Chapter 5 Frictional Boundary Layers

Total Page:16

File Type:pdf, Size:1020Kb

Chapter 5 Frictional Boundary Layers Chapter 5 Frictional boundary layers 5.1 The Ekman layer problem over a solid surface In this chapter we will take up the important question of the role of friction, especially in the case when the friction is relatively small (and we will have to find an objective measure of what we mean by small). As we noted in the last chapter, the no-slip boundary condition has to be satisfied no matter how small friction is but ignoring friction lowers the spatial order of the Navier Stokes equations and makes the satisfaction of the boundary condition impossible. What is the resolution of this fundamental perplexity? At the same time, the examination of this basic fluid mechanical question allows us to investigate a physical phenomenon of great importance to both meteorology and oceanography, the frictional boundary layer in a rotating fluid, called the Ekman Layer. The historical background of this development is very interesting, partly because of the fascinating people involved. Ekman (1874-1954) was a student of the great Norwegian meteorologist, Vilhelm Bjerknes, (himself the father of Jacques Bjerknes who did so much to understand the nature of the Southern Oscillation). Vilhelm Bjerknes, who was the first to seriously attempt to formulate meteorology as a problem in fluid mechanics, was a student of his own father Christian Bjerknes, the physicist who in turn worked with Hertz who was the first to demonstrate the correctness of Maxwell’s formulation of electrodynamics. So, we are part of a joined sequence of scientists going back to the great days of classical physics. Another intriguing figure in the story is the great Norwegian explorer/scientist Fridtjof Nansen. Trained as a botanist, a heroic arctic explorer, Nansen was a larger than life figure who played a leading role in Norway’s independence movement (from Sweden) and later, working under the auspices of the League of Nations was involved in many of the early 20th centuries most difficult human dramas. His biography is well worth reading. His relation to our study is more homely. He came to see V. Bjerknes to describe to him some unusual observations of ice-drift on one of Nansen’s expeditions on the Norwegian vessel , the Fram. Nansen noted that the ice drifted at an angle to the wind rather than directly downwind. He realized that it must be the earth’s rotation that was the cause and he asked Bjerknes if one of his students could look into the problem. Ekman was chosen, solved the problem for his doctoral thesis ( and indeed his solution was more general and elaborate that the one presented here and usually associated with his name.) This work was done in 1902, more than 100 years ago and is one of the very first boundary layer problems worked out in fluid mechanics and is essentially contemporary with the work of Prandtl who confronted a similar issue for nonrotating fluids. Now, although we have not completed our formulation of the equations of fluid mechanics in the general case, we do have a complete set of equations for the case where the density can be considered a constant. This will give us a pretty good picture of the phenomenon of interest and we will return to the problem of the general formulation afterwards and be able to check how sensible our assumption of constant density is. The problem to be discussed is the shown in Figure 5.1.1 z U g Ω y x Figure 5.1.1 The fluid flows in the x direction with speed U far from a solid wall at z=0. 2 Chapter 5 Far from a solid, horizontal plane which we idealize as having infinite extent, a uniform flow in the x direction has the magnitude U. What can we say about the structure of the resulting flow? We will use a Cartesian coordinate frame and use the notation (x,y,z) for (x1,x2, x3) and the associated components of velocity will be (u, v, w). ! Since the density is constant the continuity equation is simply !iu = 0 or !u !v !w + + = 0 (5.1.1) !x !y !z while the momentum equation for this incompressible fluid, ! du ! ! 1 ! + 2! " u = # %p # gkˆ + &%2u (5.1.2) dt $ where k is a unit vector in the vertical direction, along the z axis and parallel to the rotation axis. The sum of the gravitational and centrifugal accelerations is written as g , the effective acceleration due to gravity and yields a force anti-parallel to the rotation. Recall that ν is the kinematic viscosity, µ/ρ. The three components of (5.1.2) are !u 1 + uu + vu + wu " fv = " p + $(u + u + u ), !t x y z # x xx yy zz !v 1 + uv + vv + wv + fu = " p + $(v + v + v ), (5.1.3 a b, c) !t x y z # y xx yy zz !w 1 + uw + vw + ww = " p " g + $(w + w + w ), !t x y z # z xx yy zz Here I have used the notation that subscripts denote partial differentiation and f = 2! is the Coriolis parameter. Let’s look for steady solutions so that the local time derivative is zero. Furthermore, since the velocity at infinity is independent of x and y, let’s look for solutions that are independent of x and y. That would imply, from (5.1.1) that !w / !z = 0 . But since w 3 Chapter 5 must vanish on the lower solid surface, w must be zero everywhere. So, with u and v independent of x and y and with w zero, (5.1.3) reduces to: 1 ! fv = ! p + #u , " x zz 1 fu = ! p + #v , (5.1.4 a, b, c) " y zz 1 0= ! p ! g, " z Note that these equations are linear, a very great simplification. Now we have to see whether we can find solutions satisfying the boundary conditions. One solution to the equations is a rather simple one, u = U, v = 0, (5.1.5 a, b, c) p = !"gz ! " fUy. You can check that this satisfies the equations exactly. The flow is, everywhere, the constant velocity in the x direction, as at large z and the pressure field balances the hydrostatic force in the vertical and the Coriolis acceleration in the horizontal. y low pressure Ω U x high pressure pressure force Figure 5.1.2 The balance between the pressure gradient and the Coriolis acceleration in the solution ((5.1.5). The principal and inescapable problem with this solution is that it does not satisfy the no- slip boundary condition at z=0 where both u and v should be zero. The solution (5.1.5) 4 Chapter 5 satisfies the equation, even with the viscous term, but the solution is wrong because it does not satisfy the boundary condition and it is wrong no matter how small µ might be. So, what is the relation between (5.1.5) , that we would naively imagine to be a good solution for small viscosity, and the correct solution that satisfies the boundary condition? Let us return to (5.1.4 a, b, c) and find solutions for the velocity components u and v that are functions of z. Note that the presence of rotation then implies that although the flow is in the x direction far from the wall, the frictional term in (5.1.4a) will force a flow in the y direction under the influence of rotation. An x or y derivative of (5.1.4 c) shows immediately that ! " !p% ! " !p% $ ' = = 0 (5.1.6) !z # !x & !z #$ !y &' so that the horizontal pressure gradients are independent of z. As z ! the horizontal pressure gradient balances the Coriolis acceleration of the uniform flow U, so, !p !p = "# fU, = 0 (5.1.7 a, b) !y !x at infinity, and by (5.1.6) this is true everywhere in the flow, even down to the lower surface. Thus, (5.1.4 a, b) are just ! fv = "uzz , (5.1.8 a,b) fu = fU + "vzz So that finally, we have two simple, ordinary differential equations for the velocity components u and v. If we eliminate u between the two equations by taking 2 z derivatives of the second equation, we obtain, 5 Chapter 5 d 4v 4 + " = 0, dz4 ! 4 (5.1.9 a,b) 1/2 # 2" & ! = $% f '( The quantity δ is the Ekman layer thickness and it depends on the kinematic viscosity and the rotation. For very small viscosity, or rapid rotation, the thickness becomes smaller and smaller. The four independent solutions o f(5.1.9 a, b ) are: !z/" !z/" z/" z/" v = C1e sin z / " + C2e cos z / " + C3e sin z / " + C4e cos z / " (5.1.10 a) while from (5.1.8b) !z/" !z/" z/" z/" u = U ! C1e cos z / " + C2e sin z / " + C3e cos z / " ! C4e sin z / " (5.1.10b) Our boundary condition as z∞ is that (u,v)(U,0). To satisfy this we clearly have to take C3 = C4 = 0 . At z =0, both u and v are zero. This determines the remaining constants, C1 = U, C2 = 0 so that our final solution is, u = U(1! e!z/" cos z / "), (5.1.11 a,b) v = Ue!z/" sin z / " When z >> ! u approaches U exponentially rapidly and v goes to zero. That is, outside a region of O(! ) the solution approaches (5.1.5 a, b, c) which is the solution we would obtain completely ignoring friction. The effect of friction is limited to a region of O(! ) near the solid boundary; this is the Ekman boundary layer.
Recommended publications
  • The Influence of Magnetic Fields in Planetary Dynamo Models
    Earth and Planetary Science Letters 333–334 (2012) 9–20 Contents lists available at SciVerse ScienceDirect Earth and Planetary Science Letters journal homepage: www.elsevier.com/locate/epsl The influence of magnetic fields in planetary dynamo models Krista M. Soderlund a,n, Eric M. King b, Jonathan M. Aurnou a a Department of Earth and Space Sciences, University of California, Los Angeles, CA 90095, USA b Department of Earth and Planetary Science, University of California, Berkeley, CA 94720, USA article info abstract Article history: The magnetic fields of planets and stars are thought to play an important role in the fluid motions Received 7 November 2011 responsible for their field generation, as magnetic energy is ultimately derived from kinetic energy. We Received in revised form investigate the influence of magnetic fields on convective dynamo models by contrasting them with 27 March 2012 non-magnetic, but otherwise identical, simulations. This survey considers models with Prandtl number Accepted 29 March 2012 Pr¼1; magnetic Prandtl numbers up to Pm¼5; Ekman numbers in the range 10À3 ZEZ10À5; and Editor: T. Spohn Rayleigh numbers from near onset to more than 1000 times critical. Two major points are addressed in this letter. First, we find that the characteristics of convection, Keywords: including convective flow structures and speeds as well as heat transfer efficiency, are not strongly core convection affected by the presence of magnetic fields in most of our models. While Lorentz forces must alter the geodynamo flow to limit the amplitude of magnetic field growth, we find that dynamo action does not necessitate a planetary dynamos dynamo models significant change to the overall flow field.
    [Show full text]
  • Brief History of the Early Development of Theoretical and Experimental Fluid Dynamics
    Brief History of the Early Development of Theoretical and Experimental Fluid Dynamics John D. Anderson Jr. Aeronautics Division, National Air and Space Museum, Smithsonian Institution, Washington, DC, USA 1 INTRODUCTION 1 Introduction 1 2 Early Greek Science: Aristotle and Archimedes 2 As you read these words, there are millions of modern engi- neering devices in operation that depend in part, or in total, 3 DA Vinci’s Fluid Dynamics 2 on the understanding of fluid dynamics – airplanes in flight, 4 The Velocity-Squared Law 3 ships at sea, automobiles on the road, mechanical biomedi- 5 Newton and the Sine-Squared Law 5 cal devices, and so on. In the modern world, we sometimes take these devices for granted. However, it is important to 6 Daniel Bernoulli and the Pressure-Velocity pause for a moment and realize that each of these machines Concept 7 is a miracle in modern engineering fluid dynamics wherein 7 Henri Pitot and the Invention of the Pitot Tube 9 many diverse fundamental laws of nature are harnessed and 8 The High Noon of Eighteenth Century Fluid combined in a useful fashion so as to produce a safe, efficient, Dynamics – Leonhard Euler and the Governing and effective machine. Indeed, the sight of an airplane flying Equations of Inviscid Fluid Motion 10 overhead typifies the laws of aerodynamics in action, and it 9 Inclusion of Friction in Theoretical Fluid is easy to forget that just two centuries ago, these laws were Dynamics: the Works of Navier and Stokes 11 so mysterious, unknown or misunderstood as to preclude a flying machine from even lifting off the ground; let alone 10 Osborne Reynolds: Understanding Turbulent successfully flying through the air.
    [Show full text]
  • In Classical Fluid Dynamics, a Boundary Layer Is the Layer I
    Atm S 547 Boundary Layer Meteorology Bretherton Lecture 1 Scope of Boundary Layer (BL) Meteorology (Garratt, Ch. 1) In classical fluid dynamics, a boundary layer is the layer in a nearly inviscid fluid next to a surface in which frictional drag associated with that surface is significant (term introduced by Prandtl, 1905). Such boundary layers can be laminar or turbulent, and are often only mm thick. In atmospheric science, a similar definition is useful. The atmospheric boundary layer (ABL, sometimes called P[lanetary] BL) is the layer of fluid directly above the Earth’s surface in which significant fluxes of momentum, heat and/or moisture are carried by turbulent motions whose horizontal and vertical scales are on the order of the boundary layer depth, and whose circulation timescale is a few hours or less (Garratt, p. 1). A similar definition works for the ocean. The complexity of this definition is due to several complications compared to classical aerodynamics. i) Surface heat exchange can lead to thermal convection ii) Moisture and effects on convection iii) Earth’s rotation iv) Complex surface characteristics and topography. BL is assumed to encompass surface-driven dry convection. Most workers (but not all) include shallow cumulus in BL, but deep precipitating cumuli are usually excluded from scope of BLM due to longer time for most air to recirculate back from clouds into contact with surface. Air-surface exchange BLM also traditionally includes the study of fluxes of heat, moisture and momentum between the atmosphere and the underlying surface, and how to characterize surfaces so as to predict these fluxes (roughness, thermal and moisture fluxes, radiative characteristics).
    [Show full text]
  • Rotating Fluids
    26 Rotating fluids The conductor of a carousel knows about fictitious forces. Moving from horse to horse while collecting tickets, he not only has to fight the centrifugal force trying to kick him off, but also has to deal with the dizzying sideways Coriolis force. On a typical carousel with a five meter radius and turning once every six seconds, the centrifugal force is strongest at the rim where it amounts to about 50% of gravity. Walking across the carousel at a normal speed of one meter per second, the conductor experiences a Coriolis force of about 20% of gravity. Provided the carousel turns anticlockwise seen from above, as most carousels seem to do, the Coriolis force always pulls the conductor off his course to the right. The conductor seems to prefer to move from horse to horse against the rotation, and this is quite understandable, since the Coriolis force then counteracts the centrifugal force. The whole world is a carousel, and not only in the metaphorical sense. The centrifugal force on Earth acts like a cylindrical antigravity field, reducing gravity at the equator by 0.3%. This is hardly a worry, unless you have to adjust Olympic records for geographic latitude. The Coriolis force is even less noticeable at Olympic speeds. You have to move as fast as a jet aircraft for it to amount to 0.3% of a percent of gravity. Weather systems and sea currents are so huge and move so slowly compared to Earth’s local rotation speed that the weak Coriolis force can become a major player in their dynamics.
    [Show full text]
  • Meteorology – Lecture 19
    Meteorology – Lecture 19 Robert Fovell [email protected] 1 Important notes • These slides show some figures and videos prepared by Robert G. Fovell (RGF) for his “Meteorology” course, published by The Great Courses (TGC). Unless otherwise identified, they were created by RGF. • In some cases, the figures employed in the course video are different from what I present here, but these were the figures I provided to TGC at the time the course was taped. • These figures are intended to supplement the videos, in order to facilitate understanding of the concepts discussed in the course. These slide shows cannot, and are not intended to, replace the course itself and are not expected to be understandable in isolation. • Accordingly, these presentations do not represent a summary of each lecture, and neither do they contain each lecture’s full content. 2 Animations linked in the PowerPoint version of these slides may also be found here: http://people.atmos.ucla.edu/fovell/meteo/ 3 Mesoscale convective systems (MCSs) and drylines 4 This map shows a dryline that formed in Texas during April 2000. The dryline is indicated by unfilled half-circles in orange, pointing at the more moist air. We see little T contrast but very large TD change. Dew points drop from 68F to 29F -- huge decrease in humidity 5 Animation 6 Supercell thunderstorms 7 The secret ingredient for supercells is large amounts of vertical wind shear. CAPE is necessary but sufficient shear is essential. It is shear that makes the difference between an ordinary multicellular thunderstorm and the rotating supercell. The shear implies rotation.
    [Show full text]
  • Rotating Convection-Driven Dynamos at Low Ekman Number
    PHYSICAL REVIEW E 66, 056308 ͑2002͒ Rotating convection-driven dynamos at low Ekman number Jon Rotvig* and Chris A. Jones School of Mathematical Sciences, University of Exeter, Exeter EX4 4QE, England ͑Received 10 October 2001; revised manuscript received 18 July 2002; published 22 November 2002͒ We present a fully 3D self-consistent convection-driven dynamo model with reference to the geodynamo. A relatively low Ekman number regime is reached, with the aim of investigating the dynamical behavior at low viscosity. This regime is computationally very demanding, which has prompted us to adopt a plane layer model with an inclined rotation vector, and to make use of efficiently parallelized code. No hyperdiffusion is used, all diffusive operators are in the classical form. Our model has infinite Prandtl number, a Rayleigh number that scales as EϪ1/3 (E being the Ekman number͒, and a constant Roberts number. The optimized model allows us to study dynamos with Ekman numbers in the range ͓10Ϫ5,10Ϫ4͔. In this regime we find strong-field dynamos where the induced magnetic fields satisfy Taylor’s constraint to good accuracy. The solutions are characterized by ͑i͒ a MAC balance within the bulk, i.e., Coriolis, pressure, Lorentz, and buoyancy forces are of comparable magnitude, while viscous forces are only significant in thin boundary layers, ͑ii͒ the Elsasser number is O(10), ͑iii͒ the strong magnetic fields cannot prevent small-scale structures from becoming dominant over the large- scale components, ͑iv͒ the Taylor-Proudman effect is detectable, ͑v͒ the Taylorization decreases as the Ekman number is lowered, and ͑vi͒ the ageostrophic velocity component makes up 80% of the flow.
    [Show full text]
  • The Atmospheric Boundary Layer (ABL Or PBL)
    The Atmospheric Boundary Layer (ABL or PBL) • The layer of fluid directly above the Earth’s surface in which significant fluxes of momentum, heat and/or moisture are carried by turbulent motions whose horizontal and vertical scales are on the order of the boundary layer depth, and whose circulation timescale is a few hours or less (Garratt, p. 1). A similar definition works for the ocean. • The complexity of this definition is due to several complications compared to classical aerodynamics: i) Surface heat exchange can lead to thermal convection ii) Moisture and effects on convection iii) Earth’s rotation iv) Complex surface characteristics and topography. Atm S 547 Lecture 1, Slide 1 Sublayers of the atmospheric boundary layer Atm S 547 Lecture 1, Slide 2 Applications and Relevance of BLM i) Climate simulation and NWP ii) Air Pollution and Urban Meteorology iii) Agricultural meteorology iv) Aviation v) Remote Sensing vi) Military Atm S 547 Lecture 1, Slide 3 History of Boundary-Layer Meteorology 1900 – 1910 Development of laminar boundary layer theory for aerodynamics, starting with a seminal paper of Prandtl (1904). Ekman (1905,1906) develops his theory of laminar Ekman layer. 1910 – 1940 Taylor develops basic methods for examining and understanding turbulent mixing Mixing length theory, eddy diffusivity - von Karman, Prandtl, Lettau 1940 – 1950 Kolmogorov (1941) similarity theory of turbulence 1950 – 1960 Buoyancy effects on surface layer (Monin and Obuhkov, 1954). Early field experiments (e. g. Great Plains Expt. of 1953) capable of accurate direct turbulent flux measurements 1960 – 1970 The Golden Age of BLM. Accurate observations of a variety of boundary layer types, including convective, stable and trade- cumulus.
    [Show full text]
  • HODOGRAPHS and SEVERE WEATHE B E. W. Mccaul. Jr.. Steve
    HODOGRAPHS AND SEVERE WEATHE B E. W. McCaul. Jr.. Steve Lazarus. Fred Carr One ingredient which is believed to be important in governing the morphology of thunderstorms is the profile of vertical shear in the environment which supports the storms. Numerous observational and theoretical studies have shown that storms which form in weakly-sheared environments tend to have non-steady circulations (multicells), while those that form in strongly-sheared environments can develop steady, persistent circulations (supereells). Although intuition suggests that strong shear can only inhibit the formation of organized updrafts, this tendency is partly offset by the ability of shear flows to generate a variety of hydrodynamical instabilities which allow for the possibility of updraft growth. Forecasters who work in areas of the country prone to outbreaks of severe weather should be prepared to recognize not only the thermodynamic conditions but also the vertical shear conditions favorable for various types of severe storms. Inspection of temperature and moisture soundings can reveal the presence of the necessary thermodynamic instability, while inspection of wind profiles (hodographs) can help diagnose whether conditions are right for the development of tornadoes. Hodographs can be drawn either as plots of u (z) vs. v (z) in Cartesian coordinates or V (z) vs. 8 (z) in cylindrical coordinates. In either case, the actual shape of the hodograph will be the same for any given wind profile. On the OU Meteorology weather computer, hodographs from any raob station may be plotted using "mchodo" in GEMPAK. Each plotted hodograph represents nothing more than the curve which connects, in order of ascending altitude, the tips of all the observed wind vectors from a sounding, with the bases of the vectors attached to a common origin.
    [Show full text]
  • Coastal Upwelling Revisited: Ekman, Bakun, and Improved 10.1029/2018JC014187 Upwelling Indices for the U.S
    Journal of Geophysical Research: Oceans RESEARCH ARTICLE Coastal Upwelling Revisited: Ekman, Bakun, and Improved 10.1029/2018JC014187 Upwelling Indices for the U.S. West Coast Key Points: Michael G. Jacox1,2 , Christopher A. Edwards3 , Elliott L. Hazen1 , and Steven J. Bograd1 • New upwelling indices are presented – for the U.S. West Coast (31 47°N) to 1NOAA Southwest Fisheries Science Center, Monterey, CA, USA, 2NOAA Earth System Research Laboratory, Boulder, CO, address shortcomings in historical 3 indices USA, University of California, Santa Cruz, CA, USA • The Coastal Upwelling Transport Index (CUTI) estimates vertical volume transport (i.e., Abstract Coastal upwelling is responsible for thriving marine ecosystems and fisheries that are upwelling/downwelling) disproportionately productive relative to their surface area, particularly in the world’s major eastern • The Biologically Effective Upwelling ’ Transport Index (BEUTI) estimates boundary upwelling systems. Along oceanic eastern boundaries, equatorward wind stress and the Earth s vertical nitrate flux rotation combine to drive a near-surface layer of water offshore, a process called Ekman transport. Similarly, positive wind stress curl drives divergence in the surface Ekman layer and consequently upwelling from Supporting Information: below, a process known as Ekman suction. In both cases, displaced water is replaced by upwelling of relatively • Supporting Information S1 nutrient-rich water from below, which stimulates the growth of microscopic phytoplankton that form the base of the marine food web. Ekman theory is foundational and underlies the calculation of upwelling indices Correspondence to: such as the “Bakun Index” that are ubiquitous in eastern boundary upwelling system studies. While generally M. G. Jacox, fi [email protected] valuable rst-order descriptions, these indices and their underlying theory provide an incomplete picture of coastal upwelling.
    [Show full text]
  • Boundary Layers
    1 I-campus project School-wide Program on Fluid Mechanics Modules on High Reynolds Number Flows K. P. Burr, T. R. Akylas & C. C. Mei CHAPTER TWO TWO-DIMENSIONAL LAMINAR BOUNDARY LAYERS 1 Introduction. When a viscous fluid flows along a fixed impermeable wall, or past the rigid surface of an immersed body, an essential condition is that the velocity at any point on the wall or other fixed surface is zero. The extent to which this condition modifies the general character of the flow depends upon the value of the viscosity. If the body is of streamlined shape and if the viscosity is small without being negligible, the modifying effect appears to be confined within narrow regions adjacent to the solid surfaces; these are called boundary layers. Within such layers the fluid velocity changes rapidly from zero to its main-stream value, and this may imply a steep gradient of shearing stress; as a consequence, not all the viscous terms in the equation of motion will be negligible, even though the viscosity, which they contain as a factor, is itself very small. A more precise criterion for the existence of a well-defined laminar boundary layer is that the Reynolds number should be large, though not so large as to imply a breakdown of the laminar flow. 2 Boundary Layer Governing Equations. In developing a mathematical theory of boundary layers, the first step is to show the existence, as the Reynolds number R tends to infinity, or the kinematic viscosity ν tends to zero, of a limiting form of the equations of motion, different from that obtained by putting ν = 0 in the first place.
    [Show full text]
  • Dynamic Similarity, the Dimensionless Science
    Dynamic similarity, the dimensionless feature science Diogo Bolster, Robert E. Hershberger, and Russell J. Donnelly Dimensional analysis, a framework for drawing physical parallels between systems of disparate scale, affords key insights into natural phenomena too expansive and too energetic to replicate in the lab. Diogo Bolster is an assistant professor of civil engineering and geological sciences at the University of Notre Dame in Notre Dame, Indiana. Robert Hershberger is a research assistant in the department of physics at the University of Oregon in Eugene. Russell Donnelly is a professor of physics at the University of Oregon. Many experiments seem daunting at first glance, in accordance with general relativity, is deflected as it passes owing to the sheer number of physical variables they involve. through the gravitational field of the Sun. Assuming the Sun To design an apparatus that circulates fluid, for instance, one can be treated as a point of mass m and that the ray of light must know how the flow is affected by pressure, by the ap- passes the mass with a distance of closest approach r, dimen- paratus’s dimensions, and by the fluid’s density and viscosity. sional reasoning can help predict the deflection angle θ.1 Complicating matters, those parameters may be temperature Expressed in terms of mass M, length L, and time T, the and pressure dependent. Understanding the role of each variables’ dimensions—denoted with square brackets—are parameter in such a high-dimensional space can be elusive or prohibitively time consuming. Dimensional analysis, a concept historically rooted in Box 1. A brief history of dimensional analysis the field of fluid mechanics, can help to simplify such prob- Going back more than 300 years, discussions of dimensional lems by reducing the number of system parameters.
    [Show full text]
  • A Laboratory Model of Thermocline Depth and Exchange Fluxes Across Circumpolar Fronts*
    656 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 34 A Laboratory Model of Thermocline Depth and Exchange Fluxes across Circumpolar Fronts* CLAUDIA CENEDESE Physical Oceanography Department, Woods Hole Oceanographic Institution, Woods Hole, Massachusetts JOHN MARSHALL Department of Earth, Atmospheric and Planetary Sciences, Massachusetts Institute of Technology, Cambridge, Massachusetts J. A. WHITEHEAD Physical Oceanography Department, Woods Hole Oceanographic Institution, Woods Hole, Massachusetts (Manuscript received 6 January 2003, in ®nal form 7 July 2003) ABSTRACT A laboratory experiment has been constructed to investigate the possibility that the equilibrium depth of a circumpolar front is set by a balance between the rate at which potential energy is created by mechanical and buoyancy forcing and the rate at which it is released by eddies. In a rotating cylindrical tank, the combined action of mechanical and buoyancy forcing builds a strati®cation, creating a large-scale front. At equilibrium, the depth of penetration and strength of the current are then determined by the balance between eddy transport and sources and sinks associated with imposed patterns of Ekman pumping and buoyancy ¯uxes. It is found 2 that the depth of penetration and transport of the front scale like Ï[( fwe)/g9]L and weL , respectively, where we is the Ekman pumping, g9 is the reduced gravity across the front, f is the Coriolis parameter, and L is the width scale of the front. Last, the implications of this study for understanding those processes that set the strati®cation and transport of the Antarctic Circumpolar Current (ACC) are discussed. If the laboratory results scale up to the ACC, they suggest a maximum thermocline depth of approximately h 5 2 km, a zonal current velocity of 4.6 cm s21, and a transport T 5 150 Sv (1 Sv [ 106 m3 s 21), not dissimilar to what is observed.
    [Show full text]