Sei sulla pagina 1di 13

Adiabatic process

From Wikipedia, the free encyclopedia

In thermodynamics, an adiabatic process is one that occurs without transfer of heat or matter between a
thermodynamic system and its surroundings. In an adiabatic process, energy is transferred to its surroundings only
as work.[1][2] The adiabatic process provides a rigorous conceptual basis for the theory used to expound the rst law
of thermodynamics, and as such it is a key concept in thermodynamics.

Some chemical and physical processes occur so rapidly that they may be conveniently described by the term
"adiabatic approximation", meaning that there is not enough time for the transfer of energy as heat to take place to
or from the system.[3]

By way of example, the adiabatic ame temperature is an idealization that uses the "adiabatic approximation" so as
to provide an upper limit calculation of temperatures produced by combustion of a fuel. The adiabatic ame
temperature is the temperature that would be achieved by a ame if the process of combustion took place in the
absence of heat loss to the surroundings.

Contents
1 Description
1.1 Various applications of the adiabatic assumption
2 Adiabatic heating and cooling
3 Ideal gas (reversible process)
3.1 Example of adiabatic compression
3.2 Adiabatic free expansion of a gas
3.3 Derivation of PV relation for adiabatic heating and cooling
3.4 Derivation of PT relation for adiabatic heating and cooling
3.5 Derivation of discrete formula
4 Graphing adiabats
5 Etymology
6 Conceptual signicance in thermodynamic theory
7 Divergent usages of the word adiabatic
8 See also
9 References
10 External links

Description
A process that does not involve the transfer of heat or matter into or out of a system, so that Q=0, is called an
adiabatic process, and such a system is said to be adiabatically isolated.[4][5] The assumption that a process is
adiabatic is a frequently made simplifying assumption. For example, the compression of a gas within a cylinder of
an engine is assumed to occur so rapidly that on the time scale of the compression process, little of the system's
energy can be transferred out as heat to the surroundings. Even though the cylinders are not insulated and are quite
conductive, that process is idealized to be adiabatic. The same can be said to be true for the expansion process of
such a system.
The assumption of adiabatic isolation of a system is a useful one, and is often combined with others so as to make
the calculation of the system's behaviour possible. Such assumptions are idealizations. The behaviour of actual
machines deviates from these idealizations, but the assumption of such "perfect" behaviour provide a useful rst
approximation of how the real world works. According to Laplace, when sound travels in a gas, there is no time for
heat conduction in the medium and so the propagation of sound is adiabatic. For such an adiabatic process, the
modulus of elasticity (Young's modulus) can be expressed as E=P, where is the ratio of specic heats at
C
constant pressure and at constant volume (= Cp ) and P is the pressure of the gas .
v

Various applications of the adiabatic assumption

For a closed system, one may write the rst law of thermodynamics as: U=Q+W, where U denotes the
change of the system's internal energy, Q the quantity of energy added to it as heat, and W the work done on it by
its surroundings.

If the system has rigid walls such that work cannot be transferred in or out (W=0), and the walls of the
system are not adiabatic and energy is added in the form of heat (Q>0), and there is no phase change, the
temperature of the system will rise.
If the system has rigid walls such that pressurevolume work cannot be done, and the system walls are
adiabatic (Q=0), but energy is added as isochoric work in the form of friction or the stirring of a viscous
uid within the system (W>0), and there is no phase change, the temperature of the system will rise.
If the system walls are adiabatic (Q=0), but not rigid (W0), and, in a ctive idealized process, energy is
added to the system in the form of frictionless, non-viscous pressurevolume work, and there is no phase
change, the temperature of the system will rise. Such a process is called an isentropic process and is said to
be "reversible". Fictively, if the process is reversed, the energy added as work can be recovered entirely as
work done by the system. If the system contains a compressible gas and is reduced in volume, the uncertainty
of the position of the gas is reduced, and seemingly would reduce the entropy of the system, but the
temperature of the system will rise as the process is isentropic (S=0). Should the work be added in such a
way that friction or viscous forces are operating within the system, then the process is not isentropic, and if
there is no phase change, then the temperature of the system will rise, the process is said to be "irreversible",
and the work added to the system is not entirely recoverable in the form of work.
If the walls of a system are not adiabatic, and energy is transferred in as heat, entropy is transferred into the
system with the heat. Such a process is neither adiabatic nor isentropic, having Q>0, and S>0
according to the second law of thermodynamics.

Naturally occurring adiabatic processes are irreversible (entropy is produced).

The transfer of energy as work into an adiabatically isolated system can be imagined as being of two idealized
extreme kinds. In one such kind, there is no entropy produced within the system (no friction, viscous dissipation,
etc.), and the work is only pressure-volume work (denoted by PdV). In nature, this ideal kind occurs only
approximately, because it demands an innitely slow process and no sources of dissipation.

The other extreme kind of work is isochoric work (dV=0), for which energy is added as work solely through
friction or viscous dissipation within the system. A stirrer that transfers energy to a viscous uid of an adiabatically
isolated system with rigid walls, without phase change, will cause a rise in temperature of the uid, but that work is
not recoverable. Isochoric work is irreversible.[6] The second law of thermodynamics observes that a natural
process, of transfer of energy as work, always consists at least of isochoric work and often both of these extreme
kinds of work. Every natural process, adiabatic or not, is irreversible, with S>0, as friction or viscosity are
always present to some extent.

Adiabatic heating and cooling


The adiabatic compression of a gas causes a rise in temperature of the gas. Adiabatic expansion against pressure, or
a spring, causes a drop in temperature. In contrast, free expansion is an isothermal process for an ideal gas.

Adiabatic heating occurs when the pressure of a gas is increased from work done on it by its surroundings, e.g., a
piston compressing a gas contained within a cylinder and raising the temperature where in many practical situations
heat conduction through walls can be slow compared with the compression time. This nds practical application in
diesel engines which rely on the lack of heat dissipation during the compression stroke to elevate the fuel vapor
temperature sufciently to ignite it.

Adiabatic heating occurs in the Earth's atmosphere when an air mass descends, for example, in a katabatic wind,
Foehn wind, or chinook wind owing downhill over a mountain range. When a parcel of air descends, the pressure
on the parcel increases. Due to this increase in pressure, the parcel's volume decreases and its temperature increases
as work is done on the parcel of air, thus increasing its internal energy, which manifests itself by a rise in the
temperature of that mass of air. The parcel of air can only slowly dissipate the energy by conduction or radiation
(heat), and to a rst approximation it can be considered adiabatically isolated and the process an adiabatic process.

Adiabatic cooling occurs when the pressure on an adiabatically isolated system is decreased, allowing it to expand,
thus causing it to do work on its surroundings. When the pressure applied on a parcel of air is reduced, the air in the
parcel is allowed to expand; as the volume increases, the temperature falls as its internal energy decreases.
Adiabatic cooling occurs in the Earth's atmosphere with orographic lifting and lee waves, and this can form pileus
or lenticular clouds.

Adiabatic cooling does not have to involve a uid. One technique used to reach very low temperatures (thousandths
and even millionths of a degree above absolute zero) is via adiabatic demagnetisation, where the change in
magnetic eld on a magnetic material is used to provide adiabatic cooling. Also, the contents of an expanding
universe can be described (to rst order) as an adiabatically cooling uid. (See heat death of the universe.)

Rising magma also undergoes adiabatic cooling before eruption, particularly signicant in the case of magmas that
rise quickly from great depths such as kimberlites.[7]

Such temperature changes can be quantied using the ideal gas law, or the hydrostatic equation for atmospheric
processes.

In practice, no process is truly adiabatic. Many processes rely on a large difference in time scales of the process of
interest and the rate of heat dissipation across a system boundary, and thus are approximated by using an adiabatic
assumption. There is always some heat loss, as no perfect insulators exist.

Ideal gas (reversible process)


The mathematical equation for an ideal gas undergoing a reversible (i.e., no entropy generation) adiabatic process
can be represented by the polytropic process equation[3]

where P is pressure, V is volume, and for this case n= where

CP being the specic heat for constant pressure, CV being the specic heat for constant volume, is the adiabatic
index, and f is the number of degrees of freedom (3 for monatomic gas, 5 for diatomic gas and collinear molecules
e.g. carbon dioxide).
5
For a monatomic ideal gas, = 3 , and for a diatomic
gas (such as nitrogen and oxygen, the main components
7
of air) = 5 .[8] Note that the above formula is only
applicable to classical ideal gases and not BoseEinstein
or Fermi gases.

For reversible adiabatic processes, it is also true that


[3]

where T is an absolute temperature. This can also be


written as
[3] For a simple substance, during an adiabatic process in which
the volume increases, the internal energy of the working
substance must decrease
Example of adiabatic compression
The compression stroke in a gasoline engine can be used
as an example of adiabatic compression. The model assumptions are: the uncompressed volume of the cylinder is
one litre (1l = 1000cm3 = 0.001m3 ); the gas within is the air consisting of molecular nitrogen and oxygen only
7
(thus a diatomic gas with ve degrees of freedom and so = 5 ); the compression ratio of the engine is 10:1 (that
is, the 1l volume of uncompressed gas is reduced to 0.1l by the piston); and the uncompressed gas is at
approximately room temperature and pressure (a warm room temperature of ~27C or 300K, and a pressure of
1bar = 100kPa, i.e. typical sea-level atmospheric pressure).

so our adiabatic constant for this example is about 6.31Pam4.2.

The gas is now compressed to a 0.1l (0.0001m3) volume (we will assume this happens quickly enough that no
heat can enter or leave the gas through the walls). The adiabatic constant remains the same, but with the resulting
pressure unknown

so solving for P:

or 25.1bar. Note that this pressure increase is more than a simple 10:1 compression ratio would indicate; this is
because the gas is not only compressed, but the work done to compress the gas also increases its internal energy
which manifests itself by a rise in the gas's temperature and an additional rise in pressure above what would result
from a simplistic calculation of 10 times the original pressure.
We can solve for the temperature of the compressed gas in the engine cylinder as well, using the ideal gas law,
PV=nRT (n is amount of gas in mol and R the gas constant for that gas). Our initial conditions being 100kPa of
pressure, 1l volume, and 300K of temperature, our experimental constant (=nR) is:

We know the compressed gas has V= 0.1l and P= 2.51 106Pa, so we can solve for temperature:

That is a nal temperature of 753K, or 479C, or 896F, well above the ignition point of many fuels. This is why
a high-compression engine requires fuels specially formulated to not self-ignite (which would cause engine
knocking when operated under these conditions of temperature and pressure), or that a supercharger with an
intercooler to provide a pressure boost but with a lower temperature rise would be advantageous. A diesel engine
operates under even more extreme conditions, with compression ratios of 20:1 or more being typical, in order to
provide a very high gas temperature which ensures immediate ignition of the injected fuel.

Adiabatic free expansion of a gas


For an adiabatic free expansion of an ideal gas, the gas is contained in an insulated container and then allowed to
expand in a vacuum. Because there is no external pressure for the gas to expand against, the work done by or on the
system is zero. Since this process does not involve any heat transfer or work, the rst law of thermodynamics then
implies that the net internal energy change of the system is zero. For an ideal gas, the temperature remains constant
because the internal energy only depends on temperature in that case. Since at constant temperature, the entropy is
proportional to the volume, the entropy increases in this case, therefore this process is irreversible.

Derivation of PV relation for adiabatic heating and cooling

The denition of an adiabatic process is that heat transfer to the system is zero, Q=0. Then, according to the rst
law of thermodynamics,

where dU is the change in the internal energy of the system and W is work done by the system. Any work (W)
done must be done at the expense of internal energy U, since no heat Q is being supplied from the surroundings.
Pressurevolume work W done by the system is dened as

However, P does not remain constant during an adiabatic process but instead changes along with V.

It is desired to know how the values of dP and dV relate to each other as the adiabatic process proceeds. For an
ideal gas the internal energy is given by

where is the number of degrees of freedom divided by two, R is the universal gas constant and n is the number of
moles in the system (a constant).

Differentiating Equation (3) and use of the ideal gas law, PV=nRT, yields
Equation (4) is often expressed as dU=nCVdT because CV=R.

Now substitute equations (2) and (4) into equation (1) to obtain

factorize PdV:

and divide both sides by PV:

After integrating the left and right sides from V0 to V and from P0 to P and changing the sides respectively,

+1
Exponentiate both sides, and substitute with , the heat capacity ratio

and eliminate the negative sign to obtain

Therefore,

and

Derivation of PT relation for adiabatic heating and cooling


Substituting the ideal gas law into the above, we obtain

which simplies to
Derivation of discrete formula

The change in internal energy of a system, measured from state 1 to state 2, is equal to

At the same time, the work done by the pressure-volume changes as a result from this process, is equal to

Since we require the process to be adiabatic, the following equation needs to be true

By the previous derivation,

Rearranging (4) gives

Substituting this into (2) gives

Integrating,

+1
Substituting = ,

Rearranging,

Using the ideal gas law and assuming a constant molar quantity (as often happens in practical cases),

By the continuous formula,


Or,

Substituting into the previous expression for W,

Substituting this expression and (1) in (3) gives

Simplifying,

Simplifying,

Simplifying,

Graphing adiabats
An adiabat is a curve of constant entropy in a diagram. Some properties of adiabats on a PV diagram are indicated.
These properties may be read from the classical behaviour of ideal gases, except in the region where PV becomes
small (low temperature), where quantum effects become important.

1. Every adiabat asymptotically approaches both the V axis and the P axis (just like isotherms).
2. Each adiabat intersects each isotherm exactly once.
3. An adiabat looks similar to an isotherm, except that during an expansion, an adiabat loses more pressure than
an isotherm, so it has a steeper inclination (more vertical).
4. If isotherms are concave towards the north-east direction (45), then adiabats are concave towards the east
north-east (31).
5. If adiabats and isotherms are graphed at regular intervals of entropy and temperature, respectively (like
altitude on a contour map), then as the eye moves towards the axes (towards the south-west), it sees the
density of isotherms stay constant, but it sees the density of adiabats grow. The exception is very near
absolute zero, where the density of adiabats drops sharply and they become rare (see Nernst's theorem).

The following diagram is a PV diagram with a superposition of adiabats and isotherms:

The isotherms are the red curves and the adiabats are the black curves.

The adiabats are isentropic.

Volume is the horizontal axis and pressure is the vertical axis.

Etymology
The term adiabatic (/dibtk/) is an anglicization of the Greek term "impassable" (used by
Xenophon of rivers). It is used in the thermodynamic sense by Rankine (1866),[9][10] and adopted by Maxwell in
1871 (explicitly attributing the term to Rankine).[11] The etymological origin corresponds here to an impossibility
of transfer of energy as heat and of transfer of matter across the wall.

The Greek word is formed from privative - ("not") and , "passable", in turn deriving from
("through"), and ("to walk, go, come").[12]

Conceptual signicance in thermodynamic theory


The adiabatic process has been important for thermodynamics since its early days. It was important in the work of
Joule, because it provided a way of nearly directly relating quantities of heat and work.

For a thermodynamic system that is enclosed by walls that do not allow mass transfer, energy can pass in and out
only as heat or work. Thus a quantity of work can be related almost directly to an equivalent quantity of heat in a
cycle of two limbs. The rst is an isochoric adiabatic work process that adds to the system's internal energy. Then
an isochoric and workless heat transfer returns the system to its original state. The rst limb adds a denite amount
of energy and the second removes it. Accordingly, Rankine measured quantity of heat in units of work, rather than
as a calorimetric quantity .[13] In 1854, Rankine used a quantity that he called "the thermodynamic function" that
later was called entropy, and at that time he wrote also of the "curve of no transmission of heat",[14] which he later
called an adiabatic curve.[9] Besides it two isothermal limbs, Carnot's cycle has two adiabatic limbs.

For the foundations of thermodynamics, the conceptual importance of this was emphasized by Bryan,[15] by
Carathodory,[1] and by Born.[16] The reason is that calorimetry presupposes a type of temperature as already
dened before the statement of the rst law of thermodynamics, such as one based on empirical scales. Such a
presupposition involves making the distinction between empirical temperature and absolute temperature. Rather,
the denition of absolute thermodynamic temperature is best left till the second law is available as a conceptual
basis.[17]

In the eighteenth century, the law of conservation of energy was yet to be fully formulated or established, and the
nature of heat was debated. One approach to these problems was to regard heat, measured by calorimetry, as a
primary substance that is conserved in quantity. By the middle of the nineteenth century, it was recognized as a
form of energy, and the law of conservation of energy was thereby also recognized. The view that eventually
established itself, and is currently regarded as right, is that the law of conservation of energy is a primary axiom,
and that heat is to be analyzed as consequential. In this light, heat cannot be a component of the total energy of a
single body because it is not a state variable, but, rather, is a variable that describes a process of transfer between
two bodies. The adiabatic process is important because it is a logical ingredient of this current view.[17]

Divergent usages of the word adiabatic


This present article is written from the viewpoint of macroscopic thermodynamics, and the word adiabatic is used
in this article in the traditional way of thermodynamics, introduced by Rankine. It is pointed out in the present
article that, for example, if a compression of a gas is rapid, then there is little time for heat transfer to occur, even
when the gas is not adiabatically isolated by a denite wall. In this sense, a rapid compression of a gas is sometimes
approximately or loosely said to be adiabatic, though often far from isentropic, even when the gas is not
adiabatically isolated by a denite wall.

Quantum mechanics and quantum statistical mechanics, however, use the word adiabatic in a very different sense,
one that can at times seem almost opposite to the classical thermodynamic sense. In quantum theory, the word
adiabatic can mean something perhaps near isentropic, or perhaps near quasi-static, but the usage of the word is
very different between the two disciplines.

On the one hand, in quantum theory, if a perturbative element of compressive work is done almost innitely slowly
(that is to say quasi-statically), it is said to have been done adiabatically. The idea is that the shapes of the
eigenfunctions change slowly and continuously, so that no quantum jump is triggered, and the change is virtually
reversible. While the occupation numbers are unchanged, nevertheless there is change in the energy levels of one-
to-one corresponding, pre- and post-compression, eigenstates. Thus a perturbative element of work has been done
without heat transfer and without introduction of random change within the system. For example, Max Born writes
"Actually, it is usually the 'adiabatic' case with which we have to do: i.e. the limiting case where the external force
(or the reaction of the parts of the system on each other) acts very slowly. In this case, to a very high approximation
that is, there is no probability for a transition, and the system is in the initial state after cessation of the perturbation.
Such a slow perturbation is therefore reversible, as it is classically."[18]

On the other hand, in quantum theory, if a perturbative element of compressive work is done rapidly, it randomly
changes the occupation numbers of the eigenstates, as well as changing their shapes. In that theory, such a rapid
change is said not to be adiabatic, and the contrary word diabatic is applied to it. One might guess that perhaps
Clausius, if he were confronted with this, in the now-obsolete language he used in his day, would have said that
"internal work" was done and that 'heat was generated though not transferred'.

In classical thermodynamics, such a rapid change would still be called adiabatic because the system is adiabatically
isolated, and there is no transfer of energy as heat. The strong irreversibility of the change, due to viscosity or other
entropy production, does not impinge on this classical usage.

Thus for a mass of gas, in macroscopic thermodynamics, words are so used that a compression is sometimes
loosely or approximately said to be adiabatic if it is rapid enough to avoid heat transfer, even if the system is not
adiabatically isolated. But in quantum statistical theory, a compression is not called adiabatic if it is rapid, even if
the system is adiabatically isolated in the classical thermodynamic sense of the term. The words are used differently
in the two disciplines, as stated just above.

See also
Cyclic process
First law of thermodynamics
Heat burst
Isobaric process
Isenthalpic process
Isentropic process
Isochoric process
Isothermal process
Polytropic process
Entropy (classical thermodynamics)
Quasistatic process
Total air temperature
Magnetic refrigeration

References
1. Carathodory, C. (1909). "Untersuchungen ber die Grundlagen der Thermodynamik". Mathematische
Annalen. 67: 355386. doi:10.1007/BF01450409 (https://doi.org/10.1007%2FBF01450409).. A translation
may be found here (http://neo-classical-physics.info/uploads/3/0/6/5/3065888/caratheodory_-_thermodynami
cs.pdf). Also a mostly reliable translation is to be found (https://books.google.com/books?id=xwBRAAAAM
AAJ&q=Investigation+into+the+foundations) in Kestin, J. (1976). The Second Law of Thermodynamics.
Stroudsburg, PA: Dowden, Hutchinson & Ross.
2. Bailyn, M. (1994). A Survey of Thermodynamics. New York, NY: American Institute of Physics Press. p.21.
ISBN0-88318-797-3.
3. Bailyn, M. (1994), pp. 5253.
4. Tisza, L. (1966). Generalized Thermodynamics. Cambridge, MA: MIT Press. p.48. "(adiabatic partitions
inhibit the transfer of heat and mass)"
5. Mnster, A. (1970), p. 48: "mass is an adiabatically inhibited variable."
6. Mnster, A. (1970). Classical Thermodynamics. Translated by Halberstadt, E. S. London: Wiley
Interscience. p.45. ISBN0-471-62430-6.
7. Kavanagh, J. L.; Sparks, R. S. J. (2009). "Temperature changes in ascending kimberlite magmas" (http://mon
ash.academia.edu/JanineKavanagh/Papers/114092/Temperature_changes_in_ascending_kimberlite_magma).
Earth and Planetary Science Letters. Elsevier. 286 (34): 404413. Bibcode:2009E&PSL.286..404K (http://
adsabs.harvard.edu/abs/2009E&PSL.286..404K). doi:10.1016/j.epsl.2009.07.011 (https://doi.org/10.1016%2
Fj.epsl.2009.07.011). Retrieved 18 February 2012.
8. Adiabatic Processes (http://hyperphysics.phy-astr.gsu.edu/hbase/thermo/adiab.html)
9. Rankine, W.J.McQ. (1866). On the theory of explosive gas engines, The Engineer, July 27, 1866; at page 467
of the reprint in Miscellaneous Scientic Papers (https://archive.org/details/miscellaneoussci00rank), edited
by W.J. Millar, 1881, Charles Grifn, London.
10. Partington, J. R. (1949), An Advanced Treatise on Physical Chemistry., volume 1, Fundamental Principles.
The Properties of Gases, London: Longmans, Green and Co., p.122
11. Maxwell, J. C. (1871), Theory of Heat (https://archive.org/details/theoryheat04maxwgoog) (rst ed.),
London: Longmans, Green and Co., p.129
12. Liddell, H.G., Scott, R. (1940). A Greek-English Lexicon, Clarendon Press, Oxford UK.
13. Rankine, W. J. M. (1854). "On the geometrical representation of the expansive action of heat, and theory of
thermodynamic engines". Proc. Roy. Soc. 144: 115175. Miscellaneous Scientic Papers p. 339 (https://archi
ve.org/stream/miscellaneoussci00rank#page/340/mode/1up)
14. Rankine, W. J. M. (1854). "On the geometrical representation of the expansive action of heat, and theory of
thermodynamic engines". Proc. Roy. Soc. 144: 115175. Miscellaneous Scientic Papers p. 341 (https://archi
ve.org/stream/miscellaneoussci00rank#page/341/mode/1up/search/transmission).
15. Bryan, G. H. (1907). Thermodynamics. An Introductory Treatise dealing mainly with First Principles and
their Direct Applications (https://archive.org/details/ost-physics-thermodynamicsin00bryauoft). Leipzig: B.
G. Teubner.
16. Born, M. (1949). "Natural Philosophy of Cause and Chance" (https://archive.org/details/naturalphilosoph032
159mbp). London: Oxford University Press.
17. Bailyn, M. (1994). "Chapter 3". A Survey of Thermodynamics. New York, NY: American Institute of Physics.
ISBN0-88318-797-3.
18. Born, M. (1927). "Physical aspects of quantum mechanics". Nature. 119: 354357.
Bibcode:1927Natur.119..354B (http://adsabs.harvard.edu/abs/1927Natur.119..354B). doi:10.1038/119354a0
(https://doi.org/10.1038%2F119354a0). (Translation by Robert Oppenheimer.)

General

Silbey, Robert J.; et al. (2004). Physical chemistry. Hoboken: Wiley. p.55. ISBN978-0-471-21504-2.
Broholm, Collin. "Adiabatic free expansion." Physics & Astronomy @ Johns Hopkins University. N.p., 26
Nov. 1997. Web. 14 Apr. *Nave, Carl Rod. "Adiabatic Processes." HyperPhysics. N.p., n.d. Web. 14 Apr.
2011. [1] (http://hyperphysics.phy-astr.gsu.edu/hbase/thermo/adiab.html).
Thorngren, Dr. Jane R.. "Adiabatic Processes." Daphne A Palomar College Web Server. N.p., 21 July 1995.
Web. 14 Apr. 2011. [2] (http://daphne.palomar.edu/jthorngren/adiabatic_processes.htm).

External links
Media related to Adiabatic process at Wikimedia Commons
Article in HyperPhysics Encyclopaedia (http://hyperphysics.phy-astr.gsu.edu/hbase/thermo/adiab.html#c1:)

Retrieved from "https://en.wikipedia.org/w/index.php?title=Adiabatic_process&oldid=800325428"

This page was last edited on 12 September 2017, at 20:47.


Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply.
By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia is a registered trademark of
the Wikimedia Foundation, Inc., a non-prot organization.

Potrebbero piacerti anche