Enhancement of turbulence through lateral spreading in a stratified-shear flow: Development and assessment of a conceptual model

Size: px
Start display at page:

Download "Enhancement of turbulence through lateral spreading in a stratified-shear flow: Development and assessment of a conceptual model"

Transcription

1 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117,, doi: /2011jc007484, 2012 Enhancement of turbulence through lateral spreading in a stratified-shear flow: Development and assessment of a conceptual model Daniel G. MacDonald 1 and Fei Chen 1,2 Received 1 August 2011; revised 19 March 2012; accepted 27 March 2012; published 16 May [1] A conceptual model for the impact of lateral spreading processes on local stratified-shear turbulence is presented and evaluated using field data from the rapidly spreading near-field region of the Merrimack River plume. The conceptual model addresses increases in turbulent kinetic energy associated with the stretching of a Kelvin-Helmholz billow along its rotational axis. A non-dimensional relationship is derived with two fit parameters, including a parameter representing the magnitude of mixing under the no spreading case, and a non-dimensional time scale. For the model to be viable, this time scale, representing the period of time that a Kelvin-Helmholz billow is subject to stretching should be of the same order as the evolution time scale for the billow. Observational data from field experiments are used to evaluate the two fit parameters, indicating a stretching time scale of the same order as evolution time scales reported in the literature. In addition, application of the model to the field data reduces observed variability. We conclude that spreading is an important mechanism capable of enhancing local stratified-shear turbulence. Citation: MacDonald, D. G., and F. Chen (2012), Enhancement of turbulence through lateral spreading in a stratified-shear flow: Development and assessment of a conceptual model, J. Geophys. Res., 117,, doi: /2011jc Introduction [2] Stratified-shear flows are ubiquitous in nature, providing the context for much of the important mixing in the natural environment, and particularly in the ocean environment. Our understanding of these flows has increased dramatically over the last several decades, and a variety of mechanisms for the generation of turbulence in these environments have been identified. Among these, Kelvin- Helmholtz [e.g., Thorpe, 1973; Gregg, 1987] instabilities are perhaps the most studied and significant mechanism for a majority of geophysical flows. Kelvin-Helmholtz (K-H) instabilities take the form of a series of rolled vortices separated by thin braids. As these coherent billows grow, they are eventually arrested by stratification, and subsequently break down to form turbulent patches. Thorpe [1971, 1973] performed the first laboratory visualizations of K-H billows, and defined much of the essential dynamics of the K-H mechanism, which is believed to be the primary 1 Department of Estuarine and Ocean Sciences, School for Marine Science and Technology, University of Massachusetts Dartmouth, Fairhaven, Massachusetts, USA. 2 Now at ExxonMobil Upstream Research Company, Houston, Texas, USA. Corresponding author: D. G. Macdonald, Department of Estuarine and Ocean Sciences, School for Marine Science and Technology, University of Massachusetts Dartmouth, Fairhaven, MA 02719, USA. (dmacdonald@umassd.edu) Copyright 2012 by the American Geophysical Union /12/2011JC mechanism of generation for most of the turbulence observed in the oceanic thermocline [Gregg, 1987]. Smyth et al. [2001] obtained results from numerical model and microstructure observations, lending strong support to the hypothesis that K-H billows represent a valid model for observed turbulent patches. [3] Shear-stratified flows are usually considered in a two dimensional context, defined by the vertical and mean flow directions. However, in many geophysical environments, such flows may also be impacted by lateral forcing mechanisms, resulting in a flow that is spreading laterally as it propagates forward in the direction of mean flow. The impacts of such lateral forcing on the dynamics of stratifiedshear turbulence has not been addressed, although there has been some speculation [MacDonald et al., 2007] that turbulent mixing may be enhanced in regions of significant spreading. Near-field river plumes represent one such region, where lateral spreading and mixing are arguably the two most important processes affecting plume evolution [Hetland, 2010]. [4] The spreading rate in a near-field plume is proportional to the local internal long wave speed (c i ¼ ffiffiffiffiffiffi g h, p where g = gdr/r is the local reduced gravity, Dr is the density difference between the plume and ambient waters, r is a reference density, and h represents the plume thickness) due to strong horizontal density gradients along the edges of the plume [Wright and Coleman, 1971; Hetland and MacDonald, 2008]. Hetland and MacDonald [2008] and Chen et al. [2009] also discuss the important role of mixing in modifying the plume density, resulting in reduced 1of12

2 internal wave speeds and a slowing of the spreading rate. Thus, it is clear that mixing directly impacts lateral spreading. An understanding of the impacts of lateral spreading on the local turbulent field would provide an understanding of processes acting in the opposite direction, and yield a much clearer understanding of the dynamics of developing stratified-shear layers. [5] The dynamics of stratified-shear turbulence has been the focus of significant research effort over the last several decades, particularly within the estuarine and ocean science communities. Early studies on stratified entrainment [e.g., Ellison and Turner, 1959] laid the groundwork for Richardson number (Ri) dependence, and the advancement of observational [e.g., Gregg, 1987] and numerical techniques [e.g., Smyth et al., 2001] have yielded details of the dynamics of stratified turbulence at small scales. Recently, acoustic and microstructure observations of a stratified shear layer in an estuary have provided insight into the selfsimilar organization of turbulent structures at high Reynolds numbers [Geyer et al., 2010]. Here, this significant base of knowledge in stratified-shear turbulence is extended, as we propose a mechanism for lateral spreading to enhance the available turbulent kinetic energy (TKE) through the stretching of developing K-H billows. [6] We begin with an assessment of the important variables in a stratified-shear flow, and the creation of appropriate non-dimensional parameters to represent the dynamics. Here we assume the flow exhibits values of ɛ/nn 2 sufficiently high to insure isotropic turbulence. Following MacDonald and Geyer [2004] and Imberger and Ivey [1991] the important variables in a two-layer stratified-shear environment, considered in the classic two-dimensional context are the velocity difference, Du, the reduced gravity, g, across the shear layer, the shear layer thickness, h i, and the mean TKE dissipation rate per unit mass within the layer, ɛ. Although other scaling approaches can be envisioned, the use of Du and g are straightforward, and easy to obtain from traditional oceanographic measurement techniques. [7] From these variables, two independent non-dimensional parameters can be created: a bulk Richardson number, Ri B = g h i /(Du) 2, and a mixing parameter, ɛ/g Du. The mixing parameter represents the efficiency with which mean flow energy is converted to turbulent energy, and can be Ri B 1 Ri f ɛ Dug, related to an interfacial drag coefficient, C Di ¼ ð Þ where Ri f is the flux Richardson number. Given only two independent dimensionless parameters, the mixing parameter must be a function solely of Ri B. MacDonald and Geyer [2004] suggested a constant value for this mixing parameter on the order of for values of Ri B on the order of [8] Following additional observations in the Merrimack River plume (Massachusetts), extending further into the near-field plume region, MacDonald et al. [2007] report that the skill of this scaling appears to decrease with increasing distance from the estuary mouth. In MacDonald et al. [2007], ɛ is surrendered as the turbulent variable in place of the buoyancy flux, B ¼ ðg=rþr w. Not only is B a more direct indicator of mixing, but it is also more readily calculated with the control volume technique employed [MacDonald and Geyer, 2004]. Assuming a constant flux Richardson number, Ri f = BP 1, where P is the shear production of TKE, and homogenous and stationary turbulence (P = B + ɛ), B and ɛ are directly proportional. They further define the mixing parameter as x ¼ B g Du. Here, we also use B instead of ɛ, and focus on the mixing parameter x. The results from MacDonald et al. [2007] as well as additional data from May 2007 are shown in Figure 1. The Merrimack River data sets include multiple estimates of B from a number of longitudinal passes through a highly energetic discharge plume. In general, B, Du, and g were all found to decay with distance from the river mouth. However, B was found to decrease at a faster rate than would be predicted by the assumption of a constant value of x, but little correlation of x with Ri B was observed. The MacDonald and Geyer [2004] scaling was derived from data in the Fraser River plume just seaward of the river mouth and the bottom attached salt front, where the outgoing river water lifts off from the bottom. This scaling was found to be most effective in a similar lift-off location with regards to the Merrimack River data set. MacDonald et al. [2007] speculated that the decrease of x with seaward distance may be related to lateral spreading, as spreading is maximal near the river mouth and decreases as the plume propagates seaward due to the integrated effect of mixing decreasing the density difference between the plume and ambient waters. [9] In this manuscript, we extend the scaling of MacDonald and Geyer [2004] to include the effects of lateral spreading, and develop an analytical model to explain the relationship between lateral spreading and turbulent mixing for a stratified-shear flow, testing the relationship with data collected from the near-field of the Merrimack River plume (MA). [10] In the next section we will describe the theoretical scaling followed by development of the analytical model in section 3. Section 4 describes the field experiments, with a discussion in section 5, and a summary in section Theoretical Scaling [11] The parameters in the upper (unshaded) portion of Figure 2 represent the scaling associated with the two-layer stratified-shear problem within the traditional two dimensional context, as discussed in section 1. Several reasons for the failure of this scaling through the evolution of the nearfield plume, as shown in Figure 1, can be postulated. One possibility is that the mixing parameter, x, decreases as a function of Ri B [e.g., Ellison and Turner, 1959; Christodoulou, 1986]. However, this is not supported by the calculated values of Ri B, as shown in Figure 3. Values of Ri B range from approximately 0.05 to 0.8. However, no clear relationship is observed between x and Ri B across this range. [12] Alternatively, it could be argued that the observed decrease in x with distance from the mouth could be driven by acceleration in the upper layer. Kelvin-Helmholz instabilities are driven by shear strong enough to overcome stratification and create a coherent overturning structure [e.g., Thorpe, 1987]. The overturn structure grows in size as it evolves within the sheared environment, before it eventually breaks down into three-dimensional turbulence. Acceleration in the upper layer resulting in a more strongly sheared environment at the moment of initiation of a K-H billow would be accounted for in the local value of Ri B,as discussed above, which does not appear to be correlated with changes in x. However, significant acceleration in the 2of12

3 Figure 1. Plot of x = B/(g Du) divided by C = [e.g., MacDonald and Geyer, 2004]. This plot shows a similar trend to Figure 11 from MacDonald et al. [2007], with data from both May 2006 and May Error bars represent a subset of potential errors in the control volume method [MacDonald et al., 2007], particularly those errors which represent relative uncertainties between separate data points, and not errors expected to be correlated across the entire range of data. The bold solid line is a best fit (least squares) to the data, shown with 90% confidence limits (bold dashed lines). In general, the skill of the assumption x = C is shown to decrease with seaward distance in the plume. upper layer at time scales consistent with the evolution of an individual K-H billow, could conceivably affect the growth of the instability, resulting in a profile of x that is not correlated with mean values of Ri B. To test this hypothesis, we can estimate a scaling for the evolution of a K-H billow based on direct numerical simulations (DNS) performed by Smyth et al. [2001]. The time required for the evolution of their K-H billow from conception to decay (based on their Figure 1) is approximately 4000 s. Normalizing this value by the local shear present in the DNS model yields a non-dimensional time scale of approximately 140 (i.e., t = T KH ( u/ z) 140, where t and T KH represent the non-dimensional, and dimensional KH billow evolution timescales, respectively). The results of Thorpe [1973], as shown in his Figure 1, and reiterated in Thorpe [2005, equation (3.10)], suggest similarly that t , with the lower limit representing the disappearance of visually coherent billows and the upper limit representing the complete collapse of turbulence. Similarly, we can normalize a time scale for spatial acceleration in the Merrimack nearfield plume, ( u/ x) 1, by the shear, ( u/ z), yielding values ranging from 10 3 to 10 4 or higher. Thus, there is a clear disparity between the K-H evolution timescale and the Figure 2. turbulence. Summary of dimensional and independent non-dimensional variables for shear-stratified 3of12

4 Figure 3. Plot of x versus Ri B for data from 2006 and 2007 passes. Error bars similar to Figure 1. acceleration timescale, and it is unlikely that these two processes significantly interact. [13] We should remark here that we use the shear time scale, ( u/ z) 1 h i /Du, as a representative time scale for turbulent processes. Other logical turbulence time scales might include N 1 [Gregg, 1987] and Du/g [Thorpe, 1973; Koop and Browand, 1979], both of which are related as a function of Ri B (e.g., h i /Du= N 1 Ri 1/2 B =(Du/g )Ri B ). Thus, under the assumption of constant Ri B, all of these time scales are dynamically similar, varying only by a constant. The KH billow evolution time scales from Thorpe [1973, 2005] presented above, were originally reported as a function of Du/g, and have been transformed based on reported Ri B values (0.06 from Thorpe [1971, 2005] and from Thorpe [1973]). [14] Finally, we consider the direct impact of spreading on the developing K-H billows by developing a method to parameterize spreading and its effects. To accomplish this, we enhance the two-dimensional scaling approach of MacDonald and Geyer [2004] and Imberger and Ivey [1991] by the inclusion of an additional variable, d, to represent lateral spreading, as shown in the lower portion of Figure 2. The variable d = v/ y, where v and y are cross-plume velocity and distance, respectively, represents the lateral strain rate of the plume as proposed by Hetland and MacDonald [2008], who also showed empirically that d = DW W u r Dr, where W represents plume width, u r represents the radial velocity within the plume, and Dr represents an increment of radial (i.e., along plume) distance, across which the change in width, DW, is observed. This expression yields a time scale that can be interpreted as the time required for the plume to double in width. To non-dimensionalize this new spreading variable, we choose to normalize by our turbulent time scale, the reciprocal of the vertical shear (Du/h i ), yielding the spreading parameter, f, shown in Figure 2. [15] Thus, the five variables in Figure 2 yield three independent dimensionless parameters, denoted as x, f, and Ri B. The variable x represents the efficiency with which energy is extracted from the mean flow and converted into turbulent energy, and is similar to the ratio for which MacDonald and Geyer [2004] proposed a constant value of The MacDonald and Geyer [2004] value is based on the maximum value of ɛ in a given vertical profile. In the present definition of x, the vertical mean of B for a given profile should be used, as suggested in MacDonald et al. [2007]. They suggest that the mean value should scale as approximately 0.6 times the maximum value. Given this adjustment, and assuming a constant Ri f = 0.18, would yield a constant value of x based on the MacDonald and Geyer [2004] conclusions of approximately , which is used to scale the data in Figure 1. [16] The non-dimensional quantity, f, represents plume spreading, which can be interpreted as a ratio of the turbulent time scale to the lateral expansion time scale. Reanalysis of the MacDonald and Geyer [2004] data suggests an average value of f for their data of approximately [17] Within this new framework, it is implied that x = f(f, Ri B ), so that under constant Ri B, spreading alone controls the value of the mixing parameter. In the next section, a conceptual and analytical model is developed to define the form of the x = f(f, Ri B ) relationship. 3. Conceptual Model [18] Consider a turbulent field characterized by independent and isolated turbulent patches, presumably arising from K-H billows. The prevalence of such instabilities has been observed in similar environments [e.g., Geyer and Farmer, 1989; Tedford et al., 2009] and in direct numerical simulations [e.g., Smyth et al., 2001]. In addition, recent work based on the same data set used later in this analysis 4of12

5 Figure 4. Cartoon of spreading, spinning and squeezing of a K-H billow. b is width of the billow and l is diameter of the billow. compares mixing estimates derived from overturn scales indicative of K-H type instabilities [e.g., Thorpe, 1977], and control volume estimates of mixing, suggesting that the observed overturns are primarily responsible for the observed mixing (D. G. MacDonald et al., manuscript in preparation, 2012). [19] Such vortices may be represented as horizontal cylinders oriented in the cross stream direction (Figure 4), with a diameter, l, representing the outer length scale of the turbulence which should scale with the Ozmidov scale (L o ¼ ɛ 1 2 N 3 2 ), and a velocity scale, u, which is the characteristic turbulent velocity. The cylinders also have a length, b, in the cross-stream direction (note that b need not be the full width of the plume but it is likely that b l), that is increased at the same rate as the overall lateral expansion rate of the plume. This is a critical assumption, which is supported by the modeling study of Hetland and MacDonald [2008], where it was found that most important plume variables, including the spreading rate, showed very little variability in the cross-plume direction, within the core of the plume. We hypothesize here that such stretching of these K-H cylinders, particularly during the pre-turbulent roll-up stage, may have significant impacts on the diameter and velocity scales, thus impacting the energetics of the resulting turbulence. A similar mechanism, albeit with significant differences in scale and generation dynamics, was discussed conceptually by Farmer et al. [2002] to explain an increase in turbulence associated with boils observed in Haro Strait. [20] The rate of energy supply to the large-scale eddies, and thus the TKE dissipation rate, can be scaled [Taylor, 1932] as: ɛ u 3 l [21] In our model, the KH cylinders are stretched at the same rate as the plume when macro-scale horizontal expansion occurs. The volume of an individual cylinder, V, can be assumed constant, and expressed as, V = 0.25pl 2 b. Thus, l b 1 2 ð1þ ð2þ Equation (2) describes changes in the vertical length scale of K-H billows as the instability undergoes lateral stretching. The relationship in (2) does not rely on the absolute size of the instability but indicates only the nature of the proportionality. [22] In order to express (1) in terms of the spreading parameter, f, we apply conservation of angular momentum to the hypothetical cylinders. The angular momentum, ~H, is defined as ~H ¼ R ~r ~udm, where dm is an element of mass, and r is the radius of the vector from the center axis. Applying the appropriate scales associated with the cylindrical KH tubes, including a maximum radii of 0.5l, and the rotation speed at the edge equal to the turbulent velocity, u!, into the conservation of H, then we have ~H ¼ Z l=2 0 r 2u r ð2rpbrþdr ð3þ l As ~H is constant, the integral above leads to u b 1 l 3. With (2) and (1), we obtain ɛ b 2. Since b is increased at the same rate as the overall lateral expansion rate of the plume width, W, we can also argue that ɛ W 2. [23] Note that the validity of the assumption of conserved angular momentum could be compromised by the effect of a baroclinic gradient producing additional torque on the evolving billow. To assess the magnitude of this mechanism, a time scale can be generated by comparing the torque resulting from this mechanism (i.e., R h l=2 l=2 b R i 0 h rgzdz xdx) to the total angular momentum expressed by (3). Substituting reasonable values from the Merrimack plume region yields a time scale on the order of 30 min, an order of magnitude higher than billow evolution times predicted by the results of Smyth et al. [2001] and Thorpe [1973, 2005]. This time scale suggests that baroclinic torque is ineffective at producing significant changes in angular momentum in the near-field plume, and that our assumption of conservation is reasonable. [24] Consider a time period for which a K-H cylinder is subjected to the stretching processes caused by lateral spreading, which could be scaled as T KH = Dr/Du, where Dr represents the distance traveled by the billow during the stretching process, and Du u r assuming no velocity in the ambient ocean water below the shear layer. Here, we assume that a billow may be subject to stretching processes throughout the duration of its evolution, and thus use the evolution time scale, T KH, as the stretching time scale. The change in plume width over this time period can be expressed as DW =(dt KH )W o, where W o is the initial width. Applying the spreading parameter f yields DW W o ¼ fdu h i T KH ð4þ Assuming a constant flux Richardson number throughout the evolution of the billow, we can then derive an expression for changes in x due to lateral spreading: x x o ¼ ɛ ɛ o W 2 ¼ 1 þ DW W o W o 2 ¼ ð1 þ tfþ 2 ð5þ 5of12

6 which can be rewritten as ¼ Dr h i x ¼ x o ð1 þ tfþ 2 ð6þ where t ¼ DuT KH h i is a non-dimensional time scale of K-H billow evolution as discussed in section 2, and x o is an initial, or non-spreading, value of x. [25] Note that we have assumed constant values of g and Du, reflecting the fact that the model follows a single roll-up or instability, created under specific conditions at a single point in time. Equation (6) then describes the evolution of that roll-up due to the nature of its laterally spreading environment. [26] A key aspect of (6) is the two fit parameters, x o, and t, both of which have a distinct physical meaning. The value of x o in (6) represents the initial value of x, or the value of x when there is no lateral spreading, (i.e., f = 0). This can be interpreted as representing the turbulent potential of the rollup at the moment of its conception, which is then further modified through the stretching effects of lateral spreading (i.e., the (1 + tf) 2 term) during the course of its evolution. Note that the lateral stretching affects all billows similarly regardless of initial size, so that (6) can be used to characterize the mean dissipation across a region of constant f as effectively as it characterizes a single roll-up. The nonspreading case should be comparable to the pioneering laboratory studies of two-layer entrainment by Ellison and Turner [1959] and other entrainment studies as summarized by Christodoulou [1986]. Unfortunately, the nature of these entrainment experiments, compared to the buoyancy flux dynamics discussed here, make it difficult to use the laboratory results to propose reasonable values of x o.in addition, there are significant differences between the laboratory and geophysical scales, likely resulting in vastly different Reynolds numbers, and potentially different turbulent dynamics [e.g., Geyer et al., 2010]. [27] The second coefficient in (6) is the value t, which is a normalized time scale for K-H billow evolution, assuming that spreading impacts would affect a K-H billow through the majority of its development, from initial generation, through overturning, to ultimate turbulent decay. Thus, our value of t must be of the same order as the previous estimates of the K-H evolution timescale for this conceptual spreading model to be viable. As discussed above, based on the results of Smyth et al. [2001] and Thorpe [1973], estimates for the KH evolution time scale (normalized by the shear) are on the order of 80 to 250. [28] In the following section, we use data collected from the Merrimack River near-field plume to evaluate the fit parameters, x o and t. 4. Field Experiments and Results [29] Data collected in the near-field region of the Merrimack River plume (Figure 5) are used to estimate the fit parameters of (6), and to validate its effectiveness. The Merrimack River mouth is constrained by two jetties approximately 300 m wide, with a shallow bar (approximately 3 m deep at low tide) located about 500 m beyond the end of the jetties. During high flow periods, the river discharge creates an energetic and rapidly expanding buoyant jet in the coastal ocean, as described by MacDonald et al. [2007] and Chen et al. [2009], with a bottom attached salt front anchored on the seaward edge of the bar during ebb tides. The resulting flow is strongly sheared and stratified, with negligible influence from bottom friction, given that an essentially stagnant ambient ocean layer m thick sits below the plume. Our data (MacDonald et al., manuscript in preparation, 2012) indicates an overturn scale [e.g., Thorpe, 1977] on the order of 10 cm, consistent with estimates of the Ozmidov scale, suggesting that mixing between the plume and ambient water is driven primarily by K-H type instabilities which are forming overturns on the order of the Ozmidov scale before decaying to turbulence. This turbulence quickly establishes a two-layer system, characterized by a shear-stratified flow rather than a discrete interface, with values of Ri B of order 0.05 to 0.8. [30] The data used in this study were collected during three days of field observations: May 21, 2006, and May 9 and 11, The 2006 measurements are the same measurements reported in MacDonald et al. [2007], and occurred during unusually high discharge conditions of approximately 1,300 m 3 /s. The 2007 measurements occurred during more typical freshet conditions, with an outflow of approximately 300 m 3 /s. Mean velocity profiles were recorded with two RD Instruments 1200 khz acoustic Doppler current profilers (ADCPs). One of the ADCPs was in a downward looking position, mounted off the side of the research vessel. The other mounted on an Acrobat tow body (Sea Sciences, Inc.) in an upward looking position, towed at approximately 4 m depth. Data from the two ADCPs provides a full velocity profile, which extends to the very near surface. This is crucial to fully characterize the shallow buoyant plume. Velocity profiles were collected at a rate of approximately 1 Hz, using 0.25 m vertical bins. Ship velocity was removed via bottom tracking from the downward mounted profiler. Pressure, temperature, and conductivity were recorded using Micro-CTD (Applied Microsystems Ltd.) and XR-620 (RBR Ltd.) profilers attached to a frame towed by the research vessel, undulating between the surface and approximately 4 m depth. [31] Hydrographic data were collected along passes in the along-stream direction, as shown in Figure 5. All passes occurred during the second half of the ebb tide, with a bottom attached salt front established at the bar. On both sampling days, the weather was calm, with winds generally less than 5 m/s as measured at the NDBC Isles of Shoals buoy, approximately 25 km northeast of the field site. Winds on May and May were generally westerly, and directed offshore, while winds on May were from the south. [32] Figure 6 depicts the local radial velocity, u r (assuming a radial coordinate system centered at the river mouth), and density fields from pass 3 on May The saltdominated density stratification is stable, with warmer, fresher water over cooler, saltier water. Velocity differences between plume water and the surrounding water are of order 1 m/s, with maximum differences in density of approximately 15 s T units. The density and velocity structures of the near-field region are characterized by a rapidly thinning and mixing fresh water layer seaward of the bar, resulting in high shear and stratification. 6of12

7 Figure 5. (a) Location and (b, c) timing of sampling passes at Merrimack River (Figure 5a). Axes in Figure 5a represent coordinates (in meters) with an origin at the river mouth. The two thin dashed lines in Figure 5a indicate approximate boundaries of the plume core. [33] To estimate the three parameters identified in Figure 2, it was necessary to evaluate a number of parameters directly from the observational data. The base of the shear layer was defined based on the local fresh water flux per unit width, Q f ¼ R Fu r bdz, where b is a unit width, F is the local freshness, F =(S o S)/S o, and S o and S are the ambient ocean and local salinities, respectively. The base of the shear layer was determined by identifying the depth associated with 90% of the total vertically integrated freshwater flux. [34] In many cases, the shear layer extends from the water surface down to its base as defined by the local freshwater flux. However, in some cases, particularly in the lift-off region, the shear layer is overlain by a relatively well-mixed plume layer. To account for this, we define the top of the shear layer as the depth at which the salinity is equal to the surface salinity plus 0.3 psu (although this choice is somewhat arbitrary, it is small compared to a typical DS of 10 to 20 psu, and the results are fairly insensitive to small changes in this value). We then define the shear layer thickness, h i,as the difference between the calculated shear layer base and top depths. The upper and lower limits of the shear layer are shown on the plots in Figure 6, as an example. It should be noted that, due to extremely high stratification, the Ozmidov scale, and the vertical scale of overturns observed in the CTD tow-yo data are much smaller (on the order of 10 cm) than the shear layer thickness, which is on the order of 1 to 2 m. Thus, there is little potential for developing K-H billows to directly interact with the surface boundary, which might have the effect of limiting billow development and altering the premise of the conceptual model developed in section 3. [35] Differential quantities necessary for the analysis were defined based on the upper and lower bounds of the stratified-shear layer. For example, the shear, Du, and the reduced gravity, g, were estimated from the velocity and density differences between the top and base depths of the layer. In this study, we limit our data to the lift-off and nearfield region of the plume, which we constrain to distances from the river mouth between 500 and 3000 m. Beyond this region, resolution issues hamper the ability to accurately estimate h i due to limited plume depth, and an ADCP resolution of 0.25 m. 7of12

8 Figure 6. Contours of (left) density and (right) velocity through the near-field region, representing data from pass 3 on May 11, The dotted and dashed lines represent the top and bottom of the stratified shear layer, respectively, as defined in the text. The interface depth (h i ) is the vertical distance between the dotted and dashed lines. [36] Turbulent mixing within the plume is represented by x. Buoyancy flux, B, was estimated along observational passes on all sampling days using a control volume approach [e.g., MacDonald and Geyer, 2004; MacDonald et al., 2007; Chen and MacDonald, 2006; McCabe et al., 2008]. In this case, the control volume approach utilized data from along plume transects to compare fluxes approximately 400 m apart. Localized spreading rates (i.e., DW) are derived from the difference in vertically integrated freshwater flux. Diapycnal volume and density fluxes are evaluated from differences in vertically integrated horizontal volume and density fluxes above a specified isopycnal. Calculation of these quantities allows the vertical advective density transport to be extracted from the total vertical density flux following a Reynolds decomposition, leading directly to estimates of Reynolds salt flux (i.e., r w ¼ rw rw), and thus of B. Use of successive isopycnals allows a vertical of profile of B to be generated for each control volume. The value of B used in this analysis represents the vertically averaged buoyancy flux observed within a given control volume through the defined upper layer. MacDonald et al. [2007] described the use of the control volume method to measure B within the Merrimack River plume, and compared the results to autonomous underwater vehicle (AUV) microstructure data and Regional Ocean Modeling System (ROMS) numerical model output. The control volume results and numerical model output were in excellent agreement, consistent with the integrative nature of both the observational technique and the closure schemes inherent in the model. The microstructure results, while highly variable, were generally an order of magnitude or more lower than the other estimates, suggesting a high degree of heterogeneity in the turbulent structure, consistent with the assumption of K-H billows as the primary turbulent mechanism in the shear-stratified layer [e.g., Smyth et al., 2001; MacDonald et al., manuscript in preparation, 2012]. This paper and others utilizing variants of the control volume technique [e.g., MacDonald and Geyer, 2004; Chen and MacDonald, 2006; McCabe et al., 2008], provide strong support for the consistency of the method with other estimates of turbulence, and we use the method here to provide estimates of both buoyancy flux and plume spreading. [37] The value of f was determined from observations using (4), with calculated width ratios (i.e., DW/W o ) generated from the control volume analysis: f ¼ DW W o h i Dr CV Here, Dr CV and DW represent the along plume length of the control volume, and the change in width across the control volume, respectively. This differs from (4) where the length scale (DuT KH ) and width expansion were defined across a distance representing the evolution of a KH billow. In either case, the expression represents an average value of f over the chosen distance. The use of the control volume method to constrain spreading rates was investigated in Chen et al. [2009], where control volume estimates agreed favorably with both drifter analyses and numerical model output. Note that our calculated values of DW/W o represent a depthaveraged value of the plume spreading rate, which may exhibit significant vertical structure, with enhanced spreading near the surface and less spreading near the base of the plume. It should also be recognized that DW/W o is a nondimensional plume spreading rate, with increase in width normalized by the total plume width. Thus, even with no decrease in the interfacial wave speed, which defines the rate of lateral expansion due to buoyancy differences, f will tend to decrease as the local plume width increases in the seaward direction. [38] Ri B was estimated directly from observations using the equation in Figure 2, with the value of Ri B plotted against the parameters x and f in Figures 3 and 7, respectively. No clear trend between Ri B and the other two dimensionless parameters is apparent on these plots. ð7þ 8of12

9 Figure 7. [39] In Figure 8, x is plotted directly against f. The observations show a clear relationship between the two dimensionless parameters. With distance increasing from the river mouth, the plume data enters the graph in Figure 8 at Plot of f versus Ri B for data from 2006 and 2007 passes. the top right, with strong initial mixing, and high spreading rates, as indicated by high values of both x and f, and then move down and to the left as the plume propagates seaward. Figure 8. Measured lateral spreading parameter, f, versus the mixing parameter, x, based on observational data with error bars (see Figure 1). A best fit curve based on the conceptual mechanism described in the text is shown by the black line. The gray arrow indicates the direction of plume water propagation through the parameter space. Discharged water enters the graph at the top right, and then moves down and to the left as the plume propagates seaward. A point approximating average values from MacDonald and Geyer [2004] is also shown as the filled star at the upper right corner. 9of12

10 Figure 9. Plot of x CV /x S, with x S defined by equation (6) and constants as defined in Figure 8, with best fit line and 90% confidence limits (bold solid and dashed lines, respectively). This plot is similar to Figure 1, with the exception that spreading processes are now considered in the scaled estimate of x. The best fit line and confidence limits from Figure 1 are shown in gray for reference. Although considerable scatter remains in the data, the clear trend seen in Figure 1 of decreasing x with seaward distance in the plume is less apparent, suggesting that this decrease was primarily a function of plume spreading. An approximate marker for the conditions described by MacDonald and Geyer [2004] is also plotted for reference. 5. Discussion [40] The values of f and x in Figure 8, can be used to evaluate the relationship described by (6), by fitting an appropriate quadratic relationship to the data. The best quadratic fit for the combined 2006/2007 data sets is shown in Figure 8, with values of x o = and t = 360, and an R 2 value of The fit is considerably improved by considering the 2007 data set only, yielding x o =5 10 5, t = 386, and an R 2 value of Despite the increase in R 2 observed by excluding the 2006 data set, both fits yield very similar values of x o and t. [41] The value of t 360 best describing the data is within a factor of 1.5 to 4 of the estimates based on Smyth et al. [2001] and Thorpe [1973, 2005], which are on the order of The fact that the time required for the degree of turbulent enhancement seen in the Merrimack near-field plume is of the same order as the plume evolution time scale derived from laboratory and DNS results suggests that the distinct processes of lateral plume spreading, and turbulence generation within the stratified shear layer can and do interact. Our higher observed value of t suggests that developing roll-ups might be subject to enhancement through lateral expansion throughout the majority of their life-cycle, even beyond the loss of visually coherent billow structures. Other explanations for the discrepancy between our value of t and those derived from the literature might include the vertical structure of the spreading profile, which is unaccounted for in our average value of f. Given the quadratic nature of (6), values of f that are enhanced near the surface would result in lower values of t than reported here, depending on the nature of the vertical structure of spreading. Our higher value of t may also be due to unresolved processes related to high Re flows, or to an increased duration of coherency of the billow structure related directly to the presence of lateral spreading. Geyer et al. [2010] have recently shown that the structure of turbulence at high Re values is substantially different than that observed in laboratory experiments and DNS simulations at low Re. [42] In the Merrimack plume the shear layer thickness is on the order of 1 m, with plume velocities on the order of 1 m/s (Figure 6), implying a value of T KH of approximately 6 min, based on the calculated value of t = 360. This value suggests that the life cycle of a roll up is on the order of 6 min for an instability to be generated, begin to overturn and transition to turbulence. [43] The best fit value of x o is approximately , indicating an interfacial drag coefficient C Di of the same order (C Di ¼ Ri B Ri f x, where the two Richardson numbers are of the same order). As would be expected, this C Di value is significantly smaller than a bottom drag coefficient, which is typically on the order of MacDonald and Geyer [2004] obtained an interfacial drag coefficient of , larger by an order of magnitude, but consistent with the fact that it was from a region subject to significant lateral spreading, while the value of x o represents the base value for the no-spreading case. The entrainment studies of Ellison and Turner [1959] and Christodoulou [1986] suggest entrainment coefficients on the order of With some assumptions, this value can be converted to a comparable 10 of 12

11 value of x o, which is also on the order of 10 2, two to three orders of magnitude higher than the value obtained here (and in MacDonald and Geyer [2004]). These observed differences may be due to Reynolds number effects, as the entrainment studies were carried out in the laboratory. A full analysis of these discrepancies is beyond the scope of the present manuscript, but the results from this study can provide a benchmark for future studies regarding shearstratified mixing at high Reynolds numbers in non-spreading environments. [44] Finally, we return to the motivation for this study, the observed failure of the two-dimensional scaling for stratified shear turbulence observed by MacDonald et al. [2007] and represented in Figure 1 of this manuscript. To evaluate the effectiveness of our enhanced scaling technique we consider x the ratio, rather than x x o ð1þtfþ 2 C as shown in Figure 1, plotted as a function of distance from the mouth in the Merrimack plume in Figure 9. Although scatter remains in the data shown in Figure 9, a comparison of the data in Figures 1 and 9, along with their respective best fit lines and 90% confidence intervals, suggests that the skill of the scaling approach has been improved by inclusion of the lateral spreading mechanism for turbulent enhancement. For example, the slope of the linear fit through the data in Figure 1 (log(x/c) versus distance) yields a value of approximately 0.24 km 1, and an R 2 value of 0.20, compared to a similarly calculated slope through the data of Figure 9 of 0.08 km 1 (R 2 = 0.03). Thus, the best fit slope has been reduced by a factor of 3. Likewise, the R 2 values should be compared by considering that the values represent the improvement of the fit over the assumption of a constant value. Thus, the minimal R 2 value for the data in Figure 9 suggests that a slope of 0 would be equally appropriate, while the R 2 value for Figure 1 suggests a downward trend, but one weakened by the degree of scatter in the data. [45] The fact that there is considerable scatter remaining in the data displayed on Figure 9, and that R 2 values associated with the fit of (6) as shown in Figure 8 are only on the order of is not surprising and is reflective of the fact that there are a variety of factors playing a role in the determination of mixing. We do not suggest that spreading is the only important process. As discussed above, the effects of vertical structure in the spreading profile may introduce scatter as well as overestimating the value of t. Additionally, variability in mixing efficiency, as quantified by the value of Ri f may also play a role. However, the fact that the value of t derived from the field data matches reasonably well with laboratory and DNS estimates, and that application of the spreading relationship removes the downward trend apparent in Figure 1, strongly suggests that spreading is an important process, with significant impacts on mixing. 6. Summary [46] This manuscript presents estimates of five key variables related to near-field plume structure and evolution, including vertical shear, stratification, turbulent buoyancy flux, lateral strain rate of the plume and the interface depth. Three dimensionless parameters are derived, representing the turbulent mixing, x, the lateral spreading rate, f, and the instability of the flow, Ri B. Evaluation of these parameters using observed field data from the near-field Merrimack River plume suggests that the mixing and spreading parameters are clearly related, but that neither parameter is well correlated with Ri B. A conceptual model based on the longitudinal stretching of individual K-H billows provides a direct mechanism for lateral plume spreading to impact the energetics of the local turbulent field, resulting in a quadratic relationship between the mixing and spreading parameters. The fit of the quadratic relationship to data from the Merrimack River near-field plume explains a majority of the observed variability, and yields two values, the value of x in the nonspreading limit, referred to as x o, and a turbulence evolution time scale, t. The value of x o differs significantly from laboratory-derived values, perhaps suggesting a strong Reynolds number dependence. However, the value of t is reasonably consistent with other estimates of the turbulent evolution time scale based on laboratory and DNS results and lends credence to the viability of the lateral spreading mechanism as a means of turbulent enhancement. Further investigation of this relationship in a controlled laboratory setting or through direct numerical simulations may be warranted to more accurately define the quadratic relationship and independently measure the evolution time scale. Ultimately, these findings may have significant implications for near-field plume dynamics, as well as other shear-stratified flows in the ocean, atmosphere, and even industrial settings. [47] Acknowledgments. We thank W. Geyer, L. Goodman, R. Hetland, and A. Horner-Devine for useful discussions on this topic. This work was funded by National Science Foundation grants OCE and OCE This manuscript is contribution in the SMAST Contribution Series, School for Marine Science and Technology, University of Massachusetts Dartmouth. References Chen, F., and D. G. MacDonald (2006), Role of mixing in the structure and evolution of a buoyant discharge plume, J. Geophys. Res., 111, C11002, doi: /2006jc Chen, F., D. G. MacDonald, and R. D. Hetland (2009), Lateral spreading of a near-field river plume: Observations and numerical simulations, J. Geophys. Res., 114, C07013, doi: /2008jc Christodoulou, G. C. (1986), Interfacial mixing in stratified flows, J. Hydraul. Res., 24(2), 77 92, doi: / Ellison, T. H., and J. S. Turner (1959), Turbulent entrainment in stratified flows, J. Fluid Mech., 6(03), , doi: /s Farmer, D., R. Pawlowicz, and R. Jiang (2002), Tilting separation flows: A mechanism for intense vertical mixing in the coastal ocean, Dyn. Atmos. Oceans, 36(1 3), 43 58, doi: /s (02) Geyer, W. R., and D. M. Farmer (1989), Tide-induced variation of the dynamics of a salt wedge estuary, J. Phys. Oceanogr., 19(8), , doi: / (1989)019<1060:tivotd>2.0.co;2. Geyer, W. R., A. C. Lavery, M. E. Scully, and J. H. Trowbridge (2010), Mixing by shear instability at high Reynolds number, Geophys. Res. Lett., 37, L22607, doi: /2010gl Gregg, M. C. (1987), Diapycnal mixing in the thermocline: A review, J. Geophys. Res., 92(C5), , doi: /jc092ic05p Hetland, R. D. (2010), The effects of mixing and spreading on density in near-field river plumes, Dyn. Atmos. Oceans, 49, 37 53, doi: /j. dynatmoce Hetland, R. D., and D. G. MacDonald (2008), Spreading in the near-field Merrimack River plume, Ocean Modell., 21(1 2), 12 21, doi: /j. ocemod Imberger, J., and G. N. Ivey (1991), On the nature of turbulence in a stratified fluid. Part II: Application to lakes, J. Phys. Oceanogr., 21(5), , doi: / (1991)021<0659:otnoti>2.0.co;2. Koop, C. G., and F. K. Browand (1979), Instability and turbulence in a stratified fluid with shear, J. Fluid Mech., 93, , doi: / S MacDonald, D. G., and W. R. Geyer (2004), Turbulent energy production and entrainment at a highly stratified estuarine front, J. Geophys. Res., 109(C5), C05004, doi: /2003jc of 12

12 MacDonald, D. G., L. Goodman, and R. D. Hetland (2007), Turbulent dissipation in a near-field river plume: A comparison of control volume and microstructure observations with a numerical model, J. Geophys. Res., 112, C07026, doi: /2006jc McCabe, R., B. M. Hickey, and P. MacCready (2008), Observational estimates of entrainment and vertical salt flux in the interior of a spreading river plume, J. Geophys. Res., 113, C08027, doi: /2007jc Smyth, W. D., J. N. Moum, and D. R. Caldwell (2001), The efficiency of mixing in turbulent patches: Inferences from direct simulations and microstructure observations, J. Phys. Oceanogr., 31(8), , doi: / (2001)031<1969:teomit>2.0.co;2. Taylor, G. I. (1932), The transport of vorticity and heat through fluids in turbulent motion, Proc. R. Soc. London, Ser. A, 135, , doi: /rspa Tedford, E. W., J. R. Carpenter, R. Pawlowicz, R. Pieters, and G. A. Lawrence (2009), Observation and analysis of shear instability in the Fraser River estuary, J. Geophys. Res., 114(C11), C11006, doi: /2009jc Thorpe, S. A. (1971), Experiments on the instability of stratified shear flows: Miscible fluids, J. Fluid Mech., 46, , doi: / S Thorpe, S. A. (1973), Experiments on instability and turbulence in a stratified shear flow, J. Fluid Mech., 61, , doi: / S Thorpe, S. A. (1977), Turbulence and mixing in a Scottish loch, Philos. Trans. R. Soc. London, Ser. A, 286, , doi: / rsta Thorpe, S. A. (1987), Transitional phenomena and the development of turbulence in stratified fluids: A review, J. Geophys. Res., 92(C5), , doi: /jc092ic05p Thorpe, S. A. (2005), The Turbulent Ocean, 439 pp., Cambridge Univ. Press, Cambridge, U. K. Wright, L. D., and J. M. Coleman (1971), Effluent expansion and interfacial mixing in the presence of a salt wedge, Mississippi River Delta, J. Geophys. Res., 76(36), , doi: /jc076i036p of 12

Mixing by shear instability at high Reynolds number

Mixing by shear instability at high Reynolds number GEOPHYSICAL RESEARCH LETTERS, VOL. 37,, doi:10.1029/2010gl045272, 2010 Mixing by shear instability at high Reynolds number W. R. Geyer, 1 A. C. Lavery, 1 M. E. Scully, 2 and J. H. Trowbridge 1 Received

More information

Turbulent dissipation in a near field river plume: A comparison of control volume and microstructure observations with a numerical model.

Turbulent dissipation in a near field river plume: A comparison of control volume and microstructure observations with a numerical model. 5 Turbulent dissipation in a near field river plume: A comparison of control volume and microstructure observations with a numerical model. Daniel G. MacDonald 1, Louis Goodman 1 and Robert D. Hetland

More information

Temporal and spatial variability of vertical salt flux in a highly stratified estuary

Temporal and spatial variability of vertical salt flux in a highly stratified estuary Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113,, doi:10.1029/2007jc004620, 2008 Temporal and spatial variability of vertical salt flux in a highly stratified estuary Daniel G. MacDonald

More information

Temporal and spatial variability of vertical salt flux in a highly stratified estuary.

Temporal and spatial variability of vertical salt flux in a highly stratified estuary. Temporal and spatial variability of vertical salt flux in a highly stratified estuary. 5 Daniel G. MacDonald 1 and Alexander R. Horner-Devine 2 10 15 1 Department of Estuarine and Ocean Sciences School

More information

Kinematic Effects of Differential Transport on Mixing Efficiency in a Diffusively Stable, Turbulent Flow

Kinematic Effects of Differential Transport on Mixing Efficiency in a Diffusively Stable, Turbulent Flow Iowa State University From the SelectedWorks of Chris R. Rehmann January, 2003 Kinematic Effects of Differential Transport on Mixing Efficiency in a Diffusively Stable, Turbulent Flow P. Ryan Jackson,

More information

Buoyancy Fluxes in a Stratified Fluid

Buoyancy Fluxes in a Stratified Fluid 27 Buoyancy Fluxes in a Stratified Fluid G. N. Ivey, J. Imberger and J. R. Koseff Abstract Direct numerical simulations of the time evolution of homogeneous stably stratified shear flows have been performed

More information

The estimation of vertical eddy diffusivity in estuaries

The estimation of vertical eddy diffusivity in estuaries The estimation of vertical eddy diffusivity in estuaries A. ~temad-~hahidi' & J. 1rnberger2 l Dept of Civil Eng, Iran University of Science and Technology, Iran. 2 Dept of Env. Eng., University of Western

More information

Role of mixing in the structure and evolution of a buoyant discharge plume

Role of mixing in the structure and evolution of a buoyant discharge plume Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 111,, doi:10.1029/2006jc003563, 2006 Role of mixing in the structure and evolution of a buoyant discharge plume Fei Chen 1 and Daniel G.

More information

Turbulent Mixing During an Admiralty Inlet Bottom Water Intrusion. Coastal and Estuarine Fluid Dynamics class project, August 2006

Turbulent Mixing During an Admiralty Inlet Bottom Water Intrusion. Coastal and Estuarine Fluid Dynamics class project, August 2006 Turbulent Mixing During an Admiralty Inlet Bottom Water Intrusion Coastal and Estuarine Fluid Dynamics class project, August 2006 by Philip Orton Summary Vertical turbulent mixing is a primary determinant

More information

Shear instabilities in a tilting tube

Shear instabilities in a tilting tube Abstract Shear instabilities in a tilting tube Edmund Tedford 1, Jeff Carpenter 2 and Greg Lawrence 1 1 Department of Civil Engineering, University of British Columbia ttedford@eos.ubc.ca 2 Institute of

More information

The role of turbulence stress divergence in decelerating a river plume

The role of turbulence stress divergence in decelerating a river plume JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117,, doi:10.1029/2011jc007398, 2012 The role of turbulence stress divergence in decelerating a river plume Levi F. Kilcher, 1 Jonathan D. Nash, 1 and James N. Moum

More information

Modeling the Columbia River Plume on the Oregon Shelf during Summer Upwelling. 2 Model

Modeling the Columbia River Plume on the Oregon Shelf during Summer Upwelling. 2 Model Modeling the Columbia River Plume on the Oregon Shelf during Summer Upwelling D. P. Fulton August 15, 2007 Abstract The effects of the Columbia River plume on circulation on the Oregon shelf are analyzed

More information

GFD 2013 Lecture 10: Gravity currents on slopes and in turbulent environments

GFD 2013 Lecture 10: Gravity currents on slopes and in turbulent environments GFD 2013 Lecture 10: Gravity currents on slopes and in turbulent environments Paul Linden; notes by Gregory Wagner and Barbara Zemskova June 28, 2013 1 Introduction Natural gravity currents are often found

More information

Tilting Shear Layers in Coastal Flows

Tilting Shear Layers in Coastal Flows DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. Tilting Shear Layers in Coastal Flows Karl R. Helfrich Department of Physical Oceanography, MS-21 Woods Hole Oceanographic

More information

For example, for values of A x = 0 m /s, f 0 s, and L = 0 km, then E h = 0. and the motion may be influenced by horizontal friction if Corioli

For example, for values of A x = 0 m /s, f 0 s, and L = 0 km, then E h = 0. and the motion may be influenced by horizontal friction if Corioli Lecture. Equations of Motion Scaling, Non-dimensional Numbers, Stability and Mixing We have learned how to express the forces per unit mass that cause acceleration in the ocean, except for the tidal forces

More information

Impact of Offshore Winds on a Buoyant River Plume System

Impact of Offshore Winds on a Buoyant River Plume System DECEMBER 2013 J U R I S A A N D C H A N T 2571 Impact of Offshore Winds on a Buoyant River Plume System JOSEPH T. JURISA* AND ROBERT J. CHANT Rutgers, The State University of New Jersey, New Brunswick,

More information

2013 Annual Report for Project on Isopycnal Transport and Mixing of Tracers by Submesoscale Flows Formed at Wind-Driven Ocean Fronts

2013 Annual Report for Project on Isopycnal Transport and Mixing of Tracers by Submesoscale Flows Formed at Wind-Driven Ocean Fronts DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. 2013 Annual Report for Project on Isopycnal Transport and Mixing of Tracers by Submesoscale Flows Formed at Wind-Driven

More information

Plumes and jets with time-dependent sources in stratified and unstratified environments

Plumes and jets with time-dependent sources in stratified and unstratified environments Plumes and jets with time-dependent sources in stratified and unstratified environments Abstract Matthew Scase 1, Colm Caulfield 2,1, Stuart Dalziel 1 & Julian Hunt 3 1 DAMTP, Centre for Mathematical Sciences,

More information

Turbulent Mixing Parameterizations for Oceanic Flows and Student Support

Turbulent Mixing Parameterizations for Oceanic Flows and Student Support DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. Turbulent Mixing Parameterizations for Oceanic Flows and Student Support Subhas Karan Venayagamoorthy Department of Civil

More information

Structure and composition of a strongly stratified, tidally pulsed river plume

Structure and composition of a strongly stratified, tidally pulsed river plume Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 114,, doi:10.1029/2008jc005036, 2009 Structure and composition of a strongly stratified, tidally pulsed river plume Jonathan D. Nash, 1

More information

A note on mixing due to surface wave breaking

A note on mixing due to surface wave breaking JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113,, doi:10.1029/2008jc004758, 2008 A note on mixing due to surface wave breaking Jan Erik H. Weber 1 Received 31 January 2008; revised 14 May 2008; accepted 18 August

More information

SMS 303: Integrative Marine

SMS 303: Integrative Marine SMS 303: Integrative Marine Sciences III Instructor: E. Boss, TA: A. Palacz emmanuel.boss@maine.edu, 581-4378 5 weeks & topics: diffusion, mixing, tides, Coriolis, and waves. Pre-class quiz. Mixing: What

More information

Internal hydraulics and mixing in a highly stratified estuary

Internal hydraulics and mixing in a highly stratified estuary JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 105, NO. C6, PAGES 14,215 14,222, JUNE 15, 2000 Internal hydraulics and mixing in a highly stratified estuary Robert J. Chant Institute for Marine and Coastal Sciences,

More information

Remote Sensing of Temperature and Salinity Microstructure in Rivers and Estuaries Using Broadband Acoustic Scattering Techniques

Remote Sensing of Temperature and Salinity Microstructure in Rivers and Estuaries Using Broadband Acoustic Scattering Techniques DISTRIBUTION STATEMENT A: Approved for public release; distribution is unlimited. Remote Sensing of Temperature and Salinity Microstructure in Rivers and Estuaries Using Broadband Acoustic Scattering Techniques

More information

A Comparison of Predicted Along-channel Eulerian Flows at Cross- Channel Transects from an EFDC-based Model to ADCP Data in South Puget Sound

A Comparison of Predicted Along-channel Eulerian Flows at Cross- Channel Transects from an EFDC-based Model to ADCP Data in South Puget Sound A Comparison of Predicted Along-channel Eulerian Flows at Cross- Channel Transects from an EFDC-based Model to ADCP Data in South Puget Sound Skip Albertson, J. A. Newton and N. Larson Washington State

More information

High-Frequency Acoustic Propagation in Shallow, Energetic, Highly-Salt- Stratified Environments

High-Frequency Acoustic Propagation in Shallow, Energetic, Highly-Salt- Stratified Environments DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. High-Frequency Acoustic Propagation in Shallow, Energetic, Highly-Salt- Stratified Environments Andone C. Lavery Department

More information

Frontal Structures in the Columbia River Plume Nearfield A Nonhydrostatic Coastal Modeling Study

Frontal Structures in the Columbia River Plume Nearfield A Nonhydrostatic Coastal Modeling Study DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. Frontal Structures in the Columbia River Plume Nearfield A Nonhydrostatic Coastal Modeling Study Tian-Jian Hsu, Fengyan

More information

Summary Results from Horizontal ADCP tests in the Indiana Harbor Canal and the White River

Summary Results from Horizontal ADCP tests in the Indiana Harbor Canal and the White River Summary Results from Horizontal ADCP tests in the Indiana Harbor Canal and the White River This report summarizes results of tests of horizontally deployed ADCPs in the Indiana Harbor Canal and the White

More information

Tidal and spring-neap variations in horizontal dispersion in a partially mixed estuary

Tidal and spring-neap variations in horizontal dispersion in a partially mixed estuary Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113,, doi:10.1029/2007jc004644, 2008 Tidal and spring-neap variations in horizontal dispersion in a partially mixed estuary W. R. Geyer,

More information

Main issues of Deltas

Main issues of Deltas Global sediment supply to coastal seas and oceans; location of major river deltas RIVER DELTAS Depositional processes - Course Coastal Morphodynamics GEO3-436; lecture 4 Nile Delta, Egypt Solo Delta, Java,

More information

Northern Arabian Sea Circulation Autonomous Research (NASCar) DRI: A Study of Vertical Mixing Processes in the Northern Arabian Sea

Northern Arabian Sea Circulation Autonomous Research (NASCar) DRI: A Study of Vertical Mixing Processes in the Northern Arabian Sea DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. Northern Arabian Sea Circulation Autonomous Research (NASCar) DRI: A Study of Vertical Mixing Processes in the Northern

More information

Lecture 2. Turbulent Flow

Lecture 2. Turbulent Flow Lecture 2. Turbulent Flow Note the diverse scales of eddy motion and self-similar appearance at different lengthscales of this turbulent water jet. If L is the size of the largest eddies, only very small

More information

Submesoscale Routes to Lateral Mixing in the Ocean

Submesoscale Routes to Lateral Mixing in the Ocean DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. Submesoscale Routes to Lateral Mixing in the Ocean Amit Tandon Physics Department, UMass Dartmouth 285 Old Westport Rd

More information

Internal Wave Driven Mixing and Transport in the Coastal Ocean

Internal Wave Driven Mixing and Transport in the Coastal Ocean DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. Internal Wave Driven Mixing and Transport in the Coastal Ocean Subhas Karan Venayagamoorthy Department of Civil and Environmental

More information

Hydrodynamics in Shallow Estuaries with Complex Bathymetry and Large Tidal Ranges

Hydrodynamics in Shallow Estuaries with Complex Bathymetry and Large Tidal Ranges DISTRIBUTION STATEMENT A: Approved for public release; distribution is unlimited. Hydrodynamics in Shallow Estuaries with Complex Bathymetry and Large Tidal Ranges Stephen G. Monismith Dept of Civil and

More information

Donald Slinn, Murray D. Levine

Donald Slinn, Murray D. Levine 2 Donald Slinn, Murray D. Levine 2 Department of Civil and Coastal Engineering, University of Florida, Gainesville, Florida College of Oceanic and Atmospheric Sciences, Oregon State University, Corvallis,

More information

( ) = 1005 J kg 1 K 1 ;

( ) = 1005 J kg 1 K 1 ; Problem Set 3 1. A parcel of water is added to the ocean surface that is denser (heavier) than any of the waters in the ocean. Suppose the parcel sinks to the ocean bottom; estimate the change in temperature

More information

Understanding and modeling dense overflows. Sonya Legg Princeton University AOMIP/FAMOS school for young scientists 2012

Understanding and modeling dense overflows. Sonya Legg Princeton University AOMIP/FAMOS school for young scientists 2012 Understanding and modeling dense overflows Sonya Legg Princeton University AOMIP/FAMOS school for young scientists 2012 What is an overflow? Dense water formation on shelf or marginal sea Dense water accelerates

More information

Entrainment and flushing time in the Fraser River estuary and plume from a steady salt balance analysis

Entrainment and flushing time in the Fraser River estuary and plume from a steady salt balance analysis JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 116,, doi:10.1029/2010jc006793, 2011 Entrainment and flushing time in the Fraser River estuary and plume from a steady salt balance analysis M. Halverson 1 and R.

More information

Relating River Plume Structure to Vertical Mixing

Relating River Plume Structure to Vertical Mixing SEPTEMBER 2005 H E T LAND 1667 Relating River Plume Structure to Vertical Mixing ROBERT D. HETLAND Department of Oceanography, Texas A&M University, College Station, Texas (Manuscript received 30 March

More information

Inertial Range Dynamics in Density-Stratified Turbulent Flows

Inertial Range Dynamics in Density-Stratified Turbulent Flows Inertial Range Dynamics in Density-Stratified Turbulent Flows James J. Riley University of Washington Collaborators: Steve debruynkops (UMass) Kraig Winters (Scripps IO) Erik Lindborg (KTH) Workshop on

More information

Observations of a Kelvin-Helmholtz Billow in the Ocean

Observations of a Kelvin-Helmholtz Billow in the Ocean Journal of Oceanography, Vol. 57, pp. 709 to 721, 2001 Observations of a Kelvin-Helmholtz Billow in the Ocean HUA LI and HIDEKATSU YAMAZAKI* Department of Ocean Sciences, Tokyo University of Fisheries,

More information

Lecture 3. Turbulent fluxes and TKE budgets (Garratt, Ch 2)

Lecture 3. Turbulent fluxes and TKE budgets (Garratt, Ch 2) Lecture 3. Turbulent fluxes and TKE budgets (Garratt, Ch 2) The ABL, though turbulent, is not homogeneous, and a critical role of turbulence is transport and mixing of air properties, especially in the

More information

centrifugal acceleration, whose magnitude is r cos, is zero at the poles and maximum at the equator. This distribution of the centrifugal acceleration

centrifugal acceleration, whose magnitude is r cos, is zero at the poles and maximum at the equator. This distribution of the centrifugal acceleration Lecture 10. Equations of Motion Centripetal Acceleration, Gravitation and Gravity The centripetal acceleration of a body located on the Earth's surface at a distance from the center is the force (per unit

More information

Generation and Evolution of Internal Waves in Luzon Strait

Generation and Evolution of Internal Waves in Luzon Strait DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. Generation and Evolution of Internal Waves in Luzon Strait Ren-Chieh Lien Applied Physics Laboratory University of Washington

More information

The Stable Boundary layer

The Stable Boundary layer The Stable Boundary layer the statistically stable or stratified regime occurs when surface is cooler than the air The stable BL forms at night over land (Nocturnal Boundary Layer) or when warm air travels

More information

Asymmetric Tidal Mixing due to the Horizontal Density Gradient*

Asymmetric Tidal Mixing due to the Horizontal Density Gradient* 418 J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y VOLUME 38 Asymmetric Tidal Mixing due to the Horizontal Density Gradient* MING LI Horn Point Laboratory, University of Maryland Center for

More information

Sediment Transport at Density Fronts in Shallow Water: a Continuation of N

Sediment Transport at Density Fronts in Shallow Water: a Continuation of N DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. Sediment Transport at Density Fronts in Shallow Water: a Continuation of N00014-08-1-0846 David K. Ralston Applied Ocean

More information

Homework 5: Background Ocean Water Properties & Stratification

Homework 5: Background Ocean Water Properties & Stratification 14 August 2008 MAR 110 HW5: Ocean Properties 1 Homework 5: Background Ocean Water Properties & Stratification The ocean is a heterogeneous mixture of water types - each with its own temperature, salinity,

More information

Efficiency of Mixing Forced by Unsteady Shear Flow

Efficiency of Mixing Forced by Unsteady Shear Flow 1150 J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y VOLUME 39 Efficiency of Mixing Forced by Unsteady Shear Flow RYUICHIRO INOUE Department of Physics and Astronomy, University of Victoria,

More information

3.3 Classification Diagrams Estuarine Zone Coastal Lagoons References Physical Properties and Experiments in

3.3 Classification Diagrams Estuarine Zone Coastal Lagoons References Physical Properties and Experiments in Contents 1 Introduction to Estuary Studies... 1 1.1 Why to Study Estuaries?.... 1 1.2 Origin and Geological Age... 4 1.3 Definition and Terminology... 7 1.4 Policy and Actions to Estuary Preservation....

More information

Estuarine Boundary Layer Mixing Processes: Insights from Dye Experiments*

Estuarine Boundary Layer Mixing Processes: Insights from Dye Experiments* JULY 2007 C H A N T E T A L. 1859 Estuarine Boundary Layer Mixing Processes: Insights from Dye Experiments* ROBERT J. CHANT Institute of Marine and Coastal Sciences, Rutgers, The State University of New

More information

Laboratory Studies of Turbulent Mixing

Laboratory Studies of Turbulent Mixing Laboratory Studies of Turbulent Mixing J.A. Whitehead Woods Hole Oceanographic Institution, Woods Hole, MA, USA Laboratory measurements are required to determine the rates of turbulent mixing and dissipation

More information

Laboratory experiments on diapycnal mixing in stratified fluids

Laboratory experiments on diapycnal mixing in stratified fluids Laboratory experiments on diapycnal mixing in stratified fluids M.E. Barry, G.N. Ivey, K.B. Winters 2, and J. Imberger Centre for Water Research, The University of Western Australia, Australia 2 Applied

More information

On the turbulent Prandtl number in homogeneous stably stratified turbulence

On the turbulent Prandtl number in homogeneous stably stratified turbulence J. Fluid Mech. (2010), vol. 644, pp. 359 369. c Cambridge University Press 2010 doi:10.1017/s002211200999293x 359 On the turbulent Prandtl number in homogeneous stably stratified turbulence SUBHAS K. VENAYAGAMOORTHY

More information

The Atmospheric Boundary Layer. The Surface Energy Balance (9.2)

The Atmospheric Boundary Layer. The Surface Energy Balance (9.2) The Atmospheric Boundary Layer Turbulence (9.1) The Surface Energy Balance (9.2) Vertical Structure (9.3) Evolution (9.4) Special Effects (9.5) The Boundary Layer in Context (9.6) Fair Weather over Land

More information

arxiv: v1 [physics.flu-dyn] 12 Mar 2014

arxiv: v1 [physics.flu-dyn] 12 Mar 2014 Generated using version 3.2 of the official AMS L A TEX template The effect of Prandtl number on mixing in low Reynolds number arxiv:1403.3113v1 [physics.flu-dyn] 12 Mar 2014 Kelvin-Helmholtz billows Mona

More information

Ocean Dynamics. The Great Wave off Kanagawa Hokusai

Ocean Dynamics. The Great Wave off Kanagawa Hokusai Ocean Dynamics The Great Wave off Kanagawa Hokusai LO: integrate relevant oceanographic processes with factors influencing survival and growth of fish larvae Physics Determining Ocean Dynamics 1. Conservation

More information

HIGH RESOLUTION SEDIMENT DYNAMICS IN SALT-WEDGE ESTUARIES

HIGH RESOLUTION SEDIMENT DYNAMICS IN SALT-WEDGE ESTUARIES HIGH RESOLUTION SEDIMENT DYNAMICS IN SALT-WEDGE ESTUARIES Philip Orton, Dept. of Environmental Science and Engineering, Oregon Graduate Institute Douglas Wilson, Dept. of Environmental Science and Engineering,

More information

Internal Wave Generation and Scattering from Rough Topography

Internal Wave Generation and Scattering from Rough Topography Internal Wave Generation and Scattering from Rough Topography Kurt L. Polzin Corresponding author address: Kurt L. Polzin, MS#21 WHOI Woods Hole MA, 02543. E-mail: kpolzin@whoi.edu Abstract Several claims

More information

Wind driven mixing below the oceanic mixed layer

Wind driven mixing below the oceanic mixed layer Wind driven mixing below the oceanic mixed layer Article Published Version Grant, A. L. M. and Belcher, S. (2011) Wind driven mixing below the oceanic mixed layer. Journal of Physical Oceanography, 41

More information

Testing Turbulence Closure Models Against Oceanic Turbulence Measurements

Testing Turbulence Closure Models Against Oceanic Turbulence Measurements Testing Turbulence Closure Models Against Oceanic Turbulence Measurements J. H. Trowbridge Woods Hole Oceanographic Institution Woods Hole, MA 02543 phone: 508-289-2296 fax: 508-457-2194 e-mail: jtrowbridge@whoi.edu

More information

High-Resolution Dynamics of Stratified Inlets

High-Resolution Dynamics of Stratified Inlets DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. High-Resolution Dynamics of Stratified Inlets W. Rockwell Geyer Woods Hole Oceanographic Institution Woods Hole, MA 02543

More information

Turbulence and Energy Transfer in Strongly-Stratified Flows

Turbulence and Energy Transfer in Strongly-Stratified Flows Turbulence and Energy Transfer in Strongly-Stratified Flows James J. Riley University of Washington Collaborators: Steve debruynkops (UMass) Kraig Winters (Scripps IO) Erik Lindborg (KTH) First IMS Turbulence

More information

Small scale mixing in coastal areas

Small scale mixing in coastal areas Spice in coastal areas 1, Jody Klymak 1, Igor Yashayaev 2 University of Victoria 1 Bedford Institute of Oceanography 2 19 October, 2015 Spice Lateral stirring on scales less than the Rossby radius are

More information

Island Wakes in Shallow Water

Island Wakes in Shallow Water Island Wakes in Shallow Water Changming Dong, James C. McWilliams, et al Institute of Geophysics and Planetary Physics, University of California, Los Angeles 1 ABSTRACT As a follow-up work of Dong et al

More information

Capabilities of Ocean Mixed Layer Models

Capabilities of Ocean Mixed Layer Models Capabilities of Ocean Mixed Layer Models W.G. Large National Center for Atmospheric Research Boulder Co, USA 1. Introduction The capabilities expected in today s state of the art models of the ocean s

More information

Structure and dynamics of the Columbia River tidal plume front

Structure and dynamics of the Columbia River tidal plume front Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115,, doi:10.1029/2009jc006066, 2010 Structure and dynamics of the Columbia River tidal plume front Levi F. Kilcher 1 and Jonathan D. Nash

More information

Turbulence Instability

Turbulence Instability Turbulence Instability 1) All flows become unstable above a certain Reynolds number. 2) At low Reynolds numbers flows are laminar. 3) For high Reynolds numbers flows are turbulent. 4) The transition occurs

More information

Water mass structure of wind forced river plumes

Water mass structure of wind forced river plumes Water mass structure of wind forced river plumes ROBERT D. HETLAND Department of Oceanography, Texas A&M University, College Station, TX Submitted to J. Physical Oceanogr. on March 15, 2004 Abstract The

More information

The similarity solution for turbulent mixing of two-layer stratified fluid

The similarity solution for turbulent mixing of two-layer stratified fluid Environ Fluid Mech (28) 8:551 56 DOI 1.17/s1652-8-976-5 ORIGINAL ARTICLE The similarity solution for turbulent mixing of two-layer stratified fluid J. A. Whitehead Received: 1 March 28 / Accepted: 29 May

More information

FRICTION-DOMINATED WATER EXCHANGE IN A FLORIDA ESTUARY

FRICTION-DOMINATED WATER EXCHANGE IN A FLORIDA ESTUARY FRICTION-DOMINATED WATER EXCHANGE IN A FLORIDA ESTUARY By KIMBERLY ARNOTT A THESIS PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE

More information

Atrium assisted natural ventilation of multi storey buildings

Atrium assisted natural ventilation of multi storey buildings Atrium assisted natural ventilation of multi storey buildings Ji, Y and Cook, M Title Authors Type URL Published Date 005 Atrium assisted natural ventilation of multi storey buildings Ji, Y and Cook, M

More information

Variations of turbulent flow with river discharge in the Altamaha River Estuary, Georgia

Variations of turbulent flow with river discharge in the Altamaha River Estuary, Georgia JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112,, doi:10.1029/2006jc003763, 2007 Variations of turbulent flow with river discharge in the Altamaha River Estuary, Georgia Daniela Di Iorio 1 and Ki Ryong Kang

More information

Synthesis of TKE Dissipation Rate Observations in the Ocean's Wave Zone

Synthesis of TKE Dissipation Rate Observations in the Ocean's Wave Zone Synthesis of TKE Dissipation Rate Observations in the Ocean's Wave Zone Ayal Anis Department of Oceanography Texas A&M University Galveston, TX 77551 phone: (409) 740-4987 fax: (409) 740-4786 email: anisa@tamug.tamu.edu

More information

Salt intrusion response to changes in tidal amplitude during low river flow in the Modaomen Estuary, China

Salt intrusion response to changes in tidal amplitude during low river flow in the Modaomen Estuary, China IOP Conference Series: Earth and Environmental Science PAPER OPEN ACCESS Salt intrusion response to changes in tidal amplitude during low river flow in the Modaomen Estuary, China To cite this article:

More information

Glider measurements of overturning in a Kelvin-Helmholtz billow train

Glider measurements of overturning in a Kelvin-Helmholtz billow train Journal of Marine Research, 70, 119 140, 2012 Glider measurements of overturning in a Kelvin-Helmholtz billow train by W. D. Smyth 1,2 and S. A. Thorpe 1 ABSTRACT The prospects for glider-based measurement

More information

PHYS 432 Physics of Fluids: Instabilities

PHYS 432 Physics of Fluids: Instabilities PHYS 432 Physics of Fluids: Instabilities 1. Internal gravity waves Background state being perturbed: A stratified fluid in hydrostatic balance. It can be constant density like the ocean or compressible

More information

5. Two-layer Flows in Rotating Channels.

5. Two-layer Flows in Rotating Channels. 5. Two-layer Flows in Rotating Channels. The exchange flow between a marginal sea or estuary and the open ocean is often approximated using two-layer stratification. Two-layer models are most valid when

More information

Structure and Generation of Turbulence at Interfaces Strained by Internal Solitary Waves Propagating Shoreward over the Continental Shelf

Structure and Generation of Turbulence at Interfaces Strained by Internal Solitary Waves Propagating Shoreward over the Continental Shelf Structure and Generation of Turbulence at Interfaces Strained by Internal Solitary Waves Propagating Shoreward over the Continental Shelf J.N. Moum 1, D.M. Farmer 2, W.D. Smyth 1, L. Armi 3, S. Vagle 4

More information

Estimates of Diapycnal Mixing Using LADCP and CTD data from I8S

Estimates of Diapycnal Mixing Using LADCP and CTD data from I8S Estimates of Diapycnal Mixing Using LADCP and CTD data from I8S Kurt L. Polzin, Woods Hole Oceanographic Institute, Woods Hole, MA 02543 and Eric Firing, School of Ocean and Earth Sciences and Technology,

More information

Evolution of an initially turbulent stratified shear layer

Evolution of an initially turbulent stratified shear layer Evolution of an initially turbulent stratified shear layer Kyle A. Brucker and Sutanu Sarkar Citation: Physics of Fluids (1994-present) 19, 105105 (2007); doi: 10.1063/1.2756581 View online: http://dx.doi.org/10.1063/1.2756581

More information

Analysis of Near-Surface Oceanic Measurements Obtained During CBLAST-Low

Analysis of Near-Surface Oceanic Measurements Obtained During CBLAST-Low Analysis of Near-Surface Oceanic Measurements Obtained During CBLAST-Low John H. Trowbridge Woods Hole Oceanographic Institution, MS#12, Woods Hole, MA 02543 phone: (508) 289-2296 fax: (508) 457-2194 email:

More information

Contents. Parti Fundamentals. 1. Introduction. 2. The Coriolis Force. Preface Preface of the First Edition

Contents. Parti Fundamentals. 1. Introduction. 2. The Coriolis Force. Preface Preface of the First Edition Foreword Preface Preface of the First Edition xiii xv xvii Parti Fundamentals 1. Introduction 1.1 Objective 3 1.2 Importance of Geophysical Fluid Dynamics 4 1.3 Distinguishing Attributes of Geophysical

More information

meters, we can re-arrange this expression to give

meters, we can re-arrange this expression to give Turbulence When the Reynolds number becomes sufficiently large, the non-linear term (u ) u in the momentum equation inevitably becomes comparable to other important terms and the flow becomes more complicated.

More information

DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited.

DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. FI N A L R E PO R T : Re l a t i o n s h i p b e t we e n P h y s i c a l a n d B i o l o g i c a l p r o p e r t i e s o n th e M i c r o s c a l e : A c r o s s -com p ari son b et w een D i f f eri

More information

The layered structure in exchange flows between two basins

The layered structure in exchange flows between two basins Int. J. Mar. Sci. Eng., 1(1), 13-22, Autumn 211 IRSEN, CEERS, IAU The layered structure in exchange flows between two basins (Middle and Southern basins of the Caspian Sea) 1* A. A. Bidokhti; 2 A. Shekarbaghani

More information

Prototype Instabilities

Prototype Instabilities Prototype Instabilities David Randall Introduction Broadly speaking, a growing atmospheric disturbance can draw its kinetic energy from two possible sources: the kinetic and available potential energies

More information

Modeling of Coastal Ocean Flow Fields

Modeling of Coastal Ocean Flow Fields Modeling of Coastal Ocean Flow Fields John S. Allen College of Oceanic and Atmospheric Sciences Oregon State University 104 Ocean Admin Building Corvallis, OR 97331-5503 phone: (541) 737-2928 fax: (541)

More information

A modified law-of-the-wall applied to oceanic bottom boundary layers

A modified law-of-the-wall applied to oceanic bottom boundary layers JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 110,, doi:10.1029/2004jc002310, 2005 A modified law-of-the-wall applied to oceanic bottom boundary layers A. Perlin, J. N. Moum, J. M. Klymak, 1 M. D. Levine, T. Boyd,

More information

Roughness Sub Layers John Finnigan, Roger Shaw, Ned Patton, Ian Harman

Roughness Sub Layers John Finnigan, Roger Shaw, Ned Patton, Ian Harman Roughness Sub Layers John Finnigan, Roger Shaw, Ned Patton, Ian Harman 1. Characteristics of the Roughness Sub layer With well understood caveats, the time averaged statistics of flow in the atmospheric

More information

Eddy-Mixed Layer Interactions in the Ocean

Eddy-Mixed Layer Interactions in the Ocean Eddy-Mixed Layer Interactions in the Ocean The MIT Faculty has made this article openly available. Please share how this access benefits you. Your story matters. Citation As Published Publisher Ferrari,

More information

Performance of the Nortek Aquadopp Z-Cell Profiler on a NOAA Surface Buoy

Performance of the Nortek Aquadopp Z-Cell Profiler on a NOAA Surface Buoy Performance of the Nortek Aquadopp Z-Cell Profiler on a NOAA Surface Buoy Eric Siegel NortekUSA Annapolis, USA Rodney Riley & Karen Grissom NOAA National Data Buoy Center Stennis Space Center, USA Abstract-Observations

More information

The Relationship between Flux Coefficient and Entrainment Ratio in Density Currents

The Relationship between Flux Coefficient and Entrainment Ratio in Density Currents DECEMBER 2010 W E L L S E T A L. 2713 The Relationship between Flux Coefficient and Entrainment Ratio in Density Currents MATHEW WELLS University of Toronto, Toronto, Ontario, Canada CLAUDIA CENEDESE Woods

More information

Lecture 9+10: Buoyancy-driven flow, estuarine circulation, river plume, Tidal mixing, internal waves, coastal fronts and biological significance

Lecture 9+10: Buoyancy-driven flow, estuarine circulation, river plume, Tidal mixing, internal waves, coastal fronts and biological significance Lecture 9+10: Buoyancy-driven flow, estuarine circulation, river plume, Tidal mixing, internal waves, coastal fronts and biological significance Thermohaline circulation: the movement of water that takes

More information

Coastal ocean wind fields gauged against the performance of an ocean circulation model

Coastal ocean wind fields gauged against the performance of an ocean circulation model GEOPHYSICAL RESEARCH LETTERS, VOL. 31, L14303, doi:10.1029/2003gl019261, 2004 Coastal ocean wind fields gauged against the performance of an ocean circulation model Ruoying He, 1 Yonggang Liu, 2 and Robert

More information

ENVIRONMENTAL FLUID MECHANICS

ENVIRONMENTAL FLUID MECHANICS ENVIRONMENTAL FLUID MECHANICS Turbulent Jets http://thayer.dartmouth.edu/~cushman/books/efm/chap9.pdf Benoit Cushman-Roisin Thayer School of Engineering Dartmouth College One fluid intruding into another

More information

Direct numerical simulation of salt sheets and turbulence in a double-diffusive shear layer

Direct numerical simulation of salt sheets and turbulence in a double-diffusive shear layer Direct numerical simulation of salt sheets and turbulence in a double-diffusive shear layer Satoshi Kimura, William Smyth July 3, 00 Abstract We describe three-dimensional direct numerical simulations

More information

Goals of this Chapter

Goals of this Chapter Waves in the Atmosphere and Oceans Restoring Force Conservation of potential temperature in the presence of positive static stability internal gravity waves Conservation of potential vorticity in the presence

More information

Turbulence is a ubiquitous phenomenon in environmental fluid mechanics that dramatically affects flow structure and mixing.

Turbulence is a ubiquitous phenomenon in environmental fluid mechanics that dramatically affects flow structure and mixing. Turbulence is a ubiquitous phenomenon in environmental fluid mechanics that dramatically affects flow structure and mixing. Thus, it is very important to form both a conceptual understanding and a quantitative

More information