The present study extends this research in three important areas by including (1) a very

large volume discharge (about twice the maximum used in other models); (2) ambient flow in a

direction opposite to that of the propagation of coastally trapped waves; and (3) a sequence of

wind direction reversals. The magnitude of the ambient flow, wind stress, estuary width and river

outflow are based on typical values for the Columbia River and the Washington coast. The

horizontal and vertical resolutions are among the finest used in previous studies to ensure that all

pertinent features are adequately resolved.

2. Numerical Model

2.1. Description of the model

The numerical model used in the study is ECOM3d, a three-dimensional, sigma

coordinate, hydrostatic, primitive equation model derived from the Princeton Ocean Model

(Blumberg and Mellor, 1987). Since this model has been widely used, we refer the reader to

Kourafalou (1996a) or Fong (1998) for a more thorough description of the model details. We

focus here only on the details specific to our study.

The model domain is rectangular (100 x 200 grid cells of size 1.5 km x 2 km, with finer

resolution in the cross-shore direction) with a coastal wall on the eastern side and three open

boundaries (Fig. 2a). Freshwater at 10°C is introduced uniformly throughout the top half of the

water column into the two grid cells at the head of an estuary, located at y = 120 km for winter

runs and y = 200 km for summer runs. The estuary is 4 km wide, 10 km long and 20 m deep. An

Page 5: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

5

estuary is included to generate an estuarine circulation that inputs not only freshwater but also

momentum to the coastal ocean. Discharge rate is kept constant throughout a model run although

discharge rate is varied for different experiments. The coastal ocean is initialized with a

homogeneous temperature of 10°C and a salinity of 33 psu.

The 22 sigma layers in the vertical fall on Chebyshev collocation points so as to resolve

the surface and bottom boundary layers as well as the surface-trapped density plume (Fig. 2b).

This results in a vertical resolution better than 1.5 m near the surface across the entire domain.

The bottom topography is a uniform slope α = 2x10-3. This slope is roughly that of the

Washington shelf and does not include the shelf break and continental slope. The reasons for not

having more realistic bottom topography are two-fold: first, we are trying to resolve a surface-

trapped plume that thins out in the offshore direction; the sigma levels follow the opposite

trend—inclusion of the continental slope would decrease the near surface resolution by a factor

of at least five. Second, a larger slope (thus, larger offshore depths) would increase the external

wave speed. The time step required to resolve this larger speed would then have to be smaller by

about a factor of three to keep the model numerically stable, therefore significantly increasing

the time it takes to run the model.

The model includes a mode splitting technique for computational efficiency. The external

and internal time steps are 10 seconds and 7 minutes respectively, in compliance with the CFL

criterion. The horizontal mixing coefficients of salt, temperature and momentum are

parameterized using the Smagorinski (1963) formula, while the vertical mixing coefficients are

parameterized using the 2.5-level closure scheme of Mellor and Yamada (1982). The Coriolis

parameter corresponds to a latitude of 46°N for the majority of runs and is kept constant since

the β-effect is negligible for the spatial and temporal scales examined.

The boundary conditions at the sea surface are zero salt and heat fluxes. In experiments

with wind, the surface stress is set by an alongshore wind stress as described by Blumberg and

Mellor (1987). Winds are applied in the alongshore direction and are constant in the cross-shore

direction and, after an initial ramping, in time. At the bottom, the momentum is balanced by a

quadratic bottom stress with a bottom drag coefficient given by the "law of the wall"; salt and

heat fluxes and vertical velocity are zero. The coastal wall boundary is impenetrable,

impermeable and no-slip.

Page 6: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

6

On the open boundaries, the boundary conditions are such that the bore triggered by the

plume passes through the boundaries. The surface elevation is clamped to zero on the offshore

boundary and radiated on the northern boundary. On the southern boundary the normal external

velocity is set to the specified ambient flow and the surface elevation allowed to adjust

geostrophically to the specified flow. For the tangential external velocities a no-slip condition is

applied. Internal velocities are radiated on all open boundaries following Orlanski (1976).

Temperature and salinity on the open boundaries are relaxed to specified boundary values (those

of the coastal ocean) for inflow, and the existing gradients are advected out of the grid for

outflow. On the northern boundary a sponge layer is implemented over the last 20 km on both

temperature and salinity to absorb the excess river water. For model runs with an ambient flow a

barotropic velocity is imposed at the southern boundary. The model does not include tides since

we will focus on the subtidal response of the buoyant plume to variable wind stress, river

discharge and ambient flow.

2.2. Sensitivity to numerical details

Before proceeding to experiments with variable winds and ambient flows we investigated

the sensitivity of the model to numerical details such as the advection scheme, vertical resolution

and vertical diffusivities of momentum, salt and temperature. In particular, the hydrodynamic

stability of the modeled plume proved to be extremely sensitive to both the advection scheme

and vertical resolution. The spatial structure of the modeled plume proved to be highly sensitive

to vertical diffusivity.

a) Advection scheme and vertical resolution

The use of a centered difference scheme was eliminated as a choice from the outset since

it can lead to negative salinities, especially at the river mouth (Fennel and Mutzke, 1997). Our

next choice was an advection algorithm based on an upwind scheme with an “anti-diffusion”

velocity to correct for the numerical diffusion introduced by the upwind advection scheme

(Smolarkiewicz, 1984; Smolarkiewicz and Clarke, 1986; Smolarkiewicz and Grabowski, 1990).

The resulting scheme is then second order positive definite. We implemented a version with two

consecutive corrective steps (here designated Smolar_2) as well as a computationally more

Page 7: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

7

demanding version in which the anti-diffusion velocity equivalent to applying an infinite number

of corrective steps is estimated using a recursion relationship (here designated Smolar_r).

Surface salinity contours at 14 days for model runs with different vertical resolutions and

different advection schemes are shown in Figure 3. For all three runs a freshwater discharge rate

of 7000 m3 s-1 is used, and winds and ambient flow are zero. Vertical diffusivities are calculated

by the Mellor-Yamada closure scheme with a background value of 10-6 m2 s-1, and the bottom

slope is 2x10-3. Model runs with 12 layers in the vertical with a higher resolution near the surface

and the Smolar_2 advection scheme exhibited instabilities developing around the fringe of the

plume (Fig. 3a). Employing the same advection scheme, but increasing the vertical resolution to

22 layers (so that the surface layer is 0.6 m thick 50 km offshore instead of 0.74 m) delayed the

appearance and growth rate of those instabilities but did not eliminate them (Fig. 3b). Adequate

vertical resolution of the surface-trapped plume was crucial to the stability of the result for the

Smolar_2 advection scheme. Fong (1998) notes that horizontal resolution is also very important

and that under-resolution in the horizontal can also produce wave-like meanders around the

bulge. However, increasing horizontal resolution (grid cells of size 500 x 500 m) using the

Smolar_2 scheme produced no significant changes in our results (not shown).

Because increasing either vertical or horizontal resolution failed to eliminate model

instabilities, the computationally more demanding Smolar_r advection scheme was implemented,

with the result that bulge instabilities disappeared (Fig. 3c). The Smolar_r scheme also reduced

the offshore extent of the bulge by about 20% while increasing the width of the downshelf plume

to accommodate the additional transport. All runs mentioned hereafter use the Smolar_r

advection scheme and 22 layers in a Chebyshev distribution.

b) Vertical mixing coefficients

The model also proved very sensitive to vertical mixing coefficients for salt and heat (KH)

and for momentum (KM). In general, as mentioned above, a Mellor-Yamada level 2.5 turbulence

closure scheme was used to calculate the vertical mixing coefficients, and a background value of

10 -6 m2 s-1 was employed for both coefficients. However, when this background value (UMOL)

was increased to 10 -4 m2 s-1 (Fig. 4a) and especially to 10-3 m2 s-1 (Fig. 4b), the plume's spatial

structure changed dramatically. Garvine (1999) suggests that the turbulence closure scheme shuts

down for high Richardson numbers such as observed near the front of a river plume, so that

Page 8: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

8

vertical mixing coefficients revert to the background value set by the user. In fact, holding both

vertical mixing coefficients at a constant value of 10-4 m2 s-1 and 10-3 m2 s-1 instead of using the

closure scheme for the two runs mentioned above yielded virtually identical results (not shown).

The effects of increasing the vertical mixing coefficient were especially noticeable in the

"upshelf" (sensu Garvine, 1999) penetration of the plume as well as in the offshore extent of the

bulge (Fig. 4). The upshelf (here, south of the estuary) intrusion of the freshwater plume in some

numerical models has been briefly addressed by Garvine (1999) and, in more detail, by

McCreary et al. (1997) and Yankovsky (2000). Kourafalou et al. (1996a) attributed the upshelf

intrusion to the sloping bottom. Our results show, however, that for the base case (the unforced

plume with a bottom slope of 2 x 10-3, 22 layers in a Chebyshev distribution, Smolar_r advection

scheme and a turbulence closure scheme to calculate the vertical mixing coefficients with

background diffusivity of 10-6 m2 s-1, shown in Fig. 3c) little upshelf penetration occurs even

though the bottom has a significant slope. If the background vertical diffusivity is increased to

10-4 m2 s-1 (Fig. 4a), more upstream intrusion is noticeable (compare to the base case with a

background of 10-6 m2 s-1). If the diffusivity is further increased to 10-3 m2 s-1, the bulge is

dramatically diminished in size as the plume water is mixed both upshelf and downshelf of the

river mouth. An anticyclonic eddy develops and propagates upshelf (Fig. 4b).

In the absence of bottom slope, some upshelf intrusion still occurs, but there is no

obvious upshelf propagation of an eddy as with a sloping bottom for the same UMOL and the

bulge is again wider (Fig. 4c). This is consistent with previous numerical model results for flat

bottom cases (Oey and Mellor, 1993; Kourafalou, 1996a; Garvine, 1999). Addition of a

downshelf ambient flow also inhibits the upshelf plume penetration as shown in model results

from Yankovsky and Chapman (1997) who purposefully added a downshelf ambient flow of 4

cm s-1 to prevent upshelf penetration. Similarly, when a northward ambient flow of 10 cm s-1 was

added to the case portrayed in Figure 4b (UMOL = 10-3 m2 s-1) the upshelf penetration of the

plume was eliminated (not shown).

Since upshelf propagation of the type noted here has not been reported in observations of

the Columbia River plume, we have elected to use Mellor-Yamada 2.5 with molecular

background viscosity for our model study (Fig. 3c). However, we note that the choice of vertical

mixing coefficients (or of a background value if using Mellor-Yamada) is clearly non-trivial,

suggesting that model coefficients, configurations and turbulence closure schemes (as shown by

Page 9: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

9

Xing and Davies, 1999) be carefully examined before making model to model or model to data

comparisons.

3. Results

In Section 3.1 plumes are first allowed to develop for about 13 days with ambient shelf

flow. Northward ambient flow simulates winter conditions over the Washington shelf; southward

ambient flow simulates summer conditions. The river discharge is 7000 m3 s-1, the long-term

annual average for the Columbia River. In order to illustrate the response of plume orientation

and surface structure to realistic wind reversals (Sec. 3.1a) these pre-existing plumes formed

with a coastal ambient flow were subjected to 6 days of downwelling-favorable wind stress,

followed by 6 days of upwelling-favorable wind. The magnitudes of the ambient flow and wind

stress are 10 cm s-1 and 0.5 dynes cm-2, respectively. To illustrate plume vertical structure (Sec.

3.1b) the pre-existing plumes formed with the northward and southward ambient flows are

subjected to either several days of upwelling wind stress or several days of downwelling wind

stress. For those cases, wind stress of both 0.5 and 1.4 dynes cm-2 are used. The effect of wind

stress magnitude and direction on freshwater transport are discussed in more detail in Section

3.2. In Section 3.3, where we examine the formation of freshwater pools, discharge rate and

ambient flow are also varied.

3.1. Response to wind stress and ambient flow

a) Plume surface structure

The base case consists of an unforced plume; i.e., freshwater discharges into the coastal

ocean which is at rest (Fig. 3c). The plume structure consists of an anticyclonic bulge off the

mouth of the estuary and a coastal current that propagates as a bore in the same direction as

coastal-trapped waves, as described by others for lower discharge volume cases (Kourafalou,

1996a; Fong, 1998). As Fong (1998) points out, this is a non-steady problem in the sense that the

bulge keeps growing without limit due to the fact that the coastal current is unable to

immediately transport such a large river discharge. Some authors (Garvine, 1987; Kourafalou,

1996a) have classified these plumes as supercritical by analogy with hydraulic theory.

Page 10: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

10

With the addition of an ambient flow in the downshelf direction (northward in this case)

the problem becomes quasi-steady (Fong, 1998). The bulge is advected to the north with the

ambient flow and its offshore extent reaches a limit as the transport in the coastal current is

increased by the ambient flow (Fig. 5a). The maximum offshore extent of the fresher water

occurs several kilometers downstream of the river mouth. We have extended Fong’s study to

include an ambient flow in the opposite direction. In this case some of the freshwater of the

bulge off the mouth of the estuary is advected upshelf (southward) to form an elongated bulge;

however, a downshelf coastal current is still observed (Fig. 6a).

For northward ambient flow (the winter case) with the onset of downwelling-favorable

winds the pre-existing freshwater plume is pushed onshore against the coast (Fig. 5b) so that by

the sixth day of this wind event the freshwater is trapped in a very narrow region adjacent to the

coast (Fig. 5c). With the onset of subsequent upwelling-favorable winds the plume moves

offshore (Fig. 5d). By the end of this upwelling wind event the plume adopts a northwestward

orientation with a northward downshelf tail ~60-100 km offshore (Fig. 5f) reminiscent of the

satellite-derived SST image from winter 1991 (Fig. 1a).

For southward ambient flow (the summer case) the south-southwestward oriented plume

moves onshore at the onset of downwelling-favorable winds so that freshwater plumes are found

both north and south of the river mouth (Figs. 6b and 6c). Both north and south of the river

mouth the freshwater is transported northward in a narrow band next to the coast. By the end of

six days of downwelling winds, the plume resembles the “winter” plume hugging the coast

although a small remnant of freshwater from the southwestward plume is still observed south of

the river mouth (Fig. 6d). In the next episode of upwelling winds the freshwater, including the

remnant, is carried offshore (Fig. 6e) and after six days the plume is in a predominantly

southwest “summer” position (Fig. 6f) reminiscent of the summer ocean color image in Figure

1b. However, because of the recent downwelling, a northward tending “residual” tail remains

approximately 60-80 km offshore north of the river mouth.

The velocity field associated with the low salinity structures seen in Figures 5 and 6 is

highly three-dimensional. For example, the surface velocity field associated with plumes under

northward and southward ambient flow conditions is shown in Figure 7. For these examples, in

which plumes have been allowed to develop for 28 days, distinct pools of low salinity are

observed for ambient flow in both directions and strong geostrophic flows have developed

Page 11: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

11

around the light pools. Such pools can also be seen in the shorter model runs of Figures 5 and 6.

Pool formation and structure will be discussed in more detail in Section 3.3. For northward

ambient flow conditions, the plume reinforces the ambient flow producing a northward jet at its

outer edge where the density front is strongest. Flow is weak along the axis of the plume and a

southward jet is observed near the coast on the shoreward side of the plume (Fig. 7a). For

southward ambient flow, at locations south of the river mouth the southward ambient flow

seaward of the plume reverses to a narrow northward jet at the edge of the plume. Flow is weak

along the plume axis and a strong southward jet is observed on the coastal side of the plume.

North of the river mouth a northward coastal jet occurs with maximum velocities near the coast;

the flow reverses to the ambient direction within about 20 km of the coast (Fig. 7b).

The three-dimensional velocity structure becomes even more complex when wind stress

forcing is added. However, for the standard 0.5 dynes cm-2 forcing used in most of the model

runs, the effect of wind forcing is confined primarily to the top and bottom frictional layers (see

next subsection). This flow is too weak to reverse the ambient flow at the surface although

direction reversals can occur for higher wind stress. Geostrophic wind-driven flows develop only

with the larger stress (see next subsection), but these can reverse the ambient flow and add

substantial complexity to the velocity field.

b) Vertical plume structure

Cross-shelf sections of salinity and east-west and north-south velocity components 40 km

north or south of the river mouth for northward or southward ambient flow conditions,

respectively, are displayed in Figures 8a-c. Results are shown for the base case of no wind, as

well as 3 days after the onset of either downwelling or upwelling-favorable winds of magnitude

0.5 dynes cm-2. In Figures 9a-c the vertical structure obtained with the standard wind stress is

compared with that for greater wind stress (1.4 vs. 0.5 dynes cm-2).

Results show that the plumes are strongly surface trapped with respect to salinity (Fig.

8a). Plume depth is shallow (<10-15 m) whether in the absence of winds or with the standard

wind forcing of 0.5 dynes cm-2. Thus, plumes do not make bottom contact at this location for this

relatively weak wind stress. Under conditions of either no winds or upwelling-favorable winds,

plumes are generally deeper for northward ambient flows than for southward ambient flows and

stratification is weaker. However, this is not the case for downwelling winds (Fig. 8a, compare

Page 12: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

12

middle panels). Although stratification is weaker for northward tending plumes, surface salinity

is usually several psu lower. Plumes deepen and stratification decreases as wind stress increases

for both upwelling and downwelling winds (Fig. 9a), likely due to increased vertical mixing.

The cross-shelf salinity sections also clearly demonstrate the onshore and offshore

movement of the plumes in response to wind stress. For example, the leading edge of the plume

with either northward or southward ambient flows moves onshore almost 30 km after three days

of downwelling and offshore about 60 km after 3 days of upwelling with a wind stress of 0.5

dynes cm-2 in each case (compare location of seaward front in left three panels in Fig. 8a). The

movement across the shelf increases to about 40 km onshore and 100 km offshore with a wind

stress almost three times as great (Fig. 9a).

One of the important questions with respect to buoyant plumes is the manner and extent

to which they modify the regional circulation. We are interested in the magnitudes of velocities

in plume-affected regions in comparison to the more chronic wind-driven velocities as well as

whether a plume significantly affects flow beneath it. To explore these issues in detail we present

velocity on vertical cross-shelf transects (Figs. 8b, c and 9b, c) as well as time series at locations

inside and outside the plume in the surface layer and in the interior, i.e., outside surface and

bottom boundary layers (Fig. 10). For the cross-shelf component of flow, results are shown for

the entire water column so that surface and bottom Ekman layers are captured (Fig. 8b). Details

in the upper 20 m are shown in Figure 9b.

Cross-shelf velocity sections illustrate the presence of bottom Ekman layers in opposite

directions for the oppositely directed ambient flows across the entire shelf (Fig. 8b). During

wind-driven episodes, a surface Ekman layer is also visible, with weak velocities in the upper

~5-10 m. The onshore or offshore flow in the direction of surface Ekman transport is more than

an order of magnitude greater within the plume in most cases. For example, frictional velocities

of about 2.5 cm s-1 occur outside the plume in contrast with about 10 or 50 cm s-1 inside the

plume for a wind stress of 0.5 dynes cm-2, with the larger velocities during upwelling. Note that

for greater downwelling wind stress under northward ambient flow conditions, surface velocities

are reduced in the plume near the coast, likely because the plume is closer to the coastal wall

(Fig. 9c, lower left). A classic upwelling pattern of cross-shelf flow—offshore at the surface and

onshore in the bottom boundary layer—is observed near the coastal wall for both directions of

ambient flow with the modest wind stress of 0.5 dynes cm-2. In the northward ambient flow case,

Page 13: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

13

onshore flow beneath the plume can occur because alongshore flow reverses to southward near

the coast once the plume separates from the coast (Fig. 8b, lower left). For downwelling

conditions, on the other hand, the return flow is not purely confined to the bottom boundary layer

(Fig. 8b, middle panels), unlike the classic downwelling pattern. Reversal of the bottom

boundary layer cross-shelf flow across the entire shelf would only be expected in cases where

wind stress is sufficiently large to reverse the direction of the alongshore ambient flow all across

the shelf.

Alongshelf velocities greater than ambient are largely confined to the upper 10 m of the

water column, indicating that plume related currents are essentially surface trapped (Fig. 8c). The

magnitude of velocities in the plume is several times ambient. In general, plume velocities are

greater for northward ambient flow than for southward ambient flow (~40 vs. 10 cm s-1) except

during upwelling conditions. For the unforced case with both ambient flow directions and for the

upwelling case with northward ambient flow, counterflow due to the geostrophic flow around the

plume is observed—on the coastal side of the plume for northward ambient flow and on the

seaward side of the plume for southward ambient flow (see plan view velocity maps in Fig. 7 and

related discussion). For a wind stress of 0.5 dynes cm-2 wind-driven alongshelf flow adds or

subtracts (depending on wind and ambient current directions) about 5 cm s-1 to the flow field, but

only in the upper 5 m or less (Fig. 8c). When wind stress increases both the magnitude and depth

of influence of the wind-driven frictional flow increase (Fig. 9c, lower panels).

To separate contributions to the alongshelf velocity from buoyancy forcing, wind forcing

and ambient flow, time series of velocity near the surface and in the interior are compared at sites

in the coastal current within (“nearshore”) and outside (“offshore”) the plume (Fig. 10).

Measurement depths are the surface grid point and either 22 m or 85 m in bottom depths of 36

and 140 m at nearshore and offshore sites, respectively, about 140 km downstream of the river

mouth (see locations in Figure 2). Note that surface measurement depths are within the surface

Ekman layer whereas the deeper measurement depths are well outside the bottom Ekman layer.

The heavy lines in the second row of figures display model results for runs with an ambient flow

but without a plume—a “control” case. These data illustrate that nearshore flow is about 3 cm s-1

below ambient both at the surface and in the interior even in the absence of a plume. This

decrease is due to frictional drag on the coastal wall. Thus, in every case we can expect a slight

frictional reduction of the nearshore velocity in comparison to that farther offshore.

Page 14: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

14

The time series show that the majority of the variance occurs in the nearshore surface

layers and that this variance increases when ambient flows or wind driving are added to the

forcing. In contrast to the surface layers, the plume appears to have little effect on the interior

velocity field—offshore and nearshore time series are virtually (with the small deficit due to

frictional effects nearshore as mentioned above). In the absence of wind driving or ambient flow

(top row) alongshelf currents of about 40 cm s-1 are generated by the freshwater plume at this

location 140 km from the river mouth. The total velocity increases by the value of the ambient

flow when ambient flow is added (second row). When wind stress is added to the pre-existing

plumes formed with ambient flow at day 13.3, velocities decrease in the weaker wind case and

then increase dramatically when the buoyancy from the bulge region reaches the measurement

site (third row). When this freshwater passes the velocities return to levels attained in the absence

of wind forcing. The rapid increase is due primarily to the increasing lateral density

gradient—hence geostrophic flow—as the freshwater is moved onshore in the surface Ekman

layer (see Fig. 8a). The increase occurs sooner and is greater (up to 115 cm s-1 above ambient)

with greater wind stress (compare the 1.4 dynes cm-2 case with the 0.5 dynes cm-2 case in Figure

10) .

Wind-driven contributions to the variance are evident in the surface layer offshore of the

plume as a slight increase (~5 cm s-1) over the ambient flow that remains constant over the 3 days

shown (Fig. 10, third row, left; note level of dashed line relative to arrow). This increase is not

observed in the interior (third row, right; dashed line), an indication that the flow is entirely

frictional. For greater stress, a velocity increase with time in both the interior (bottom right,

dashed line) and surface layer outside the plume (bottom left, dashed line) suggests that this

increase (10 cm s-1 over 2 days) is geostrophic rather than frictional.

3.2. Effect of wind stress magnitude on freshwater transport

To study the mechanisms responsible for the advection of freshwater as a function of

applied wind stress, the freshwater transport, Q, was calculated across a rectangular control

volume enclosing a region off the river mouth (shown in Fig. 2a) normalized by the river

discharge. Thus,

Page 15: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

15

QQ

S SS

u n dsest

amb

amb= −

∫∫1

' • '

where Qest is the river discharge, S is the salinity, Samb is the background salinity (33 psu in our

case), u' is the total velocity (for the “total transport”) or the geostrophic velocity obtained from

the pressure field (for the “geostrophic transport”), and n' is the unit vector normal to the surface

of the control volume with a surface element ds (= dxdz or dydz, depending on the transect

chosen). The sides of the control volume were located 80 km north of the estuary (northern

transect), 40 km south (southern transect) and 50 km west (western transect). The control volume

was chosen so that it framed the bulge in the base case at the time at which the winds are added

in the different experiments. The total transport as well as the transport due to geostrophic flow

through each element were obtained as a function of time. The calculations were performed for

both northward and southward ambient flows of 10 cm s-1 and for several wind stress

magnitudes. All runs were spun-up with the average river discharge (7000 m3 s-1) and ambient

flow for 13.6 days, after which a uniform alongshore wind stress (either upwelling or

downwelling) was applied. The run with the lowest wind stress (τ = 0.5 dynes cm-2) had a

duration of almost 6 days; the run with τ = 1.4 dynes cm-2 had a duration of about 3 days; and the

highest wind stress run (τ = 3 dynes cm-2) terminated after about one day due to a violation of the

CFL criterion.

Results show that downwelling winds enhance northward transport of freshwater across

the northern transect by as much as a factor of ten over the plume formed under no wind

conditions for northward ambient flow, and a factor of five for southward ambient flow (Fig. 11,

compare two top panels). In general, northward and westward transports are much greater for

northward ambient flow (top two left panels) than for southward ambient flow (top two right

panels). The transport increases immediately when wind stress is applied, with a more rapid rate

of increase for higher stress. The transport peaks in less than 1.75 days in all cases shown, with

shorter times associated with higher stress. The peak in transport occurs when the primary

freshwater bulge passes the transect. Note that the maximum occurs somewhat earlier here than

in the velocity time series in Figure 10 due to the closer proximity of the transect to the river

mouth. The discussion in the last section demonstrates that the onset of downwelling-favorable

wind stress is accompanied by large flows in the surface Ekman layer which transport the

freshwater toward shore. The increased cross-shelf density gradients generate enhanced

Page 16: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

16

alongshelf flow which produces the transport peak across the northern transect in Figure 11

(upper left). After the bulge of freshwater passes, transports decrease almost to their pre-wind

levels. However, the ongoing Ekman transport confines the freshwater near the coast so that

geostrophic alongshelf currents and hence total transports remain higher than in the no wind

case. Comparison between total transport and geostrophic transport through the northern transect

shows that the transport is primarily geostrophic (Fig. 11). Small but significant differences are

observed between total and geostrophic transport across this transect and also, at times, across

the southern transect. However, the differences disappear once the bulge passes the transect. This

indicates that the ageostrophic flow is due to non-linear advection in the bulge area rather than to

near surface wind-driven frictional flow. Frictional flow is clearly inefficient at driving transport

in the alongshelf direction.

On the other hand, frictional ageostrophic flow is the dominant transport mechanism for

cross-shelf flow. For both northward and southward ambient flow conditions, cross-shelf

transport of freshwater through the western transect occurs only under upwelling conditions (Fig.

11, middle panels; cases with zero transport are not shown). For both northward and southward

ambient flows, the cross-shelf ageostrophic transport moves about as much freshwater as the

alongshelf, primarily geostrophic transport.

Our results have shown that a plume generated under a constant ambient flow will tend in

the direction of the ambient flow (Fig. 7). An important question with respect to plume dynamics

is under what wind conditions the direction of a plume can be reversed. Freshwater transports

show that with northward ambient flow conditions only strong and persistent upwelling winds

generate significant freshwater transport across a transect 40 km south of the river (Fig. 11,

bottom left). The transport must be sufficient to overcome the natural tendency of the plume to

turn northward as well as the northward ambient flow. For our particular model configuration,

such an event required a wind stress of over 1.4 dynes cm-2 lasting longer than 2 days. With

southward ambient flow, on the other hand, the freshwater transport is easily reversed with the

weakest wind stress (Fig. 11, top and bottom right panels).

For northward ambient flow twice as strong, freshwater transports are very similar to the

case with the ambient flow shown in Figure 11 except that the coastal current transports nearly

double the amount of freshwater (as noted also by Fong, 1998). Maximum transport for

downwelling winds, although the same in magnitude, is achieved sooner than with a weaker

Page 17: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

17

ambient flow (not shown). For upwelling winds with stronger northward ambient flow, more

freshwater is transported through the northern transect and less through the western transect than

for lower ambient flow (not shown).

3.3. Freshwater pool formation

As mentioned previously, the structure of the plume differs remarkably when an ambient

flow is added to a freshwater discharge. The most outstanding result, apart from the deformation

of the bulge, is the formation of distinct freshwater pools that detach from the bulge and are

advected with the ambient flow (Fig. 7). These pools all have anticyclonic motion, producing

counterflows adjacent to the coast (for northward ambient flow; Fig. 7a) or in the region offshore

of the coast (for southward ambient flow; Fig. 7b). Oey and Mellor (1993) described the

formation and detachment of freshwater pools in their plume model with no ambient flow.

However, Fong (1998) attributed the features in their model to instabilities around the fringe of

the plume resulting from lack of horizontal resolution. More recently, Yankovsky (2000)

described the periodic shedding of anticyclones in the presence of a weak downshelf ambient

flow. As in the present model Yankovsky (2000) has a sloping bottom (exponential) and constant

discharge.

For a river discharge of 7000 m3 s-1, larger, more energetic pools containing fresher water

at a given time are formed under northward ambient flow conditions than under southward

ambient flow conditions of the same magnitude (compare Figs. 7a and 7b). Also, we note that

pools produced with southward ambient flow conditions are located further offshore than pools

produced under northward flow conditions. Along-plume salinity sections about 15 km from the

coast illustrate differences in depth structure of the plume and its pools formed under ambient

flows of the same magnitude but opposite directions (Fig. 12). In particular, the pools formed

under northward ambient flow conditions are fresher than those formed with southward ambient

flows. They also have a greater thickness, although weaker stratification, at a given time after

formation than pools formed under southward flow conditions. This is due to the fact that with

northward ambient flow the freshwater transport is composed of the buoyancy-induced coastal

current and the ambient flow; on the other hand, for southward ambient flow, part of the

freshwater is transported northward by the coastal current.

Page 18: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

18

The rate of formation and spatial structure of the fresh pools could be affected by the

strength and direction of the ambient flow as well as by river discharge rate. To investigate the

characteristics of the pools in more detail fifteen experiments were performed with a variety of

discharge rates and ambient flows. Results indicate that formation characteristics differ between

northward and southward ambient flow cases. For northward ambient flow the rate of pool

formation depends primarily on the magnitude of the ambient flow. For example, with an

ambient flow of 4 cm s-1, one pool is formed with a variety of discharge rates (compare Figs. 13a

and 13b); with an ambient flow of 10 cm s-1, three pools are formed with a variety of discharge

rates (compare Figs. 7a and 13c).

For southward ambient flow pool formation rate is primarily determined by the

magnitude of the ambient flow as it was for northward ambient flow. However, different

formation regimes were identified for low and high discharge rates. For a discharge rate of 7,000

m3 s-1 pools form within the thin upshelf elongation of the bulge (Fig. 7b). In contrast, for a

discharge rate of 14,000 m3 s-1 the whole bulge detaches to generate a distinct pool (Fig. 13d). In

this case the discharge rate is sufficiently large to rapidly form another bulge about to detach

from the river mouth at 28 days (Fig. 13d). Pools formed in this manner are larger, deeper and

have stronger velocities than their lower discharge counterparts.

In contrast to the rate of pool formation, the cross-shore scale of the pools appears to be a

function of both river discharge rate and ambient flow—pools are wider with greater discharge

and with lower ambient flow (compare examples in Figs. 7 and 13). Typical pool scales are

about 10-30 km. A parameter analysis of the spatial scale, L, of the freshwater pools shows that

pool width depends on the characteristic velocity of the pools, U, as well as the Coriolis

parameter, f, and estuary width. The model runs used to determine the scaling are listed in

Table 1. All runs have the same northward ambient flow (10 cm s-1). Southward ambient flow

runs were excluded because of the extra variable that the two different pool formation regimes

introduces. The velocity scale was measured as the average of the magnitude of the velocity

within the largest pool (the one that was first formed) at about 28 days. The length scale was

measured as the distance between the maximum and minimum alongshore velocity along a

cross-shore transect through the center of the largest pool at about 28 days. The length scale was

also measured from the salinity field (as the radius of the 32 psu contour comprising the pool)

with analogous results. Results show that for the same ambient flow and discharge rate pool

Page 19: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

19

scale decreases significantly with latitude, e.g., from 30 km at 20˚N to 15 km at 60˚N (Table 1).

The width scale of the pools also decreases significantly with increasing estuary width. In

general, pool width scales as U/f (Fig. 14). The near linear relationship between the scale of the

pools and U/f shows that the pools formed in the different model runs all have a similar Rossby

number, U/fL. Values range between 0.3 and 0.5, with a best linear least-square fit of 0.41. This

shows that although rotation is dominant, nonlinear advection of momentum is significant within

the pools.

Preliminary calculations of the form drag on isopycnal surfaces within the pools suggest

that for a northward ambient flow the form drag differentially accelerates the fluid in the

horizontal to reduce the pre-existing horizontal shear (not shown). Thus the pools are formed as a

result of barotropic instabilities, gaining energy from the lateral shear. For southward ambient

flow the instabilities are baroclinic, gaining energy from the vertical shear. A complete analysis

of the formation of the anticyclonic pools is a subject of ongoing research and is beyond the

scope of this paper.

4. Summary and Discussion

A numerical model was used to study the factors affecting the spatial structure and

variability of a high discharge buoyant plume over a sloping shelf. Response of the plume to

ambient flows (both northward and southward), periods of wind reversals and a variety of

freshwater discharge rates were examined. The model parameters are based on the Columbia

River plume and its oceanographic environment, and, in spite of being a somewhat idealized

model (no continental slope and no stratification in the coastal ocean), the model appears to

capture much of the reported seasonal behavior of the Columbia River plume. The model has

also distinguished features hitherto unreported in the literature—separating pools with both

directions of ambient flow and a “dual-mode” plume structure (i.e., branches both upstream and

downstream of the river mouth) in the summer season.

Various numerical details were found to affect model results. In particular, a fine vertical

resolution of the surface-trapped plume along with the advection scheme Smolar_2 proved

critical to avoiding instabilities around the fringe of the plume’s bulge. These instabilities were

completely eliminated by employing a recursive version of this advection scheme (Smolar_r).

The choice of vertical mixing coefficients (or a background value if calculating coefficients with

Page 20: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

20

the Mellor-Yamada 2.5 turbulence closure scheme) had a strong influence on the upshelf

penetration of the plume. The degree of penetration increases as the vertical mixing coefficient

increases. Some penetration occurs even when the bottom is flat, although the propagation of an

eddy-like feature derived from the bulge was observed only with a sloping bottom.

The model plume's response to winds is very rapid (several hours) as is the case for the

Columbia River plume, which responds to winds in a matter of 3-6 hours (Hickey et al., 1998).

The plume is moved onshore during periods of downwelling wind and offshore during periods of

upwelling wind, regardless of the direction of the ambient flow typically northward in winter and

southward in summer off the Washington coast. Both alongshelf velocity and freshwater

transport increase by as much as a factor of 10 for a brief period following the onset of wind

stress when the low salinity water in the bulge moves onshore, increasing lateral gradients and

hence alongshelf geostrophic flow. Once the bulge water has been transported through the area,

transports and velocities return almost to their pre-wind levels for weak wind stress, although

transport remains higher than in the no wind case due to the onshore trapping of the continuous

supply of light water by the Ekman transport. Transport increases with increasing wind stress;

with higher stress an alongshelf geostrophic wind-driven component adds to the total flow and

enhances the alongshelf transport.

Plume-related velocities are highly three dimensional, with counterflows occurring

frequently as a result of near-geostrophic transport around features of lower salinity water as well

as the competing effects of the ambient flow and the plume. Plume-induced velocities are

generally confined to the near surface region of lower salinity, as observed in the Columbia

plume (Hickey et al., 1998). Both the width and depth of the plumes during the winter

simulations are similar to those shown in Hickey et al. (1998); namely, for the majority of the

freshwater, 5-20 km width and 10-40 m depth for downwelling and ~70-100 km width and less

than 10 m depth for upwelling.

The model results challenge longstanding notions about the Columbia plume: first, that

the plume orientation is in a relatively stable southwest position in summer (see e.g., Barnes et

al., 1972; Hickey, 1989). The model results show that with average discharge conditions (7,000

m3 s-1) a summertime downwelling event of typical magnitude and duration can erode and advect

away the bulk of the southwestward plume to the north of the river mouth over several days. The

surface Ekman transport first pushes the southwestward tending plume of freshwater against the

Page 21: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

21

coast; once there, the geostrophic flow associated primarily with the lateral density gradients

transports the plume northward until it reaches almost a typical “winter downwelling” position.

Only a weak plume remnant is left off the coast south of the river mouth. The return to upwelling

conditions moves the plume offshore north of the river mouth as it does in winter with northward

ambient flow conditions, but the southward ambient flow directs the newly formed bulge region

south-southwest as opposed to north-northwest in winter. A plume originating from a larger

discharge event (e.g., 14,000 m3 s-1) would be expected to have higher stratification and a larger

volume and hence would be more difficult to mix or displace. The presence of ambient

stratification, which was not included in our model, might also inhibit erosion of plume water by

limiting vertical mixing.

A second traditional notion of the Columbia plume is that the plume is only oriented

southwest in summer (Barnes et al., 1972; Hickey, 1989). The model results show that even

when the plume tends southwest due to the ambient southward flow, a narrow plume hugs the

coast north of the river mouth. Thus, the plume frequently has both northward and southward

branches at the same time (Fig. 7b). Careful inspection of available observations show that this

does indeed appear to be the case. In every available survey, light water is observed in summer

off the Washington coast—over the mid to outer shelf during upwelling events and next to the

coast during downwelling events (e.g., Hermann et al., 1989, Figs. 6.7b and 6.7c; Horner et al.,

2000). The presence of this buoyant coastal current is also consistent with the mean northward

flow that has been reported in this region in summer near the coast (Hickey, 1989). This

phenomenon has been overlooked in previous studies—the presence of light water was usually

attributed to an isolated newly emerging plume rather than a persistent plume as suggested by the

model results.

Observations of the Columbia plume in winter showed that in spite of periods of

persistent upwelling the plume never changed direction from generally north-northwestward to

southwestward (Hickey et al., 1998). The model results suggest that the difficulty in reversing

the plume direction in winter is due in part to the northward direction of the mean ambient flow,

which is in the same direction as the natural rotational tendency of the plume. For the model’s

winter conditions the plume develops substantial reversals, ones that reach the measurement

section 40 km south of the river mouth only for southward wind stress greater than 1.4 dynes

cm-2 blowing for at least two days. Such strong upwelling wind events in the winter are rare.

Page 22: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

22

Another important result from the model is the demonstration that the addition of an

ambient flow to the river plume model elongates the freshwater bulge in the direction of the

ambient flow leading to the generation of distinct freshwater anticyclonic pools that detach from

the bulge. Such pools have been observed in both northward and southward tending plumes from

the Columbia. In particular, winter observations revealed a strong counterflow next to the coast

downstream of the mouth consistent with such pool formation (Hickey et al., 1998). For very

large discharges and southward ambient flow conditions typical of summer in the Pacific

Northwest (14,000 m3 s-1) plumes are more likely to form detached eddies. Distinct low salinity

pools have been observed in the southwest tending Columbia plume (e.g., Barnes et al., 1972;

Fiedler and Laurs, 1990; Hickey 1989) and have sometimes been attributed to tidal flows (see

Fig. 1b). Our model results provide another possible mechanism for pool formation, one that

does not depend on a time-variable source.

The rate at which pools form is a function of the strength of the ambient flow in either

direction but not the river discharge rate. For example, for northward ambient flow conditions,

only one pool formed with an ambient flow of 4 cm s-1; three pools formed with an ambient flow

of 10 cm s-1 for a variety of discharge rates. For southward ambient flow, the structure of the

pools does vary with discharge rate. Low discharge rates produce thin, weakly stratified pools;

large discharge rates result in more vigorous, deeper and more stratified pools because the entire

bulge can completely detach from the river mouth. The size of the pools depends not only on the

magnitude of the ambient flow, but also on discharge rate, width of the estuary mouth and

latitude.

River plumes are highly important for distribution of sediment as well as for

phytoplankton growth and transport, larval transport and movement of juvenile fish. A high

volume river plume such as we model here can impact thousands of square kilometers of a

coastal region. The plume affects stratification, provides fronts that may act as lateral boundaries

and induces local currents that can be several times the strictly wind-driven flow. The presence

of a plume enhances cross-shelf movement of suspended and dissolved material by an order of

magnitude over that in a region of wind-driven transport alone. This model has taken an

important step in examining the variability of such plumes in realistic conditions of ambient flow

and changeable wind driving. However, several questions remain. For example, what are the

formation mechanisms for the anticyclonic pools that form with the ambient flows; how would

Page 23: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

23

ambient stratification and realistic bottom slope affect plume structure and variability; what are

the respective roles of mixing and advection in distributing freshwater? These and other

questions are the subject of our continuing research.

Acknowledgements

We would like to thank Alan Blumberg for making the numerical code available to us

and Derek Fong, Rich Signell and John Klinck for their assistance with the model

implementation. This work was funded in part by grants to B. Hickey (the Pacific Northwest

Coastal Ecosystem Regional Research program, a National Oceanic and Atmospheric

Administration Coastal Ocean Program grant #NA960PO238), Washington Sea Grant (grant

#NA76RG0119 and NA76RG0119 AM08) and the National Science Foundation (grant

#OCE968186). Ms. García Berdeal was funded primarily by a fellowship from the Fundación

Pedro Barrié de la Maza, A Coruña (Spain).

Page 24: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

24

References

Barnes, C. A., A. C. Duxbury, and B.-A. Morse, Circulation and selected properties of theColumbia River effluent at sea, in The Columbia River Estuary and Adjacent OceanWaters, edited by A. T. Pruter and D. L. Alverson, pp. 41-80, Univ. of Wash. Press,Seattle, 1972.

Blumberg, A. F., and G. L. Mellor, A description of a three-dimensional coastal oceancirculation model, in Three-Dimensional Coastal Ocean Models, Coastal Estuarine Sci.,vol. 4, edited by N. S. Heaps, pp.1-16, AGU, Washington, D. C., 1987.

Chao, S.-Y., Wind-driven motions of estuarine plumes, J. Phys. Oceanogr., 18, 1144-1166,1988a.

Chao, S.-Y., River-forced estuarine plumes, J. Phys. Oceanogr., 18, 72-88, 1988b.

Chao, S.-Y., and W. C. Boicourt, Onset of estuarine plumes, J. Phys. Oceanogr., 16, 2137-2149,1986.

Fiedler, P. C., and R. M. Laurs, Variability of the Columbia River plume observed in visible andinfrared satellite imagery, Int. J. Remote Sensing, 11, 999-1010, 1990.

Fennel, W., and A. Mutzke, The initial evolution of a buoyant plume, J. Mar. Syst., 12, 53-68,1997.

Fong, D. A., Dynamics of freshwater plumes: Observations and numerical modeling of the wind-forced response and along-shore freshwater transport. Ph.D. thesis, MIT/WHOI JointProgram in Oceanography, Woods Hole, 1998.

Fong, D. A., and W. R. Geyer, Response of a river plume during an upwelling favorable windevent, J. Geophys. Res., 106, 1067-1084, 2001.

Garvine, R. W., Penetration of buoyant coastal discharge onto the continental shelf: A numericalmodel experiment, J. Phys. Oceanogr., 29, 1892-1909, 1999.

Grimes, C. B., and M. J. Kingsford, How do riverine plumes of different sizes influence fishlarvae: do they enhance recruitment?, Mar. Freshwater Res. 47, 191-208, 1996.

Hermann, A., B. M. Hickey, M. R. Landry, and D. Winter, Coastal upwelling dynamics, inCoastal Oceanography of Washington and Oregon, edited by M. R. Landry and B. M.Hickey, pp. 211-253, Elsevier Science, Amsterdam, The Netherlands, 1989.

Hickey, B. M., Patterns and processes of circulation over the shelf and slope, in CoastalOceanography of Washington and Oregon, edited by M. R. Landry and B. M. Hickey,pp. 41-109, Elsevier Science, Amsterdam, The Netherlands, 1989.

Page 25: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

25

Hickey, B. M., L. J. Pietrafesa, D. A. Jay, and W. C. Boicourt, The Columbia River plume study:Subtidal variability in the velocity and salinity fields, J. Geophys. Res., 103, 10,339-10,368, 1998.

Hickey, B. M., J. Newton, and N. Banas, Event scale physical processes in a coastal plainestuary, Willapa Bay, Washington (abstract), Eos Trans. AGU, 80(49), 1999.

Horner, R., B. M. Hickey, and J. Postel, Harmful algal blooms off Washington State, USA, inProceedings of International Symposium and Workshop on Harmful Algal Blooms in theBenguela Current and other Upwelling Ecosystems, Swakopmund, Namibia, 5-6November, 1998, in press, 2000.

Kourafalou, V. H., L. Oey, J. Wang, and T. N. Lee, The fate of river discharge on the continentalshelf, 1, Modeling the river plume and inner shelf coastal current, J. Geophys. Res., 101,3415-3434, 1996a.

Kourafalou, V.H., T.N. Lee, L. Oey, and J. Wang, The fate of river discharge on the continentalshelf, 2, Transport of coastal low-salinity waters under realistic wind and tidal forcing, J.Geophys. Res., 101, 3435-3456, 1996b.

McCreary, J. P., S. Zhang, and S. R. Shetye, Coastal circulation driven by river outflow in avariable-density 11/2-layer model, J. Geophys. Res., 102, 15,535-15,554, 1997.

Mellor, G. L., and T. Yamada, Development of a turbulent closure model for geophysical fluidproblems, Rev. Geophys. Space Phys., 20, 851-875, 1982.

Oey, L.-Y., and G. L. Mellor, Subtidal variability of estuarine outflow, plume, and coastalcurrent: A model study, J. Phys. Oceanogr., 23, 164-171, 1993.

Smolarkiewicz, P. K., A fully multidimensional positive definite advection transport algorithmwith small implicit diffusion, J. Comput. Phys., 54, 325-362, 1984.

Smolarkiewicz, P. K., and T. L. Clarke, The multidimensional positive definite advectiontransport algorithm: Further development and applications, J. Comput. Phys., 67, 396-438, 1986.

Smolarkiewicz, P. K., and W. W. Grabowski, The multidimensional positive definite advectiontransport algorithm: Nonoscillatory opinion, J. Comput. Phys., 86, 355-375, 1990.

Yankovsky, A. E., The cyclonic turning and propagation of buoyant coastal discharge along theshelf, J. Mar. Res, 58, 585-607, 2000.

Yankovsky, A. E., and D. C. Chapman, A simple theory for the fate of buoyant coastaldischarges, J. Phys. Oceanogr., 27, 1386-1401, 1997.

Page 26: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

26

Xing J., and A. M. Davies, The effect of wind direction and mixing upon the spreading of abuoyant plume in a non-tidal regime, Cont. Shelf Res., 19, 1437-1483, 1999.

Page 27: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

27

Table 1. Parameter values for the model runs used to study the length scale L of the freshwaterpools. Q is the river discharge, f is the Coriolis parameter, w is the width of the estuary and U isthe scale of the velocity within the pools. Each run was made for 28 days. Ambient flow wasnorthward at10 cm s-1 for each case.

Model Run Q (m3 s-1) f (s-1) x104 w (km) U (m s-1) L (km)B4 1500 1.046 4 0.5 10B5 7000 1.046 4 0.8 20B6 14000 1.046 4 0.8 25C1 7000 0.497 4 0.6 30C2 7000 1.260 4 0.7 18C3 7000 1.046 14 0.6 15

Page 28: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

28

Figure Captions

Figure 1. (a) SST image of the Washington-Oregon coast from 23 February 1991, illustrating

the Columbia River plume during an upwelling-favorable wind event in winter. (b) CZCS image

from 24 May 1982 illustrating the Columbia River plume during peak river flow in early summer

with southward winds (adapted from Fiedler and Laurs, 1990).

Figure 2. (a) Model domain and horizontal resolution in the x-y plane. Location of the estuary is

y=120 km for winter runs and y=280 km for summer runs. Symbols denote locations for data

used in time series in Figure 10. The rectangular region marked in the figure indicates the control

volume used in the freshwater transport calculations in Section 3.2. (b) Model domain and

resolution in the x-z plane. Dots denote the center of the grids.

Figure 3. Model sensitivity to vertical resolution and advection scheme. Surface salinity (psu)

contours at t=14 days for (a) 12 layers in the vertical and Smolar_2, (b) 25 layers in the vertical

and Smolar_2, (c) 22 layers in the vertical and Smolar_r. For all three runs a freshwater

discharge rate of 7000 m3 s-1 is used, the diffusivities are calculated by the Mellor-Yamada

closure scheme with a background value of 10-6 m3 s-1 and the bottom slope is 2x10-3. Contours

are drawn from 31 to 13 psu in units of 3.

Figure 4. Model sensitivity to vertical mixing coefficients and bottom slope. Surface salinity

(psu) contours at t=14 days for a background vertical mixing of (a) 10-4 m2 s-1 and bottom slope

of 2 x 10-3, (b) 10-3 m2 s-1 and a bottom slope of 2 x 10-3, and (c) 10-3 m2 s-1 with a flat bottom. The

freshwater discharge rate is 7000 m3 s-1 and winds and ambient flow are zero. Contours are

drawn from 32 to 12 psu in units of 2.

Figure 5. Evolution of surface salinity (psu) for northward ambient flow conditions in response

to 6 days of downwelling-favorable winds followed by 6 days of upwelling-favorable winds at

(a) 13 days, (b) 15 days, (c) 19 days, (d) 21 days, (e) 22 days and (f) 25 days with a northward

ambient flow of 10 cm s-1. Winds change direction after 19 days immediately after (c). The

distance between tick marks is 20 km.

Page 29: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

29

Figure 6. Evolution of surface salinity (psu) for southward ambient flow conditions in response

to 6 days of downwelling-favorable winds, followed by 6 days of upwelling-favorable winds at

(a) 13 days, (b) 15 days, (c) 16 days, (d) 19 days, (e) 21 days and (f) 25 days with a southward

ambient flow of 10 cm s-1. Winds change direction after 19 days immediately after (d). The

distance between tick marks is 20 km.

Figure 7. Surface salinity (psu) contours and surface velocity vectors (m s-1) at t = 28 days for

(a) northward ambient flow of 10 cm s-1 and (b) southward ambient flow of 10 cm s-1. River

discharge for both cases is 7000 m3 s-1.

Figure 8a. Salinity contours at 2 psu intervals for a cross-shore section 40 km north of the river

mouth for northward ambient flow conditions (left panels) and 40 km south of the river mouth

for southward ambient flow conditions (right panels). Results are shown for ambient flow only

(top panels), after 13.3 days of ambient flow and 3 days of downwelling winds (middle panels),

and after 13.3 days of ambient flow and 3 days of upwelling winds (bottom panels).

Figure 8b. Cross-shelf velocity structure (m s-1) corresponding to panels in Figure 8a. Offshore

flow is shaded. Note that to show the bottom boundary layer, the vertical sale differs from that

used in Figures 8a and 8c. To aid visualization of this model “snapshot” data were smoothed

with a 5-7 point binomial filter. Contours are drawn at 0.0 and +/- 0.05, 0.1, 0.3, 0.5 and 0.7 m

s-1. Dashed contour levels fall between those in the list. In this figure only contour maxima are

labeled.

Figure 8c. Alongshelf velocity structure (m s-1) corresponding to panels in Figures 8a and 8b.

Southward flow is shaded. To aid visualization of this model “snapshot” data were smoothed

with a 5-7 point binomial filter. Contours are drawn at 0.0 and +/- 0.1, 0.2, 0.4, 0.6, and 0.8 m s-1.

Dashed contour levels fall between those in the list. Ambient flow is 0.1 m s-1.

Figure 9a, 9b, 9c. Comparison of (a) salinity, (b) cross-shelf velocity and (c) alongshelf

velocity for two different values of downwelling-favorable wind stress, applied for 3 days

following 13.3 days of ambient flow. Contour intervals, smoothing and shading are as in Figure

Page 30: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

30

8. Dramatic structure in the velocity fields near the left margin are strictly numerical and are due

to the presence of the boundary at 150 km.

Figure 10. Time series of velocity at sites within the plume (nearshore) and outside the plume

(offshore) for surface and interior flow during northward ambient flow. Locations are shown in

Figure 2. The sites are in 36 and 140 m bottom depths; the measurement depths are the surface

grid point and 22 or 85 m, the latter depth being well above the bottom boundary layer. Heavy

lines and dashes show results with an ambient flow but no plume. Wind was applied after 13.3

days of forcing with ambient flow (data near the end of the second row). The initial spike before

day 14 in the bottom two left panels is a transient response to the onset of wind forcing and can

be ignored in the context of the discussion. Arrows on the side of the panels indicate ambient

flow magnitude. Note halving of vertical scale in bottom 2 rows.

Figure 11. Total freshwater transport (heavy lines) and geostrophic transport (faint lines) across

northern (N), western (W) or southern (S) transects for northward (left panels) and southward

(right panels) ambient flows of 10 cm s-1 and either downwelling or upwelling-favorable wind

stress. The magnitude of the wind stress (in dynes cm-2) is indicated next to the corresponding

curve. The zero winds case is labeled as “0.0”. Note change in y-axis scale for transport across

the southern transect in bottom two panels.

Figure 12. Along-plume salinity structure15 or 18 km from the coast with ambient flows in both

northward and southward directions but with no winds applied. River discharge rate is 7000 m3

s-1. To aid visualization of this model “snapshot” data were smoothed with a 5 point binomial

filter.

Figure 13. Surface salinity (psu) contours and surface velocity vectors (m s-1) at t = 28 days for

selected runs with various discharge rates and ambient flows.

Figure 14. Relationship between the characteristic length scale (L) of the freshwater pools and

the velocity scale of the pools (U) divided by the Coriolis parameter (f). The straight line shows

Page 31: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

31

the linear, least squares fit. The inverse of the slope (the Rossby number) is 0.41 (r2 = 0.96;

correlations exceeding 0.75 are significant at the 95% level).

Page 32: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

_ 4

6o N

4

6o N -

47o N

- 125o W

124

o W

48o N

-

45o N

-

(a)

(b)

Figu

re 1

125

o W12

4o W

Page 33: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

120 100 80 60 40 20 0 0

20

40

60

80

100

120

140

160

180

200

220

240

260

280

300

320

340

360

380

400

x (km)

y (k

m)

Plan view

120 100 80 60 40 20 0 −300

−200

−100

0

x (km)

z (m

)

Section (x−z)(a) (b)

x

Figure 2

x

Page 34: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

120 80 40 0

80

120

160

200

240

280

x (km)

y (k

m)

31

25

19 13

Smolar_2

12 layers

t=14 d

120 80 40 0 x (km)

3125

19

13

Smolar_2

22 layers

t=14 d

120 80 40 0 x (km)

3125 19

13

Smolar_r

22 layers

t=14 d

a) b) c)

Figure 3

Page 35: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

120 80 40 0

80

120

160

200

240

280

x (km)

y (k

m)

3228

24

UMOL=10−4 m2 s−1

α=2 x 10−3

t=14 d

120 80 40 0 x (km)

32

28

UMOL=10−3 m2 s−1

α=2 x 10−3

t=14 d

120 80 40 0 x (km)

32

28

24

UMOL=10−3 m2 s−1

α=0

t=14 d

a) b) c)

Figure 4

Page 36: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

(a) No wind (b) Downwelling (c) Downwelling

(d) Upwelling (e) Upwelling (f) Upwelling

Northward ambient flow

t = 13 days t = 15 days t = 19 days

t = 21 days t = 22 days t = 25 days

40 km

Figure 5

Page 37: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

Southward ambient flow

(a) No wind (b) Downwelling (c) Downwelling

(d) Downwelling (e) Upwelling (f) Upwelling

t = 13 days t = 15 days t = 16 days

t = 19 days t = 21 days t = 25 days

40 km

Figure 6

Page 38: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

(a)

Nor

thw

ard

ambi

ent f

low

, no

win

ds

(b)

Sout

hwar

d am

bien

t flo

w, n

o w

inds

Q =

700

0 m

3 s-1, v

amb=

10

cm s

-1

Q =

700

0 m

3 s-1

, , v am

b= -

10 c

m s

-1

x (k

m)

x (k

m)

1

.0 m

s-1

1.

0 m

s-1

t = 2

8 da

ys

t =

28

days

1

20

80

40

1

20

80

40

-

40 -

40

0

0

4

0

40

80

80

1

20 1

20

1

60 1

60

y (km)

y (km)

Fig

ure

7

Page 39: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

−20

−15

−10

−5

0

z(m

)

Northward flow, no wind

32

28

24

20

32.75

t=13.3 d

Southward flow, no wind

32

2824

32.75

t=13.3 d

−20

−15

−10

−5

0

z(m

)

Northward flow, downwelling, τ = 0.5

3228

24

32.75

t=16.4 d

Southward flow, downwelling, τ = 0.5

32

32.75

t=16.4 d

120 100 80 60 40 20 −20

−15

−10

−5

0

x(km)

z(m

)

Northward flow, upwelling, τ = −0.5

32

28

32.75

t=16.4 d

120 100 80 60 40 20 x(km)

Southward flow, upwelling, τ = −0.5

32

2832.75

t=16.4 d

28

24 24

~33.0

~33.0 ~33.0

~33.0 ~33.0

Figure 8a

~33.0

Page 40: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

z(m

)

t=13.3 d

−300

−250

−200

−150

−100

−50

0

t=13.3 d

z(m

)

t=16.4 d

−300

−250

−200

−150

−100

−50

0

t=16.4 d

x(km)

t=16.4 d

140 120 100 80 60 40 20 x(km)

z(m

)

t=16.4 d

0

140 120 100 80 60 40 20 −300

−250

−200

−150

−100

−50

0

−0.1 −0.05

0.1 0.1

−0.5 −0.3

Northward flow, no wind Southward flow, no wind

Northward flow, downwelling, τ = 0.5 Southward flow, downwelling, τ = 0.5

Northward flow, upwelling, τ = −0.5 Southward flow, upwelling, τ = −0.5

Figure 8b

Page 41: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

z(m

)

Northward flow, no wind

0.10.2 0.4

0

0.6

t=13.3 d

−20

−15

−10

−5

0Southward flow, no wind

0−0.1

−0.05

t=13.3 d

z(m

)

Northward flow, downwelling, τ = 0.5

0.1

0.20.4

0.15

t=16.4 d

−20

−15

−10

−5

0Southward flow, downwelling, τ = 0.5

0

0.1

−0.075

−0.05

t=16.4 d

x (km)

z(m

)

Northward flow, upwelling, τ = −0.5

0

−0.1−0.2

0.2

0.1

0.1

0.1

t=16.4 d

120 100 80 60 40 20 −20

−15

−10

−5

0

x (km)

Southward flow, upwelling, τ = −0.5

−0.

1

−0.2−0.1 −0.15

t=16.4 d

120 100 80 60 40 20

−0.1 −0.2

−0.4

Figure 8c

Page 42: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

−20

−15

−10

−5

0

z(m

)

Northward flow, downwelling, τ = 0.5

3228

2432.75

t=16.4 d

140 120 100 80 60 40 20 −20

−15

−10

−5

0

x(km)

z(m

)

Northward flow, downwelling, τ = 1.4

322832.75

t=16.4 d

Northward flow, upwelling, τ = −0.5

32

28

32.75

t=16.4 d

140 120 100 80 60 40 20 x(km)

Northward flow, upwelling, τ = −1.4

32

28

32.75

t=16.4 d

~33.0

24 24

~33.0

~33.0 ~33.0

Figure 9a

Page 43: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

z(m

)

Northward flow, downwelling, τ = 0.5

0.050.1

0

0.025

t=16.4 d

−20

−15

−10

−5

0

x(km)

z(m

)

Northward flow, downwelling, τ = 1.4

0.05

0.06

0.02

5t=16.4 d

140 120 100 80 60 40 20 −20

−15

−10

−5

0

Northward flow, upwelling, τ = −0.5

−0.05

−0.1−0.3

0

0

−0.025

t=16.4 d

x(km)

Northward flow, upwelling, τ = −1.4

−0.

05

−0.1−0

.3

−0.5−0.3

00

0

−0.0

25

t=16.4 d

140 120 100 80 60 40 20

0.05

−0.1

−0.5

−0.1

Figure 9b

Page 44: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

z(m

)

Northward flow, downwelling, τ = 0.5

0.1

0.20.4

0.15

t=16.4 d

−20

−15

−10

−5

0

x (km)

z(m

)

Northward flow, downwelling, τ = 1.40.4

0.3

0.2

0.2

t=16.4 d

140 120 100 80 60 40 20 −20

−15

−10

−5

0

Northward flow, upwelling, τ = −0.5

0.1

0.1

0.2

0.1 0

−0.1

t=16.4 d

x (km)

Northward flow, upwelling, τ = −1.4

0−

0.1−0.2

0.10.2

0.1

0 0

0.1

t=16.4 d

140 120 100 80 60 40 20

−0.05

−0.2

Figure 9c

Page 45: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

0 5 10 15

0

0.2

0.4

0.6v

(m s

−1 )

0 5 10 15

0

0.2

0.4

0.6

0 5 10 15

0

0.2

0.4

0.6

v (m

s−

1 )

0 5 10 15

0

0.2

0.4

0.6

14 16 18

0

0.4

0.8

1.2

v (m

s−

1 )

14 16 18

0

0.4

0.8

1.2

14 16 18 0

0.4

0.8

1.2

time (days)

v (m

s−

1 )

14 16 18 0

0.4

0.8

1.2

time (days)

Uamb

=0

Uamb

=10

Uamb

=10

Uamb

=10

SURFACE

nearshore

offshore τ = 0

τ = 0

τ = 0.5

τ = 1.4

no plume

INTERIOR

Figure 10

Page 46: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

12 14 16 18 200

2

4

Northward ambient flow

Nor

mal

ized

FW

Tra

ns.

12 14 16 18 20

−4

−2

0

Nor

mal

ized

FW

Tra

ns.

12 14 16 18 20−0.1

−0.05

0

time (days)

Nor

mal

ized

FW

Tra

ns.

12 14 16 18 200

2

4

Southward ambient flow

12 14 16 18 20

−4

−2

0

12 14 16 18 20

−1

0

1

time (days)

N

W

S

downwelling

downwelling upwelling

upwelling

0.0

3.0

1.4

0.5

1.4

0.5

3.0

1.4

0.5 1.4

0.5

3.0

1.4 0.5

1.4 0.5

0.0

0.0

0.0

0.0

0.0

Figure 11

N

W

S

downwelling

upwelling

Page 47: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

z(m

)

Northward flow, no wind

33

32

28 2420

16

32.75

15 km offshoret=28 d

−80 −40 0 40 80 120 160 200−35

−30

−25

−20

−15

−10

−5

0

Distance from mouth (km)

z(m

)

Southward flow, no wind

33

3228

24 16

32.75

18 km offshoret=28 d

−160 −120 −80 −40 0 40 80 120 −35

−30

−25

−20

−15

−10

−5

0

Figure 12

Page 48: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

1.00 m s-1 1.00 m s-1

1.00 m s-1 1.00 m s-1

(a) Q = 1500 m3 s-1, vamb

= 4 cm s-1 (b) Q = 7000 m3 s-1, vamb

= 4 cm s-1

(c) Q = 14,000 m3 s-1, vamb

= 10 cm s-1 (d) Q = 14,000 m3 s-1, vamb

= -10 cm s-1

0 0

40 40

80 80

120 120

160 160

-40 -40

120 80 40 120 80 40 x (km) x (km)

y (

km)

y (

km)

y (

km)

y (

km)

120 80 40 x (km)

120 80 40 x (km)

-40

0

40

80

120

160 80

40

0

-40

-80

-120

Figure 13

Page 49: Influence of Wind Stress and Ambient Flow on a High ...coast.ocean.washington.edu/coastfiles/GarciaBerdeal_et_al-plume.pdf · Influence of Wind Stress and Ambient Flow on a High Discharge

0 5 10 150

5

10

15

20

25

30

35

U/f (km)

L (k

m)

Figure 14