Fact-checked by Grok 2 weeks ago

Control volume

A control volume is a fixed, arbitrary in space used in and to analyze open systems where , , and can flow across its boundaries, known as the control surface. This approach employs an Eulerian description, focusing on properties at fixed points in space rather than tracking individual particles. Unlike a , which consists of a fixed with no net flow across its boundaries, a control volume allows for inflows and outflows, making it essential for studying processes like and . The fundamental tool for control volume analysis is the , which relates the time rate of change of an extensive property (such as , , or ) within the control volume to the corresponding rate for a moving of particles. This theorem enables the application of laws to open systems by accounting for fluxes across the control surface, expressed mathematically as \frac{dN}{dt} = \frac{\partial}{\partial t} \int_V \eta \rho \, dV + \int_S \eta \rho \mathbf{u} \cdot d\mathbf{S}, where N is the extensive property, \eta is the intensive property per unit , \rho is , \mathbf{u} is , and the integrals are over the volume V and surface S. For , the equation simplifies to the net equaling the rate of change of inside the volume: \frac{dm_{cv}}{dt} = \dot{m}_{in} - \dot{m}_{out}. Control volume formulations are widely applied in engineering contexts, such as analyzing steady-flow devices like jet engines, turbines, and pumps, where assumptions of uniform flow and steady state often simplify calculations. In steady-state conditions, the mass inside the control volume remains constant, so inflow equals outflow (\dot{m}_{in} = \dot{m}_{out}), and the first law of thermodynamics yields the steady flow energy equation: \dot{Q}_{cv} - \dot{W}_{cv} = \dot{m} \left[ (h_e + \frac{c_e^2}{2} + gz_e) - (h_i + \frac{c_i^2}{2} + gz_i) \right], incorporating h, , and terms for inlet (i) and exit (e) conditions. These principles extend to unsteady processes and momentum balances, providing a versatile framework for both theoretical derivations and practical simulations in fields like and .

Fundamentals

Definition

A control volume is an arbitrary, fixed region in space selected for the analysis of fluid flow or problems, where properties such as , , and are accounted for through their rates of change within the volume and the net fluxes across its boundaries. This concept enables the application of laws to practical scenarios, such as flow through pipes or around objects, by focusing on the interactions at the boundaries rather than tracking individual fluid particles. The key characteristics of a control volume include its well-defined boundaries, known as the control surface, which may be real (e.g., solid walls) or imaginary and is typically stationary relative to a chosen . Control volumes can represent open systems, permitting mass to enter and exit via designated inlet and outlet ports, or closed systems with no net across the surface, though they are predominantly used for open systems in to capture convective transport effects. In contrast to infinitesimal control volumes, which are differential elements used to develop local governing equations, finite control volumes facilitate integral analyses that provide global insights into system behavior. Visually, a control volume is often depicted as an enclosed geometric domain—such as a , , or the within a device—with the control surface outlined and ports marked for inflows and outflows, illustrated by arrows representing fluid motion across those boundaries. This spatial fixity corresponds to the Eulerian approach in fluid description.

Distinction from control mass

A control mass, also referred to as a , consists of a fixed of bounded by a surface that moves with the material itself, preventing any mass from crossing the boundary while allowing energy exchanges such as and work. This approach tracks a specific parcel of or as it evolves, making it suitable for analyses where the identity of the material remains constant over time. In distinction, a control volume defines a fixed in space—often aligned with geometries like or reactors—through which mass freely enters and exits, enabling straightforward application of principles to open systems. This method offers key advantages by simplifying the study of continuous flows in stationary setups, as it inherently incorporates mass fluxes across boundaries without requiring the complex tracking of deforming surfaces inherent to control mass formulations. For instance, it facilitates efficient integral balances for devices involving steady or unsteady throughput, such as turbines or nozzles, where boundary conditions remain constant. However, control volumes have limitations in that they do not follow the paths of individual particles, rendering them less appropriate for detailed tracking of microscopic phenomena or dispersion within specific material elements. Instead, they excel in macroscopic, analyses that prioritize overall system behavior over particle-level details. The , central to control mass descriptions, contrasts with this by quantifying rates of change along paths, though control volume methods adapt such concepts for spatial fixedness. The origins of control volume analysis lie in 19th-century thermodynamic studies of open systems, such as those by on efficiency, which laid groundwork for handling and transfers in practical devices. This evolved through 20th-century engineering innovations, particularly in , where it emerged as a pragmatic alternative to physics-oriented differential methods, emphasizing empirical balances for complex flows.

Theoretical Framework

Eulerian and Lagrangian approaches

In continuum mechanics, both the Eulerian and Lagrangian approaches rely on the continuum hypothesis, which posits that matter can be treated as a continuous medium with smoothly varying properties, disregarding its discrete molecular structure. This assumption enables the mathematical description of fluid and solid behaviors at macroscopic scales, allowing properties like density and velocity to be defined at every point in space and time. The approach describes the motion of a by tracking individual particles or parcels over time, using coordinates that label particles based on their initial positions. In this framework, the position of a particle is expressed as a function of its coordinates and time, making it particularly suitable for analyzing deformable bodies where boundaries move with the . However, for fluids, the method becomes computationally complex due to the intricate, often entangled trajectories of particles in three-dimensional flows, limiting its practical use despite its natural alignment with conservation laws applied to fixed sets of particles, as in control mass analysis. In contrast, the Eulerian approach observes the from fixed points , employing spatial coordinates to describe properties as that vary with position and time. This perspective is the foundation for control volume analysis in , where a fixed —known as a control volume—allows monitoring of how enters and exits, facilitating the application of principles without tracking individual particles. The Eulerian method is especially advantageous for problems involving fixed boundaries, such as or channels, as it simplifies the formulation of equations for steady or unsteady flows by focusing on local rather than global particle paths. Control volumes thus serve as a practical implementation of Eulerian methods, bridging theoretical descriptions with applications.

Reynolds transport theorem

The Reynolds transport theorem serves as the fundamental mathematical bridge between the description, which follows a material system of fluid particles, and the Eulerian description, which analyzes flow at fixed points in space within a control volume. Named after Osborne Reynolds (1842–1912), it enables the formulation of conservation laws for open systems by relating the time rate of change of an extensive property in a material system to changes within and across the boundaries of a control volume. The theorem states that the rate of change of an extensive property B for a system coincides instantaneously with a fixed or moving volume equals the of B within the volume plus the net convective of B across the control surface. This formulation accounts for both the accumulation of the property inside the control volume and its transport due to fluid motion through the boundaries. The derivation starts with an arbitrary extensive property B of a material , expressed as B = \int_{V(t)} \rho b \, dV, where \rho is the , b is the corresponding intensive property per unit mass (e.g., , , or ), and V(t) is the time-dependent volume occupied by the system. The time of B is then \frac{dB}{dt} = \frac{d}{dt} \int_{V(t)} \rho b \, dV. Applying the Leibniz rule for under the integral sign with variable limits—generalized to three dimensions for a deforming volume—this expands to include a local time term within the volume and a surface term capturing the property's across the boundary due to the \mathbf{v}. For a control volume that may be fixed, moving, or deforming but coinciding with the at the instant of interest, the relative motion between the control surface and the fluid contributes to the . In its general form for an arbitrary control volume (fixed, moving, or deforming), the is \frac{dB}{dt} = \frac{\partial}{\partial t} \int_{CV} \rho b \, dV + \int_{CS} \rho b (\mathbf{v}_{rel} \cdot \mathbf{n}) \, dA, where CV denotes the , CS the , \mathbf{v}_{rel} = \mathbf{v} - \mathbf{v}_s is the fluid velocity relative to the control surface (\mathbf{v} absolute fluid velocity, \mathbf{v}_s control surface velocity), and \mathbf{n} the outward-pointing unit normal vector to the surface element dA. For fixed control volumes, \mathbf{v}_s = 0 so \mathbf{v}_{rel} = \mathbf{v}. The first term on the right represents the rate of accumulation or depletion within the control volume (the "" in Eulerian terms), while the second term is the net outward flux due to . This form assumes the control volume has a well-defined , the fluid properties are continuous and differentiable, and any deformation of the volume is accounted for through the relative velocity. Special cases arise by specifying b for particular properties, assuming the control volume encloses the system instantaneously and the properties are extensive and additive. For , B = m and b = 1, so the theorem becomes \frac{dm}{dt} = \frac{\partial}{\partial t} \int_{CV} \rho \, dV + \int_{CS} \rho (\mathbf{v}_{rel} \cdot \mathbf{n}) \, dA, which for conserved yields the integral . For linear , B is the total and b = \mathbf{v}, leading to \frac{d}{dt} \int_{sys} \rho \mathbf{v} \, dV = \frac{\partial}{\partial t} \int_{CV} \rho \mathbf{v} \, dV + \int_{CS} \rho \mathbf{v} (\mathbf{v}_{rel} \cdot \mathbf{n}) \, dA, relating changes to convective . For total , B is the system's and b = e ( per unit , including kinetic, potential, and internal forms), giving \frac{dE}{dt} = \frac{\partial}{\partial t} \int_{CV} \rho e \, dV + \int_{CS} \rho e (\mathbf{v}_{rel} \cdot \mathbf{n}) \, dA, which captures flux across the surface. These cases hold under the same general assumptions of arbitrary control volume shape and continuous flow fields.

Conservation Laws

Continuity equation

The continuity equation expresses the conservation of mass for a control volume by applying the Reynolds transport theorem to the total mass within the system. The theorem relates the time rate of change of a system property B to control volume integrals, and for mass conservation, B = m (total mass) with the specific property b = 1 (mass per unit mass), yielding \frac{dm_{sys}}{dt} = 0 since mass is conserved. Substituting into the Reynolds transport theorem for a fixed control volume gives the integral form: \frac{d}{dt} \int_{CV} \rho \, dV + \int_{CS} \rho (\mathbf{v} \cdot \mathbf{n}) \, dA = 0 where \rho is the fluid density, \mathbf{v} is the velocity vector, \mathbf{n} is the outward unit normal to the control surface (CS), and the integrals are over the control volume (CV) and its bounding surface, respectively. This equation balances the time rate of change of mass inside the control volume with the net mass flux across its boundary. To obtain the , apply the to the surface integral term in the , converting it to a over the control volume. For an arbitrary control volume, the integrand must vanish pointwise, resulting in: \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \mathbf{v}) = 0 This local form describes mass conservation at every point in the flow field. The continuity equation has key implications depending on flow conditions. For steady-state flows, where properties do not vary with time, the partial derivative term vanishes, reducing the integral form to \int_{CS} \rho (\mathbf{v} \cdot \mathbf{n}) \, dA = 0, indicating zero net mass flux across the control surface. In incompressible flows, where density \rho is constant, the equation simplifies further to \nabla \cdot \mathbf{v} = 0 in differential form or \int_{CS} (\mathbf{v} \cdot \mathbf{n}) \, dA = 0 in integral form, enforcing volume conservation rather than mass. Compressible flows, by contrast, allow \rho to vary, requiring the full form to account for density changes.

Momentum equation

The momentum equation for a control volume provides a statement of Newton's second law applied to an arbitrary fixed or moving region in a , accounting for the accumulation, transport, and external influences on . It is derived by applying the to the extensive property of , where the total momentum B = \int_{CV} \rho \mathbf{v} \, dV and the intensive property b = \mathbf{v} (velocity ). The theorem relates the time rate of change of in a material system to that within the control volume plus the net flux across the control surface, yielding the integral conservation form: \frac{d}{dt} \int_{CV} \rho \mathbf{v} \, dV + \int_{CS} \rho \mathbf{v} (\mathbf{v} \cdot \mathbf{n}) \, dA = \sum \mathbf{F} Here, the left-hand side represents the local rate of momentum accumulation within the control volume and the convective momentum flux through the control surface (with \mathbf{n} as the outward unit normal), while the right-hand side sums all forces acting on the fluid within the volume. This form is fundamental for analyzing momentum balance in open systems, such as flows through devices or boundaries. The forces \sum \mathbf{F} include body forces (e.g., gravity, \int_{CV} \rho \mathbf{g} \, dV), pressure forces on the control surface (\int_{CS} -p \mathbf{n} \, dA), viscous shear forces (\int_{CS} \boldsymbol{\tau} \cdot \mathbf{n} \, dA, where \boldsymbol{\tau} is the viscous stress tensor), and any external forces applied directly to the control surface. The convective term \int_{CS} \rho \mathbf{v} (\mathbf{v} \cdot \mathbf{n}) \, dA captures the transport of momentum due to bulk fluid motion across the surface, often dominating in high-speed or confined flows. For consistency with mass conservation, the density \rho aligns with the continuity equation, ensuring the analysis remains coupled. To obtain the differential form, the control volume is reduced to an infinitesimal element, applying the to convert surface integrals to volume integrals and taking limits as the volume shrinks. This process results in the , which for a forms the basis of the Navier-Stokes equations: \rho \left( \frac{\partial \mathbf{v}}{\partial t} + \mathbf{v} \cdot \nabla \mathbf{v} \right) = -\nabla p + \nabla \cdot \boldsymbol{\tau} + \rho \mathbf{g} The left-hand side includes local acceleration \rho \frac{\partial \mathbf{v}}{\partial t} and convective acceleration \rho (\mathbf{v} \cdot \nabla \mathbf{v}), while the right-hand side balances forces -\nabla p, viscous diffusion \nabla \cdot \boldsymbol{\tau}, and body forces \rho \mathbf{g}. Pressure forces arise from the normal on the infinitesimal surfaces, and the convective term reflects the nonlinear inherent to fluid motion. This underpins detailed simulations and theoretical analyses of viscous flows.

Energy equation

The energy equation for a control volume expresses the within a fixed or moving region of space, accounting for the rates of change of inside the volume, the net of across its surface, and the addition or removal of through and work. This equation is derived from of applied to an open system, where the total per unit mass, e, includes u, \frac{1}{2} v^2, and gz. Unlike closed-system analyses, the control volume formulation incorporates convective transport of by motion across the boundaries. In integral form, the energy balance for a control volume CV with control surface CS is given by \frac{d}{dt} \int_{CV} e \rho \, dV + \int_{CS} e \rho (\mathbf{v} \cdot \mathbf{n}) \, dA = \dot{Q}_{CV} - \dot{W}_{CV}, where the left-hand side represents the time rate of change of energy within the control volume plus the net efflux of energy through the surface (with \mathbf{n} as the outward unit normal), \dot{Q}_{CV} is the net rate of heat addition to the control volume, and \dot{W}_{CV} is the net rate of work done by the control volume on its surroundings. The work term \dot{W}_{CV} typically includes shaft work \dot{W}_s (e.g., from turbines or pumps) and other forms like pressure work, while the energy flux term often uses specific enthalpy h = u + pv for incompressible flows to simplify flow work contributions. This form assumes no nuclear or chemical energy changes unless specified. The integral energy equation is derived by applying the to the total of a material volume (system) and relating it to the control volume. The theorem states that for an arbitrary extensive property B = \int_{sys} b \rho \, dV, the is \frac{D}{Dt} \int_{sys} b \rho \, dV = \frac{d}{dt} \int_{CV} b \rho \, dV + \int_{CS} b \rho (\mathbf{v} \cdot \mathbf{n}) \, dA, where b = e is the specific total . Invoking for the system, \frac{D}{Dt} (E_{sys}) = \dot{Q} - \dot{W}, and substituting the theorem yields the control volume form, bridging Lagrangian (system-following) and Eulerian (fixed-volume) perspectives. This derivation highlights how the convective term arises from motion relative to the fixed control surface. For microscopic analysis, the differential form of the energy equation is obtained by applying the to the integral form and taking the as the control volume shrinks to size. This yields the local balance \rho \frac{De}{Dt} = -\nabla \cdot \mathbf{q} + \nabla \cdot (\boldsymbol{\tau} \cdot \mathbf{v}) - p (\nabla \cdot \mathbf{v}) + \rho \phi + \rho \mathbf{g} \cdot \mathbf{v}, where \frac{De}{Dt} is the substantial derivative of specific total , \mathbf{q} is the vector (governed by Fourier's law \mathbf{q} = -k \nabla T), \boldsymbol{\tau} is the contributing to mechanical work, p (\nabla \cdot \mathbf{v}) accounts for compression/expansion work, \phi = \boldsymbol{\tau} : \nabla \mathbf{v} is the viscous dissipation rate (converting mechanical to ), and \rho \mathbf{g} \cdot \mathbf{v} is the gravitational work term (often negligible). This form emphasizes the interplay of conduction, viscous effects, and dissipation in transport. Boundary conditions for the energy equation specify the treatment of heat and work across the control surface. The heat flux \mathbf{q} enters as a surface integral \int_{CS} \mathbf{q} \cdot \mathbf{n} \, dA in the integral form, representing conduction or radiation into the volume. Shaft work \dot{W}_s is modeled as \dot{W}_s = \int_{CS} \mathbf{v}_s \cdot (\mathbf{T} \cdot \mathbf{n}) \, dA or simplified for rotating machinery as \dot{W}_s = \omega T (torque times angular velocity), where positive \dot{W}_s indicates work extracted from the fluid. These conditions ensure compatibility with momentum flux for pressure-related flow work.

Applications and Examples

Steady-state flow analysis

In steady-state flow analysis, the conservation laws for a volume simplify significantly because the time-dependent terms vanish, reducing the equations to balances between the fluxes of , , and energy across the control surface boundaries. For the , the rate of accumulation within the volume is zero, leading to the condition that the entering the volume equals the exiting it, expressed as ∫ ρ (v · n) dA = 0, where ρ is , v is , n is the outward normal, and the is over the entire control surface. This simplification assumes no temporal variations in properties, allowing engineers to focus solely on and outlet conditions for steady processes like flow through pipes or nozzles. The equation in steady-state flow equates the on the control volume to the net across its boundaries, which is particularly useful for calculating in systems. The force F generated by a exhaust, for instance, arises from the imbalance, given by F = ∫ ρ v (v · n) dA over the control surface, where the integral captures the efflux of from the minus any . In a typical or analysis, if the is negligible compared to the exhaust v_e, the approximates to ṁ v_e, with ṁ as the , enabling direct computation of propulsive forces without time-dependent terms. This approach relies on the general conservation of principle applied to fixed control volumes. For energy analysis under steady-state conditions, the first law of thermodynamics for the control volume reduces to a balance where the net equals the rate of work done, often leading to extensions of the for incompressible flows. In such cases, assuming inviscid, steady along a streamline, the becomes \frac{p}{\rho} + \frac{v^2}{2} + gz = \text{constant}, where p is , g is , and z is ; this form highlights the trade-off between pressure, kinetic, and potential energies without shaft work or heat transfer complications. This Bernoulli extension is derived by integrating the energy over a stream tube, treating the as one-dimensional and at cross-sections, and is widely applied to problems like in venturi meters or over airfoils. The assumptions of properties at inlets and outlets, along with one-dimensional approximations, further streamline calculations by averaging velocities and pressures across sections.

Unsteady flow and transient processes

In unsteady control volume analysis, the inclusion of with respect to time terms accounts for the accumulation or depletion of quantities within the volume, distinguishing these processes from steady-state conditions where such terms vanish. This approach is essential for modeling time-dependent phenomena in systems, where inflows, outflows, and internal changes evolve dynamically. For mass conservation, the unsteady equation takes the form \frac{dM}{dt} = \dot{m}_{\text{in}} - \dot{m}_{\text{out}}, where M denotes the within the control volume, and \dot{m}_{\text{in}} and \dot{m}_{\text{out}} are the respective inlet and outlet flow rates. This formulation applies directly to scenarios like filling or emptying, where the control volume varies as enters or exits at rates that may change over time, such as during startup or . For instance, in a being filled from a , the rising level leads to increasing M, balanced by the net positive . Unsteady momentum analysis incorporates transient terms to capture the rate of change of linear inside the control volume, often arising from time-varying velocities or forces. In systems, startup transients or oscillating flows exemplify this, where initial of generates inertial forces. A prominent case is during sudden valve closure, producing a pressure surge with magnitude \Delta p = \rho a V_0 for instantaneous stops, where \rho is , a is the pressure wave speed, and V_0 is the initial ; for gradual closure over time t_c in a of length L, it approximates \Delta p = \frac{\rho L V_0}{t_c}. These effects can cause significant structural stresses in pipelines, highlighting the need for transient balances. For energy conservation in transients, the equation features a \frac{\partial}{\partial t} term for the total energy accumulation, coupled with time-varying heat transfer rates \dot{Q}. This is critical in processes like fluid heating or cooling within enclosures, where external heat addition or removal alters internal energy, kinetic energy, and potential energy over time. For example, in a tank undergoing convective heating, \dot{Q} fluctuates with temperature differences, driving unsteady changes in the fluid's thermal state. Viscous dissipation may also contribute significantly in high-shear transients, as seen in rapidly accelerating flows. Numerical simulation of these unsteady control volume problems relies on the in (CFD), which divides the domain into discrete control volumes and integrates the governing equations to enforce local . This discretization handles transient terms through time-stepping schemes, such as implicit or explicit methods, ensuring accurate resolution of accumulation and variations in applications like turbulent pipe transients or thermal cycling. The method's conservative nature makes it robust for capturing shocks or waves in unsteady flows.

References

  1. [1]
    2.5 Control volume form of the conservation laws - MIT
    The rate of change of mass inside the volume is given by the difference between the mass flow rate in and the mass flow rate out.
  2. [2]
    Control Volume (Integral) Technique
    Recall from thermo class, that a system is defined as a volume of mass of fixed identity. · Let m = the mass of the system. Let V = the velocity of the system
  3. [3]
    [PDF] Review of fluid dynamics
    It can be likened to a closed system in thermodynamics. • A Control Volume is an arbitrary volume of space through which fluid can flow.
  4. [4]
    [PDF] COPYRIGHTED MATERIAL
    A control volume is an arbitrary volume in space through which fluid flows. The geometric boundary of the control volume is called the control surface. The ...
  5. [5]
    [PDF] Control Volume
    An important fundamental step in setting up the equations of fluid flow in any particular context is the establishment of a well-defined control volume.
  6. [6]
    None
    ### Summary of Control Volume Definition and Characteristics
  7. [7]
    Lecture 1
    Closed Systems (Control Mass). - A fixed quantity of matter. - There can be no transfer of mass across its boundary. But energy, in the form of heat or work, ...
  8. [8]
    [PDF] Introduction to Thermodynamics Definitions - Purdue Engineering
    A system (aka closed system, aka control mass) is a particular quantity of matter chosen for study. ... For example, kinetic energy and mass are extensive ...
  9. [9]
    [PDF] Fluids – Lecture 6 Notes - MIT
    One concept is the control volume, which can be either finite or infinitesimal. Two types of control volumes can be employed:
  10. [10]
    [PDF] A Physical Introduction to Fluid Mechanics
    Dec 19, 2013 · the x-momentum of the fluid leaving the control volume is (ρ2V2A2∆t)V2 cos θ. ... In fluid mechanics, a “system” is a fixed mass of fluid. The “ ...
  11. [11]
    W. J. M. Rankine and the Rise of Thermodynamics
    Jan 5, 2009 · In the history of thermodynamics, two dates stand out as especially important: 1824, when Sadi Carnot's brilliant memoir Réflexions sur la ...Missing: control | Show results with:control
  12. [12]
    A Difference in Thinking between Engineering and Physics
    Jun 6, 2023 · Control-Volume Analysis: A Difference in Thinking between Engineering and Physics. Walter G. Vincenti; Technology and Culture; Johns Hopkins ...
  13. [13]
    [PDF] History of Continuum Mechanics
    In 1776, Euler introduced and interpreted the tensor of rate of deformation or stretching and with minor changes would yield the description of infinitesimal ...
  14. [14]
    [PDF] Lagrangian and Eulerian Representations of Fluid Flow: Kinematics ...
    Jun 7, 2006 · Summary: This essay introduces the two methods that are widely used to observe and analyze fluid flows, either by observing the trajectories ...
  15. [15]
    [PDF] Lagrangian and Eulerian Representations of Fluid Flow, Part 1
    Both methods are used widely in the analysis of the atmosphere and oceans, and in fluid mechanics generally. This essay aims to investigate both systems, and ...
  16. [16]
  17. [17]
    [PDF] REYNOLDS TRANSPORT THEOREM In this lesson, we will
    The Reynolds transport theorem is used to transform from system to control volume for integral analysis. Page 2. Another way to think about the RTT is that it ...
  18. [18]
    Continuity Equation – Introduction to Aerospace Flight Vehicles
    Know how to derive the most general form of the continuity equation in its control volume or integral form. Learn how to solve simple flow problems using the ...
  19. [19]
    [PDF] Fluid Mechanics - Internet Archive
    gross fluid property, a result called the Reynolds transport theorem. ... application of the Reynolds transport theorem gives the linear momentum relation for.
  20. [20]
    [PDF] The Reynolds Transport Theorem (RTT) (Section 4-6)
    Sep 12, 2012 · the control volume approach: See text for detailed derivation of the RTT. Here are some highlights: • Let B represent any extensive property ...
  21. [21]
    None
    ### Summary of Differential Form of Momentum Equation from Control Volume Analysis
  22. [22]
    [PDF] Derivation of the Navier–Stokes equations - UC Davis Math
    Jan 4, 2012 · The control volume can remain fixed in space or can move with the fluid. The material derivative. Main article: material derivative. Changes in ...Missing: differential | Show results with:differential
  23. [23]
    The Energy Equation for Control Volumes
    The Energy Equation for Control Volumes. Recall, the First Law of Thermodynamics: where = rate of change of total energy of the system, = rate of heat added ...
  24. [24]
    [PDF] Chapter 3 Integral Relations for a Control Volume
    The first is the “differential” approach and is developed in Chap. 4. We first develop the concept of the control volume, in nearly the same manner as one does ...
  25. [25]
    [PDF] Integral Balance Equations - Research
    Equation (16) is the integral (control volume) equation of energy conservation. The term on the left hand side is the rate of accumulation of energy in the ...
  26. [26]
    [PDF] CHAPTER 6 THE CONSERVATION EQUATIONS
    Apr 15, 2013 · Use the Reynolds transport theorem to replace the left-hand-side of (6.23) and the. Gauss theorem to convert the surface integral to a volume ...
  27. [27]
    [PDF] Control Volume Derivations for Thermodynamics
    This document will give a summary of the necessary mathematical operations necessary to cast the conservation of mass and energy principles in a traditional ...
  28. [28]
    [PDF] Conservation of Energy Equation
    The control volume is arbitrary, hence the sum of all the integrands must be zero to satisfy the equilibrium. Finally, we obtain the differential form of the ...
  29. [29]
    Energy Equation & Bernoulli's Equation – Introduction to Aerospace ...
    A finite control volume fixed in space is used to establish the energy equation in its most general form. The objective is to develop an analogous equation to ...
  30. [30]
    9.1.1 Conservation of Mass for a Control Volume
    A process during which a fluid flows through a control volume steadily. The fluid properties can change from point to point within the control volume.
  31. [31]
    Momentum Equation – Introduction to Aerospace Flight Vehicles
    Equation 10 is the momentum equation in its integral form, first established by Leonhard Euler. It will be apparent that it is a vector equation, so three ...Missing: citation | Show results with:citation
  32. [32]
    Conservation of Mass using Control Volumes
    It turns out that = mass flow rate outward through a surface so, if we integrate over the entire surface, and the conservation of mass equation becomes:
  33. [33]
    Time-Dependent Flows – Introduction to Aerospace Flight Vehicles
    Unsteady effects, known as water hammer, are produced when a valve is closed in a pipeline. Nikolay Joukowsky laid the foundation of the water hammer theory ...
  34. [34]
    [PDF] Governing Equations of Fluid Dynamics
    Fig. 2.1(a). This control volume is always made up of the same fluid particles as it.