Isentropic process

Last updated

In thermodynamics, an isentropic process is an idealized thermodynamic process that is both adiabatic and reversible. [1] [2] [3] [4] [5] [6] The work transfers of the system are frictionless, and there is no net transfer of heat or matter. Such an idealized process is useful in engineering as a model of and basis of comparison for real processes. [7] This process is idealized because reversible processes do not occur in reality; thinking of a process as both adiabatic and reversible would show that the initial and final entropies are the same, thus, the reason it is called isentropic (entropy does not change). Thermodynamic processes are named based on the effect they would have on the system (ex. isovolumetric: constant volume, isenthalpic: constant enthalpy). Even though in reality it is not necessarily possible to carry out an isentropic process, some may be approximated as such.

Contents

The word "isentropic" can be interpreted in another way, since its meaning is deducible from its etymology. It means a process in which the entropy of the system remains unchanged; as mentioned, this could occur if the process is both adiabatic and reversible. However, this could also occur in a system where the work done on the system includes friction internal to the system, and heat is withdrawn from the system in just the right amount to compensate for the internal friction, so as to leave the entropy unchanged. [8] However, in relation to the universe, the entropy of the universe would increase as a result, in agreement with the Second Law of Thermodynamics.

Background

The second law of thermodynamics states [9] [10] that

where is the amount of energy the system gains by heating, is the temperature of the surroundings, and is the change in entropy. The equal sign refers to a reversible process, which is an imagined idealized theoretical limit, never actually occurring in physical reality, with essentially equal temperatures of system and surroundings. [11] [12] For an isentropic process, if also reversible, there is no transfer of energy as heat because the process is adiabatic; δQ = 0. In contrast, if the process is irreversible, entropy is produced within the system; consequently, in order to maintain constant entropy within the system, energy must be simultaneously removed from the system as heat.

For reversible processes, an isentropic transformation is carried out by thermally "insulating" the system from its surroundings. Temperature is the thermodynamic conjugate variable to entropy, thus the conjugate process would be an isothermal process, in which the system is thermally "connected" to a constant-temperature heat bath.

Isentropic processes in thermodynamic systems

T-s (entropy vs. temperature) diagram of an isentropic process, which is a vertical line segment Isentropic.jpg
T–s (entropy vs. temperature) diagram of an isentropic process, which is a vertical line segment

The entropy of a given mass does not change during a process that is internally reversible and adiabatic. A process during which the entropy remains constant is called an isentropic process, written or . [13] Some examples of theoretically isentropic thermodynamic devices are pumps, gas compressors, turbines, nozzles, and diffusers.

Isentropic efficiencies of steady-flow devices in thermodynamic systems

Most steady-flow devices operate under adiabatic conditions, and the ideal process for these devices is the isentropic process. The parameter that describes how efficiently a device approximates a corresponding isentropic device is called isentropic or adiabatic efficiency. [13]

Isentropic efficiency of turbines:

Isentropic efficiency of compressors:

Isentropic efficiency of nozzles:

For all the above equations:

is the specific enthalpy at the entrance state,
is the specific enthalpy at the exit state for the actual process,
is the specific enthalpy at the exit state for the isentropic process.

Isentropic devices in thermodynamic cycles

CycleIsentropic stepDescription
Ideal Rankine cycle 1→2Isentropic compression in a pump
Ideal Rankine cycle 3→4Isentropic expansion in a turbine
Ideal Carnot cycle 2→3Isentropic expansion
Ideal Carnot cycle 4→1Isentropic compression
Ideal Otto cycle 1→2Isentropic compression
Ideal Otto cycle 3→4Isentropic expansion
Ideal Diesel cycle 1→2Isentropic compression
Ideal Diesel cycle 3→4Isentropic expansion
Ideal Brayton cycle 1→2Isentropic compression in a compressor
Ideal Brayton cycle 3→4Isentropic expansion in a turbine
Ideal vapor-compression refrigeration cycle1→2Isentropic compression in a compressor
Ideal Lenoir cycle 2→3Isentropic expansion

Note: The isentropic assumptions are only applicable with ideal cycles. Real cycles have inherent losses due to compressor and turbine inefficiencies and the second law of thermodynamics. Real systems are not truly isentropic, but isentropic behavior is an adequate approximation for many calculation purposes.

Isentropic flow

In fluid dynamics, an isentropic flow is a fluid flow that is both adiabatic and reversible. That is, no heat is added to the flow, and no energy transformations occur due to friction or dissipative effects. For an isentropic flow of a perfect gas, several relations can be derived to define the pressure, density and temperature along a streamline.

Note that energy can be exchanged with the flow in an isentropic transformation, as long as it doesn't happen as heat exchange. An example of such an exchange would be an isentropic expansion or compression that entails work done on or by the flow.

For an isentropic flow, entropy density can vary between different streamlines. If the entropy density is the same everywhere, then the flow is said to be homentropic.

Derivation of the isentropic relations

For a closed system, the total change in energy of a system is the sum of the work done and the heat added:

The reversible work done on a system by changing the volume is

where is the pressure, and is the volume. The change in enthalpy () is given by

Then for a process that is both reversible and adiabatic (i.e. no heat transfer occurs), , and so All reversible adiabatic processes are isentropic. This leads to two important observations:

Next, a great deal can be computed for isentropic processes of an ideal gas. For any transformation of an ideal gas, it is always true that

, and

Using the general results derived above for and , then

So for an ideal gas, the heat capacity ratio can be written as

For a calorically perfect gas is constant. Hence on integrating the above equation, assuming a calorically perfect gas, we get

that is,

Using the equation of state for an ideal gas, ,

(Proof: But nR = constant itself, so .)

also, for constant (per mole),

and

Thus for isentropic processes with an ideal gas,

or

Table of isentropic relations for an ideal gas

Derived from

where:

= pressure,
= volume,
= ratio of specific heats = ,
= temperature,
= mass,
= gas constant for the specific gas = ,
= universal gas constant,
= molecular weight of the specific gas,
= density,
= specific heat at constant pressure,
= specific heat at constant volume.

See also

Notes

  1. Partington, J. R. (1949), An Advanced Treatise on Physical Chemistry., vol. 1, Fundamental Principles. The Properties of Gases, London: Longmans, Green and Co., p. 122.
  2. Kestin, J. (1966). A Course in Thermodynamics, Blaisdell Publishing Company, Waltham MA, p. 196.
  3. Münster, A. (1970). Classical Thermodynamics, translated by E. S. Halberstadt, Wiley–Interscience, London, ISBN   0-471-62430-6, p. 13.
  4. Haase, R. (1971). Survey of Fundamental Laws, chapter 1 of Thermodynamics, pages 1–97 of volume 1, ed. W. Jost, of Physical Chemistry. An Advanced Treatise, ed. H. Eyring, D. Henderson, W. Jost, Academic Press, New York, lcn 73–117081, p. 71.
  5. Borgnakke, C., Sonntag., R.E. (2009). Fundamentals of Thermodynamics, seventh edition, Wiley, ISBN   978-0-470-04192-5, p. 310.
  6. Massey, B. S. (1970), Mechanics of Fluids, Section 12.2 (2nd edition) Van Nostrand Reinhold Company, London. Library of Congress Catalog Card Number: 67-25005, p. 19.
  7. Çengel, Y. A., Boles, M. A. (2015). Thermodynamics: An Engineering Approach, 8th edition, McGraw-Hill, New York, ISBN   978-0-07-339817-4, p. 340.
  8. Çengel, Y. A., Boles, M. A. (2015). Thermodynamics: An Engineering Approach, 8th edition, McGraw-Hill, New York, ISBN   978-0-07-339817-4, pp. 340–341.
  9. Mortimer, R. G. Physical Chemistry, 3rd ed., p. 120, Academic Press, 2008.
  10. Fermi, E. Thermodynamics, footnote on p. 48, Dover Publications,1956 (still in print).
  11. Guggenheim, E. A. (1985). Thermodynamics. An Advanced Treatment for Chemists and Physicists, seventh edition, North Holland, Amsterdam, ISBN   0444869514, p. 12: "As a limiting case between natural and unnatural processes[,] we have reversible processes, which consist of the passage in either direction through a continuous series of equilibrium states. Reversible processes do not actually occur..."
  12. Kestin, J. (1966). A Course in Thermodynamics, Blaisdell Publishing Company, Waltham MA, p. 127: "However, by a stretch of imagination, it was accepted that a process, compression or expansion, as desired, could be performed 'infinitely slowly'[,] or as is sometimes said, quasistatically." P. 130: "It is clear that all natural processes are irreversible and that reversible processes constitute convenient idealizations only."
  13. 1 2 Cengel, Yunus A., and Michaeul A. Boles. Thermodynamics: An Engineering Approach. 7th Edition ed. New York: Mcgraw-Hill, 2012. Print.

Related Research Articles

<span class="mw-page-title-main">Adiabatic process</span> Thermodynamic process in which no mass or heat is exchanged with surroundings

In thermodynamics, an adiabatic process is a type of thermodynamic process that occurs without transferring heat or mass between the thermodynamic system and its environment. Unlike an isothermal process, an adiabatic process transfers energy to the surroundings only as work. As a key concept in thermodynamics, the adiabatic process supports the theory that explains the first law of thermodynamics.

<span class="mw-page-title-main">Ideal gas</span> Mathematical model which approximates the behavior of real gases

An ideal gas is a theoretical gas composed of many randomly moving point particles that are not subject to interparticle interactions. The ideal gas concept is useful because it obeys the ideal gas law, a simplified equation of state, and is amenable to analysis under statistical mechanics. The requirement of zero interaction can often be relaxed if, for example, the interaction is perfectly elastic or regarded as point-like collisions.

In electrochemistry, the Nernst equation is a chemical thermodynamical relationship that permits the calculation of the reduction potential of a reaction from the standard electrode potential, absolute temperature, the number of electrons involved in the redox reaction, and activities of the chemical species undergoing reduction and oxidation respectively. It was named after Walther Nernst, a German physical chemist who formulated the equation.

<span class="mw-page-title-main">Second law of thermodynamics</span> Physical law for entropy and heat

The second law of thermodynamics is a physical law based on universal experience concerning heat and energy interconversions. One simple statement of the law is that heat always moves from hotter objects to colder objects, unless energy in some form is supplied to reverse the direction of heat flow. Another definition is: "Not all heat energy can be converted into work in a cyclic process."

<span class="mw-page-title-main">Otto cycle</span> Thermodynamic cycle for spark ignition piston engines

An Otto cycle is an idealized thermodynamic cycle that describes the functioning of a typical spark ignition piston engine. It is the thermodynamic cycle most commonly found in automobile engines.

<span class="mw-page-title-main">Third law of thermodynamics</span> Law of physics

The third law of thermodynamics states, regarding the properties of closed systems in thermodynamic equilibrium:

The entropy of a system approaches a constant value when its temperature approaches absolute zero.

<span class="mw-page-title-main">Isothermal process</span> Thermodynamic process in which temperature remains constant

In thermodynamics, an isothermal process is a type of thermodynamic process in which the temperature T of a system remains constant: ΔT = 0. This typically occurs when a system is in contact with an outside thermal reservoir, and a change in the system occurs slowly enough to allow the system to be continuously adjusted to the temperature of the reservoir through heat exchange (see quasi-equilibrium). In contrast, an adiabatic process is where a system exchanges no heat with its surroundings (Q = 0).

<span class="mw-page-title-main">Isobaric process</span> Thermodynamic process in which pressure remains constant

In thermodynamics, an isobaric process is a type of thermodynamic process in which the pressure of the system stays constant: ΔP = 0. The heat transferred to the system does work, but also changes the internal energy (U) of the system. This article uses the physics sign convention for work, where positive work is work done by the system. Using this convention, by the first law of thermodynamics,

<span class="mw-page-title-main">Joule expansion</span>

The Joule expansion is an irreversible process in thermodynamics in which a volume of gas is kept in one side of a thermally isolated container, with the other side of the container being evacuated. The partition between the two parts of the container is then opened, and the gas fills the whole container.

<span class="mw-page-title-main">Heat capacity ratio</span> Thermodynamic quantity

In thermal physics and thermodynamics, the heat capacity ratio, also known as the adiabatic index, the ratio of specific heats, or Laplace's coefficient, is the ratio of the heat capacity at constant pressure to heat capacity at constant volume. It is sometimes also known as the isentropic expansion factor and is denoted by γ (gamma) for an ideal gas or κ (kappa), the isentropic exponent for a real gas. The symbol γ is used by aerospace and chemical engineers.

Rayleigh flow refers to frictionless, non-adiabatic flow through a constant area duct where the effect of heat addition or rejection is considered. Compressibility effects often come into consideration, although the Rayleigh flow model certainly also applies to incompressible flow. For this model, the duct area remains constant and no mass is added within the duct. Therefore, unlike Fanno flow, the stagnation temperature is a variable. The heat addition causes a decrease in stagnation pressure, which is known as the Rayleigh effect and is critical in the design of combustion systems. Heat addition will cause both supersonic and subsonic Mach numbers to approach Mach 1, resulting in choked flow. Conversely, heat rejection decreases a subsonic Mach number and increases a supersonic Mach number along the duct. It can be shown that for calorically perfect flows the maximum entropy occurs at M = 1. Rayleigh flow is named after John Strutt, 3rd Baron Rayleigh.

<span class="mw-page-title-main">Thermodynamic cycle</span> Linked cyclic series of thermodynamic processes

A thermodynamic cycle consists of linked sequences of thermodynamic processes that involve transfer of heat and work into and out of the system, while varying pressure, temperature, and other state variables within the system, and that eventually returns the system to its initial state. In the process of passing through a cycle, the working fluid (system) may convert heat from a warm source into useful work, and dispose of the remaining heat to a cold sink, thereby acting as a heat engine. Conversely, the cycle may be reversed and use work to move heat from a cold source and transfer it to a warm sink thereby acting as a heat pump. If at every point in the cycle the system is in thermodynamic equilibrium, the cycle is reversible. Whether carried out reversible or irreversibly, the net entropy change of the system is zero, as entropy is a state function.

<span class="mw-page-title-main">Polytropic process</span>

A polytropic process is a thermodynamic process that obeys the relation:

In classical thermodynamics, entropy is a property of a thermodynamic system that expresses the direction or outcome of spontaneous changes in the system. The term was introduced by Rudolf Clausius in the mid-nineteenth century from the Greek word τρoπή (transformation) to explain the relationship of the internal energy that is available or unavailable for transformations in form of heat and work. Entropy predicts that certain processes are irreversible or impossible, despite not violating the conservation of energy. The definition of entropy is central to the establishment of the second law of thermodynamics, which states that the entropy of isolated systems cannot decrease with time, as they always tend to arrive at a state of thermodynamic equilibrium, where the entropy is highest. Entropy is therefore also considered to be a measure of disorder in the system.

<span class="mw-page-title-main">Lenoir cycle</span>

The Lenoir cycle is an idealized thermodynamic cycle often used to model a pulse jet engine. It is based on the operation of an engine patented by Jean Joseph Etienne Lenoir in 1860. This engine is often thought of as the first commercially produced internal combustion engine. The absence of any compression process in the design leads to lower thermal efficiency than the more well known Otto cycle and Diesel cycle.

<span class="mw-page-title-main">Carnot cycle</span> Idealized thermodynamic cycle

A Carnot cycle is an ideal thermodynamic cycle proposed by French physicist Sadi Carnot in 1824 and expanded upon by others in the 1830s and 1840s. By Carnot's theorem, it provides an upper limit on the efficiency of any classical thermodynamic engine during the conversion of heat into work, or conversely, the efficiency of a refrigeration system in creating a temperature difference through the application of work to the system.

<span class="mw-page-title-main">Prandtl–Meyer expansion fan</span> Phenomenon in fluid dynamics

A supersonic expansion fan, technically known as Prandtl–Meyer expansion fan, a two-dimensional simple wave, is a centered expansion process that occurs when a supersonic flow turns around a convex corner. The fan consists of an infinite number of Mach waves, diverging from a sharp corner. When a flow turns around a smooth and circular corner, these waves can be extended backwards to meet at a point.

Fanno flow is the adiabatic flow through a constant area duct where the effect of friction is considered. Compressibility effects often come into consideration, although the Fanno flow model certainly also applies to incompressible flow. For this model, the duct area remains constant, the flow is assumed to be steady and one-dimensional, and no mass is added within the duct. The Fanno flow model is considered an irreversible process due to viscous effects. The viscous friction causes the flow properties to change along the duct. The frictional effect is modeled as a shear stress at the wall acting on the fluid with uniform properties over any cross section of the duct.

In astrophysics, what is referred to as "entropy" is actually the adiabatic constant derived as follows.

<span class="mw-page-title-main">Entropy production</span> Development of entropy in a thermodynamic system

Entropy production is the amount of entropy which is produced in any irreversible processes such as heat and mass transfer processes including motion of bodies, heat exchange, fluid flow, substances expanding or mixing, anelastic deformation of solids, and any irreversible thermodynamic cycle, including thermal machines such as power plants, heat engines, refrigerators, heat pumps, and air conditioners.

References