Lecture Notes in Fluid Mechanics: from Basics to the Millennium Problem

Total Page:16

File Type:pdf, Size:1020Kb

Lecture Notes in Fluid Mechanics: from Basics to the Millennium Problem Lecture notes in fluid mechanics: From basics to the millennium problem Laurent Schoeffel, CEA Saclay. These lecture notes have been prepared as a first course in fluid mechanics up to the presentation of the millennium problem listed by the Clay Mathematical Institute. Only a good knowledge of classical Newtonian mechanics is assumed. We start the course with an elementary derivation of the equations of ideal fluid mechanics and end up with a discussion of the millennium problem related to real fluids. With this document, our primary goal is to debunk this beautiful problem as much as possible, without assuming any previous knowledge neither in fluid mechanics of real fluids nor in the mathematical formalism of Sobolev’s inequalities. All these items are introduced progressively through the document with a linear increase in the difficulty. Some rigorous proofs of important partial results concerning the millennium problem are presented. At the end, a very modern aspect of fluid mechanics is covered concerning the subtle issue of its application to high energetic hadronic collisions. Lecture notes in fluid mechanics: From basics to the millennium problem / Laurent Schoeffel 1 §1. Introduction §2. Continuum hypothesis §3. Mathematical functions that define the fluid state §4. Limits of the continuum hypothesis §5. Closed set of equations for ideal fluids §6. Boundary conditions for ideal fluids §7. Introduction to nonlinear differential equations §8. Euler’s equations for incompressible ideal fluids §9. Potential flows for ideal fluids §10. Real fluids and Navier-Stokes equations §11. Boundary conditions for real fluids §12. Reynolds number and related properties §13. The millennium problem of the Clay Institute §14. Bounds and partial proofs §15. Fluid mechanics in relativistic Heavy-Ions collisions Notations: Bold letters will represent vectors: for example x is equivalent to the 3 Cartesian coordinates (x, y, z) in 3- dimensional space and to (x, y) in 2-dimensional space. Consequently, we refer as vx, vy and vz the 3 Cartesian components of a vector v. We may also use the notation (v1, v2, v3) or generically (vi) where ‘i’ runs over x, y, z or 1, 2, 3. The norm of v is written as ‖ ‖, with ‖ ‖ v v v ∑ v . For differential operators, we use the standard notations for gradient, divergence and curl (rotational) as explained along the text. In some circumstances, we can also use the nabla notation, when it helps for the clarity of formulae. Lecture notes in fluid mechanics: From basics to the millennium problem / Laurent Schoeffel 2 §1. Introduction Fluid mechanics concerns the study of the motion of fluids (in general liquids and gases) and the forces acting on them. Like any mathematical model of the real world, fluid mechanics makes some basic assumptions about the materials being studied. These assumptions are turned into equations that must be satisfied if the assumptions are to be held true. Modern fluid mechanics, in a well-posed mathematical form, was first formulated in 1755 by Euler for ideal fluids. With the laws of fluid mechanics, it will be possible to solve exactly the following problem: a bubble of radius “a” is formed at initial time in a fluid. How long does it take for the fluid to fill up the bubble, knowing the pressure (P), temperature of the fluid and its mass density? In order to give a first hint to this problem, we can guess that this duration is proportional to the radius or a power of it and is decreasing with P is increasing. Then, we can write a priori the expression of the duration to fill up the bubble as: a . Laws of fluid mechanics will allow us to show that: n 1, 1 2. Interestingly, it can be shown that the laws of fluid mechanics cover more materials than standard liquid and gases. Indeed, the idea of exploiting the laws of ideal fluid mechanics to describe the expansion of the strongly interacting nuclear matter that is formed in high energetic hadronic collisions was proposed in 1953 by Landau. This theory has been developed extensively in the last 60 years and is still an active field of research. This gives a very simple 3-steps picture of a non-trivial phenomenon observed in hot nuclear matter after the collision of high energetic heavy ions, composed of a large collection of charged particles. (i) After the collision a nuclear medium, a zone of high density of charges, is formed with high pressure in the middle (center of the collision). Lecture notes in fluid mechanics: From basics to the millennium problem / Laurent Schoeffel 3 (ii) According to the laws of fluid mechanics, as we shall prove them, this implies that an acceleration field is generated from high pressures to low pressures. (iii) This implies that particles will flow in a certain transverse direction, as indicated on the figure. This is known as the transverse flow property, well established experimentally. We come back on these ideas and their developments in the last section of this document. It requires a relativistic formulation of fluid mechanics. Up to this section, we always assume that the dynamics is non- relativistic. §2. Continuum hypothesis Fluid mechanics is supposed to describe motion of fluids and related phenomena at macroscopic scales, which assumes that a fluid can be regarded as a continuous medium. This means that any small volume element in the fluid is always supposed so large that it still contains a very great number of molecules. Accordingly, when we consider infinitely small elements of volume, we mean very small compared with the volume of the body under consideration, but large compared with the distances between the molecules. The expressions fluid particle and point in a fluid are to be understood in this sense. That is, properties such as density, pressure, temperature, and velocity are taken to be well-defined at infinitely small points. These properties are then assumed to vary continuously and smoothly from one point to another. Consequently, the fact that the fluid is made up of discrete molecules is ignored. If, for example, we deal with the displacement of some fluid particle, we do mean not the displacement of an individual molecule, but that of a volume element containing many molecules, though still regarded as a point in space. That’s why fluid mechanics is a branch of continuum mechanics, which models matter from a macroscopic viewpoint without using the information that it is made out of molecules (microscopic viewpoint). Exercise: Recall the orders of magnitude for the scales in a typical fluid. Verify the continuum hypothesis. Solution: We note L* the molecular size scale, characterized by the mean free path of molecules between collisions, d the scale of a fluid point (fluid particle) and L the scale of the fluid volume considered. In general, the orders of magnitude are: L*=10nm, d=1m and L=1m. Then, we verify the relation: L*<<d<<L. Lecture notes in fluid mechanics: From basics to the millennium problem / Laurent Schoeffel 4 §3. Mathematical functions that define the fluid state Following the continuous assumption, the mathematical description of the state of a moving fluid can be characterized by functions of the coordinates x, y, z and of the time t. These functions of (x, y, z, t) are related to the quantities defined for the fluid at a given point (x, y, z) in space and at a given time t, which refers to fixed points in space and not to fixed particles of the fluid. For example, we can consider the mean local velocity v(x, y, z, t) of fluid particles or fluid points, called the fluid velocity, and two thermodynamic quantities that characterize the fluid state, the pressure p(x, y, z, t) and the mass density (x, y, z, t), the mass per unit volume of fluid. Following the discussion of §2, two remarks can be done at this stage: i. The fluid is assumed to be a continuum. This implies that all space-time derivatives of all dependent variables exist in some reasonable sense. In other words, local properties such as density pressure and velocity are defined as averages over elements large compared with the microscopic structure of the fluid but small enough in comparison with the scale of the macroscopic. This allows the use of differential calculus to describe such a system. ii. All the thermodynamic quantities are determined by the values of any two of them, together with the equation of state. Therefore, if we are given five quantities, namely the three components of the velocity v, the pressure p and the mass density , the state of the moving fluid is completely determined. We recall that only if the fluid is close to thermodynamic equilibrium, its thermodynamic properties, such as pressure, density, temperature are well-defined. This requires (as a very former hypothesis) that local relaxation times towards thermal equilibrium are much shorter than any macroscopic dynamical time scale. In particular, microscopic collision time scale (between elementary constituents of the fluid) needs to be much shorter than any macroscopic evolution time scales. This hypothesis is almost a tautology for standard fluids build up by molecules at reasonable density, but becomes not trivial in the case of some hot dense matter state created in high energetic hadronic collisions. In the following, we prove that these five unknown quantities describe completely the case of what we define as ideal fluids, in which we take no account of processes of energy dissipation. Energy dissipation may occur in a moving fluid as a consequence of internal friction (or viscosity) within the fluid and heat exchange between different parts of it. Neglecting this phenomenon, we can find a set of five equations that are sufficient obtain a closed system: 5 equations for 5 unknown quantities.
Recommended publications
  • Derivation of Fluid Flow Equations
    TPG4150 Reservoir Recovery Techniques 2017 1 Fluid Flow Equations DERIVATION OF FLUID FLOW EQUATIONS Review of basic steps Generally speaking, flow equations for flow in porous materials are based on a set of mass, momentum and energy conservation equations, and constitutive equations for the fluids and the porous material involved. For simplicity, we will in the following assume isothermal conditions, so that we not have to involve an energy conservation equation. However, in cases of changing reservoir temperature, such as in the case of cold water injection into a warmer reservoir, this may be of importance. Below, equations are initially described for single phase flow in linear, one- dimensional, horizontal systems, but are later on extended to multi-phase flow in two and three dimensions, and to other coordinate systems. Conservation of mass Consider the following one dimensional rod of porous material: Mass conservation may be formulated across a control element of the slab, with one fluid of density ρ is flowing through it at a velocity u: u ρ Δx The mass balance for the control element is then written as: ⎧Mass into the⎫ ⎧Mass out of the ⎫ ⎧ Rate of change of mass⎫ ⎨ ⎬ − ⎨ ⎬ = ⎨ ⎬ , ⎩element at x ⎭ ⎩element at x + Δx⎭ ⎩ inside the element ⎭ or ∂ {uρA} − {uρA} = {φAΔxρ}. x x+ Δx ∂t Dividing by Δx, and taking the limit as Δx approaches zero, we get the conservation of mass, or continuity equation: ∂ ∂ − (Aρu) = (Aφρ). ∂x ∂t For constant cross sectional area, the continuity equation simplifies to: ∂ ∂ − (ρu) = (φρ) . ∂x ∂t Next, we need to replace the velocity term by an equation relating it to pressure gradient and fluid and rock properties, and the density and porosity terms by appropriate pressure dependent functions.
    [Show full text]
  • Lagrangian and Eulerian Representations of Fluid Flow: Part I, Kinematics and the Equations of Motion
    Lagrangian and Eulerian Representations of Fluid Flow: Part I, Kinematics and the Equations of Motion James F. Price MS 29, Clark Laboratory Woods Hole Oceanographic Institution Woods Hole, MA, 02543 http://www.whoi.edu/science/PO/people/jprice [email protected] Version 7.4 September 13, 2005 Summary: This essay introduces the two methods that are commonly used to describe fluid flow, by observing the trajectories of parcels that are carried along with the flow or by observing the fluid velocity at fixed positions. These yield what are commonly termed Lagrangian and Eulerian descriptions. Lagrangian methods are often the most efficient way to sample a fluid domain and the physical conservation laws are inherently Lagrangian since they apply to specific material parcels rather than points in space. It happens, though, that the Lagrangian equations of motion applied to a continuum are quite difficult, and thus almost all of the theory (forward calculation) in fluid dynamics is developed within the Eulerian system. Eulerian solutions may be used to calculate Lagrangian properties, e.g., parcel trajectories, which is often a valuable step in the description of an Eulerian solution. Transformation to and from Lagrangian and Eulerian systems — the central theme of this essay — is thus the foundation of most theory in fluid dynamics and is a routine part of many investigations. The transformation of the Lagrangian conservation laws into the Eulerian equations of motion requires three key results. (1) The first is dubbed the Fundamental Principle of Kinematics; the velocity at a given position and time (the Eulerian velocity) is identically the velocity of the parcel (the Lagrangian velocity) that occupies that position at that time.
    [Show full text]
  • Continuity Equation in Pressure Coordinates
    Continuity Equation in Pressure Coordinates Here we will derive the continuity equation from the principle that mass is conserved for a parcel following the fluid motion (i.e., there is no flow across the boundaries of the parcel). This implies that δxδyδp δM = ρ δV = ρ δxδyδz = − g is conserved following the fluid motion: 1 d(δM ) = 0 δM dt 1 d()δM = 0 δM dt g d ⎛ δxδyδp ⎞ ⎜ ⎟ = 0 δxδyδp dt ⎝ g ⎠ 1 ⎛ d(δp) d(δy) d(δx)⎞ ⎜δxδy +δxδp +δyδp ⎟ = 0 δxδyδp ⎝ dt dt dt ⎠ 1 ⎛ dp ⎞ 1 ⎛ dy ⎞ 1 ⎛ dx ⎞ δ ⎜ ⎟ + δ ⎜ ⎟ + δ ⎜ ⎟ = 0 δp ⎝ dt ⎠ δy ⎝ dt ⎠ δx ⎝ dt ⎠ Taking the limit as δx, δy, δp → 0, ∂u ∂v ∂ω Continuity equation + + = 0 in pressure ∂x ∂y ∂p coordinates 1 Determining Vertical Velocities • Typical large-scale vertical motions in the atmosphere are of the order of 0. 01-01m/s0.1 m/s. • Such motions are very difficult, if not impossible, to measure directly. Typical observational errors for wind measurements are ~1 m/s. • Quantitative estimates of vertical velocity must be inferred from quantities that can be directly measured with sufficient accuracy. Vertical Velocity in P-Coordinates The equivalent of the vertical velocity in p-coordinates is: dp ∂p r ∂p ω = = +V ⋅∇p + w dt ∂t ∂z Based on a scaling of the three terms on the r.h.s., the last term is at least an order of magnitude larger than the other two. Making the hydrostatic approximation yields ∂p ω ≈ w = −ρgw ∂z Typical large-scale values: for w, 0.01 m/s = 1 cm/s for ω, 0.1 Pa/s = 1 μbar/s 2 The Kinematic Method By integrating the continuity equation in (x,y,p) coordinates, ω can be obtained from the mean divergence in a layer: ⎛ ∂u ∂v ⎞ ∂ω ⎜ + ⎟ + = 0 continuity equation in (x,y,p) coordinates ⎝ ∂x ∂y ⎠ p ∂p p2 p2 ⎛ ∂u ∂v ⎞ ∂ω = − ⎜ + ⎟ ∂p rearrange and integrate over the layer ∫p ∫ ⎜ ⎟ 1 ∂x ∂y p1⎝ ⎠ p ⎛ ∂u ∂v ⎞ ω(p )−ω(p ) = (p − p )⎜ + ⎟ overbar denotes pressure- 2 1 1 2 ⎜ ⎟ weighted vertical average ⎝ ∂x ∂y ⎠ p To determine vertical motion at a pressure level p2, assume that p1 = surface pressure and there is no vertical motion at the surface.
    [Show full text]
  • Chapter 3 Newtonian Fluids
    CM4650 Chapter 3 Newtonian Fluid 2/5/2018 Mechanics Chapter 3: Newtonian Fluids CM4650 Polymer Rheology Michigan Tech Navier-Stokes Equation v vv p 2 v g t 1 © Faith A. Morrison, Michigan Tech U. Chapter 3: Newtonian Fluid Mechanics TWO GOALS •Derive governing equations (mass and momentum balances •Solve governing equations for velocity and stress fields QUICK START V W x First, before we get deep into 2 v (x ) H derivation, let’s do a Navier-Stokes 1 2 x1 problem to get you started in the x3 mechanics of this type of problem solving. 2 © Faith A. Morrison, Michigan Tech U. 1 CM4650 Chapter 3 Newtonian Fluid 2/5/2018 Mechanics EXAMPLE: Drag flow between infinite parallel plates •Newtonian •steady state •incompressible fluid •very wide, long V •uniform pressure W x2 v1(x2) H x1 x3 3 EXAMPLE: Poiseuille flow between infinite parallel plates •Newtonian •steady state •Incompressible fluid •infinitely wide, long W x2 2H x1 x3 v (x ) x1=0 1 2 x1=L p=Po p=PL 4 2 CM4650 Chapter 3 Newtonian Fluid 2/5/2018 Mechanics Engineering Quantities of In more complex flows, we can use Interest general expressions that work in all cases. (any flow) volumetric ⋅ flow rate ∬ ⋅ | average 〈 〉 velocity ∬ Using the general formulas will Here, is the outwardly pointing unit normal help prevent errors. of ; it points in the direction “through” 5 © Faith A. Morrison, Michigan Tech U. The stress tensor was Total stress tensor, Π: invented to make the calculation of fluid stress easier. Π ≡ b (any flow, small surface) dS nˆ Force on the S ⋅ Π surface V (using the stress convention of Understanding Rheology) Here, is the outwardly pointing unit normal of ; it points in the direction “through” 6 © Faith A.
    [Show full text]
  • CONTINUITY EQUATION  Another Principle on Which We Can Derive a New Equation Is the Conservation of Mass
    ESCI 342 – Atmospheric Dynamics I Lesson 7 – The Continuity and Additional Equations Suggested Reading: Martin, Chapter 3 THE SYSTEM OF EQUATIONS IS INCOMPLETE The momentum equations in component form comprise a system of three equations with 4 unknown quantities (u, v, p, and ). Du1 p fv (1) Dt x Dv1 p fu (2) Dt y p g (3) z They are not a closed set, because there are four dependent variables (u, v, p, and ), but only three equations. We need to come up with some more equations in order to close the set. DERIVATION OF THE CONTINUITY EQUATION Another principle on which we can derive a new equation is the conservation of mass. The equation derived from this principle is called the mass continuity equation, or simply the continuity equation. Imagine a cube at a fixed point in space. The net change in mass contained within the cube is found by adding up the mass fluxes entering and leaving through each face of the cube.1 The mass flux across a face of the cube normal to the x-axis is given by u. Referring to the picture below, these fluxes will lead to a rate of change in mass within the cube given by m u yz u yz (4) t x xx 1 A flux is a quantity per unit area per unit time. Mass flux is therefore the rate at which mass moves across a unit area, and would have units of kg s1 m2. The mass in the cube can be written in terms of the density as m xyz so that m x y z .
    [Show full text]
  • 1 the Continuity Equation 2 the Heat Equation
    1 The Continuity Equation Imagine a fluid flowing in a region R of the plane in a time dependent fashion. At each point (x; y) R2 it has a velocity v = v (x; y; t) at time t. Let ρ = ρ(x; y; t) be the density of the fluid 2 −! −! at (x; y) at time t. Let P be any point in the interior of R and let Dr be the closed disk of radius r > 0 and center P . The mass of fluid inside Dr at any time t is ρ dxdy: ZZDr If matter is neither created nor destroyed inside Dr, the rate of decrease of this quantity is equal to the flux of the vector field −!J = ρ−!v across Cr, the positively oriented boundary of Dr. We therefore have d ρ dxdy = −!J −!N ds; dt − · ZZDr ZCr where N~ is the outer normal and ds is the element of arc length. Notice that the minus sign is needed since positive flux at time t represents loss of total mass at that time. Also observe that the amount of fluid transported across a small piece ds of the boundary of Dr at time t is ρ−!v −!N ds. Differentiating under the integral sign on the left-hand side and using the flux form of Green’s· Theorem on the right-hand side, we get @ρ dxdy = −!J dxdy: @t − r · ZZDr ZZDr Gathering terms on the left-hand side, we get @ρ ( + −!J ) dxdy = 0: @t r · ZZDr If the integrand was not zero at P it would be different from zero on Dr for some sufficiently small r and hence the integral would not be zero which is not the case.
    [Show full text]
  • Chapter 2 Wave Propagation in Viscous Fluid
    Chapter 2 Wave Propagation in Viscous Fluid This chapter summarizes with the derivation of the mathematical form of the acoustic wave propagation in the fluid. Before we derive the final form of the wave propagation equation in viscous fluid, we first look at two conservation (mass and momentum) of equations and state equation in the fluid. Detailed derivations can be found in the literatures [4, 6, 27, 28]. We limit our discussion only on the lossy 1-dimensional plane wave. 2.1 One-dimensional Viscous Wave Equation 2.1.1 One-dimensional Continuity Equation (Conversation of Mass) Consider the flow of a compressible fluid through a duct of arbitrary cross section (area S) in one dimension. The control volume (CV), is the segment between x and x + ∆x . 5 We want to know the rate at which the inside mass changes. First, we made two assumptions: 1. The CV is fixed in space 2. The flow is one-dimensional, so the mass flow only depends on t and x. Mass flow in Mass flow out = ρ uS |x = ρ uS |x+∆x x x+∆x , where ρ, u are the average mass flow density and average mass flow speed, respectively. For ∆x →0 , ρ and u will become a true point function. The time rate of increase of mass inside the CV is equal to net mass flow into the CV through the CV surfaces or in mathematical terms, ∂ (Sρ ∆x)= ρ uS | −ρ uS | ( 1 ) ∂t x x+∆x Since S is a constant and ∆x is not a function of time, Eq.
    [Show full text]
  • Lectures in Elementary Fluid Dynamics: Physics, Mathematics and Applications James M
    University of Kentucky UKnowledge Mechanical Engineering Textbook Gallery Mechanical Engineering 2009 Lectures In Elementary Fluid Dynamics: Physics, Mathematics and Applications James M. McDonough University of Kentucky, [email protected] Right click to open a feedback form in a new tab to let us know how this document benefits oy u. Follow this and additional works at: https://uknowledge.uky.edu/me_textbooks Part of the Mechanical Engineering Commons Recommended Citation McDonough, James M., "Lectures In Elementary Fluid Dynamics: Physics, Mathematics and Applications" (2009). Mechanical Engineering Textbook Gallery. 1. https://uknowledge.uky.edu/me_textbooks/1 This Book is brought to you for free and open access by the Mechanical Engineering at UKnowledge. It has been accepted for inclusion in Mechanical Engineering Textbook Gallery by an authorized administrator of UKnowledge. For more information, please contact [email protected]. LECTURES IN ELEMENTARY FLUID DYNAMICS: Physics, Mathematics and Applications J. M. McDonough Departments of Mechanical Engineering and Mathematics University of Kentucky, Lexington, KY 40506-0503 c 1987, 1990, 2002, 2004, 2009 Contents 1 Introduction 1 1.1 ImportanceofFluids.............................. ...... 1 1.1.1 Fluidsinthepuresciences. ...... 2 1.1.2 Fluidsintechnology .. .. .. .. .. .. .. .. .... 3 1.2 TheStudyofFluids ................................ .... 4 1.2.1 Thetheoreticalapproach . ..... 5 1.2.2 Experimentalfluiddynamics . ..... 6 1.2.3 Computationalfluiddynamics . ..... 6 1.3 OverviewofCourse...............................
    [Show full text]
  • An Essay on Lagrangian and Eulerian Kinematics of Fluid Flow
    Lagrangian and Eulerian Representations of Fluid Flow: Kinematics and the Equations of Motion James F. Price Woods Hole Oceanographic Institution, Woods Hole, MA, 02543 [email protected], http://www.whoi.edu/science/PO/people/jprice June 7, 2006 Summary: This essay introduces the two methods that are widely used to observe and analyze fluid flows, either by observing the trajectories of specific fluid parcels, which yields what is commonly termed a Lagrangian representation, or by observing the fluid velocity at fixed positions, which yields an Eulerian representation. Lagrangian methods are often the most efficient way to sample a fluid flow and the physical conservation laws are inherently Lagrangian since they apply to moving fluid volumes rather than to the fluid that happens to be present at some fixed point in space. Nevertheless, the Lagrangian equations of motion applied to a three-dimensional continuum are quite difficult in most applications, and thus almost all of the theory (forward calculation) in fluid mechanics is developed within the Eulerian system. Lagrangian and Eulerian concepts and methods are thus used side-by-side in many investigations, and the premise of this essay is that an understanding of both systems and the relationships between them can help form the framework for a study of fluid mechanics. 1 The transformation of the conservation laws from a Lagrangian to an Eulerian system can be envisaged in three steps. (1) The first is dubbed the Fundamental Principle of Kinematics; the fluid velocity at a given time and fixed position (the Eulerian velocity) is equal to the velocity of the fluid parcel (the Lagrangian velocity) that is present at that position at that instant.
    [Show full text]
  • Dynamics of Ideal Fluids
    2 Dynamics of Ideal Fluids The basic goal of any fluid-dynamical study is to provide (1) a complete description of the motion of the fluid at any instant of time, and hence of the kinematics of the flow, and (2) a description of how the motion changes in time in response to applied forces, and hence of the dynamics of the flow. We begin our study of astrophysical fluid dynamics by analyzing the motion of a compressible ideal fluid (i.e., a nonviscous and nonconducting gas); this allows us to formulate very simply both the basic conservation laws for the mass, momentum, and energy of a fluid parcel (which govern its dynamics) and the essentially geometrical relationships that specify its kinematics. Because we are concerned here with the macroscopic proper- ties of the flow of a physically uncomplicated medium, it is both natural and advantageous to adopt a purely continuum point of view. In the next chapter, where we seek to understand the important role played by internal processes of the gas in transporting energy and momentum within the fluid, we must carry out a deeper analysis based on a kinetic-theory view; even then we shall see that the continuum approach yields useful results and insights. We pursue this line of inquiry even further in Chapters 6 and 7, where we extend the analysis to include the interaction between radiation and both the internal state, and the macroscopic dynamics, of the material. 2.1 Kinematics 15. Veloci~y and Acceleration 1n developing descriptions of fluid motion it is fruitful to work in two different frames of reference, each of which has distinct advantages in certain situations.
    [Show full text]
  • ESCI 342 – Atmospheric Dynamics I Lesson 10 – Vertical Motion, Pressure Coordinates
    ESCI 342 – Atmospheric Dynamics I Lesson 10 – Vertical Motion, Pressure Coordinates Reading: Martin, Section 4.1 PRESSURE COORDINATES Pressure is often a convenient vertical coordinate to use in place of altitude. If the hydrostatic approximation is used, the relationship between pressure and altitude is given by the hydrostatic equation, p g . (1) z In height coordinates the vertical velocity is defined as w Dz Dt . In pressure coordinates the vertical velocity is defined as Dp , (2) Dt and is commonly called simply omega. The units of are Pa/s (often microbars per second, b/s, is also used). Since pressure decreases upward, a negative omega means rising motion, while a positive omega means subsiding motion. w and are related as follows: Dp p p p p u v w . (3) Dt t x y z On the synoptic scale we can assume hydrostatic vertical balance, so that (3) becomes Dp p p p u v gw gw , (4) Dt t x y since the local pressure tendencies and horizontal pressure advection terms are much, much smaller in magnitude than the vertical pressure advection terms. The total derivative in pressure coordinates is D uv .1 (5) Dt t x y p The conversion of a height derivative to a pressure derivative is accomplished using the chain rule as follows z . (6) p p z g z In pressure coordinates, the directions of the unit vectors ( iˆ , ˆj , and kˆ ) are the same as in height coordinates. The x and y axes are still horizontal, and not oriented along the constant pressure surface.2 The vertical axis is still vertical (perpendicular to x and y.) 1 If you take the total derivative of pressure you end up with the seeming absurdity that p p p uv .
    [Show full text]
  • Fluid Mechanics – Continuity Equation Bernoulli’S Equation
    PART II Continuity and Conservation of Mass • Some applications of fluid mechanics • Flow through a pipe: • . Conservation. of mass for steady state (no storage) says m in = m out . m m _1A1Vm1 = _ 2A2Vm2 • For incompressible fluids, density does not changes (_ 1 = _ 2) so A1Vm1 = A2Vm2 = Q Fluid Mechanics – Continuity Equation Bernoulli’s equation • The equation of continuity states that for an Y incompressible fluid flowing in a tube of varying 2 A V cross-sectional area (A), the mass flow rate is the 2 2 same everywhere in the tube: . m1= m2 _ 1A1V1 = _ 2A2V2 _1A1V1= _2 A2V2 • Generally, the density stays constant and then it's Y For incompressible flow simply the flow rate (Av) that is constant. 1 A1 V1 A1V1= A2V2 Assume steady flow, V parallel to streamlines & no viscosity 1 Bernoulli Equation – energy Bernoulli Equation – work • Consider energy terms for steady flow: • Consider work done on the system is Force x distance • We write terms for KE and PE at each point • We write terms for force in terms of Pressure and area Y Y 2 2 Wi = FiVi dt =PiViAi dt A2 V2 A2 V2 E = KE + PE i i i Note ViAi dt = mi/_i 1 2 E1 = 2 m&1V1 + gm&1 y1 W1 = P1m&1 / ρ1 2 Y 1 Y 1 E2 = 2 m&2V2 + gm&2 y2 1 W2 = −P2m&2 / ρ2 A1 V1 A1 V1 As the fluid moves, work is being done by the external Now we set up an energy balance on the system.
    [Show full text]