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

Similar documents
Laboratory Studies of Turbulent Mixing

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

Buoyancy Fluxes in a Stratified Fluid

Grid-generated turbulence, drag, internal waves and mixing in stratified fluids

On the Velocity Gradient in Stably Stratified Sheared Flows. Part 2: Observations and Models

Turbulent Mixing Parameterizations for Oceanic Flows and Student Support

Double-diffusive lock-exchange gravity currents

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

Atrium assisted natural ventilation of multi storey buildings

Laboratory experiments on diapycnal mixing in stratified fluids

A process study of tidal mixing over rough topography

Supporting Information for Nonlinear Response of Iceberg Side Melting to Ocean Currents

Mixing Efficiency in a Lock Exchange Experiment

Numerical Simulations of a Stratified Oceanic Bottom Boundary Layer. John R. Taylor - MIT Advisor: Sutanu Sarkar - UCSD

SMAST Technical Report The Performance of a Coupled 1-D Circulation and Bottom Boundary Layer Model with Surface Wave Forcing

Journal of Fluid Science and Technology

Internal Wave Driven Mixing and Transport in the Coastal Ocean

Shear instabilities in a tilting tube

Internal Wave Generation and Scattering from Rough Topography

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

Influence of Lock Aspect Ratio upon the Evolution of an Axisymmetric Intrusion

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

Numerical studies on natural ventilation flow in an enclosure with both buoyancy and wind effects

Mean potential energy change in stratified grid turbulence

Numerical and laboratory generation of internal waves from turbulence

A note on mixing due to surface wave breaking

Submesoscale Routes to Lateral Mixing in the Ocean

n v molecules will pass per unit time through the area from left to

Explicit algebraic Reynolds stress models for internal flows

Sediment resuspension from shoaling internal solitary waves. Daniel Bourgault, Institut des sciences de la mer de Rimouski, Rimouski, QC, Canada

Lecture 25: Ocean circulation: inferences from geostrophic and thermal wind balance

Community Ocean Vertical Mixing (CVMix) Parameterizations

SMS 303: Integrative Marine

VERTICAL TURBULENT BUOYANT HELIUM JET CFD MODELING AND VALIDATION

On the Stratification of Turbulent Mixed Layers

1. Comparison of stability analysis to previous work

Salt of the Sea. Miscellaneous Notes

ES265 Order of Magnitude Phys & Chem Convection

Deep-Water Flow over the Lomonosov Ridge in the Arctic Ocean

The estimation of vertical eddy diffusivity in estuaries

Tilting Shear Layers in Coastal Flows

Physics of Aquatic Systems

Transient natural ventilation of a room with a distributed heat source

Internal Wave Generation in Straits

Dynamics of Atmospheres and Oceans

Investigation of the Behavior of a Jet Issued into Two-Layer Density-Stratified Fluid

Donald Slinn, Murray D. Levine

Detrainment Fluxes for Multi-Phase Plumes in Quiescent Stratification

Ocean Dynamics. Stability in the water column

Testing Turbulence Closure Models Against Oceanic Turbulence Measurements

Prototype Instabilities

Community Ocean Vertical Mixing (CVMix) Parameterizations

meters, we can re-arrange this expression to give

Generation and Evolution of Internal Waves in Luzon Strait

Physical Oceanographic Context of Seamounts. Pierre Dutrieux Department of Oceanography, University of Hawai'i at Manoa, Honolulu, Hawai'i

Turbulence and Energy Transfer in Strongly-Stratified Flows

Collapse of turbulence in atmospheric flows turbulent potential energy budgets in a stably stratified couette flow

( ) = 1005 J kg 1 K 1 ;

Tutorial for the heated pipe with constant fluid properties in STAR-CCM+

The atmospheric boundary layer: Where the atmosphere meets the surface. The atmospheric boundary layer:

Tide-Topography Interactions: Asymmetries in Internal Wave Generation due to Surface Trapped Currents

7. Basics of Turbulent Flow Figure 1.

Internal boundary layers in the ocean circulation

2.3 The Turbulent Flat Plate Boundary Layer

Turbulence in Strongly-Stratified Flows

Dynamics of turbulence under the effect of stratification and internal waves

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

SUPPLEMENTARY INFORMATION

A two-fluid model of turbulent two-phase flow for simulating turbulent stratified flows

Phase Analysis of the Stretching Cycles of the Head of Unsteady Gravity Currents Developing over Smooth and Rough Beds

5. Two-layer Flows in Rotating Channels.

Chapter 10: Boiling and Condensation 1. Based on lecture by Yoav Peles, Mech. Aero. Nuc. Eng., RPI.

On the turbulent Prandtl number in homogeneous stably stratified turbulence

Diffusion: Molecular and Turbulent. K = the molecular diffusion coefficient, and = (1.98 x 10-5 cm 2 /sec) for oxygen dissolved in water,

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

METHODOLOGY (3) where, x o is the heat source separation and α is the. entrainment coefficient α.

Differential diffusion: an oceanographic primer

Estimates of Diapycnal Mixing Using LADCP and CTD data from I8S

Inertial Range Dynamics in Density-Stratified Turbulent Flows

GFD 2012 Lecture 1: Dynamics of Coherent Structures and their Impact on Transport and Predictability

Stability of meridionally-flowing grounded abyssal currents in the ocean

Local generation of internal solitary waves in an oceanic pycnocline

Convection Induced by Cooling at One Side Wall in Two-Dimensional Non-Rotating Fluid Applicability to the Deep Pacific Circulation

Prediction of solar activity cycles by assimilating sunspot data into a dynamo model

PHYS 432 Physics of Fluids: Instabilities

ScienceDirect. Buoyancy driven turbulence of moist air

Parameterization of turbulent fluxes and scales using homogeneous sheared stably stratified turbulence simulations

C C C C 2 C 2 C 2 C + u + v + (w + w P ) = D t x y z X. (1a) y 2 + D Z. z 2

NONLINEAR FEATURES IN EXPLICIT ALGEBRAIC MODELS FOR TURBULENT FLOWS WITH ACTIVE SCALARS

EXAMPLE SHEET FOR TOPIC 3 AUTUMN 2013

15. Physics of Sediment Transport William Wilcock

OCEAN MODELING II. Parameterizations

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

A note concerning the Lighthill sandwich model of tropical cyclones

Laboratory experiments on the effect of baroclinic eddies on a dense plume in a rotating stratified fluid

The Impact of Topographic Steepness on Tidal Dissipation at Bumpy Topography

The Effect of Endplates on Rectangular Jets of Different Aspect Ratios

Applying Gerris to Mixing and Sedimentation in Estuaries

Chapter 1. Governing Equations of GFD. 1.1 Mass continuity

Transcription:

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 28 / Published online: 19 June 28 The Author(s) 28 Abstract Experiments are reviewed in which a two-layer salt-stratified tank of water was mixed by turbulence. The density profile began as a single step and evolved to a smooth mixed profile. The turbulence was generated by many excursions of a horiontally moving vertical rod with Richardson number Ri >.9 and Reynolds Number Re > 6. There was almost perfect collapse of all the profiles to one universal profile as a function of a similarity variable. We develop a theoretical model for a simple mixing law with a buoyancy flux that is a function of internal Richardson number Rii. A similarity equation is found. A flux law that increases with small Rii and decreases with large Rii is considered next. Since no analytical solution is known, the similarity concept is tested by numerically integrating the equations in space and time. With buoyancy flux monotonically increasing with internal Richardson number, the similarity approach is valid for a profile starting from a slightly smoothed step. However, a shock forms for a mixing law with higher initial Rii (so that buoyancy flux decreases with Richardson number) and the similarity approach is invalid for those initial conditions. Keywords Turbulence Mixing Stratified Similarity solution Layered fluid 1 Introduction This paper reviews some experimental results obtained by Ian Stevenson [1]. The work itself involves mixing by turbulence that is produced uniformly throughout all depths in layered water. The experiment is motivated by the fact that strong currents spontaneously develop near saddle points in the ocean floor [2,3]. They exhibit high values of shear and mixing which has renewed our interest in mixing within layered fluid. Many saddle points are distributed throughout all depths of the ocean because of the complex topography of ocean J. A. Whitehead (B) Department of Physical Oceanography, MS#21, Woods Hole Oceanographic Institution, Wood Hole, MA 2543, USA e-mail: jwhitehead@whoi.edu

552 Environ Fluid Mech (28) 8:551 56 ridges, therefore, the interest here is to have turbulence produced equally at all depths. The density field evolves as it is stirred by turbulence starting from two layers of water of differing salinity. There are two governing dimensionless parameters: a Reynolds number Re = Ud/ν for the turbulence, and either an initial Richardson number Ri = g ρd/ρ U 2 for layered fluid, or an internal and local Richardson number Rii = N 2 d 2 /U 2 for smoothly stratified fluid. In these formulas, the acceleration of gravity is g, ρ is the initial density difference of two layers, ρ is horiontally averaged density of the fluid, ρ is its average throughout the depth of the tank, d is the length scale of the turbulence source, U is the speed of the turbulence source, ν is the viscosity of the fluid, and N 2 = (g/ρ ) dρ/d is local stratification frequency. Molecular diffusivity of the salt is neglected by assuming it is not of first order importance to the mixing dynamics if Reynolds number is high enough. The experiment extends Linden s experiments [4] where a grid is dropped through two layers with differing salinity. The turbulence in the wake mixed the fluid. Buoyancy flux (gravity times density flux) was easily estimated by measuring the evolution of the density field. Here, we generate many such stirring events instead of only one and then we document the evolution of the density field using density profiles between successive stirring events. Using a similarity variable, the evolving density profile collapses to one profile. We then describe the implications in terms of a theoretical model in slightly more general terms than discussed previously. To our knowledge, this collapse for turbulent mixing is not previously reported. 2 The experiment An acrylic tank that is.6 m high and with a horiontal cross-section.1 m.2 m was initially filled with approximately.25 m of salt water (NaCl solution, density ρ varied in the 11 experiments from 1, 3 mg/m 3 to 1, 2 kg/m 3 ). Then, an equal thickness of fresh water was carefully added using a float-mechanism. This procedure created two distinct layers of differing density with a sharp interface between them. Both layers were close to room temperature so that heat transfer effects between the layers could be neglected. A descending conductivity probe measured density profiles at a speed of.3 m/s. The sharpness of the initial density profiles of all the runs showed that this value of probe speed gave excellent fine-scale resolution of the profile. A mixing rod traversed back and forth along the wide direction across a.2 m long region, and the wake produced relatively uniform turbulence throughout the depth of the tank. The mixer was started after we recorded the initial density profile. After a certain number of rod excursions back and forth, mixing stopped and after a 2 min settling period, another profile was taken. Then, the same number of mixing excursions followed by a profile continued cyclically. A total of up to 3 rod excursions constituted an experiment. Rod diameter d is the primary imposed length for the turbulence and therefore we take this to be the length scale for turbulence in this experiment. The values of d used were 1.27, 1.8, and 2.68 1 2 m, and the range of the values of U was.5.71 m/s. Using the values for reference density and viscosity at 2 C, the 11 experimental runs had initial Richardson numbers ranging between.93 < Ri < 4.84 and Reynolds numbers extending between 638 < Re < 1765. At the beginning of a typical run, the measured profile almost exactly duplicates a step profile and because of the mixing subsequent profiles are more spread out. Correspondingly, the initial local values of Rii (proportional to vertical derivative) are very large in the vicinity of the interface and virtually ero in each layer. After mixing starts, the maximum value of Rii at the interface decreases for each excursion, and Rii is measurable from the data throughout

Environ Fluid Mech (28) 8:551 56 553.6.4.2 (m) -.2 -.4 -.6 1 15 11 115 12 125 13 135 ρ (kg/m3) Fig. 1 Density profiles [1] throughout a typical run at equal mixing intervals (The first profile is a thick dashed curve and the last is a thin dashed curve.) the fluid. A typical sie of Rii for each experiment can be estimated as Rii = d Ri/D, where D is water depth (.5 m). The range for these experiments is.24 < d Ri/D <.174. These values caused us to expect that the experiment lies outside the range Rii > 1 where layer formation was found in previous experiments [5,6]. Each run produced about 8 profiles, with a typical set shown in Fig. 1. Drift of the calibration constants that convert voltage to density was inconsequential for the duration of all runs. Instrumental noise was small (of order 1 4 of the maximum range measured) but since an analysis of buoyancy flux required calculation of vertical derivative and the subtraction of two records (see next section), the scatter of the differences between closely spaced points or sequential profiles was considerable. The density values had precision better than.1%. An error also arises from small drifts in the position of the probe. To account for these errors we centered the profiles by adding a small uniform shift, a common practice in these types of experiments [1,4 6]. The centered profiles were averaged over 5 points taken at.1 m intervals to smooth noise, and the centered density profiles yielded values of density as a function of depth, from which we calculated buoyancy frequency and from subsequent profiles determined buoyancy flux and change in potential energy. In the experiment, time is taken to be linearly proportional to the number of excursions. All our laboratory experiments showed strikingly good data collapse using the similarity variable of depth divided by the square root of the number of excursions n. The collapse for actual data is remarkable (Figs. 2 and 3). Thirteen runs in all exhibited similar collapse, although not all are as good as these examples (probably from experimental effects such as not exactly equal layer depths, or slight effects from room temperature influences, for example). Stevenson found that the first few profiles had apparently not yet adjusted to the exact experimental similarity profile shape. Because of this, the best collapse excluded the first few

554 Environ Fluid Mech (28) 8:551 56.5.5 (a).4.3.3.2.2.1.1 (m) -.1 n1/2 (m) -.1 -.2 -.2 -.3 -.3 -.4 -.4 -.5 1 11 12 -.5 1 13.5 12 13 (d).4.3.3.2.2.1.1 -.1 n1/2 (m) -.1 -.2 -.2 -.3 -.3 -.4 -.4 -.5 1 11.5 (c).4 (m) (b).4 11 12 -.5 1 13 11 12 13 3 ρ (kg/m ) Fig. 2 Selected laboratory experimental profiles (left) and their collapse [1] when divided by the square root of the number of traverses (right). The traverse numbers for panels (a) and (b) are: 2 (thin line); 41 84 (cluster of lines); 3 (thick line). The traverse numbers for panels (c) and (d) are: 14 (thin line); 4 6 (cluster of lines); and 336 (thick line) profiles. We also saw poorer fit to a perfect collapse for some of the final profiles. Possibly those later profiles exhibited effects from the top and bottom boundaries of the chamber. 3 A simple model The conservation of density obeys Fd ρ = t (3.1) with Fd the density flux in the direction, and is positive upward. Whitehead and Stevenson, show results for specific models of Fd. Here we illustrate the general consequences of turbulence driven by a source with velocity and length scales that are independent of depth.

Environ Fluid Mech (28) 8:551 56 555 Fig. 3 Close-up of traverses number 2 (black) and 3 (grey) with depth divided by n.1.5 1/2 n (m) -.5 -.1 1 11 12 13 ρ (kg/m 3 ) Taking a vertical derivative, 2 ρ F d t = 2 2 (3.2) and multiplying by g/ρ the equation is N 2 = g 2 t ρ 2 F d. (3.3) Next, let F d = ρ g UdF (Rii) and make the equation non-dimensional using the time scale d/u and vertical length scale d. The result is Rii t = 2 F (Rii). (3.4) 2 This equation can be converted to an ordinary differential equation using the similarity variable η = 2 t and the governing equation becomes η d dη Rii = 1 F (Rii). (3.5) 2 dη2 We don t know of any analytical solutions for (3.5), but the error function is a well-known solution for (3.1) with such a similarity variable. It is found for the special case of F d = κ ρ (the diffusion equation) with the initial condition of two overlying uniform layers of different density. Equation (3.5) will not have such an initial condition in general for the case of two layers since it applies to a vertical derivative of density rather than to density itself. In spite of this, numerical calculations [1] have found an example for Eq. 3.4 that collapsed to such a similarity variable using the function F (Rii) = D Rii.ThevariableD is an equivalent 1+Rii diffusion for this nonlinear flux law. Those previous calculations 2 started with two overlying d 2

556 Environ Fluid Mech (28) 8:551 56 4 (a) 8 6 (b) 2 4 2 2t 1/2-2 -2-4 -6-4.2.4.6.6.4 (c).8-8.2.4.6.8 1..2 2t 1/2 -.2 -.4 -.6 -.8.2.4.6 Fig. 4 Theoretical density profiles with D =.1 for various dimensionless times (1, 5, 1, 2 (long dashed grey), 4 (thin black dashed), 8 (gray) and 16 (thick black dashed). (a) Profiles versus depth. (b) The same profiles plotted with the similarity variable. (c) Close-up for the four longest times plotted with the similarity variable. These overlap significantly layers of 5 grid points each with an initial region of 1 grid points with smoothed density change between the two layers. Although the results verified the similarity scaling for small D, results for D > 1raiseda question concerning the robustness of the similarity scaling. To be specific, it was mentioned that there was a breakdown to a shock for D > 1. Unfortunately their numerical calculation had only 1 levels and no details of the shock formation were possible. To determine more details of the shock, additional calculations have now been done with a higher resolution for large D. The domain has 1, vertical points. The initial condition

Environ Fluid Mech (28) 8:551 56 557 4 (a) 8 6 (b) 2 4 2 2t 1/2-2 -2-4 -4-6.6.4.2.4.6.8 1. (c) 8.2.4.6.8 1..2 2t 1/2 -.2 -.4 -.6 -.8.2.4.6.8 1. Fig. 5 Theoretical density profiles with D = 2. for various dimensionless times. The dimensionless times and panels are the same as in Fig. 2 is Rii(, t = ) = δ() where δ is a linearly changing distribution of stratification (with a peak in the middle) spread over the central 5 grid points. First, the case with D =.1 shows that, after an initial period, the distribution (Fig. 4a) settles down to agreement with similarity theory (Fig. 4b,c). In contrast, for D = 2 the calculations do not verify a similarity theory. Instead, a shock forms near mid-depth and within the shock are steps that remain froen for long periods of time. Figure 5 shows results for the same parameters as in the panels of Fig. 4 except that D = 2.. The collapse of profiles is better in real time than with scaled time.

1 558 Environ Fluid Mech (28) 8:551 56 Fig. 6 Comparison of calculations and measurements. A similarity error function is shown by large black dashes, the numerical similarity calculation is a wide gray curve, and the experiment curves are the same as in Fig. 2b ρ 4 Application of the similarity scaling to data Next, we compared theory and data [1] by overlaying the collapsed theoretical similarity curve in Fig. 4b and the collapsed data in Fig. 2b. To make the best fit, the depth for the theoretical curve was stretched to have the same slope as the experimental data at mid-depth (Fig. 6). The theory has moderately good agreement with the data but certainly the fit is not perfect. The theory has more curvature away from the mid-point than the data. Therefore, the theoretical curve is not the same as the experiment in detail. Using the similarity variable, the thickness of the transition region between the upper and lower layer in Fig. 2 can be used to infer a value of eddy diffusivity. In dimensional units, the similarity variable is η = /2 dut so that Ud serves as a substitute for the diffusivity. Results show that the region with 9% of the change between the upper and lower layer is found within about η =±.5. Let us compare this to the complementary error function. In that case, we can define the transition layer to lie within about η =±1. For our present experiment to have the same dimensionless thickness, we should use the similarity variable η = /2.5 2 dut. In other words, the experiment shows that the effective diffusivity is about.5 2 Ud = 2.5 1 3 Ud. This is somewhat smaller than the value κ ef f = 6.7 1 3 U M found by Holford and Linden [5]whereM is the grid spacing. Since there is not a direct geometrical correspondence between the rod used in this experiment and the grids used by Holford and Linden, our two results agree as well as can be expected.

Environ Fluid Mech (28) 8:551 56 559 5 (a) 5 (b) d d -5.8-5.4.8.1.2.3.4.5.6.7 Fu Rii (c) Fu.4.1.2.3.4.5.6.7 Rii Fig. 7 (a) Measured flux as a function of depth. The flux is not total flux (see text). (b) Flux as a function of Richardson number, and (c) flux versus Richardson Number Rii from one experimental run. Note that F u monotonically increases with Rii 5 Discussion Buoyancy flux and potential energy evolution were also measured but results are incomplete since the probe was unable to traverse the entire depth of the tank. Buoyancy flux F u was determined by integrating Eq. 1 from the bottom of the measurement region to find F d and then multiplying by gravity g. Thus, it is not the total buoyancy flux through the region, but only a deviation from the flux that comes from the unmeasured deep part of the tank. Generally, the measurement (Fig. 5) showed a monotonic increase of buoyancy flux with Rii, and no decrease in flux at large Rii. The lack of fine structure in these experiments and the numerical model is thus consistent with the regime diagram of Holford and Linden [3], who report nolayers in that range ofrii. Finally, estimates of the amount of turbulent kinetic energy going to mixing processes show that a maximum of approximately 5% of the kinetic energy generation rate by the moving rod noise. No experiments have yet recovered the formation of a shock that is found in the numerical calculations in which F attains a peak value with an increasing Rii and which then decreases for greater value. Moreover, we do not observe any evidence of such a decrease in flux when plotting flux versus stratification from the data. Unfortunately, those measurements are not definitive because the profiles did not extend to the very bottom and top of the tank. In addition, the values of Rii are not large enough to produce layering that is seen in experiments with smooth profiles [5,6].

56 Environ Fluid Mech (28) 8:551 56 6 Conclusions The application of similarity variables to the data from stratified turbulent mixing shows excellent collapse over a wide range of times. Specifically, the collapse of data when depth is divided by the square root of the number of excursions is remarkably good over a span of starting at excursion number 8 and extending to excursion 3. The similarity approach here has been developed subject to the simplification that flux is a monotonically increasing function of Rii alone. These experimental results differ from the results of Holford and Linden [5], who concluded that their experiment required an additional dimensionless number. Finally, this laboratory observation of such a similarity-based approach to turbulently mixed data gives support to the use of scaling in oceanography [7]. The experiment obtained reasonable results but quantitative precision of flux laws was not obtained because the density was not sampled throughout the entire depth of the tank. Acknowledgements Ian Stevenson was supported as a summer student fellow of Woods Hole Oceanographic Institution. The laboratory expenses were supported by WHOI s Ocean and Climate Change Institute. Open Access This article is distributed under the terms of the Creative Commons Attribution Noncommercial License which permits any noncommercial use, distribution, and reproduction in any medium, provided the original author(s) and source are credited. References 1. Whitehead JA, Stevenson I (27) Turbulent mixing of two-layer stratified fluid. Phys Fluids 19:12514. doi:1.163/1.2821913 2. Whitehead JA (1989) Internal hydraulic control in rotating fluids applications to oceans. Geophys Astrophys Fluid Dyn 48:169 192. doi:1.18/39192898219532 3. Thurnherr AM, St. Laurent LC, Speer KG, Toole JM, Ledwell JR (25) Mixing associated with sills in a canyon on the midocean ridge flank. J Phys Oceanogr 35:137 1381. doi:1.1175/jpo2773.1 4. Linden PF (198) Mixing across a density interface produced by grid turbulence. J Fluid Mech 1:691 73. doi:1.117/s221128136x 5. Holford JM, Linden PF (1999) Turbulent mixing in a stratified fluid. Dyn Atmos Oceans 3:173 198. doi:1.116/s377-265(99)25-1 6. Park Y-G, Whitehead JA, Gnanadesikan A (1994) Turbulent mixing in stratified fluids: layer formation and energetics. J Fluid Mech 279:279 311. doi:1.117/s22112943915 7. Ledwell JR, Watson AJ, Law CS (1998) Mixing of a tracer in the pycnocline. J Geophys Res O 13:21499 21529. doi:1.129/98jc1738