Docsity
Docsity

Prepare for your exams
Prepare for your exams

Study with the several resources on Docsity


Earn points to download
Earn points to download

Earn points by helping other students or get them with a premium plan


Guidelines and tips
Guidelines and tips

Deriving Fluid Temperature Equation: Thermodynamics & Motion, Study notes of Thermodynamics

Continuum MechanicsMeteorologyFluid MechanicsThermodynamics

The thermodynamics of fluids and derives the equation of state for a fluid in thermodynamic equilibrium. It also discusses the first law of thermodynamics and the budget equation for a fluid's energy. The document derives the governing equation for the temperature of a fluid using the thermodynamic potential temperature and potential vorticity.

What you will learn

  • What is the first law of thermodynamics for a fluid?
  • What is the thermodynamic potential temperature and how is it used in fluid dynamics?
  • How is the potential vorticity derived and what is its significance in fluid dynamics?
  • What is the equation of state for a fluid in thermodynamic equilibrium?
  • How is the budget equation for a fluid's energy derived?

Typology: Study notes

2021/2022

Uploaded on 09/07/2022

zaafir_ij
zaafir_ij 🇦🇪

4.4

(60)

888 documents

1 / 14

Toggle sidebar

Related documents


Partial preview of the text

Download Deriving Fluid Temperature Equation: Thermodynamics & Motion and more Study notes Thermodynamics in PDF only on Docsity! Chapter 6 Thermodynamics and the Equations of Motion 6.1 The first law of thermodynamics for a fluid and the equation of state. We noted in chapter 4 that the full formulation of the equations of motion required additional information to deal with the state variables density and pressure and that we were one equation short of matching unknowns and equations. In both meteorology and oceanography the variation of density and hence buoyancy is critical in many phenomenon such cyclogenesis and the thermohaline circulation, to name only two. To close the system we will have to include the thermodynamics pertinent to the fluid motion. In this course we will examine a swift review of those basic facts from thermodynamics we will need to complete our dynamical formulation. In actuality, thermodynamics is a misnomer. Classical thermodynamics deals with equilibrium states in which there are no variations of the material in space or time, hardly the situation of interest to us. However, we assume that we can subdivide the fluid into regions small enough to allow the continuum field approximation but large enough, and changing slowly enough so that locally thermodynamic equilibrium is established allowing a reasonable definition of thermodynamic state variables like pressure, density and pressure. We have already noted that for some quantities, like the pressure for molecules with more than translational degrees of freedom, the departures from thermodynamic equilibrium have to be considered. Generally, such considerations are of minor importance in the fluid mechanics of interest to us. If the fluid is in thermodynamic equilibrium any thermodynamic variable for a pure substance, like pure water, can be written in terms of any two other thermodynamic variables✿, i.e. p = p(ρ,T ) (6.1.1) ✿ For sea water, the presence of salt renders the equation of state very complex. There are tomes written on the subject and we will slide over this issue entirely in this course. Chapter 6 2 where the functional relationship in depends on the substance. Note that, as discussed before, (6.1.11) does not necessarily yield a pressure that is the average normal force on a fluid element. The classic example of and equation of state is the perfect gas law; p = ρRT (6.1.2) which is appropriate for dry air. The constant R is the gas constant and is a property of the material that must be specified. For air (from Batchelor) R = 2.870x103cm2 / sec2 degC (6.1.3) One of the central results of thermodynamics is the specification of another thermodynamic state variable e(ρ,T) which is the internal energy per unit mass and is, in fact, defined by a statement of the first law of thermodynamics. Consider a fluid volume, V, of fixed mass (Figure 6.1.1) Figure 6.1.1 A fixed mass of fluid, of volume V subject to body force F, a surface flux of heat (per unit surface area) out of the volume, K, and the surface force per unit area due to the surface stress tensor. The first law of thermodynamics states that the rate of change of the total energy of the fixed mass of fluid in V, i.e. the rate of change of the sum of the kinetic energy and V Fi Kj σ ijn j dS Chapter 6 5 σ ij ∂ui ∂x j = σ ji ∂uj ∂xi = σ ij ∂uj ∂xi = σ ij 1 2 ∂ui ∂x j + ∂uj ∂xi ⎛ ⎝⎜ ⎞ ⎠⎟ = σ ijeij (6.1.11) The first step in (6.1.11) is just a relabeled form with i and j interchanged. The second step uses the symmetry of the stress tensor and the last line rewrites the result in terms of the inner product of the stress tensor and the rate of strain tensor. Since, (3.7.15) σ ij = − pδ ij + 2µeij + λekkδ ij (6.1.12) or with the relation , λ = η − 2 3µ we have σ ij = − pδ ij + 2µ(eij − ekkδ ij ) +ηekkδ ij (6.1.13) where the pressure is the thermodynamic pressure of the equation of state and η is the coefficient relating to the deviation of that pressure from the average normal stress on a fluid element. The scalar σ ijeij = − p∇i u + 2µ eij 2 − 1 3 ekk 2⎡ ⎣⎢ ⎤ ⎦⎥ +ηekk 2 (6.1.14) The term eij 2 − 1 3 ekk 2⎡ ⎣⎢ ⎤ ⎦⎥ can be shown to be always positive (it’s is easiest to do this in a coordinate system where the rate of strain tensor is diagonalized. So this term always represents an increase of internal energy provided by the viscous dissipation of mechanical energy. Traditionally, this term is defined as the dissipation function, Φ , i.e. where, Φ = 2 µ ρ eij 2 − 1 3 ekk 2⎡ ⎣⎢ ⎤ ⎦⎥ (6.1.15) In much the same way that we approached the relation between the stress tensor and the velocity gradients, we assume that the heat flux vector depends linearly on the local value of the temperature gradient, or, in the general case Chapter 6 6 Ki = ℜij ∂T ∂x j (6.1.16) Again assuming that the medium is isotropic in terms of the relation between temperature gradient and heat flux, the tensor ℜij needs to be a second order isotropic tensor. The only such tensor is the kronecker delta so ℜij = −kδ ij ⇒ Ki = −k ∂T ∂xi , (6.1.17) The minus sign in (6.1.17) expresses our knowledge that heat flows from hot to cold, i.e. down the temperature gradient, i.e.  K = −k∇T (6.1.18) where k is the coefficient of heat conduction. For dry air at 200 C, k= 2.54 103 grams /(cm sec3degC) Putting these results together yields, ρ de dt = − p∇i u reversiblework  + ρΦ +η(∇i u)2 irreversiblework   + ρQ +∇i(k∇T ) (6.1.19) The pressure work term involves the product of the pressure and the rate of volume change; a convergence of velocity is a compression of the fluid element and so leads to an increase of internal energy but an expansion of the volume (a velocity divergence) can produce a compensating decrease of internal energy. On the other hand, the viscous terms represent an irreversible transformation of mechanical to internal energy. It is useful to separate the effects of the reversible from the irreversible work by considering the entropy. The entropy per unit mass is a state variable we shall refer to as s and satisfies for any variation δ s Tδs = δe + pδ 1 ρ( ) (6.1.20) so that, for variations with time for a fluid element, Chapter 6 7 T ds dt = de dt − p ρ2 dρ dt = de dt + p ρ ∇i u (6.1.21) Substituting for de/dt into (6.1.19) leads to an equation for the entropy in terms of the heating and the irreversible work, ρT ds dt = ρΦ +η(∇i u)2 +∇i(k∇T ) (6.1.22) Since s is a thermodynamic variable we can write, s = s(p,T ) or s = s(ρ,T ) , so that, ds = ∂s ∂p ⎞ ⎠⎟ T dp + ∂s ∂T ⎞ ⎠⎟ p dT = ∂s ∂ρ ⎞ ⎠⎟ T dρ + ∂s ∂T ⎞ ⎠⎟ ρ dT (6.1.23 a, b) Similarly, we can write (6.1.20) in two forms, Tds = de + pd(1 / ρ) = d(e + p / ρ) − 1 ρ dp (6.1.24 a, b) and we define another state variable, the enthalpy , h, as h = e + p ρ (6.1.25) It follows from (6.1.23) and (6.1.24) that we can define, Chapter 6 10 ρ cp dT dt − αT ρ dp dt ⎡ ⎣ ⎢ ⎤ ⎦ ⎥ = ρΦ +η ∇i u( )2 + ρQ +∇i(k∇T ) (6.1.35 a) while our other equations are the state equation (6.1.1) p = p(ρ,T ) (6.1.35 b) and the mass conservation equation, (2.1.11) dρ dt + ρ∇i u = 0 (6.1.35 c) and the momentum equation ( 4.1.13), ρ d u dt + ρ2  Ω × u = ρg − ∇p + µ∇2 u + (λ + µ)∇(∇i u) + (∇λ)(∇i u) + îieij ∂µ ∂x j .(6.1.35 d) Our unknowns are p,ρ,T , u 6   while we have 3 momentum equations, the thermodynamic equation, the equation of state, and the mass conservation equation, i.e. 6 equations for 6 unknowns, assuming that we can specify, in terms of these variables the thermodynamic functions α,µ,η,cp ,k which we suppose is possible. (If we were to think of the coefficients η,κ,µ as turbulent mixing coefficients it is less clear that the system can be closed in terms of the variables p,ρ,T and u ) At this point we have derived a complete set of governing equations and the formulation of our dynamical system is formally complete. But, and this is a big but, our work is just beginning. Even if we specify the nature of the fluid; air, water, syrup or galactic gas the equations we have derived are capable of describing the motion whether it deals with acoustic waves, spiral arms in hurricanes, weather waves in the atmosphere or the meandering Gulf Stream in the ocean. This very richness in the basic equations is an impediment to solving any one of those examples since for some phenomenon of interest we have included more physics than we need, for example the compressibility of water is not needed to discuss the waves in your bathtub. If the equations were simpler, especially if they were linear, it might be possible to nevertheless accept this unnecessary richness but both the momentum, thermodynamic and mass conservation equations are nonlinear because of the advective derivative so a frontal Chapter 6 11 attack on the full equations , even with the most powerful modern computers is a hopeless approach. This is both the challenge and the attraction of fluid mechanics. Mathematics must be allied with physical intuition to make progress and in the remainder of the course we will approach this in a variety of ways. Before doing so we will discuss two specializations of the thermodynamics of special interest to us as meteorologists and oceanographers. 6.2 The perfect gas The state equation (6.1.1) is appropriate for a gas like air for which R is 0.294 joule/gm deg C (1 joule =107 gm cm2/sec2). It follows that, dp p = dT T + dρ ρ (6.2.1) so that for processes which take place at constant pressure, − 1 ρ ∂ρ ∂T ⎛ ⎝⎜ ⎞ ⎠⎟ p = 1 T ≡ α (6.2.2) Thus, for a perfect gas, (6.1.35) becomes, cp dT dt − 1 ρ dp dt ⎡ ⎣ ⎢ ⎤ ⎦ ⎥ =Φ + η ρ ∇i u( )2 +Q + 1 ρ ∇i(k∇T ), (6.2.3) or, cpT 1 T dT dt − 1 cpρT dp dt ⎡ ⎣ ⎢ ⎢ ⎤ ⎦ ⎥ ⎥ =Φ + η ρ ∇i u( )2 +Q + 1 ρ ∇i(k∇T ), ⇒ cpT 1 T dT dt − R cp p dp dt ⎡ ⎣ ⎢ ⎢ ⎤ ⎦ ⎥ ⎥ =Φ + η ρ ∇i u( )2 +Q + 1 ρ ∇i(k∇T ), ⇒ cpT d dt ln T pR /cp ⎛ ⎝⎜ ⎞ ⎠⎟ ⎡ ⎣ ⎢ ⎤ ⎦ ⎥ =Φ + η ρ ∇i u( )2 +Q + 1 ρ ∇i(k∇T ), (6.2.4 a, b, c) Chapter 6 12 We define the potential temperature: θ = T po p ⎛ ⎝⎜ ⎞ ⎠⎟ R /cp (6.2.5) where p0 is an arbitrary constant. In atmospheric applications it is usually chosen to be a nominal surface pressure (1000 mb). Thus for a process at constant θ (whose pertinence we shall shortly see) a decrease in pressure, for example the elevation of the fluid to higher altitude, corresponds to a reduction in T. Our thermodynamic equation can then be written as, cpT θ dθ dt =Φ + η ρ ∇i u( )2 +Q + 1 ρ ∇i(k∇T ) ≡ Η , (6.2.6) where H is the collection of the non-adiabatic contributions to the increase of entropy. If the motion of the gas is isentropic, i.e. if we can ignore thermal effects that add heat to the fluid element either by frictional dissipation, thermal conduction or internal heat sources, then the potential vorticity is a conserved quantity following the fluid motion since in general, dθ dt = θ cpT Η (6.2.7) We can use (6.2.5) to express the gas law (6.1.1) in terms of the potential temperature. We use the thermodynamic relation R = cp − cv (6.2.8) which follows from the fact that for a perfect gas the specific heats are constants so that, e = cvT , h = cpT = e + p ρ = T (cv + R) (6.2.9) Then, p1/γ ρ = Rθ po R /cp (6.2.10)
Docsity logo



Copyright © 2024 Ladybird Srl - Via Leonardo da Vinci 16, 10126, Torino, Italy - VAT 10816460017 - All rights reserved