Abstract

Marine outfalls discharge wastewater on the inner shelf and are designed to encourage rapid effluent mixing sufficient to maintain a submerged wastefield. A high-resolution nonhydrostatic Regional Ocean Modeling System (ROMS) model was used to resolve concomitantly the intermediate- and far-field submarine plume development. ROMS simulations were validated with cross-flow laboratory experiments. Generally, results showed that the nonhydrostatic high-resolution ROMS is capable of resolving plume dynamics in typical cross-flow conditions. Top-of-plume elevation was quantified and found to be highly variable. The ROMS model is relatively insensitive to changes in horizontal effluent input parameterization. Multiple grid resolutions were tested, and good model–data agreement was achieved in low to medium cross-flow experiments. Additional resolution improved high cross-flow results. This intermediate- and far-field three-dimensional nonhydrostatic model resolves plume development over multiple spatiotemporal scales and can include natural oceanic processes currently absent in many plume models. Integrated outfall plume and marine process modeling will advance future wastewater management.

Introduction

Marine outfalls discharge municipal wastewater into coastal environments. Outfalls are designed to encourage rapid mixing and sufficient offshore effluent dilution to restrict plume rise and ultimately produce a submerged wastefield. Preventing wastewater effluent from penetrating the surface is a priority for wastewater treatment plants. Recently, publicly owned treatment works (POTWs) have begun reducing the total effluent volume discharged by extracting a portion of freshwater, which decreases volume and increases discharge pollutant concentrations and density. Initial mixing characteristics, far-field plume establishment, and plume rise heights may be affected.
Wastewater outfall modeling has been approached through various analytical and empirical, random walk particle tracking (RWPT), jet-integrated, or numerical modeling efforts depending on the region of interest (e.g, near, intermediate, or far field). Near-field mixing is defined as the immediate mixing upon release from the diffuser and its interaction with ambient ocean conditions. Turbulence generated from effluent buoyancy and momentum is the dominant process in near-field mixing and results in plume mixing, intrusion into the water surface, and lateral dispersion. The collapse of the initial turbulence generated by buoyancy forces specifies the near-field terminus (Carvalho et al. 2002). The intermediate field is described as the near-to-far-field transition in which the processes depend on near-field mixing and ambient flow that can affect ambient stratification, concentration, and gravity dispersion (Choi and Lee 2007). The intermediate field begins at the end of the jet momentum–dominated regime and ends when the wastefield is established (Zhao et al. 2011).
Near- and intermediate-field modeling traditionally has focused on analytical methods, empirical methods, or a mixed modeling strategy referred to as the jet integral method. Idealized plume and jet properties have been derived from dimensional analyses (Fischer et al. 1979) and augmented with empirical laboratory experiments deducing relative relationships for jet and plume characteristics such as velocity, concentration, and dilution as a function of initial parameters (e.g., buoyancy, concentration, flow rate, and port angle and diameter). These relationships have been extended to model environmental plumes (Carvalho et al. 2002; Frick et al. 2003; Roberts et al. 2011; Roberts and Villegas 2017). Laboratory observations are used to deduce empirical relationships between the dimensionless groups to scale to realistic, full-sized operations (Fischer et al. 1979; Zhao et al. 2011). Empirical formulas for the maximum height of buoyant plumes in stably stratified conditions were found by Morton et al. (1956). Laboratory experiments have investigated plume entrainment and terminal rise height, the elevation at which vertical momentum flux disappears, in pure jets, pure plumes, particle-laden plumes, and different ambient stratifications to form dimensional analyses (Konstantinidou and Papanicolaou 2003; Mirajkar et al. 2015; Mirajkar and Balasubramanian 2017). Empirical relationships of interactions of cross-flow, parallel flow, and port spacing on buoyant plumes also have been developed (Roberts et al. 1989a, b).
An alternative and popular approach is the jet integral method, which assumes a Gaussian jet profile based on the pioneering experimental work of Reichardt (1941). Morton et al. (1956) established the jet entrainment hypothesis, which states that ambient nonturbulent fluid is entrained into the edge of the turbulent jet zone with a mean velocity proportional to the mean centerline velocity (Morton et al. 1956; Jirka 2004). The fundamental equations of conservation of momentum, mass, buoyancy, and concentration form the basis of the jet integral method. These jet integral models may be Lagrangian, in which the independent variable is time, or Eulerian, in which the independent variable is distance. Ordinary differential equations are solved through integration along with the flow characteristics. The significant limitation of jet integral models is an assumption of an infinite receiving body of water that does not interact with boundaries (Bleninger and Jirka 2004). Examples of the jet integrated method are CORMIX (e.g., Doneker and Jirka 1991, 2001; Carvalho et al. 2002; Kang et al. 1999; Matos et al. 1998), VISJET (e.g., Lee and Chu 2012; Lee and Cheung 1990; Choi and Lee 2007), and Visual Plumes (e.g., Frick 1984; Roberts et al. 1989a; Baumgartner et al. 1994; Frick et al. 2003). The reader is referred to the citations for the most relevant software version.
Far-field mixing is defined as the drifting of the established wastefield, and is affected by ambient oceanic turbulence (Roberts 1991). In the established wastefield, advection and diffusion determine plume dilution. Advection is a bulk transport process dominated by the mean current, whereas diffusion is the combined effects of molecular diffusion, turbulence, and shear instabilities (Kim et al. 2000). Far-field mixing commonly is modeled using a Lagrangian approach for plume development and an Eulerian framework for flow resolution (Blumberg et al. 1996; Kim et al. 2001; Uchiyama et al. 2014). Tracer dispersion in buoyant plumes and jets (e.g., Kim et al. 2000, 2001, 2002; Israelsson et al. 2006; Roberts and Villegas 2017) often is represented using a RWPT in which the displacement of each particle is determined by an independent, random Markovian coefficient. RWPT presents key challenges: concentration field depends on particle density per grid cell and becomes less accurate with long simulations and increasing distance from source, and computing costs limit large particle number simulations (Zhao et al. 2011). Additionally, diffusion coefficients must be chosen empirically (Kim et al. 2002). Lagrangian-based RWPT methods are most applicable near point sources with high concentration gradients (Periáñez and Elliott 2002; Israelsson et al. 2006). At times, a far-field particle tracking module may be sufficient to approximate near-field mixing, as shown in the two-dimensional (2D) depth-averaged modeling in Roberts and Villegas (2017).
Eulerian numerical models directly solving the advection–diffusion equation using finite-difference, finite-element, or finite-volume methods are more appropriate in far-field modeling for long timescales and complex domains (Zhao et al. 2011). Far-field three-dimensional (3D) numerical models (e.g., Blumberg and Mellor 1987; Bouma et al. 2007) can determine plume characteristics and drifting away from pipe in an ocean environment, but lack the grid size resolution to resolve near- or intermediate-field mixing. Wastewater constituents and properties essentially are diluted within the intital grid volume (Zhang and Adams 1999). Parameterization of tracer input into a fixed grid cell volume causes uniform mixing of the tracer concentration, which can lead to overestimation of mixing. Hydrodynamic and oceanic models such as MIKE 21 (e.g., Gourbesville and Thomassin 2000; Tomicic et al. 2001; Pritchard et al. 2013), Delft3D (e.g., Morelissen et al. 2013; Roberts and Villegas 2017), and the Regional Ocean Modeling System (ROMS) (e.g., Uchiyama et al. 2014) have been used to model the far field.
Although some near-field models may include far-field modules (e.g., Doneker and Jirka 2001; Frick 2004), they are steady state and cannot account for spatiotemporal variability. Critically, near-field models do not account for many physical phenomena inherent to ocean outfall discharge and plume evolution, such as bottom topography, waves, tides, and the rotation of the Earth. Concomitantly resolving multiple (i.e., near, intermediate, and far) mixing scales is a fundamental challenge in wastewater plume modeling (Zhao et al. 2011). A high-resolution, nonhydrostatic approach tightly coupling intermediate and far-field dynamics has not been examined in the literature. The objective of this research was to develop and validate a hydrodynamic model capable of resolving both intermediate- and far-field mixing. A high-resolution (1  m), nonhydrostatic model, ROMS (Shchepetkin and McWilliams 2003, 2005; Guillaume et al. 2017) was developed, tested, and validated against a subset of the cross-flow Roberts, Snyder, and Baumgartner (RSB) experiments (Roberts et al. 1989a, b, c).

Methodology

RSB

Roberts, Snyder, and Baumgartner developed the RSB model, an empirical length scale–derived model based on towed tank experiments and applied to ocean outfalls (Roberts et al. 1989a, b, c; Frick et al. 2003). The experimental studies were done on multiport T-shaped diffusers in linearly density-stratified conditions using a line source. Complete experimental details and diagrams of their experiments and diffusers were given by Roberts et al. (1989a). The parameters examined in these studies were current speed, u, current direction, θ, port spacing, s, effluent density, ρ0, and horizontal jet velocity from either side of a T-shaped diffuser, uj. The relevant parameters for this study are presented. Discharge is characterized by the source flux per diffuser length, q (m2/s), momentum, m (m3/s2), and buoyancy flux, b (m3/s3)
q=QL
(1)
m=ujq
(2)
b=g(ρaρ0)ρaq
(3)
where Q = total discharge (m3/s); L = diffuser length(m); and ρa = ambient density at port level (kg/m3). The most important parameter controlling multiport near-field diffuser dynamics is given by the plume Froude number
F=u3b
(4)
and relates the current speed to the buoyancy flux of the source. A length scale, lb (m)
lb=b1/3N
(5)
is used to relate buoyancy flux to buoyancy frequency, or Brunt–Väisälä frequency, N (1/s) (Wright et al. 1982)
N=(gρadρdz)1/2
(6)
Parameter lb can be used to nondimensionalize the horizontal and vertical scales of different domains. Additionally
lm=mb2/3
(7)
relates the momentum to buoyancy. Empirical relationships then can be characterized as a dimensional analysis by four independent parameters
SmqNb2/3,zelb,helb,zmlb=f(lmlb,slb,F,θ)
(8)
where Sm = minimum dilution; ze = height of top of plume; he = thickness of plume; and zm = height to minimum dilution. These parameters are normalized to the buoyancy flux and lb.
The laboratory experiments of Roberts, Snyder, and Baumgartner were chosen for two primary reasons: (1) multiple far-field models have used the RSB model to initialize the established wastefield, including ROMS (Zhang and Adams 1999; Kim et al. 2001; Uchiyama et al. 2014), and (2) the EPA has adopted the RSB model (based on the RSB experiments) to meet wastewater treatment National Pollutant Discharge Elimination System (NPDES) dilution permits (Roberts et al. 2011). Additionally, Orange County Sanitation District (OCSD) relies upon the model to examine the impact of water conservation (i.e., reducing total effluent volume by extraction freshwater) on outfall plume characteristics.
RSB series 3 and 4 experiments were chosen as validation data for ROMS because of low jet momentum flux, strong dependence on Froude number, and genesis of RSB data fit curves. These series correspond to perpendicular (cross) flow experiments at Froude numbers 0, 0.1, 1, 10, and 100. The source is moved along the length of the tank at speeds corresponding to the appropriate Froude number to mimic cross-flow. Detailed experimental descriptions were given by Roberts et al. (1989a). The specific parameters in these experiments have lm/lb<0.2, the lowest momentum fluxes, and port spacing s=5  cm. These experiments are categorized as a line plume in which port spacing and jet momentum flux are negligible, and plume development is dominated by buoyancy; Sm, ze, he, and zm have the greatest dependence on Froude number compared with the other setups (i.e., parallel currents or oblique currents over linear diffuser), and therefore the perpendicular current experiments provide a rigorous test for the ROMS model. Furthermore, Series 3 and 4 line plume experimental results exclusively are shown in the RSB height characterization figures [Roberts et al. (1989a), Figs. 10–12] and are the basis of the RSB curves and equations.

Nonhydrostatic ROMS

ROMS is a terrain-following-coordinate, split-explicit time-stepping oceanic model that solves the hydrostatic, free-surface primitive equations in a rotating environment and uses a K-profile parameterization (KPP) for turbulence closure (Large et al. 1994) with Boussinesq approximations (Shchepetkin and McWilliams 2003, 2005). The model solves the three-dimensional momentum (Reynolds-averaged Navier–Stokes paradigm), continuity, and tracer equations. The modeling system uses the Arakawa C-grid (Arakawa and Lamb 1977) for model state variables (Haidvogel et al. 2008). Separate barotropic and baroclinic modes are calculated using the split-explicit time-stepping method (Shchepetkin and McWilliams 2003, 2005). To discretize advection, ROMS utilizes the third-order upstream, or upwind, biased advection scheme by a Uniformly Third-Order Polynomial Interpolation Algorithm (UTOPIA)-like algorithm (Shchepetkin and McWilliams 1998, 2005). Horizontal tracer advection is constructed through this finite-volume method (Haidvogel et al. 2008). This scheme accounts for eddy diffusivity by a numerical hyperdiffusion related to the horizontal advection with an effective diffusivity coefficient that decreases with the grid scale (Uchiyama et al. 2014). Vertical tracer advection is handled by a conservative parabolic spline (Haidvogel et al. 2008).
A nonhydrostatic discretization is applied to the model (Guillaume et al. 2017). No explicit eddy diffusivity parameterizations are used; therefore, KPP is turned off. The necessary small-scale viscosity and diffusivity is purely the result of the diffusive discretization error of the upwind third-order advection schemes that are not strictly monotonic (Shchepetkin and McWilliams 1998, 2005). Limitations to this advection scheme were discussed and revised by Marchesiello et al. (2009) and Lemarié et al. (2012). However, several of the issues these schemes specifically improved upon were not applicable to this study because of the idealized conditions (i.e., flat bottom, linearly stratified, and no coastal boundaries). ROMS has many applications, and a majority of simulations are large-scale O(10100  km) (Marchesiello et al. 2003; Penven et al. 2005). Recent ROMS simulations have moved from mesoscale to submesoscale (Gula et al. 2015; Molemaker et al. 2015; Dauhajre et al. 2017) and focused on local and coastal regions (Howard et al. 2014; Uchiyama et al. 2014).

High-Resolution Nonhydrostatic ROMS Validation Experiments

Validation was conducted by nondimensionalizing the ROMS parameters of cross-flow velocity, u, buoyancy frequency, N, and buoyancy flux, b, into the Froude number relationship F [Eq. (4)] and length scale lb [Eq. (5)]. Typical effluent temperature, salinity, and ammonium were chosen as 26.9°C, 1.2 practical salinity units (PSU), and 1,500 mmol N m3. A nonreactive passive tracer (Cp) was used to represent ammonium and measure dilution. Table 1 lists all ROMS simulations conducted.
Table 1. ROMS experimental parameters
ExperimentFroude No.Cross-flow velocity (ms1)Horizontal resolution (m)Effluent volume flux per grid cell (m3s1)
F00030.0033
F010.10.06730.0033
F110.14430.0033
F10100.3130.0033
F1001000.6730.0033
F01 1 m0.10.06710.00037
F01  3×3000.10.06730.011
F01  20×3000.10.06730.00167
F100 1 m1000.6710.00037
F100 10 m1000.67100.037
F100  5×3001000.6730.0067

Note: For all experiments, effluent temperature = 26.9°C; salinity = 1.2 PSU; buoyancy frequency N=0.011  s1; buoyancy flux b=0.003  m3s3; and depth of domain = 60 m. Total volume flux = 10  m3s1 for all simulations except 1-m-resolution experiments, in which volume flux = 5.56  m3s1. Bottom section of table denotes experiments discussed in the section “Quantifying Variability and Sensitivity.”

A linearly stratified, or stably stratified, vertical profile was implemented uniformly in ROMS. The RSB experiments discharged negatively buoyant effluent into linearly density-stratified water created by filling their tank with saltwater. All RSB experiments used a buoyancy flux between 9.8×105 and 1.1×104  m3s3, and buoyancy frequency of 0.3  s1. The linearly stratified density profile in ROMS was forced through both temperature and salinity profiles, and the buoyancy frequency was 0.011  s1 for all experiments. ROMS buoyancy flux was 0.003  m3s3, and lb was calculated from Eq. (5) as 13.09 m.
The flow speed for each experiment was calculated by selecting Froude numbers 0.1, 1, 10, and 100 and solving u=(Fb)1/3. The corresponding velocities were 0.067, 0.144, 0.31, and 0.67  ms1, respectively, assuming that the buoyancy is calculated from Eq. (3) with ρa=1,025.54  kgm3. Typical ocean shelf flow velocities are between 0.05 and 0.30  m·s1 (F=0.110).
ROMS experiments were run on a cluster in the UCLA Center for Earth Systems Research laboratory using 128–256 cores with message passing interface (MPI) parallelization. Each experiment was run for 11.75 h to give sufficient time for the region of interest to reach steady state. Time steps ranged between 0.3 and 7.5 s depending on cross-flow velocity. Total wall clock time for experiments was between 2 and 96 h. Output was saved every 15 min.
Results were normalized using the dimensional analyses presented by Roberts et al. (1989a, b); F=0 and 0.1 model runs were averaged over the last 1.25 h of the experiment, whereas F=1, 10, and 100 were averaged over the last 3.5 h for the diagnostics presented in the “Results” section. Lower Froude numbers require more time to reach a steady state in the region of interest. Model validation results also are averaged over the length of the pipe.

Model Configuration and Diffuser Representation

The ROMS domain was designed with realistic ocean length and height scales representative of wastewater pipe placement. Most experiments were computed using a 1,024×512×64 numerical grid. Unless specifically mentioned, the horizontal resolution used was 3 m in the horizontal directions and slightly less than 1 m in the vertical direction to resolve intermediate- and far-field dynamics. ROMS requires centimetric grid resolution to completely resolve immediate near-field mixing. The physical size of the computational domain was 3  km×1.5  km×60  m. Additional experiments at 1- and 10-m horizontal resolutions were conducted for certain Froude numbers. Open boundary conditions were chosen with a zero-gradient eastern boundary (du/dx=0) and 100-m-wide sponge layers on all boundaries to reduce reflections of internal waves at the open boundaries. The magnitude of the mixing parameter in the sponge layer was 0.1  m2s1. Drag at the oceanic bottom was turned off to allow a uniform current from the ocean surface to bottom. A volume flux of nearly fresh water was implemented at the bottom of the vertical domain to mimic a bottom-mounted pipe. Because the objective was to compare directly with RSB laboratory experiments in which Earth’s rotation played no role, we set the Coriolis force equal to zero.
A constant effluent volume flux of Qp=10  m3s1 was forced uniformly over an area 30 m wide and 900 m long in the 3- and 10-m-resolution experiments. Additional simulations with different forcing widths also were performed (Table 1). An ideal source width would parameterize initial momentum–dominated mixing by injecting effluent over multiple grid cells. The additional grid cells used to force the effluent discharge take advantage of the uniform mixing at the input to reach a dilution that is similar to a dilution after jet momentum–dominated near field mixing. A random input of volume flux on the order of 106  m3s1 was applied to each effluent input grid cell in all experiments to induce low-level noise that facilitates the occurrence of natural instabilities in the plume which otherwise could be repressed for an unrealistic period. The resolution and present configuration of the ROMS experiments were unable to represent the horizontal momentum of the diffuser ports. The 900-m-long pipe was chosen to represent a typical length of the diffuser portion of a marine outfall. Source flux calculated from Eq. (1) was q=Qp/L=1/90. The 1-m-resolution experiments had shorter pipe length because the maximum length of the pipe with the 1,024×512 grid was 512 m. A 500-m-long pipe was chosen to prevent unwanted boundary interactions, and the volume flux was reduced proportionally to 5.56  m3s1. The sponge layer width was reduced in the 1-m-resolution simulations to 33 m at F=100 and to 5 m at F=0.1. The 10-m simulation was run on a smaller grid of   256×128   to capture the same area as the 1-m resolution for comparison and to decrease computing time. Vertical resolution was kept at 64 grid points in a 60-m vertical domain regardless of horizontal resolution.
For all cross-flow experiments, the tracer input was placed one-quarter of the distance downstream into the domain (i.e., 256 grid cells). In the case of the zero cross-flow experiment (F=0) the tracer is placed in the center of domain to to allow western and eastern plume spread. Fig. 1 shows a schematic of tracer input placement in the domain for zero cross-flow and cross-flow experiments.
Fig. 1. Schematic of tracer input placement denoted by thick black line in domain for (a) F=0; and (b) 0.1F100. Arrows indicate cross-flow current direction. The x- and y-axes are in grid points. All tracer input is introduced at the bottom grid cell of the domain.
Effluent source forcing was parameterized and mixed uniformly within the input grid cells. Incoming temperature, salinity, and passive tracer concentration are subject to grid size parameterization and become uniformly mixed within the grid cells that have effluent input. The immediate turbulent mixing that occurs when the effluent first exits the pipe, therefore, was not resolved here. Roberts et al. (1989a) noted that source momentum flux and port spacing do not significantly affect normalized dilution across all experiments and parameters tested. The individual plumes from each port merged to approximate a line source. Increased momentum flux caused decreased plume rise height, although dilution was nearly constant.
Effluent tracer source forcing was shaped by diffuser geometry and was dependent on resolution. Details of the implementation were presented by Uchiyama et al. (2014). Relevant equations are the nondimensional tracer concentration equation with equivalent source P (1/s)
ct=·F+P
(9)
with
P(x,y,z,t)=Ps(t)A(x,y)H(z)
(10)
where c = pollutant concentration normalized by input pollutant concentration Cp; and F advection-mixing flux associated with resolved flow and upwind advection scheme mentioned previously in the ROMS description. Pollutant species can be represented by multiplying c fields by inflow concentration Cp. In Eq. (10), A and H are the spatial functions mimicking unresolved near field mixing above the diffusers, and have integrals equal to the source area and depth
Adxdy=As
(11)
and
Hdz=Hs
(12)
where As = horizontal area of diffuser; Hs = vertical size; and Vs=AsHs = volume. In Eq. (10), Ps is determined by effluent volume flux and model source and is equal to Qp/Vs, where Qp is volume flux (m3/s). All tracers were introduced to the domain at the bottom grid cell, i.e., Hs=H(z)=1 at z=60  m. No shape function was fitted to H in order to mimic bottom-mounted pipes. Parameter A=1 in the horizontal grid cells of the diffuser pipe that tile the bottom of the domain to uniformly force effluent volume flux over pipe area. Parameter As=Nsdx2, where Ns is the number of cells that make up the diffuser, where Ns=3,000 for dx=3  m, Ns=15,000 for dx=1  m, and Ns=270 for dx=10  m. The F=0.1 experiments with different pipe widths at 3-m resolution had Ns=900 and 6,000, and the F=100 simulation with half the source width had Ns=1,500.

Results

Flow Regime

Roberts et al. (1989a) observed two different flow regimes for buoyant plumes: forced entrainment and free plume. Clear forced entrainment was observed in simulations F=10 and 100, and free plumes were observed in F=0 and 0.1. Additionally, in the F=1 experiment, Roberts et al. (1989a) observed forced entrainment, free plume, and internal wave features. Model results were consistent with these observations.
Cross-pipe wastefields beginning at the first horizontal cell with tracer input with various Froude numbers and concentration gradients are shown in Fig. 2. The forced entrainment flow regime [Fig. 2(d)] was observed when current speed was high (F10) and the bottom of the wastefield remained at diffuser level. A buoyant plume flow pattern entraining oncoming flow could not be maintained and there was efficient mixing near the source (Cederwall 1971). These plumes exhibited evolving concentration fields and transient filaments with high concentration.
Fig. 2. Instantaneous concentration fields of ROMS experiments for (a) F = 0; (b) F = 0.1; (c) F = 1; (d) F = 10; and (e) F = 100 after 11.5 h from the center of pipe to the eastern boundary. Only the right side of the domain of the F = 0 plume is shown, with a smaller domain. Markers in Figs. 2(b and c) are examined in the section “Discussion.”
The F=0 simulation had the pipe in the middle of the domain and thus had less downstream distance from the pipe than the other experiments. The F=0 plume spread westward and eastward, but only one side of the plume is depicted. Stable stratification suppressed the vertical motion after sufficient dilution and prevents tracer from reaching the surface. At F=0.1, no lateral spreading upstream occurred and all effluent was advected downstream. As expected, thickness and plume rise height decreased with increasing F. The free plume pattern [Fig. 2(b)] occurred during low current speed and had normal plumelike characteristics, with the plume curved downstream and entrainment of the ambient flow. The F=1 simulation accurately exhibited both forced (<750  m) and free (>750  m) characteristics [Fig. 2(c)]. Froude number 1 had a strong signal of an internal wave between fluid layers called a lee wave [Fig. 2(c)]. Lee waves are generated from stably stratified flow over an obstacle, with the obstacle here consisting of the constant intrusion of a buoyant plume, and oscillate at the Brunt–Väisälä frequency (Marshall and Plumb 2008). Roberts et al. (1989a) also noted a pronounced internal wave in their F=1 experiments.
Tracer concentration decreased as current speed increased. At F=0, continuous entrainment of effluent discharge above the pipe caused lateral spreading of the plume, and some mixing occurred, but the core of the plume was relatively uniform in concentration; F=0.1 and 1 had lower concentrations from ambient flow forcing. Filaments of high concentration tracer moved through the plumes. The forced entrainment flow regime at F>1 had incidental holes with little to no effluent concentration. Portions of the plume with no tracer adjacent to high concentrations, most notably in Fig. 2(e) at x=500  m, were the result of an overshoot artifact in the presence of sharp gradients in concentration. Although these overshoots led to minor visual artifacts in which passive tracer concentrations temporarily were slightly negative, the overall implications were negligible. The tracer advection schemes were strictly conservative as a result of the control volume approach used (Shchepetkin and McWilliams 1998).
As Froude number increased, some distance was required before the plume was able to lift off the bottom of the domain, which was not observed in the RSB experiments. This distance is most noticeable in Fig. 2(e), where effluent is carried approximately 300 m downstream in the bottom cell before observable entrainment occurs. Consequently, nearly all of the tracer, and therefore the minimum dilution Sm, was trapped at the bottom of the domain, and distance for the plume to rise and develop was overestimated considerably. The plume was expected to rise off the bottom cell closer to the diffuser, which was realized in the F=100 1-m-resolution simulation.

Plume Characteristics

Table 2 lists quantitative comparisons of ROMS experiments to RSB validation data. Generally, minimum dilution was well predicted, whereas plume rise height, height to plume top, and consequently plume thickness were more variable. Top of plume height percentage divergence converged to approximately 20% for F1. The largest deviation between ROMS and RSB experiments was 42%, and the smallest was less than 1%.
Table 2. Percentage difference between ROMS simulation data and closest RSB series 3 and 4 experiments
ExperimentFroude No.Minimum dilution, SmqN/b2/3 (%)Height to minimum dilution, zm/lb (%)Height to top of plume, ze/lb (%)Plume thickness, he/lb (%)
F001.5420.4329.5218.71
F010.115.4613.862.6935.58
F111.4415.6119.414.70
F10100.270.2219.7417.56
F10010012.1342.4821.6222.28

Note: Negative percentages indicate underprediction, and positive percentages indicate overprediction.

Fig. 3 shows ROMS nondimensional dilution versus length results (solid line), plotted with RSB experimental data (symbols) for various Froude numbers. Roberts et al. (1989b) defined the established wastefield as the end of the initial mixing region xi, which is the distance to the point at which the limiting value of dilution is obtained. The dashed line in Fig. 3 signifies the location of established wastefield of the RSB laboratory experiments given in Roberts et al. (1989b) as a function of F
xilb=8.5F1/3
(13)
Fig. 3. Minimum dilutions for each Froude number at normalized distances from the pipe. Symbols are RSB experiments, and dashed line is Eq. (13) representing location of established wastefield in RSB experiments. ROMS dash-dot-dot-dash line intersects points of established wastefield in ROMS experiments. All other lines are ROMS experiments at different Froude numbers.
Fig. 3 is used to determine the end of the initial mixing region. The established wastefield locations for F=0,0.1,1,10, and 100 were x/lb=12, 12, 15, 32, and 82, respectively. These locations were used to represent the established wastefield for this metric and other metrics, zm,ze, and he, and are represented by the dash-dot-dot-dash line. Dilution increased as the effluent was advected in the down-pipe direction and was mixed by ambient forcing until a nearly constant dilution was reached. Generally, the minimum dilution was well modeled. Good agreement was observed at low Froude number flows (10). However, at high velocities (F=100), the high-resolution ROMS underpredicted dilution at 3-m resolution. The 1-m resolution improved on this underprediction. Lower cross-flow experiments (Froude numbers 0.1 and 1) still were developing at the intersection with the RSB established wastefield estimate (dashed line). Dilution initially was overestimated for F=0.1, 1, and 10, and then converged to a steady state in the down-pipe direction. Notably, some experimental data (i.e., F=1) suggested that this higher dilution near the pipe may be occurring. It is unclear if higher dilution is physical or an artifact of the parameterized effluent forcing or limited resolution. For Froude numbers 10 and 100, the distance required before the plume became established was longer and did not intersect the RSB established wastefield estimate. Contrary to Roberts et al. (1989a), dilution was affected by current, and dilution at F=0.1 was not equal to the dilution at F=0.
Fig. 4 shows normalized minimum dilution in the established wastefield as a function of Froude number. Minimum dilution values were taken at the x/lb values listed previously. Lower dilution was observed for F=100 and higher dilution was observed for low F than that of the experimental data. The scale of the x-axis of F on all graphs is linear from 0 to 102 and logarithmic above 102. Symbols represent RSB experimental data for minimum dilution at the established wastefield (Roberts et al. 1989c). The dashed curve in Fig. 4 plots dilution data fit as a function of Froude number presented by Roberts et al. (1989a)
SmqNb2/3=0.97,F0.1
(14)
and
SmqNb2/3=2.19F1/60.52,0.1<F100
(15)
Fig. 4. Normalized minimum dilution in the established wastefield for each Froude number. Solid line is ROMS, and dotted is RSB data fit curve. Symbols are RSB experimental points. The scale of the x-axis of F on all graphs is linear from 0 to 102, and logarithmic above 102.
Generally, model results and experimental data were in good agreement. Minimum dilutions for Froude numbers 1 and 10 coincided with experimental results. The RSB data fit curve [Eq. (15)] overestimated the F=10 data points. Model results indicated lower dilution with high cross-flow (F=100) and slightly higher minimum plume dilution at lower Froude number (0.1).
Vertical plume characteristics of rise height to minimum dilution, zm, top of the plume, ze, and thickness, he, at the end of the initial mixing region were found for all experiments. Roberts et al. (1989a) used 5% of the maximum concentration to delineate the plume top (ze) and thickness (he). The same method was employed here.
Fig. 5 shows the normalized height of minimum dilution (or maximum concentration) in the established wastefield. The RSB empirical equations plotted are
zmlb=1.7,F=0
(16)
and
zmlb=1.5F1/6,1F100
(17)
[Roberts et al. (1989a), Eqs. (16c) and (18), respectively]. Similar to minimum dilution, ROMS overpredicted the rise height of minimum dilution at low Froude numbers and underpredicted it at high Froude numbers. Minimum dilution height at the end of the intial mixing region changed by 6% between F=0 and 0.1. Between F=0 and 0.1 in the RSB experiments, the change was 11% and 15% for the two data points. The height of the Froude number 0 case between ROMS and RSB was overestimated by 20% and 25%. Although minimum dilutions were well predicted for F=1 (Figs. 3 and 4), the height to the minimum dilution was underestimated by 15% for the lowest data point. In addition, F=10 closely matched experimental values, whereas F=100 deviated significantly from RSB experimental results (Table 2).
Fig. 5. Normalized height to minimum dilution in the established wastefield for each Froude number.
Normalized height to top of the plume is shown in Fig. 6, and normalized plume thickness is shown in Fig. 7. RSB estimates [Roberts et al. (1989a), Eq. (16) and (17)] for these characteristics are given by
zelb=2.6;helb=1.8,F=0
(18)
and
zelb,helb=2.5F1/6,1F100
(19)
and are shown in Figs. 6 and 7, respectively. ROMS simulation height to top of the plume (ze) was consistent with experimental data for F=0.1. For higher cross-flows (F>0.1), ROMS deviated by 20% or more, whereas overprediction loccurred for no cross-flow (Table 2).
Fig. 6. Height to top of the established wastefield, defined as 5% of maximum concentration, for each Froude number.
Fig. 7. Thickness of the established wastefield, defined as 5% of maximum concentration, for each Froude number.
RSB experimental results for plume thickness, he, were dependent on the Froude number as he decreased at F10. Except for F=0.1, the ROMS model results for the plume thickness were consistent with the RSB experimental results in their dependence on he. ROMS underpredicted he for F10 compared with the RSB experimental results, which likely was a result of resolution; there was limited convergence with the F=100 1-m experiment studied in the next section.
Despite good agreement of ze for F=0.1, the difference between experimental and model he was significant (Table 2), indicating that the bottom of the plume was represented poorly. Fig. 2(b) shows a thick plume of about 35 m between 100 and 700 m from the pipe, and decreasing plume thickness farther from the pipe. The velocity fields indicate a current reversal under the F=0.1 plume with converging flow from the northern and southern open boundaries of the domain. The reasons for this current reversal and the likely effect on the plume dilution and thickness are explored in the “Discussion” section.
In ROMS, F=1 underpredicted the top and minimum dilution of the plume but was close to the laboratory experiments for plume thickness. This indicates that zm and ze were underestimated by similar heights and gave nearly the expected plume thickness. The ze and he were virtually the same as each other for F=10 and 100 due to the forced entrainment plume regime.

Quantifying Variability and Sensitivity

Additional simulations and analyses considered the differences between physical experiments and model results along with model resolution and parameter sensitivities. These investigations included quantifying top of the plume height variability, the impacts of effluent source pipe resolution, and grid resolution. Figs. 810 summarize the supplementary simulations and analyses for the parameters top of the plume height, minimum dilution, and dilution as a function of distance from the pipe, respectively.
Fig. 8. (a) Height to top of the established wastefield for 5% (solid line) and 1% (dashed-dotted line) of maximum concentration for each Froude number with error bars and shading at two standard deviations, ze of additional experiments of effluent input width and resolution are plotted as markers, and RSB line denoting data fit curve; (b) Fig. 8(a) zoomed-in to the F=0.1 experiments, with the F=0.1 1-m experiment marker on top of the F=0.13×300 marker; and (c) Fig. 8(a) zoomed-in to the F=100 experiments. The F=100 10-m marker is on top of the F=100  5×300 marker.
Fig. 9. Normalized minimum dilution in the established wastefield for each Froude number with additional source width and resolution experiments. Solid line is ROMS, and dotted is RSB data fit curve. Upright triangles and wide diamonds are RSB experimental points. Other symbols are additional experiments. The F=0.1  20×300 symbol is on top of the F=0.1  3×300 marker.
Fig. 10. As Fig. 3, with F=0.1 and 100 experiments and additional source width and resolution experiments. Symbols are RSB experiments, and dashed line is Eq. (13) representing location of established wastefield in RSB experiments. Dash-dot-dot-dash line intersects points of established wastefield in ROMS experiments with updated value for F=100 at 1 m resolution.
The top of the plume, ze, is of particular interest. This parameter determines whether the plume reaches a critical height in the water column, such as the mixed layer or euphotic zone. Model results were averaged over both over time and along pipe. Additionally, ze and he were based on 5% of the maximum concentration at the established wastefield as calculated by Roberts et al. (1989a). The variability of ze for all Froude numbers was calculated across time and along the pipe to determine whether the experimental values were within this range. The standard deviations were calculated from all height values of ze along the length of the pipe at the established wastefield for five saved outputs, equaling 1.25 h; e.g., the heights to the top of the plume were found at the distance from the pipe that was the established wastefield for every grid cell along the pipe i.e., 300 cells, for 5 saved outputs, for a total of 1,500 ze values used to calculate standard deviations. The ze at the 1% maximum concentration level also was calculated.
Fig. 8 shows two standard deviations of ze over time and distance along the pipe at 1% and 5% of the maximum concentration at the established wastefield. The ze RSB experimental values of F0.1 were within the variability across time and space of the ROMS experiments at the 5% line. The 1% line has a higher ze across all F numbers and further increases the range of variability. The no-cross-flow (F=0) ze value was greater than the RSB experimental ze values. The ROMS experiment was run for multiple hours, and the continuous buoyant influx likely caused higher plume rise by erosion of the background stratification. This zero-current-speed scenario is unlikely to occur in realistic environments. Calculating ze as 1% of the minimum dilution may be more appropriate because any amount of effluent reaching higher in the water column could affect ocean chemistry and biology in that zone.
Various effluent source widths were tested. Effluent source widths at F=0.1 (a typical flow velocity) were tested at 9 and 60 m (3 and 20 grid points, respectively) while input pipe length was held constant at 300 grid points (denoted 3×300 and 20×300). Additionally, a 15-m-wide pipe (5×300) was tested at a high cross-flow velocity, F=100. The established wastefield locations of these F=0.1 and 100 experiments were chosen to be the same as those in the previously presented experiments, x/lb=12 and 82, respectively.
Roberts et al. (1989c) stated that changing jet momentum flux (uj) primarily affects rise height and thickness, whereas dilution remains similar. Changing source pipe width in ROMS changed initial input mixing and is analogous to changing uj. ROMS similarly showed that changing source pipe width primarily affected rise height, whereas dilution generally was insensitive to pipe width. Fig. 8 shows established wastefield ze with changing source widths. As expected, the 3 grid cell (9-m-wide) pipe had higher ze because of less initial mixing, and therefore greater buoyancy force.
Fig. 9 shows the minimum dilutions at the established wastefield for the source width experiments and the resolution experiments. At typical current speeds (F=0.1), the minimum dilution reached was insensitive to the increased resolution or source pipe width. Fig. 10 shows the minimum dilutions at down-pipe distances for the additional experiments. A prominent spike in dilution above the pipe occurred in both width extremes for F=0.1. The effluent was more diluted above both of these pipe widths, potentially for two different reasons. The 60-m-wide pipe had an initial input effluent spread in 2 times volume of that in the 30-m-wide pipe, and therefore had a spike in dilution closest to the pipe compared with the other experiments. The effluent from the 9-m-wide pipe was subject to greater mixing from the increased buoyancy force. Across all parameters, the 15-m-wide pipe at F=100 had no differences from the validation results. At high Froude numbers, the cross-flow dominated plume dilution and height characteristics regardless of pipe width.
The 3-m-resolution ROMS model agreed reasonably well with the RSB experiments. However, the potential impacts of grid resolution are explored with additional simulations. The effect of increasing the resolution to 1 m for F=0.1 and 100 was quantified with a 30-m-wide pipe. A coarser grid resolution of 10 m for F=100 also was considered. The 1- and 10-m locations of the established wastefield for F=100 are chosen to be at the end of their respective domains. The location of the established wastefield for F=0.1 at 1-m resolution was the same as that in the other F=0.1 experiments. Fig. 9 shows the minimum dilution for the F=1 and 100 experiments at different resolutions. The 1-m-resolution results had an overall higher dilution and top of the plume for both Froude numbers. Higher resolution substantially improved the model data comparison at the highest cross-flow velocity (F=100) because more mixing was resolved. The 1-m F=100 experiment (Fig. 10) increased dilution in less distance than the other F=100 experiments due to better resolved mixing near the effluent input. The 10-m resolution F=100 simulation began mixing the input effluent at a much greater distance from the pipe compared with the other F=100 experiments. The coarse resolution did not allow the plume to reach an established wastefield in the current domain. Despite meaningful differences in dilution for F=100 resolution experiments, the differences in plume top rise heights in Fig. 8 for these experiments were negligible. The 1-m F=0.1 simulation had a similar pattern as the other F=0.1 experiments in dilution as the plume moved away from the pipe (Fig. 10), but the initial overdilution amplitude near the pipe was less pronounced. The dilution remained relatively close to the value of that above the pipe.

Discussion

Generally, RSB uniform perpendicular flow experiments and modeling at 3-m resolution had excellent correspondence, particularly in the low– to mid–Froude number simulations, which were consistent with typical flow velocities at depths ranging between 0.05 and 0.30  m·s1, or F=0.1 and 10. Sensitivity to forcing width was low. The F=100  1-m resolution results suggested that resolution is the primary limitation to accurate plume characteristics at high Froude numbers. Meanwhile, a resolution of 10 m was insufficient for the intermediate field. Significant differences were evident in plume height and minimum dilution height, particularly for high Froude numbers. These significant differences can be explained in part by plume variability and the idealized conditions used.
Plumes are tremendously spatially and temporally variable, making exact quantitative plume characterization challenging in physical experiments in which sampling is limited by physical constraints and sensor placement. Plume characteristic estimates (i.e., rise height and thickness) may depend on sensor placement. Numerical modeling facilitates complete spatial analyses, and results showed high spatiotemporal variability and filamentous flow structures. Figs. 2(b and c) clearly illustrate this spatial variation. High concentrations of effluent are threaded in the plume near lower concentrations. Fig. 2(b) shows a region of high tracer concentration, marked o, near a region of low concentration at the same depth, marked x. The concentration at o is nearly 2 times the concentration at x. The height of the top of the plume also varied in time and space considerably. Fig. 2(c) shows a region of plume with a depth of 29 m (x) and another nearby region with a depth of 41 m (o). The difference between these two heights was more than one standard deviation of ze at F=1.
Tank experiments may differ widely from models and the ocean due to the finite volume and physical boundaries. Current reversal and recirculation occurred in the F=0.1 ROMS simulation that were not observed at F1 and may lead to the slightly increased dilution in Fig. 4 and thicker plume in Fig. 7. This reversal and recirculation may be caused by a lee wake eddy recirculation and/or converging flow near the bottom from upward velocities caused by buoyant effluent mixing with bottom water. Lee wake eddies occur when a flowing fluid comes in contact with some obstacle and recirculation occurs behind the obstacle. The obstacle of the plume causes turbulent recirculation behind it and may increase mixing (Marshall and Plumb 2008). Alternatively, the effluent input is mixed with ambient water and is advected upward. This water must be replaced, and ambient water is pulled from the (open) northern and southern boundaries of the domain. The open boundaries in the ROMS model allow for more accurate resolution of this current reversal, which may be present in the ocean with its infinite boundaries compared with a closed-boundary tank. This pulling of ambient water causes a current reversal near the tracer input, leading to additional vertical mixing that potentially increases dilution and extends the bottom of the plume, increasing the plume thickness.
The relationship between dilution and plume rise depends on where mixing occurs. Mixing primarily at the bottom in cold, dense water leads to lower plume height because the plume density is increased. If less mixing occurs at the bottom and more is mixed higher in the water column, the density becomes lower (i.e., more buoyant), and the plume rises higher. Parameterizing effluent forcing input in the bottom cell and higher in the water column may better represent jet momentum–dominated initial mixing.
The most important factor in the model’s error versus the data is the amount of mixing of a water parcel from the bottom to its level of equilibrium. The results are sensitive to this, and impact all metrics. The amount of mixing along the path of a parcel depends on resolution. Higher resolutions encourage mixing, and as a result, make the other characteristics more consistent with experiments (Figs. 9 and 10).
The RSB experiments utilized a linearly stratified water column that is inconsistent with typical oceanic conditions in which vertical density gradients are substantially lower. The strongest stratification characteristically occurs at the thermocline, where temperature decreases rapidly in relatively little depth. Elsewhere, density gradients are minimal, especially in winter. A realistic ocean vertical profile has limited density differences between the bottom and thermocline. Zhang and Adams (1999) similarly concluded that their far-field model performs better in nonlinear stratification than in linear stratification in terms of height predictions. The potential overestimation of mixing at the bottom in these ROMS simulations would have a limited effect in typical oceanic conditions compared with the tank experiments.
The ROMS model performs well, both qualitatively and quantitatively. No calibrations or adjustments were made to the model for each Froude number. The results are robust for several perturbations, such as the initial mixing volume. Grid convergence is achieved at the low Froude numbers when the physics of the intermediate- to far-field plume and cross-flow interaction are resolved. In high cross-flows, the 3-m grid resolution limits accuracy (e.g., Fig. 10).
ROMS is able to resolve many realistic oceanic conditions applied to plume development and fate. The idealized conditions presented here were primarily for validation, and are recognized to be well-predicted by analytical and integral-type models. Numerical 3D models allow more-complex oceanic processes, singularly or in a suite, to be applied in realistic domains to study their effects on plumes (e.g., varying bathymetry, vertically variable currents, Coriolis force, nonlinear stratification, and so forth).

Conclusions

A high-resolution nonhydrostatic version of ROMS with a marine outfall was developed and tested against laboratory experiments of varying cross-flows. These modeling experiments accurately predicted buoyant plume flow regime and reasonably predicted dilution metrics. When cross-flow velocity was high, simulations required additional down-pipe distance for accurate plume development. Height parameters generally were overpredicted for low F and underpredicted for high F. Despite lack of convergence of horizontal resolution for dilution and height parameters at high Froude numbers, reasonable success was achieved at Froude numbers in typical oceanic conditions. Assessing model response to variability, pipe width, and resolution elucidated parameter sensitivity and the differences between model and tank experiments.
The validation of the high-resolution nonhydrostatic ROMS model with wastewater effluent plumes is the first step to modeling realistic ocean conditions and scenarios. Typical near- and intermediate-field plume models rely on experimental data in idealized conditions (Fischer et al. 1979; Roberts et al. 1989a; Doneker and Jirka 2001). The modeling described here facilitates high-resolution near-pipe and far-field plume simulation critical to predicting plume impacts in the coastal ocean.
Current regulation for wastewater treatment plants relies on measurement data and empirical and jet integral models to meet permit requirements. Plume measurements have been difficult to capture because of their transient nature, and current models require many approximations without the full context of plume development and intermediate-field dynamics. Three-dimensional numerical models resolving multiple spatiotemporal scales can be informative and prognostic for wastewater management. For example, this model can be used to prescribe effluent properties and range of dilution for appropriate plume dispersion and rise height to minimize coastal impacts. Local wastewater discharge impact on chemistry and ecology can be assessed by development and integration of a biogeochemical module. Impact of effluent concentration increases (e.g., from water recycling efforts), seasonal ocean conditions, and intermittent effluent flow on dilution and rise of height can be assessed easily. Wastewater plume tracking from sufficient observation data may be possible. Critically, accurate simulation facilitates prescribed effluent density and volume flux to prevent deleterious plume effects.
Applications of this high-resolution nonhydrostatic ROMS model are abundant. Increasing resolution to a centimetric level facilitates accurate simulation of near field dynamics. Realistic domains with varying bathymetry, nonuniform flows, atmospheric forcing, tides, and Earth’s rotation can alter buoyant plume fate and transport. High-resolution, nested simulations allow the near, intermediate, and far fields to be well represented. Future work includes refined initial mixing parameterizations and implementation of realistic ocean topography and conditions, such as time- and space-varying currents.

Data Availability Statement

Some or all data, models, or code generated or used during the study are available in a repository or online in accordance with funder data retention policies. The underlying model code is available at https://github.com/ducousso/CROCO-NH. Laboratory tank validation data were taken from Roberts et al. (1989a, b, c). Some or all data, models, or code that support the findings of this study are available from the corresponding author upon reasonable request. Figure-generating code, wastewater pipe module implementation, and model simulations may be made available upon request.

Acknowledgments

This work is partially funded through the Southern California Coastal Water Research Project by the Orange County Sanitation District project “Outfall Plume Modeling in Support of Orange County Sanitation District Water Reclamation Study” (P.O. No. 106480 OA). The authors acknowledge the computing resources and technical support provided by the Center for Earth Systems Research at the University of California, Los Angeles, and thank George L. Robertson for data, Martha Sutula for guidance, and Daniele Bianchi for thoughtful comments.

References

Arakawa, A., and V. Lamb. 1977. “Computational design of the basic dynamical processes of the UCLA general circulation model.” In Vol. 17 of Methods of computational physics, 173–265. New York: Academic Press.
Baumgartner, D. J., W. Frick, and P. J. W. Roberts. 1994. Dilution models for effluent discharges. Washington, DC: USEPA.
Bleninger, T., and G. Jirka. 2004. “Near-and far-field model coupling methodology for wastewater discharges.” Environ. Hydraul. Sustainable Water Manage. 1: 447–453.
Blumberg, A. F., Z.-G. Ji, and C. K. Ziegler. 1996. “Modeling outfall plume behavior using far field circulation model.” J. Hydraul. Eng. 122 (11): 610–616. https://doi.org/10.1061/(ASCE)0733-9429(1996)122:11(610).
Blumberg, A. F., and G. L. Mellor. 1987. “A description of a three-dimensional coastal ocean circulation model.” In Vol. 4 of Three-dimensional coastal ocean models, 1–16. Washington, DC: American Geophysical Union. https://doi.org/10.1029/CO004p0001.
Bouma, T. J., L. A, van Duren, S. Temmerman, T. Claverie, A. Blanco-Garcia, T. Ysebaert, and P. M. J. Herman. 2007. “Spatial flow and sedimentation patterns within patches of epibenthic structures: Combining field, flume and modelling experiments.” Cont. Shelf Res. 27 (8): 1020–1045. https://doi.org/10.1016/j.csr.2005.12.019.
Carvalho, J. L. B., P. J. W. Roberts, and J. Roldão. 2002. “Field observations of Ipanema Beach outfall.” J. Hydraul. Eng. 128 (2): 151–160. https://doi.org/10.1061/(ASCE)0733-9429(2002)128:2(151).
Cederwall, K. 1971. Buoyant slot jets into stagnant or flowing environments. Pasadena, CA: W. M. Keck Laboratory of Hydraulics and Water Resources, California Institute of Technology.
Choi, K. W., and J. H. Lee. 2007. “Distributed entrainment sink approach for modeling mixing and transport in the intermediate field.” J. Hydraul. Eng. 133 (7): 804–815. https://doi.org/10.1061/(ASCE)0733-9429(2007)133:7(804).
Dauhajre, D. P., J. C. McWilliams, and Y. Uchiyama. 2017. “Submesoscale coherent structures on the continental shelf.” J. Phys. Oceanogr. 47 (12): 2949–2976. https://doi.org/10.1175/JPO-D-16-0270.1.
Doneker, R. L., and G. H. Jirka. 1991. “Expert systems for mixing-zone analysis and design of pollutant discharges.” J. Water Resour. Plann. Manage. 117 (6): 679–697. https://doi.org/10.1061/(ASCE)0733-9496(1991)117:6(679).
Doneker, R. L., and G. H. Jirka. 2001. “CORMIX-GI systems for mixing zone analysis of brine wastewater disposal.” Desalination 139 (1): 263–274. https://doi.org/10.1016/S0011-9164(01)00318-6.
Fischer, H. B., E. J. List, R. C. Y. Koh, J. Imberger, and N. H. Brooks. 1979. Mixing in inland and coastal waters. New York: Academic Press.
Frick, W., P. Roberts, L. Davis, J. Keyes, D. Baumgartner, and K. George. 2003. Dilution models for effluent discharges. Athens, Georgia: USEPA.
Frick, W. E. 1984. “Non-empirical closure of the plume equations.” Atmos. Environ. 18 (4): 653–662. https://doi.org/10.1016/0004-6981(84)90252-X.
Frick, W. E. 2004. “Visual Plumes mixing zone modeling software.” Environ. Modell. Software 19 (7): 645–654. https://doi.org/10.1016/j.envsoft.2003.08.018.
Gourbesville, P., and B. A. Thomassin. 2000. “Coastal environment assessment procedure for sustainable wastewater management in tropical islands: The Mayotte example.” Ocean Coastal Manage. 43 (12): 997–1014. https://doi.org/10.1016/S0964-5691(00)00072-7.
Guillaume, R., J. M. Molemaker, D. Nicolas, and D. Thomas. 2017. “Compact symmetric poisson equation discretization for non-hydrostatic sigma coordinates ocean model.” Ocean Modell. 118 (Oct): 107–117. https://doi.org/10.1016/j.ocemod.2017.09.001.
Gula, J., M. J. Molemaker, and J. C. McWilliams. 2015. “Gulf stream dynamics along the southeastern US seaboard.” J. Phys. Oceanogr. 45 (3): 690–715. https://doi.org/10.1175/JPO-D-14-0154.1.
Haidvogel, D., et al. 2008. “Ocean forecasting in terrain-following coordinates: Formulation and skill assessment of the Regional Ocean Modeling System.” J. Comput. Phys. 227 (7): 3595–3624. https://doi.org/10.1016/j.jcp.2007.06.016.
Howard, M. D. A., M. Sutula, D. A. Caron, Y. Chao, J. D. Farrara, H. Frenzel, B. Jones, G. Robertson, K. McLaughlin, and A. Sengupta. 2014. “Anthropogenic nutrient sources rival natural sources on small scales in the coastal waters of the Southern California Bight.” Limnol. Oceanogr. 59 (1): 285–297. https://doi.org/10.4319/lo.2014.59.1.0285.
Israelsson, P. H., Y. D. Kim, and E. E. Adams. 2006. “A comparison of three Lagrangian approaches for extending near field mixing calculations.” Environ. Modell. Software 21 (12): 1631–1649. https://doi.org/10.1016/j.envsoft.2005.07.008.
Jirka, G. 2004. “Integral model for turbulent buoyant jets in unbounded stratified flows. Part I: Single round jet.” Environ. Fluid Mech. 4 (1): 1–56. https://doi.org/10.1023/A:1025583110842.
Kang, S.-W., B.-C. Oh, K.-S. Park, and S.-H. You. 1999. “Near-field dilution of wastewater effluents discharged from submerged ocean outfalls in Masan Bay.” KSCE J. Civ. Eng. 3 (4): 395–405. https://doi.org/10.1007/BF02830475.
Kim, Y. D., I. W. Seo, S. W. Kang, and B. C. Oh. 2001. “Modeling the mixing of wastewater effluent discharged from ocean outfalls using a hybrid model.” Coastal Eng. J. 43 (4): 259–288. https://doi.org/10.1142/S0578563401000372.
Kim, Y. D., I. W. Seo, S. W. Kang, and B. C. Oh. 2002. “Jet integral–particle tracking hybrid model for single buoyant jets.” J. Hydraul. Eng. 128 (8): 753–760. https://doi.org/10.1061/(ASCE)0733-9429(2002)128:8(753).
Kim, Y.-D., S.-W. Kang, I.-W. Seo, and B.-C. Oh. 2000. “Far-field transport of effluent plumes discharged from Masan Sea outfalls.” Ocean Polar Res. 22 (2): 69–80.
Konstantinidou, K., and P. N. Papanicolaou. 2003. “Vertical round and orthogonal buoyant jets in a linear density-stratified fluid.” In Proc., 30th IAHR Congress on Water Engineering and Research in a Learning Society: Modern Developments and Traditional Concepts, 293–300. Beijing: International Association for Hydro-Environment Engineering and Research.
Large, W. G., J. C. McWilliams, and S. C. Doney. 1994. “Oceanic vertical mixing: A review and a model with a nonlocal boundary layer parameterization.” Rev. Geophys. 32 (4): 363–403. https://doi.org/10.1029/94RG01872.
Lee, J. H. W., and V. Cheung. 1990. “Generalized Lagrangian model for buoyant jets in current.” J. Environ. Eng. 116 (6): 1085–1106. https://doi.org/10.1061/(ASCE)0733-9372(1990)116:6(1085).
Lee, J. H. W., and V. Chu. 2012. Turbulent jets and plumes: A Lagrangian approach. New York: Springer.
Lemarié, F., J. Kurian, A. F. Shchepetkin, M. J. Molemaker, F. Colas, and J. C. McWilliams. 2012. “Are there inescapable issues prohibiting the use of terrain-following coordinates in climate models?” Ocean Modell. 42: 57–79. https://doi.org/10.1016/j.ocemod.2011.11.007.
Marchesiello, P., L. Debreu, and X. Couvelard. 2009. “Spurious diapycnal mixing in terrain-following coordinate models: The problem and a solution.” Ocean Modell. 26 (3): 156–169. https://doi.org/10.1016/j.ocemod.2008.09.004.
Marchesiello, P., J. C. McWilliams, and A. Shchepetkin. 2003. “Equilibrium structure and dynamics of the California Current System.” J. Phys. Oceanogr. 33 (4): 753–783. https://doi.org/10.1175/1520-0485(2003)33%3C753:ESADOT%3E2.0.CO;2.
Marshall, J., and R. A. Plumb. 2008. Vol. 93 of Atmosphere, ocean and climate dynamics: An introductory text. Amsterdam, Netherlands: Elsevier.
Matos, J., A. Monteiro, P. Costa, R. Neves, A. Bettencourt, A. Frazao, and C. Santos. 1998. “Wastewater diffusion in the Estoril Coast: Theoretical calculations and field studies.” Water Sci. Technol. 38 (10): 337–344. https://doi.org/10.2166/wst.1998.0420.
Mirajkar, H. N., and S. Balasubramanian. 2017. “Effects of varying ambient stratification strengths on the dynamics of a turbulent buoyant plume.” J. Hydraul. Eng. 143 (7): 04017013. https://doi.org/10.1061/(ASCE)HY.1943-7900.0001297.
Mirajkar, H. N., S. Tirodkar, and S. Balasubramanian. 2015. “Experimental study on growth and spread of dispersed particle-laden plume in a linearly stratified environment.” Environ. Fluid Mech. 15 (6): 1241–1262. https://doi.org/10.1007/s10652-015-9412-5.
Molemaker, M. J., J. C. McWilliams, and W. K. Dewar. 2015. “Submesoscale instability and generation of mesoscale anticyclones near a separation of the California Undercurrent.” J. Phys. Oceanogr. 45 (3): 613–629. https://doi.org/10.1175/JPO-D-13-0225.1.
Morelissen, R., T. van der Kaaij, and T. Bleninger. 2013. “Dynamic coupling of near field and far field models for simulating effluent discharges.” Water Sci. Technol. 67 (10): 2210–2220. https://doi.org/10.2166/wst.2013.081.
Morton, B. R., G. I. Taylor, and J. S. Turner. 1956. “Turbulent gravitational convection from maintained and instantaneous sources.” Proc. R. Soc. London, Ser. A 234 (1196): 1–23. https://doi.org/10.1098/rspa.1956.0011.
Penven, P., V. Echevin, J. Pasapera, F. Colas, and J. Tam. 2005. “Average circulation, seasonal cycle, and mesoscale dynamics of the Peru Current System: A modeling approach.” J. Geophys. Res. Oceans 110 (C10): C10021. https://doi.org/10.1029/2005JC002945.
Periáñez, R., and A. Elliott. 2002. “A particle-tracking method for simulating the dispersion of non-conservative radionuclides in coastal waters.” J. Environ. Radioact. 58 (1): 13–33. https://doi.org/10.1016/S0265-931X(01)00028-5.
Pritchard, D., G. Savidge, and B. Elsäßer. 2013. “Coupled hydrodynamic and wastewater plume models of Belfast Lough, Northern Ireland: A predictive tool for future ecological studies.” Mar. Pollut. Bull. 77 (1–2): 290–299. https://doi.org/10.1016/j.marpolbul.2013.09.046.
Reichardt, H. 1941. “Ãœber eine neue Theorie der freien Turbulenz.” J. Appl. Math. Mech. 21 (5): 257–264. https://doi.org/10.1002/zamm.19410210501.
Roberts, P. J. W. 1991. “Ocean outfalls.” Crit. Rev. Environ. Control 20 (5–6): 311–339. https://doi.org/10.1080/10643389109388404.
Roberts, P. J. W., C. D. Hunt, M. J. Mickelson, and X. Tian. 2011. “Field and model studies of the Boston outfall.” J. Hydraul. Eng. 137 (11): 1415–1425. https://doi.org/10.1061/(ASCE)HY.1943-7900.0000445.
Roberts, P. J. W., W. H. Snyder, and D. J. Baumgartner. 1989a. “Ocean outfalls. I: Submerged wastefield formation.” J. Hydraul. Eng. 115 (1): 1–25. https://doi.org/10.1061/(ASCE)0733-9429(1989)115:1(1).
Roberts, P. J. W., W. H. Snyder, and D. J. Baumgartner. 1989b. “Ocean outfalls. II: Spatial evolution of submerged wastefield.” J. Hydraul. Eng. 115 (1): 26–48. https://doi.org/10.1061/(ASCE)0733-9429(1989)115:1(26).
Roberts, P. J. W., W. H. Snyder, and D. J. Baumgartner. 1989c. “Ocean outfalls. III: Effect of diffuser design on submerged wastefield.” J. Hydraul. Eng. 115 (1): 49–70. https://doi.org/10.1061/(ASCE)0733-9429(1989)115:1(49).
Roberts, P. J. W., and B. Villegas. 2017. “Modeling and design of the Buenos Aires outfalls.” J. Hydraul. Eng. 143 (2): 05016007. https://doi.org/10.1061/(ASCE)HY.1943-7900.0001244.
Shchepetkin, A. F., and J. C. McWilliams. 1998. “Quasi-monotone advection schemes based on explicit locally adaptive dissipation.” Mon. Weather Rev. 126 (6): 1541–1580. https://doi.org/10.1175/1520-0493(1998)126%3C1541:QMASBO%3E2.0.CO;2.
Shchepetkin, A. F., and J. C. McWilliams. 2003. “A method for computing horizontal pressure-gradient force in an oceanic model with a nonaligned vertical coordinate.” J. Geophys. Res. Oceans 108 (C3): 1–34. https://doi.org/10.1029/2001JC001047.
Shchepetkin, A. F., and J. C. McWilliams. 2005. “The regional oceanic modeling system (ROMS): A split-explicit, free-surface, topography-following-coordinate oceanic model.” Ocean Modell. 9 (4): 347–404. https://doi.org/10.1016/j.ocemod.2004.08.002.
Tomicic, B., A. Lützen, and O. Mark. 2001. Integrated modelling of the sewer system and the receiving waters for the Island of Ischia, 548–557. Reston, VA: ASCE.
Uchiyama, Y., E. Y. Idica, J. C. McWilliams, and K. D. Stolzenbach. 2014. “Wastewater effluent dispersal in Southern California Bays.” Cont. Shelf Res. 76 (Mar): 36–52. https://doi.org/10.1016/j.csr.2014.01.002.
Wright, S. J., D. R. Wong, K. E. Zimmerman, and R. B. Wallace. 1982. “Outfall diffuser behavior in stratified ambient fluid.” J. Hydraul. Div. 108 (4): 483–501. https://doi.org/10.1061/JYCEAJ.0005850.
Zhang, X.-Y., and E. E. Adams. 1999. “Prediction of near field plume characteristics using far field circulation model.” J. Hydraul. Eng. 125 (3): 233–241. https://doi.org/10.1061/(ASCE)0733-9429(1999)125:3(233).
Zhao, L., Z. Chen, and K. Lee. 2011. “Modelling the dispersion of wastewater discharges from offshore outfalls: A review.” Environ. Rev. 19 (NA): 107–120. https://doi.org/10.1139/a10-025.

Information & Authors

Information

Published In

Go to Journal of Hydraulic Engineering
Journal of Hydraulic Engineering
Volume 147Issue 8August 2021

History

Received: Mar 9, 2020
Accepted: Feb 25, 2021
Published online: Jun 11, 2021
Published in print: Aug 1, 2021
Discussion open until: Nov 11, 2021

Authors

Affiliations

Scientist, Dept. of Biogeochemistry, Southern California Coastal Water Research Project, Costa Mesa, CA 92626. ORCID: https://orcid.org/0000-0002-9027-3584. Email: [email protected]
Jeroen M. Molemaker [email protected]
Researcher, Dept. of Atmospheric and Oceanic Sciences, Univ. of California, Los Angeles, CA 90095. Email: [email protected]
Senior Scientist, Dept. of Biogeochemistry, Southern California Coastal Water Research Project, Costa Mesa, CA 92626. ORCID: https://orcid.org/0000-0002-8752-9200. Email: [email protected]
Professor, Dept. of Atmospheric and Oceanic Sciences, Univ. of California, Los Angeles, CA 90095. ORCID: https://orcid.org/0000-0002-1237-5008. Email: [email protected]
Assistant Professor, Dept. of Civil and Environmental Engineering, Univ. of California, Los Angeles, CA 90095 (corresponding author). ORCID: https://orcid.org/0000-0003-2905-1306. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

  • OpenFOAM-ROMS nested model for coastal flow and outfall assessment, Ocean Engineering, 10.1016/j.oceaneng.2022.112535, 264, (112535), (2022).
  • Configuration and Validation of an Oceanic Physical and Biogeochemical Model to Investigate Coastal Eutrophication in the Southern California Bight, Journal of Advances in Modeling Earth Systems, 10.1029/2020MS002296, 13, 12, (2021).

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share