Vorticity equation: Difference between revisions

From formulasearchengine
Jump to navigation Jump to search
en>Laustemil
No edit summary
 
(One intermediate revision by one other user not shown)
Line 1: Line 1:
{{pp-protected|expiry=2014-06-17 22:25:18|small=yes}}{{pp-move-indef|small=yes}}
Msvcr71.dll is an significant file which helps help Windows procedure different components of the system including important files. Specifically, the file is selected to aid run corresponding files inside the "Virtual C Runtime Library". These files are significant inside accessing any settings that help the different applications and programs in the system. The msvcr71.dll file fulfills various significant functions; though it's not spared from getting damaged or corrupted. Once the file gets corrupted or damaged, the computer may have a hard time processing plus reading components of the system. However, consumers require not panic considering this problem could be solved by following several procedures. And I usually show we several strategies about Msvcr71.dll.<br><br>You are able to reformat the computer to make it run faster. This will reset a computer to whenever you first used it. Always remember to back up all files plus programs before carrying this out since this usually remove a files from the database. Remember before we do this you need all drivers plus installation files and this should be a last resort in the event you are hunting for slow computer tips.<br><br>With RegCure to improve the begin up plus shut down of your computer. The program shows the scan progress and you shouldn't worry where it is actually functioning at which time. It shows you exactly what occurs. Dynamic link library section of the registry will cause severe application failures. RegCure restores plus repairs the registry and keeps you from DLL. RegCure can make individual corrections, so it may functions for your requires.<br><br>Always see to it which you have installed antivirus, anti-spyware plus anti-adware programs plus have them updated regularly. This can help stop windows XP running slow.<br><br>Many [http://bestregistrycleanerfix.com/registry-reviver registry reviver] s let you to download their product for free, to scan the computer yourself. That way we can see how countless mistakes it finds, where it finds them, and how it can fix them. A ideal registry cleaner will remove the registry difficulties, and optimize plus accelerate your PC, with little effort on the piece.<br><br>Reinstall Windows 7 - If nothing appears to work, reinstall Windows 7 with the installation disc which came with all the pack. Kindly backup or restore all your data to a flash drive or another hard drive/CD etc. before operating the reinstallation.<br><br>Maybe you are asking why these windows XP error messages appear. Well, for we to be able to know the fix, you need to initially recognize where those errors come from. There is this software called registry. A registry is software that shops everything on your PC from a usual configuration, setting, info, plus logs of escapades from installing to UN-installing, saving to deleting, along with a lot more alterations we do in a system pass from it and gets 'tagged' plus saved as a simple file for recovery reasons. Imagine it because a big recorder, a registrar, of all a records in your PC.<br><br>So, the number one thing to do when your computer runs slow is to buy an authentic plus legal registry repair tool that would help you eliminate all problems associated to registry plus aid we enjoy a smooth running computer.
{{Thermodynamics|cTopic=[[Laws of thermodynamics|Laws]]}}
 
The '''first law of [[thermodynamics]]''' is a version of the law of [[conservation of energy]], adapted for [[thermodynamic system]]s. The law of conservation of energy states that the total [[energy]] of an isolated system is constant; energy can be transformed from one form to another, but cannot be created or destroyed. The first law is often formulated by stating that the change in the internal energy of a [[Thermodynamic system#Closed system|closed system]] is equal to the amount of [[heat]] supplied to the system, minus the amount of [[Work (thermodynamics)|work]] done by the system on its surroundings.   Equivalently, [[perpetual motion machines]] of the first kind are impossible.
 
==History==
The process of development of the first law of thermodynamics was by way of many tries and mistakes of investigation, over a period of about half a century. The first full statements of the law came in 1850 from [[Rudolf Clausius]] and by [[William Rankine]]; Rankine's statement was perhaps not quite as clear and distinct as was Clausius'.<ref name="Truesdell, C.A. 1980"/> A main aspect of the struggle was to deal with the previously proposed [[caloric theory]] of heat.
 
[[Germain Hess]] in 1840 stated a [[Hess's Law|conservation law]] for the so-called 'heat of reaction' for chemical reactions.<ref>Hess, H. (1840). Thermochemische Untersuchungen, ''Annalen der Physik und Chemie'' (Poggendorff, Leipzig) '''126'''(6): 385-404 [http://gallica.bnf.fr/ark:/12148/bpt6k151359/f397.image.r=Annalen%20der%20Physik%20(Leipzig)%20125.langEN].</ref> His law was later recognized as a consequence of the first law of thermodynamics, but Hess's statement was not explicitly concerned with the relation between energy exchanges by heat and work.
 
According to [[Clifford Truesdell|Truesdell]] (1980), [[Julius Robert von Mayer]] in 1841 made a statement that meant that "in a process at constant pressure, the heat used to produce expansion is universally interconvertible with work", but this is not a general statement of the first law.<ref>Truesdell, C.A. (1980), pp. 157-158.</ref><ref>Mayer, Robert (1841). Paper: 'Remarks on the Forces of Nature"; as quoted in: Lehninger, A. (1971). Bioenergetics - the Molecular Basis of Biological Energy Transformations, 2nd. Ed. London: The Benjamin/Cummings Publishing Company.</ref>
 
===Original statements: the "thermodynamic approach"===
The original nineteenth century statements of the first law of thermodynamics appeared in a conceptual framework in which transfer of energy as heat was taken as a [[primitive notion]], not defined or constructed by the theoretical development of the framework, but rather presupposed as prior to it and already accepted. The primitive notion of heat was taken as empirically established, especially through calorimetry regarded as a subject in its own right, prior to thermodynamics. Jointly primitive with this notion of heat were the notions of empirical temperature and thermal equilibrium. This framework did not presume a concept of energy in general, but regarded it as derived or synthesized from the prior notions of heat and work. By one author, this framework has been called the "thermodynamic" approach.<ref name="Bailyn 79">Bailyn, M. (1994), p. 79.</ref>
 
The first explicit statement of the first law of thermodynamics, by [[Rudolf Clausius]] in 1850, referred to cyclic thermodynamic processes.
::''In all cases in which work is produced by the agency of heat, a quantity of heat is consumed which is proportional to the work done; and conversely, by the expenditure of an equal quantity of work an equal quantity of heat is produced.''<ref>Clausius, R. (1850). Ueber die bewegende Kraft der Wärme und die Gesetze, welche sich daraus für die Wärmelehre selbst ableiten lassen, ''Annalen der Physik und Chemie'' (Poggendorff, Leipzig), '''155''' (3): 368-394, particularly on page 373 [http://gallica.bnf.fr/ark:/12148/bpt6k15164w/f389.image], translation here taken from Truesdell, C.A. (1980), pp. 188-189.</ref>
 
Clausius also stated the law in another form, referring to the existence of a function of state of the system, the internal energy, and expressed it in terms of a differential equation for the increments of a thermodynamic process. This equation may described as follows:
::''In a thermodynamic process involving a closed system, the increment in the internal energy is equal to the difference between the heat accumulated by the system and the work done by it.''<ref>Clausius, R. (1850). Ueber die bewegende Kraft der Wärme und die Gesetze, welche sich daraus für die Wärmelehre selbst ableiten lassen, ''Annalen der Physik und Chemie'' (Poggendorff, Leipzig), '''155''' (3): 368-394, page 384 [http://gallica.bnf.fr/ark:/12148/bpt6k15164w/f400.image].</ref>
 
Because of its definition in terms of increments, the value of the internal energy of a system is not uniquely defined. It is defined only up to an arbitrary additive constant of integration, which can be adjusted to give arbitrary reference zero levels. This non-uniqueness is in keeping with the abstract mathematical nature of the internal energy. The internal energy is customarily stated relative to a conventionally chosen standard reference state of the system.
 
The concept of internal energy is considered by Bailyn to be of "enormous interest". Its quantity cannot be immediately measured, but can only be inferred, by differencing actual immediate measurements. Bailyn likens it to the energy states of an atom, that were revealed by Bohr's energy relation {{math|''hν'' {{=}} ''E<sub>n′''</sub> − ''E<sub>n′''</sub>}}. In each case, an unmeasurable quantity (the internal energy, the atomic energy level) is revealed by considering the difference of measured quantities (increments of internal energy, quantities of emitted or absorbed radiative energy.<ref>Bailyn, M. (1994), p. 80.</ref>
 
===Conceptual revision: the "mechanical approach"===
 
In 1907, [[George H. Bryan|G.H. Bryan]] wrote about systems between which there is no transfer of matter (closed systems): "'''Definition.''' When energy flows from one system or part of a system to another otherwise than by the performance of mechanical work, the energy so transferred is called ''heat''."<ref>[[George H. Bryan|Bryan, G.H.]] (1907), p.47. Also Bryan had written about this in the ''Enzyklopädie der Mathematischen Wissenschaften'', volume 3, p. 81. Also in 1906 [[Jean Baptiste Perrin]] wrote about it in ''Bull. de la société français de philosophie'', volume 6, p. 81.</ref>
 
Largely through the influence of [[Max Born]],<ref name="Born 1949 {{math|V}}"/> in the twentieth century, this revised conceptual approach to the definition of heat came to be preferred by many writers, including Constantin Carathéodory. It might be called the "mechanical approach".<ref>Bailyn, M. (1994), pp. 65, 79.</ref> This approach takes as its primitive notion energy transferred as work defined by mechanics. From this, it derives the notions of transfer of energy as heat, and of temperature, as theoretical developments. It regards calorimetry as a derived theory. It has an early origin in the nineteenth century, for example in the work of Helmholtz,<ref>Helmholtz, H. (1847).</ref> but also in the work of many others.<ref name="Bailyn 79"/> For this approach, it is necessary to be sure that if there is transfer of energy associated with transfer of matter, then the transfer of energy other than by transfer of matter is by a physically separate pathway, and independently defined and measured, from the transfer of energy by transfer of matter.
 
==Conceptually revised statement, according to the mechanical approach==
 
The revised statement of the law takes the notions of adiabatic mechanical work, and of non-adiabatic transfer of energy, as empirically or theoretically established primitive notions. It rests on the primitive notion of ''walls'', especially [[Adiabatic enclosure|adiabatic walls]], presupposed as physically established. Energy can pass such walls as only as adiabatic work, reversibly or irreversibly. If transfer of energy as work is not permitted between them, two systems separated by an adiabatic wall can come to their respective internal mechanical and material thermodynamic equilibrium states completely independently of one another.<ref>Bailyn, (1994), p. 82.</ref>
 
The revised statement of the law postulates that a change in the internal energy of a system due to an arbitrary process of interest, that takes the system from its specified initial to its specified final state of internal thermodynamic equilibrium, can be determined through the physical existence of a reference process, for those specified states, that occurs purely through stages of adiabatic work.
 
The revised statement is then
 
::''For a closed system, in any arbitrary process of interest that takes it from an initial to a final state of internal thermodynamic equilibrium, the change of internal energy is the same as that for a reference adiabatic work process that links those two states. This is so regardless of the path of the process of interest, and regardless of whether it is an adiabatic or a non-adiabatic process. The reference adiabatic work process may be chosen arbitrarily from amongst the class of all such processes.''
 
This statement is much less close to the empirical basis than are the original statements,<ref name="Pippard 15"/> but is often regarded as conceptually parsimonious in that it rests only on the concepts of adiabatic work and of non-adiabatic processes, not on the concepts of transfer of energy as heat and of empirical temperature that are presupposed by the original statements. Largely through the influence of [[Max Born]], it is often regarded as theoretically preferable because of this conceptual parsimony. Born particularly observes that the revised approach avoids thinking in terms of what he calls the "imported engineering" concept of heat engines.<ref name="Born 1949 {{math|V}}">[[Max Born|Born, M.]] (1949), Lecture {{math|V}}, pp. 31–45.</ref>
 
Basing his thinking on the mechanical approach, Born in 1921, and again in 1949, proposed to revise the definition of heat.<ref name="Born 1949 {{math|V}}"/><ref name="Born 1921"/> In particular, he referred to the work of [[Constantin Carathéodory]], who had in 1909 stated the first law without defining quantity of heat.<ref name="Carathéodory 1909"/> Born's definition was specifically for transfers of energy without transfer of matter, and it has been widely followed in textbooks (examples:<ref name="Münster 23 24"/><ref name="Reif 122"/><ref name="Haase 1971"/>).  Born observes that a transfer of matter between two systems is accompanied by a transfer of internal energy that cannot be resolved into heat and work components. There can be pathways to other systems, spatially separate from that of the matter transfer, that allow heat and work transfer independent of and simultaneous with the matter transfer. Energy is conserved in such transfers.
 
==Description==
The first law of thermodynamics for a closed system was expressed in two ways by Clausius. One way referred to cyclic processes and the inputs and outputs of the system, but did not refer to increments in the internal state of the system. The other way referred to any incremental change in the internal state of the system, and did not expect the process to be cyclic. A cyclic process is one that can be repeated indefinitely often and still eventually leave the system in its original state.
 
In each repetition of a cyclic process, the work done by the system is proportional to the heat consumed by the system. In a cyclic process in which the system does work on its surroundings, it is necessary that some heat be taken in by the system and some be put out, and the difference is the heat consumed by the system in the process. The constant of proportionality is universal and independent of the system and was measured by [[James Prescott Joule|James Joule]] in 1845 and 1847, who described it as the ''[[mechanical equivalent of heat]]''.
 
For a closed system, in any process, the change in the internal energy is considered due to a combination of [[heat]] added to the system and [[Work (thermodynamics)|work]] done ''by'' the system.  Taking <math>\Delta U</math> as a change in internal energy, one writes
 
:<math>\Delta U = Q\, - \, W\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,{\mathrm{(sign\,convention\,of\,Clausius\,and\,generally\,in\,this\,article)}}\, ,</math>
 
where <math>Q</math> and <math>W</math> are quantities of heat supplied to the system by its surroundings and of work done by the system on its surroundings, respectively. This sign convention is implicit in Clausius' statement of the law given above, and is consistent with the use of thermodynamics to study [[heat engine]]s, which provide useful work that is regarded as positive.
 
In modern style of teaching science, however, it is conventional to use the [[IUPAC]] convention by which the first law is formulated in terms of the work done on the system. With this alternate sign convention for work, the first law for a closed system may be written:
:<math>\Delta U = Q + W\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\mathrm{(sign\,convention\,of\,IUPAC)}\, .</math><ref>[http://media.iupac.org/publications/books/gbook/IUPAC-GB3-2ndPrinting-Online-22apr2011.pdf Quantities, Units and Symbols in Physical Chemistry (IUPAC Green Book)] See Sec. 2.11 Chemical Thermodynamics</ref>
This convention follows physicists such as [[Max Planck]],<ref>Planck, M.(1897/1903). [https://ia700200.us.archive.org/15/items/treatiseonthermo00planrich/treatiseonthermo00planrich.pdf ''Treatise on Thermodynamics'', translated by A. Ogg, Longmans, Green & Co., London.], p. 43</ref> and considers all net energy transfers to the system as positive and all net energy transfers from the system as negative, irrespective of any use for the system as an engine or other device.
 
When a system expands in a fictive [[quasistatic process]], the work done by the system on the environment is the product, {{math|''P''&nbsp;d''V''}}, &nbsp;of pressure, {{math|''P''}}, and volume change, {{math|d''V''}}, whereas the work done ''on'' the system is {{math|&nbsp;-''P''&nbsp;d''V''}}.&nbsp; Using either sign convention for work, the change in internal energy of the system is:
 
:<math>\mathrm d U = \delta Q - P \, \mathrm d V\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\text{(quasi-static process)} ,</math>
 
where {{math|''δQ''}} denotes the infinitesimal increment of heat supplied to the system from its surroundings.
 
Work and heat are expressions of actual physical processes of supply or removal of energy, while the internal energy {{math|''U''}} is a mathematical abstraction that keeps account of the exchanges of energy that befall the system. Thus the term heat for {{math|''Q''}} means "that amount of energy added or removed by conduction of heat or by thermal radiation", rather than referring to a form of energy within the system. Likewise, the term work energy for {{math|''W''}} means "that amount of energy gained or lost as the result of work". Internal energy is a property of the system whereas work done and heat supplied are not. A significant result of this distinction is that a given internal energy change {{math|Δ''U''}} can be achieved by, in principle, many combinations of heat and work.
 
==Various statements of the law for closed systems==
 
The law is of very great importance and generality and is consequently thought of from several points of view. Most careful textbook statements of the law express it for closed systems. It is stated in several ways, sometimes even by the same author.<ref name="Bailyn 79"/><ref>Münster, A. (1970).</ref>
 
For the thermodynamics of closed systems, the distinction between transfers of energy as work and as heat is central and is within the scope of the present article. For the thermodynamics of [[Thermodynamic system#Open system|open systems]], such a distinction is beyond the scope of the present article, but some limited comments are made on it in the section below headed [[First law of thermodynamics#First law of thermodynamics for open systems|'First law of thermodynamics for open systems']].
 
There are two main ways of stating a law of thermodynamics, physically or mathematically. They should be logically coherent and consistent with one another.<ref>[[John Gamble Kirkwood|Kirkwood, J.G.]], Oppenheim, I. (1961), pp. 31–33.</ref>
 
An example of a physical statement is that of [[Max Planck|Planck]] (1897/1903):
 
:It is in no way possible, either by mechanical, thermal, chemical, or other devices, to obtain perpetual motion, i.e. it is impossible to construct an engine which will work in a cycle and produce continuous work, or kinetic energy, from nothing.<ref>Planck, M.(1897/1903), p. 86.</ref>
 
This physical statement is restricted neither to closed systems nor to systems with states that are strictly defined only for thermodynamic equilibrium; it has meaning also for open systems and for systems with states that are not in thermodynamic equilibrium.
 
An example of a mathematical statement is that of Crawford (1963):
 
::::::For a given system we let {{math|Δ''E''<sup>&nbsp;kin</sup>&nbsp;{{=}}}} large-scale mechanical energy, {{math|Δ''E''<sup>&nbsp;pot</sup>&nbsp;{{=}}}} large-scale potential energy, and {{math|Δ''E''<sup>&nbsp;tot</sup>&nbsp;{{=}}}} total energy. The first two quantities are specifiable in terms of appropriate mechanical variables, and by definition
 
:::::::::<math>E^{\mathrm{tot}}=E^{\mathrm{kin}}+E^{\mathrm{pot}}+U\,\,.</math>
 
::::::For any finite process, whether reversible or irreversible,
 
:::::::::<math>\Delta E^{\mathrm{tot}}=\Delta E^{\mathrm{kin}}+\Delta E^{\mathrm{pot}}+\Delta U\,\,.</math>
 
::::::The first law in a form that involves the principle of conservation of energy more generally is
 
:::::::::<math>\Delta E^{\mathrm{tot}}=Q+W\,\,.</math>
 
::::::Here {{math|''Q''}} and {{math|''W''}} are heat and work added, with no restrictions as to whether the process is reversible, quasistatic, or irreversible.[Warner, ''Am. J. Phys.'', '''29''', 124 (1961)]<ref name ="Crawford 106">Crawford, F.H. (1963), pp. 106–107.</ref>
 
This statement by Crawford, for {{math|''W''}}, uses the sign convention of IUPAC, not that of Clausius. Though it does not explicitly say so, this statement refers to closed systems, and to internal energy {{math|''U''}} defined for bodies in states of thermodynamic equilibrium, which possess well-defined temperatures.
 
The history of statements of the law for closed systems has two main periods, before and after the work of [[George H. Bryan|Bryan]] (1907),<ref name="Bryan 1907">[[George H. Bryan|Bryan, G.H.]] (1907), p. 47.</ref> of [[Constantin Carathéodory|Carathéodory]] (1909),<ref name="Carathéodory 1909">[[Constantin Carathéodory|Carathéodory]], C. (1909).</ref> and the approval of Carathéodory's work given by [[Max Born|Born]] (1921).<ref name="Born 1921">Born, M. (1921). Kritische Betrachtungen zur traditionellen Darstellung der Thermodynamik, ''Physik. Zeitschr.'' '''22''': 218–224.</ref> The earlier traditional versions of the law for closed systems are nowadays often considered to be out of date.
 
Carathéodory's celebrated presentation of equilibrium thermodynamics<ref name="Carathéodory 1909"/> refers to closed systems, which are allowed to contain several phases connected by internal walls of various kinds of impermeability and permeability (explicitly including walls that are permeable only to heat). Carathéodory's 1909 version of the first law of thermodynamics was stated in an axiom which refrained from defining or mentioning temperature or quantity of heat transferred. That axiom stated that the internal energy of a phase in equilibrium is a function of state, that the sum of the internal energies of the phases is the total internal energy of the system, and that the value of the total internal energy of the system is changed by the amount of work done adiabatically on it, considering work as a form of energy. That article considered this statement to be an expression of the law of conservation of energy for such systems. This version is nowadays widely accepted as authoritative, but is stated in slightly varied ways by different authors.
 
The 1909 Carathéodory statement of the law in axiomatic form does not mention heat or temperature, but the equilibrium states to which it refers are explicitly defined by variable sets that necessarily include "non-deformation variables", such as pressures, which, within reasonable restrictions, can be rightly interpreted as empirical temperatures,<ref>Buchdahl, H.A. (1966), p. 34.</ref> and the walls connecting the phases of the system are explicitly defined as possibly impermeable to heat or permeable only to heat.
 
According to Münster (1970), "A somewhat unsatisfactory aspect of Carathéodory's theory is that a consequence of the Second Law must be considered at this point [in the statement of the first law], i.e. that it is not always possible to reach any state 2 from any other state 1 by means of an adiabatic process." Münster instances that no adiabatic process can reduce the internal energy of a system at constant volume.<ref name="Münster 23 24"/> Carathéodory's paper asserts that its statement of the first law corresponds exactly to Joule's experimental arrangement, regarded as an instance of adiabatic work. It does not point out that Joule's experimental arrangement performed essentially irreversible work, through friction of paddles in a liquid, or passage of electric current through a resistance inside the system, driven by motion of a coil and inductive heating, or by an external current source, which can access the system only by the passage of electrons, and so is not strictly adiabatic, because electrons are a form of matter, which cannot penetrate adiabatic walls. The paper goes on to base its main argument on the possibility of quasi-static adiabatic work, which is essentially reversible. The paper asserts that it will avoid reference to Carnot cycles, and then proceeds to base its argument on cycles of forward and backward quasi-static adiabatic stages, with isothermal stages of zero magnitude.
 
Some respected modern statements of the first law for closed systems assert the existence of internal energy as a function of state defined in terms of adiabatic work and accept the idea that heat is not defined in its own right, that is to say calorimetrically or as due to temperature difference; they define heat as a residual difference between change of internal energy and work done on the system, when that work does not account for the whole of the change of internal energy and the system is not adiabatically isolated.<ref name="Münster 23 24">Münster, A. (1970), pp. 23–24.</ref><ref name="Reif 122">Reif, F. (1965), p. 122.</ref><ref name="Haase 1971">Haase, R. (1971), pp. 24–25.</ref>
 
Sometimes the concept of internal energy is not made explicit in the statement.<ref>[[Brian Pippard|Pippard, A.B.]] (1957/1966), p. 14.</ref><ref>Reif, F. (1965), p. 82.</ref><ref>Adkins, C.J. (1968/1983), p. 31.</ref>
 
Sometimes the existence of the internal energy is made explicit but work is not explicitly mentioned in the statement of the first postulate of thermodynamics. Heat supplied is then defined as the residual change in internal energy after work has been taken into account, in a non-adiabatic process.<ref>[[Herbert Callen|Callen, H.B.]] (1960/1985), pp. 13, 17.</ref>
 
A respected modern author states the first law of thermodynamics as "Heat is a form of energy", which explicitly mentions neither internal energy nor adiabatic work. Heat is defined as energy transferred by thermal contact with a reservoir, which has a temperature, and is generally so large that addition and removal of heat do not alter its temperature.<ref name="Kittel and Kroemer 1980">Kittel, C. Kroemer, H. (1980). ''Thermal Physics'', (first edition by Kittel alone 1969), second edition, W.H. Freeman, San Francisco, ISBN 0-7167-1088-9, pp. 49, 227.</ref> A current student text on chemistry defines heat thus: "''heat'' is the exchange of thermal energy between a system and its surroundings caused by a temperature difference." The author then explains how heat is defined or measured by calorimetry, in terms of [[heat capacity]], specific heat capacity, molar heat capacity, and temperature.<ref>Tro, N.J. (2008). ''Chemistry. A Molecular Approach'', Pearson/Prentice Hall, Upper Saddle River  NJ, ISBN 0-13-100065-9, p. 246.</ref>
 
A respected text disregards the Carathéodory's exclusion of mention of heat from the statement of the first law for closed systems, and admits heat calorimetrically defined along with work and internal energy.<ref>[[John Gamble Kirkwood|Kirkwood, J.G.]], Oppenheim, I. (1961), pp. 17–18. Kirkwood & Oppenheim 1961 is recommended by Münster, A. (1970), p. 376. It is also cited by Eu, B.C. (2002), ''Generalized Thermodynamics, the Thermodynamics of Irreversible Processes and Generalized Hydrodynamics'', Kluwer Academic Publishers, Dordrecht, ISBN 1-4020-0788-4, pp. 18, 29, 66.</ref> Another respected text defines heat exchange as determined by temperature difference, but also mentions that the Born (1921) version is "completely rigorous".<ref>[[Edward A. Guggenheim|Guggenheim, E.A.]] (1949/1967). ''Thermodynamics. An Advanced Treatment for Chemists and Physicists'', (first edition 1949), fifth edition 1967, North-Holland, Amsterdam, pp. 9–10. Guggenheim 1949/1965 is recommended by Buchdahl, H.A. (1966), p. 218. It is also recommended by Münster, A. (1970), p. 376.</ref> These versions follow the traditional approach that is now considered out of date, exemplified by that of Planck (1897/1903).<ref name="Planck 1903">Planck, M.(1897/1903).</ref>
 
==Evidence for the first law of thermodynamics for closed systems==
 
The first law of thermodynamics for closed systems was originally induced from empirically observed evidence, including calorimetric evidence. It is nowadays, however, taken to provide the definition of heat via the law of conservation of energy and the definition of work in terms of changes in the external parameters of a system. The original discovery of the law was gradual over a period of perhaps half a century or more, and some early studies were in terms of cyclic processes.<ref name="Truesdell, C.A. 1980">Truesdell, C.A. (1980).</ref>
 
The following is an account in terms of changes of state of a closed system through compound processes that are not necessarily cyclic. This account first considers processes for which the first law is easily verified because of their simplicity, namely [[adiabatic process]]es (in which there is no transfer as heat) and [[Thermodynamics#System models|adynamic processes]] (in which there is no transfer as work).
 
===Adiabatic processes===
 
In an adiabatic process, there is transfer of energy as work but not as heat. For all adiabatic process that takes a system from a given initial state to a given final state, irrespective of how the work is done, the respective eventual total quantities of energy transferred as work are one and the same, determined just by the given initial and final states. The work done on the system is defined and measured by changes in mechanical or quasi-mechanical variables external to the system. Physically, adiabatic transfer of energy as work requires the existence of adiabatic enclosures.
 
For instance, in Joule's experiment, the initial system is a tank of water with a paddle wheel inside. If we isolate thermally the tank and move the paddle wheel with a pulley and a weight we can relate the increase in temperature with the height descended by the mass. Now the system is returned to its initial state, isolated again, and the same amount of work is done on the tank using different devices (an electric motor, a chemical battery, a spring,...). In every case, the amount of work can be measured independently. The return to the initial state is not conducted by doing adiabatic work on the system. The evidence shows that the final state of the water (in particular, its temperature and volume) is the same in every case. It is irrelevant if the work is [[electrical work|electrical]], mechanical, chemical,... or if done suddenly or slowly, as long as it is performed in an adiabatic way, that is to say, without heat transfer into or out of the system.
 
Evidence of this kind shows that to increase the temperature of the water in the tank, the qualitative kind of adiabatically performed work does not matter. No qualitative kind of adiabatic work has ever been observed to decrease the temperature of the water in the tank.
 
A change from one state to another, for example an increase of both temperature and volume, may be conducted in several stages, for example by externally supplied electrical work on a resistor in the body, and adiabatic expansion allowing the body to do work on the surroundings. It needs to be shown that the time order of the stages, and their relative magnitudes, does not affect the amount of adiabatic work that needs to be done for the change of state. According to one respected scholar: "Unfortunately, it does not seem that experiments of this kind have ever been carried out carefully. ... We must therefore admit that the statement which we have enunciated here, and which is equivalent to the first law of thermodynamics, is not well founded on direct experimental evidence."<ref name="Pippard 15">[[Brian Pippard|Pippard, A.B.]] (1957/1966), p. 15. According to [[Herbert Callen]], in his most widely cited text, Pippard's text gives a "scholarly and rigorous treatment"; see Callen, H.B. (1960/1985), p. 485. It is also recommended by Münster, A. (1970), p. 376.</ref>
 
This kind of evidence, of independence of sequence of stages, combined with the above-mentioned evidence, of independence of qualitative kind of work, would show the existence of a very important state variable that corresponds with adiabatic work, but not that such a state variable represented a conserved quantity. For the latter, another step of evidence is needed, which may be related to the concept of reversibility, as mentioned below.
 
That very important state variable was first recognized and denoted <math>U</math> by Clausius in 1850, but he did not then name it, and he defined it in terms not only of work but also of heat transfer in the same process. It was also independently recognized in 1850 by Rankine, who also denoted it <math>U</math> ; and in 1851 by Kelvin who then called it "mechanical energy", and later "intrinsic energy". In 1865, after some hestitation, Clausius began calling his state function <math>U</math> "energy". In 1882 it was named as the ''[[internal energy]]'' by Helmholtz.<ref>Cropper, W.H. (1986). Rudolf Clausius and the road to entropy, ''Am. J. Phys.'', '''54''': 1068–1074.</ref> If only adiabatic processes were of interest, and heat could be ignored, the concept of internal energy would hardly arise or be needed. The relevant physics would be largely covered by the concept of potential energy, as was intended in the 1847 paper of Helmholtz on the principle of conservation of energy, though that did not deal with forces that cannot be described by a potential, and thus did not fully justify the principle. Moreover, that paper was very critical of the early work of Joule that had by then been performed.<ref>Truesdell, C.A. (1980), pp. 161–162.</ref> A great merit of the internal energy concept is that it frees thermodynamics from a restriction to cyclic processes, and allows a treatment in terms of thermodynamic states.
 
In an adiabatic process, adiabatic work takes the system either from a reference state <math>O</math> with internal energy <math>U(O)</math> to an arbitrary one <math>A</math> with internal energy <math>U(A)</math>, or from the state <math>A</math> to the state <math>O</math>:
 
:<math>U(A)=U(O) - W^\mathrm{adiabatic}_{O\to A}\,\, \mathrm{or}\,\,U(O)=U(A) - W^\mathrm{adiabatic}_{A\to O}\,.</math>
 
Except under the special, and strictly speaking, fictional, condition of reversibility, only one of the processes &nbsp;&nbsp;<math>\mathrm{adiabatic},\,O\to A</math>&nbsp;&nbsp; or &nbsp;&nbsp;<math>\mathrm{adiabatic},\,{A\to O}\,</math> &nbsp;&nbsp;is empirically feasible by a simple application of externally supplied work. The reason for this is given as the second law of thermodynamics and is not considered in the present article.
 
The fact of such irreversibility may be dealt with in two main ways, according to different points of view.
since the work of Bryan (1907),
 
*To deal with it nowadays, the most accepted way, followed by Carathéodory,<ref name="Carathéodory 1909"/><ref name="Haase 1971"/><ref>Buchdahl, H.A. (1966), p. 43.</ref> is to rely on the previously established concept of quasi-static processes,<ref>[[James Clerk Maxwell|Maxwell, J. C.]] (1871). ''Theory of Heat'', Longmans, Green, and Co., London, p. 150.</ref><ref>Planck, M. (1897/1903), Section 71, p. 52.</ref><ref>Bailyn, M. (1994), p. 95.</ref> as follows. Actual physical processes of transfer of energy as work are always at least to some degree irreversible. The irreversibility is often due to mechanisms known as dissipative, that transform bulk kinetic energy into internal energy. Examples are friction and viscosity. If the process is performed more slowly, the frictional or viscous dissipation is less. In the limit of infinitely slow performance, the dissipation tends to zero and then the limiting process, though fictional rather than actual, is notionally reversible, and is called quasi-static. Throughout the course of the fictional limiting quasi-static process, the internal intensive variables of the system are equal to the external intensive variables, those that describe the reactive forces exerted by the surroundings.<ref>Adkins, C.J. (1968/1983), p. 35.</ref> This can be taken to justify the formula
:<math>W^\mathrm{adiabatic,\,quasi-static}_{A\to O} = -W^\mathrm{adiabatic,\, quasi-static}_{O\to A}\,.\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,(1)</math>
 
*Another way to deal with it is to allow that experiments with processes of heat transfer to or from the system may be used to justify the formula <math>(1)</math> just above. Moreover, it deals to some extent with the problem of lack of direct experimental evidence that the time order of stages of a process does not matter in the determination of internal energy. This way does not provide theoretical purity in terms of adiabatic work processes, but is empirically feasible, and is in accord with experiments actually done, such as the Joule experiments mentioned just above, and with older traditions.
 
The formula <math>(1)</math> above allows that to go by processes of quasi-static adiabatic work from the state <math>A</math> to the state <math>B</math> we can take a path that goes through the reference state <math>O</math>, since the quasi-static adiabatic work is independent of the path
 
:<math>-W^\mathrm{adiabatic,\,quasi-static}_{A\to B}=-W^\mathrm{adiabatic,\,quasi-static}_{A\to O}-W^\mathrm{adiabatic,\,quasi-static}_{O\to B} = W^\mathrm{adiabatic,\,quasi-static}_{O\to A}- W^\mathrm{adiabatic,\,quasi-static}_{O\to B} = -U(A) + U(B) = \Delta U</math>
 
This kind of empirical evidence, coupled with theory of this kind, largely justifies the following statement:
 
:''For all adiabatic processes between two specified states of a closed system of any nature, the net work done is the same regardless the details of the process, and determines a state function called internal energy, <math>U</math>."
 
===Adynamic processes===
{{see also|Thermodynamic processes}}
A complementary observable aspect of the first law is about [[heat transfer]]. Adynamic transfer of energy as heat can be measured empirically by changes in the surroundings of the system of interest by calorimetry. This again requires the existence of adiabatic enclosure of the entire process, system and surroundings, though the separating wall between the surroundings and the system is thermally conductive or radiatively permeable, not adiabatic. A calorimeter can rely on measurement of [[sensible heat]], which requires the existence of thermometers and measurement of temperature change in bodies of known sensible heat capacity under specified conditions; or it can rely on the measurement of [[latent heat]], through [[Calorimetry#Calorimetry through phase change|measurement of masses of material that change phase]], at temperatures fixed by the occurrence of phase changes under specified conditions in bodies of known latent heat of phase change. The calorimeter can be calibrated by adiabatically doing externally determined work on it. The most accurate method is by passing an electric current from outside through a resistance inside the calorimeter. The calibration allows comparison of calorimetric measurement of quantity of heat transferred with quantity of energy transferred as work. According to one textbook, "The most common device for measuring <math>\Delta U</math> is an '''adiabatic bomb calorimeter'''."<ref>[[Peter Atkins|Atkins, P.]], de Paula, J. (1978/2010). ''Physical Chemistry'', (first edition 1978), ninth edition 2010, Oxford University Press, Oxford UK, ISBN 978-0-19-954337-3, p. 54.</ref> According to another textbook, "Calorimetry is widely used in present day laboratories."<ref>Kondepudi, D. (2008). ''Introduction to Modern Thermodynamics'', Wiley, Chichester, ISBN 978-0-470-01598-8, p. 63.</ref> According to one opinion, "Most thermodynamic data come from calorimetry..."<ref>Gislason, E.A., Craig, N.C. (2005). Cementing the foundations of thermodynamics:comparison of system-based and surroundings-based definitions of work and heat, ''J. Chem. Thermodynamics'' '''37''': 954–966.</ref> According to another opinion, "The most common method of measuring “heat” is with a calorimeter."<ref>Rosenberg, R.M. (2010). From Joule to Caratheodory and Born: A conceptual evolution of the first law of thermodynamics, ''J. Chem. Edu.'', '''87''': 691–693.</ref>
 
When the system evolves with transfer of energy as heat, without energy being transferred as work, in an adynamic process, the heat transferred to the system is equal to the increase in its internal energy:
 
:<math>Q^\mathrm{adynamic}_{A\to B}=\Delta U\,.</math>
 
===General case for reversible processes===
 
Heat transfer is practically reversible when it is driven by practically negligibly small temperature gradients. Work transfer is practically reversible when it occurs so slowly that there are no frictional effects within the system; frictional effects outside the system should also be zero if the process is to be globally reversible. For a particular reversible process in general, the work done reversibly on the system, <math>W^{\mathrm{path}\,P_0,\, \mathrm{reversible}}_{A\to B}</math>, and the heat transferred reversibly to the system, <math>Q^{\mathrm{path}\,P_0,\, \mathrm{reversible}}_{A\to B}</math> are not required to occur respectively adiabatically or adynamically, but they must belong to the same particular process defined by its particular reversible path, <math>P_0</math>, through the space of thermodynamic states. Then the work and heat transfers can occur and be calculated simultaneously.
 
Putting the two complementary aspects together, the first law for a particular reversible process can be written
 
:<math>-W^{\mathrm{path}\,P_0,\, \mathrm{reversible}}_{A\to B} + Q^{\mathrm{path}\,P_0,\, \mathrm{reversible}}_{A\to B} = \Delta U\, .</math>
 
This combined statement is the expression the first law of thermodynamics for reversible processes for closed systems.
 
In particular, if no work is done on a thermally isolated closed system we have
 
:<math>\Delta U = 0\,</math>.
 
This is one aspect of the law of conservation of energy and can be stated:
 
:''The internal energy of an isolated system remains constant.''
 
===General case for irreversible processes===
 
If, in a process of change of state of a closed system, the energy transfer is not under a practically zero temperature gradient and practically frictionless, then the process is irreversible. Then the heat and work transfers may be difficult to calculate, and irreversible thermodynamics is called for. Nevertheless, the first law still holds and provides a check on the measurements and calculations of the work done irreversibly on the system, <math>W^{\mathrm{path}\,P_1,\, \mathrm{irreversible}}_{A\to B}</math>, and the heat transferred irreversibly to the system, <math>Q^{\mathrm{path}\,P_1,\, \mathrm{irreversible}}_{A\to B}</math>, which belong to the same particular process defined by its particular irreversible path, <math>P_1</math>, through the space of thermodynamic states.
 
:<math>-W^{\mathrm{path}\,P_1,\, \mathrm{irreversible}}_{A\to B} + Q^{\mathrm{path}\,P_1,\, \mathrm{irreversible}}_{A\to B} = \Delta U\, .</math>
 
This means that the internal energy <math>U</math> is a function of state and that the internal energy change <math>\Delta U</math> between two states is a function only of the two states.
 
===Overview of the weight of evidence for the law===
 
The first law of thermodynamics is very general and makes so many predictions that they can hardly all be directly tested by experiment. Nevertheless, very very many of its predictions have been found empirically accurate. And very importantly, no accurately and properly conducted experiment has ever detected a violation of the law. Consequently, within its scope of applicability, the law is so reliably established, that, nowadays, rather than experiment being considered as testing the accuracy of the law, it is far more practical and realistic to think of the law as testing the accuracy of experiment. An experimental result that seems to violate the law may be assumed to be inaccurate or wrongly conceived, for example due to failure to consider an important physical factor.
 
==State functional formulation for infinitesimal processes==
When the heat and work transfers in the equations above are infinitesimal in magnitude, they are often denoted by δ, rather than [[exact differential]]s denoted by "d", as a reminder that heat and work do not describe the ''state'' of any system. The integral of an inexact differential depends upon the particular path taken through the space of thermodynamic parameters while the integral of an exact differential depends only upon the initial and final states. If the initial and final states are the same, then the integral of an inexact differential may or may not be zero, but the integral of an exact differential is always zero. The path taken by a thermodynamic system through a chemical or physical change is known as a [[thermodynamic process]].
 
For a homogeneous system, with a well-defined temperature and pressure, the expression for d''U'' can be written in terms of exact differentials, if the work that the system does is equal to its pressure times the infinitesimal increase in its volume. Here one assumes that the changes are quasistatic, so slow that there is at each instant negligible departure from  thermodynamic equilibrium within the system. In other words, δ''W'' = ''-P''d''V'' where ''P'' is [[pressure]] and ''V'' is [[Volume (thermodynamics)|volume]]. As such a quasistatic process in a homogeneous system is reversible, the total amount of heat added to a closed system can be expressed as δ''Q'' =''T''d''S'' where ''T'' is the [[temperature]] and ''S'' the [[entropy]] of the system. Therefore, for closed, homogeneous systems:
 
:<math>dU=TdS-PdV.\,</math>
 
The above equation is known as the [[fundamental thermodynamic relation]], for which the independent variables are taken as ''S'' and ''V'', with respect to which ''T'' and ''P'' are partial derivatives of ''U''. While this has been derived for quasistatic changes, it is valid in general, as U can be considered as a thermodynamic state function of the independent variables S and V.
 
As an example, one may suppose that the system is initially in a state of thermodynamic equilibrium defined by ''S'' and ''V''. Then the system is suddenly perturbed so that thermodynamic equilibrium breaks down and no temperature and pressure can be defined. Eventually the system settles down again to a state of thermodynamic equilibrium, defined by an entropy and a volume that differ infinitesimally from the initial values. The infinitesimal difference in internal energy between the initial and final state satisfies the above equation. But the work done and heat added to the system do not satisfy the above expressions. Rather, they satisfy the inequalities:  δ''Q'' < ''T''d''S' and δ''W'' < ''P''d''V'.
 
In the case of a closed system in which the particles of the system are of different types and, because chemical reactions may occur, their respective numbers are not necessarily constant, the expression for d''U'' becomes:
 
:<math>dU=\delta Q-\delta W + \sum_i \mu_i dN_i\,</math>
 
where d''N''<sub>i</sub> is the (small) increase in amount of type-i particles in the reaction, and ''μ''<sub>i</sub> is known as the [[chemical potential]] of the type-i particles in the system. If d''N''<sub>i</sub> is expressed in [[Mole (unit)|mol]] then ''μ''<sub>i</sub> is expressed in J/mol. The statement of the first law, using exact differentials is now:
 
:<math>dU=TdS-PdV + \sum_i \mu_i dN_i.\,</math>
 
If the system has more external mechanical variables than just the volume that can change, the fundamental thermodynamic relation generalizes to:
 
:<math>dU = T dS - \sum_{i}X_{i}dx_{i} + \sum_{j}\mu_{j}dN_{j}.\,</math>
 
Here the ''X''<sub>i</sub> are the [[generalized forces]] corresponding to the external variables ''x''<sub>i</sub>. The parameters ''X''<sub>i</sub> are independent of the size of the system and are called intensive parameters and the ''x''<sub>i</sub> are proportional to the size and called extensive parameters.
 
For an open system, there can be transfers of particles as well as energy into or out of the system during a process. For this case, the first law of thermodynamics still holds, in the form that the internal energy is a function of state and the change of internal energy in a process is a function only of its initial and final states, as noted in the section below headed [[First law of thermodynamics#First law of thermodynamics for open systems|First law of thermodynamics for open systems]].
 
A useful idea from mechanics is that the energy gained by a particle is equal to the force applied to the particle multiplied by the displacement of the particle while that force is applied. Now consider the first law without the heating term: d''U'' = -''P''d''V''. The pressure ''P'' can be viewed as a force (and in fact has units of force per unit area) while d''V''is the displacement (with units of distance times area). We may say, with respect to this work term, that a pressure difference forces a transfer of volume, and that the product of the two (work) is the amount of energy transferred out of the system as a result of the process. If one were to make this term negative then this would be the work done on the system.
 
It is useful to view the ''T''d''S'' term in the same light: here the temperature is known as a "generalized" force (rather than an actual mechanical force) and the entropy is a generalized displacement.
 
Similarly, a difference in chemical potential between groups of particles in the system drives a chemical reaction that changes the numbers of particles, and the corresponding product is the amount of chemical potential energy transformed in process. For example, consider a system consisting of two phases: liquid water and water vapor. There is a generalized "force" of evaporation that drives water molecules out of the liquid. There is a generalized "force" of condensation that drives vapor molecules out of the vapor. Only when these two "forces" (or chemical potentials) are equal is there equilibrium, and the net rate of transfer zero.
 
The two thermodynamic parameters that form a generalized force-displacement pair are called "conjugate variables". The two most familiar pairs are, of course, pressure-volume, and temperature-entropy.
 
==Spatially inhomogeneous systems==
 
Classical thermodynamics is initially focused on closed homogeneous systems (e.g. Planck 1897/1903<ref name="Planck 1903"/>), which might be regarded as 'zero-dimensional' in the sense that they have no spatial variation. But it is desired to study also systems with distinct internal motion and spatial inhomogeneity. For such systems, the principle of conservation of energy is expressed in terms not only of internal energy as defined for homogeneous systems, but also in terms of kinetic energy and potential energies of parts of the inhomogeneous system with respect to each other and with respect to long-range external forces.<ref>Bailyn, M. (1994), 254-256.</ref> How the total energy of a system is allocated between these three more specific kinds of energy varies according to the purposes of different writers; this is because these components of energy are to some extent mathematical artefacts rather than actually measured physical quantities. For any closed homogeneous component of an inhomogeneous closed system, if <math>E</math> denotes the total energy of that component system, one may write
 
:<math>E = E^{\mathrm {kin}} + E^{\mathrm {pot}} + U</math>
 
where <math>E^{\mathrm {kin}}</math> and <math>E^{\mathrm {pot}}</math> denote respectively the total kinetic energy and the total potential energy of the component closed homogeneous system, and <math>U</math> denotes its internal energy.<ref name ="Crawford 106"/><ref>Glansdorff, P., [[Ilya Prigogine|Prigogine, I.]] (1971), page 8.</ref>
 
Potential energy can be exchanged with the surroundings of the system when the surroundings impose a force field, such as gravitational or electromagnetic, on the system.
 
A compound system consisting of two interacting closed homogeneous component subsystems has a potential energy of interaction <math>E^{\mathrm {pot}}_{12}</math> between the subsystems. Thus, in an obvious notation, one may write
 
:<math>E = E^{\mathrm {kin}}_1 + E^{\mathrm {pot}}_1 + U_1 + E^{\mathrm {kin}}_2 + E^{\mathrm {pot}}_2 + U_2 + E^{\mathrm {pot}}_{12}</math>
 
The quantity <math>E^{\mathrm {pot}}_{12}</math> in general lacks an assignment to either subsystem in a way that is not arbitrary, and this stands in the way of a general non-arbitrary definition of transfer of energy as work. On occasions, authors make their various respective arbitrary assignments.<ref>Tisza, L. (1966), p. 91.</ref>
 
The distinction between internal and kinetic energy is hard to make in the presence of turbulent motion within the system, as friction gradually dissipates macroscopic kinetic energy of localised bulk flow into molecular random motion of molecules that is classified as internal energy.<ref>Denbigh, K.G. (1951), p. 50.</ref>  The rate of dissipation by friction of kinetic energy of localised bulk flow into internal energy,<ref name="Kelvin 1852a">Thomson, William (1852 a). "[http://zapatopi.net/kelvin/papers/on_a_universal_tendency.html On a Universal Tendency in Nature to the Dissipation of Mechanical Energy]" Proceedings of the Royal Society of Edinburgh for April 19, 1852 [This version from Mathematical and Physical Papers, vol. i, art. 59, pp. 511.]</ref><ref name="Kelvin 1852b">Thomson, W. (1852 b). On a universal tendency in nature to the dissipation of mechanical energy, ''Philosophical Magazine'' 4: 304-306.</ref><ref>Helmholtz, H. (1869/1871). Zur Theorie der stationären Ströme in reibenden Flüssigkeiten, ''Verhandlungen des naturhistorisch-medizinischen Vereins zu Heidelberg'', Band '''V''': 1-7. Reprinted in Helmholtz, H. (1882), ''Wissenschaftliche Abhandlungen'', volume 1, Johann Ambrosius Barth, Leipzig, pages 223-230 [http://echo.mpiwg-berlin.mpg.de/ECHOdocuViewfull?url=/mpiwg/online/permanent/einstein_exhibition/sources/QWH2FNX8/index.meta&start=231&viewMode=images&pn=237&mode=texttool]</ref> whether in turbulent or in streamlined flow, is an important quantity in [[non-equilibrium thermodynamics]]. This is a serious difficulty for attempts to define entropy for time-varying spatially inhomogeneous systems.
 
==First law of thermodynamics for open systems==
 
For the first law of thermodynamics, there is no trivial passage of physical conception from the closed system view to an open system view.<ref name="Münster 51">Münster A. (1970), Sections 14, 15, pp. 45–51.</ref><ref>Landsberg, P.T. (1978), p. 78.</ref> For closed systems, the concepts of an adiabatic enclosure and of an adiabatic wall are fundamental. Matter and internal energy cannot permeate or penetrate such a wall. For an open system, there is a wall that allows penetration by matter. In general, matter in diffusive motion carries with it some internal energy, and some microscopic potential energy changes accompany the motion. An open system is not adiabatically enclosed.
 
There are some cases in which a process for an open system can, for particular purposes, be considered as if it were for a closed system. In an open system, by definition hypothetically or potentially, matter can pass between the system and its surroundings. But when, in a particular case, the process of interest involves only hypothetical or potential but no actual passage of matter, the process can be considered as if it were for a closed system.
 
===Internal energy for an open system===
 
Since the revised and more rigorous definition of the internal energy of a closed system rests upon the possibility of processes by which adiabatic work takes the system from one state to another, this leaves a problem for the definition of internal energy for an open system, for which adiabatic work is not in general possible. According to [[Max Born]], the transfer of matter and energy across an open connection "cannot be reduced to mechanics".<ref>[[Max Born|Born, M.]] (1949), p. 44.</ref> In contrast to the case of closed systems, for open systems, in the presence of diffusion, there is no unconstrained and unconditional physical distinction between convective transfer of internal energy by bulk flow of matter, the transfer of internal energy without transfer of matter (usually called heat conduction and work transfer), and change of various potential energies.<ref>Denbigh, K.G. (1951), p. 56. Denbigh states in a footnote that he is indebted to correspondence with [[Edward A. Guggenheim|Professor E.A. Guggenheim]] and with Professor N.K. Adam. From this, Denbigh concludes "It seems, however, that when a system is able to exchange both heat and matter with its environment, it is impossible to make an unambiguous distinction between energy transported as heat and by the migration of matter, without already assuming the existence of the 'heat of transport'."</ref><ref>Fitts, D.D. (1962), p. 28.</ref><ref>Denbigh, K. (1954/1971), pp. 81–82.</ref> The older traditional way and the conceptually revised (Carathéodory) way agree that there is no physically unique definition of heat and work transfer processes between open systems.<ref>Münster, A. (1970), p. 50.</ref><ref>Haase, R. (1963/1969), p. 15.</ref><ref>Haase, R. (1971), p. 20.</ref><ref name="Smith 1980">Smith, D.A. (1980). Definition of heat in open systems, [http://www.publish.csiro.au/paper/PH800095.htm ''Aust. J. Phys.'', '''33''': 95–105.]</ref><ref name="Bailyn, M. 1994, p. 308">Bailyn, M. (1994), p. 308.</ref>
 
In particular, between two otherwise isolated open systems an adiabatic wall is by definition impossible.<ref>Münster, A. (1970), p. 46.</ref> This problem is solved by recourse to the principle of [[conservation of energy]]. This principle allows a composite isolated system to be derived from two other component non-interacting isolated systems, in such a way that the total energy of the composite isolated system is equal to the sum of the total energies of the two component isolated systems. Two previously isolated systems can be subjected to the [[thermodynamic operation]] of placement between them of a wall permeable to matter and energy, followed by a time for establishment of a new thermodynamic state of internal equilibrium in the new single unpartitioned system.<ref>Tisza, L. (1966), p. 41.</ref> The internal energies of the initial two systems and of the final new system, considered respectively as closed systems as above, can be measured.<ref name="Münster 51"/> Then the law of conservation of energy requires that
:<math>\Delta U_s+\Delta U_o=0\, ,</math><ref name="Callen 54">Callen H.B. (1960/1985), p. 54.</ref><ref name="Tisza 110">Tisza, L. (1966), p. 110.</ref>
 
where {{math|Δ''U''<sub>s</sub>}} and {{math|Δ''U''<sub>o</sub>}} denote the changes in internal energy of the system and of its surroundings respectively. This is a statement of the first law of thermodynamics for a transfer between two otherwise isolated open systems,<ref>Tisza, L. (1966), p. 111.</ref> that fits well with the conceptually revised and rigorous statement of the law stated above.
 
For the thermodynamic operation of adding two systems with internal energies {{math|''U''<sub>1</sub>}} and {{math|''U''<sub>2</sub>}}, to produce a new system with internal energy {{math|''U''}}, one may write {{math|''U'' {{=}} ''U''<sub>1</sub> + ''U''<sub>2</sub>}}; the reference states for  {{math|''U''}}, {{math|''U''<sub>1</sub>}} and {{math|''U''<sub>2</sub>}} should be specified accordingly, maintaining also that the internal energy of a system be proportional to its mass, so that the internal energies are [[Intensive and extensive properties|extensive variables]].<ref name="Münster 51"/><ref>[[Ilya Prigogine|Prigogine, I.]], (1955/1967), p. 12.</ref>
 
There is a sense in which this kind of additivity expresses a fundamental postulate that goes beyond the simplest ideas of classical closed system thermodynamics; the extensivity of some variables is not obvious, and needs explicit expression; indeed one author goes so far as to say that it could be recognized as a fourth law of thermodynamics, though this is not repeated by other authors.<ref>Landsberg, P.T. (1961), pp. 142, 387.</ref><ref>Landsberg, P.T. (1978), pp. 79,102.</ref>
 
Also of course
:<math>\Delta N_s+\Delta N_o=0\, ,</math><ref name="Callen 54"/><ref name="Tisza 110"/>
 
where {{math|Δ''N''<sub>s</sub>}} and {{math|Δ''N''<sub>o</sub>}} denote the changes in mole number of a component substance of the system and of its surroundings respectively. This is a statement of the law of [[conservation of mass]].
 
===Process of transfer of matter between an open system and its surroundings===
 
A system connected to its surroundings only through contact by a single permeable wall, but otherwise isolated, is an open system. If it is initially in a state of contact equilibrium with a surrounding subsystem, a [[thermodynamic process]] of transfer of matter can be made to occur between them if the surrounding subsystem is subjected to some thermodynamic operation, for example, removal of a partition between it and some further surrounding subsystem. The removal of the partition in the surroundings initiates a process of exchange between the system and its contiguous surrounding subsystem.
 
An example is evaporation. One may consider an open system consisting of a collection of liquid, enclosed except where it is allowed to evaporate into or to receive condensate from its vapor above it, which may be considered as its contiguous surrounding subsystem, and subject to control of its volume and temperature.
 
A thermodynamic process might be initiated by a thermodynamic operation in the surroundings, that mechanically increases in the controlled volume of the vapor. Some mechanical work will be done within the surroundings by the vapor, but also some of the parent liquid will evaporate and enter the vapor collection which is the contiguous surrounding subsystem. Some internal energy will accompany the vapor that leaves the system, but it will not make sense to try to uniquely identify part of that internal energy as heat and part of it as work. Consequently, the energy transfer that accompanies the transfer of matter between the system and its surrounding subsystem cannot be uniquely split into heat and work transfers to or from the open system. The component of total energy transfer that accompanies the transfer of vapor into the surrounding subsystem is customarily called 'latent heat of evaporation', but this use of the word heat is a quirk of customary historical language, not in strict compliance with the thermodynamic definition of transfer of energy as heat. In this example, kinetic energy of bulk flow and potential energy with respect to long-range external forces such as gravity are both considered to be zero. The first law of thermodynamics refers to the change of internal energy of the open system, between its initial and final states of internal equilibrium.
 
===Open system with multiple contacts===
 
An open system can be in contact equilibrium with several other systems at once.<ref name="Carathéodory 1909"/><ref>Prigogine, I. (1947), p. 48.</ref><ref>Born, M. (1949), Appendix 8, pp. 146–149.</ref><ref>Aston, J.G., Fritz, J.J. (1959), Chapter 9.</ref><ref>Kestin, J. (1961).</ref><ref>Landsberg, P.T. (1961), pp. 128–142.</ref><ref>Tisza, L. (1966), p. 108.</ref><ref>Tschoegl, N.W. (2000), p. 201.</ref>
 
This includes cases in which there is contact equilibrium between the system, and several subsystems in its surroundings, including separate connections with subsystems through walls that are permeable to the transfer of matter and internal energy as heat and allowing friction of passage of the transferred matter, but immovable, and separate connections through adiabatic walls with others, and separate connections through diathermic walls impermeable to matter with yet others. Because there are physically separate connections that are permeable to energy but impermeable to matter, between the system and its surroundings, energy transfers between them can occur with definite heat and work characters. Conceptually essential here is that the internal energy transferred with the transfer of matter is measured by a variable that is mathematically independent of the variables that measure heat and work.<ref>[[Max Born|Born, M.]] (1949), pp. 146–147.</ref>
 
With such independence of variables, the total increase of internal energy in the process is then determined as the sum of the internal energy transferred from the surroundings with the transfer of matter through the walls that are permeable to it, and of the internal energy transferred to the system as heat through the diathermic walls, and of the energy transferred to the system as work through the adiabatic walls, including the energy transferred to the system by long-range forces. These simultaneously transferred quantities of energy are defined by events in the surroundings of the system. Because the internal energy transferred with matter is not in general uniquely resolvable into heat and work components, the total energy transfer cannot in general be uniquely resolved into heat and work components.<ref>Haase, R. (1971), p. 35.</ref> Under these conditions, the following formula can describe the process in terms of externally defined thermodynamic variables, as a statement of the first law of thermodynamics:
:<math>\Delta U_0 \,=\,Q\, -\, W\, -\, \sum_{i=1}^m \Delta U_i \, \,\,\,\, \text {(suitably defined surrounding subsystems, general process, quasi-static or irreversible),}\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, (1)</math>
 
where {{math|Δ''U''<sub>0</sub>}} denotes the change of internal energy of the system, and {{math|Δ''U<sub>i''</sub>}} denotes the change of internal energy of the {{math|''i''th}} of the {{math|''m''}} surrounding subsystems that are in open contact with the system, due to transfer between the system and that {{math|''i''th}} surrounding subsystem, and {{math|''Q''}} denotes the internal energy transferred as heat from the heat reservoir of the surroundings to the system, and {{math|''W''}} denotes the energy transferred from the system to the surrounding subsystems that are in adiabatic connection with it. The case of  a wall that is permeable to matter and can move so as to allow transfer of energy as work is not considered here.
 
====Combination of first and second laws====
 
If the system is described by the energetic fundamental equation, {{math|''U''<sub>0</sub> {{=}}''U''<sub>0</sub>(''S'', ''V'', ''N<sub>j''</sub>)}}, and if the process can be described in the quasi-static formalism, in terms of the internal state variables of the system, then the process can also be described by a combination of the first and second laws of thermodynamics, by the formula
:<math>\mathrm d U_0 \,=\, T\, \mathrm d S\, -\, P\, \mathrm d V\, +\, \sum_{j=1}^n \mu _j \, \mathrm d N_j \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, (2)</math>
 
where there are {{math|''n''}} chemical constituents of the system and permeably connected surrounding subsystems, and where {{math|''T''}}, {{math|''S''}}, {{math|''P''}}, {{math|''V''}}, {{math|''N<sub>j''</sub>}}, and {{math|''μ<sub>j''</sub>}}, are defined as above.<ref>[[Herbert Callen|Callen, H.B]], (1960/1985), p. 35.</ref>
 
For a general natural process, there is no simple termwise correspondence between equations (1) and (2), because they describe the process in different conceptual frames.
 
Nevertheless, for the special fictive case of quasi-static transfers, there is a simple correspondence.<ref name="Aston Fritz account">Aston, J.G., Fritz, J.J. (1959), Chapter 9. This is an unusually explicit account of some of the physical meaning of the Gibbs formalism.</ref> For this, it is supposed that the system has multiple areas of contact with its surroundings. There are pistons that allow adiabatic work, purely diathermal walls, and open connections with surrounding subsystems of completely controllable chemical potential (or equivalent controls for charged species). Then, for a suitable fictive quasi-static transfer, one can write
:<math>\delta Q \,=\, T\, \mathrm d S\,\text{  and    }\delta W \,=\, P\, \mathrm d V\,\, \,\,\,\, \text {(suitably defined surrounding subsystems, quasi-static transfers of energy)}\, .</math>
 
For fictive quasi-static transfers for which the chemical potentials in the connected surrounding subsystems are suitably controlled, these can be put into equation (2) to yield
:<math>\mathrm d U_0 \,=\, \delta Q\, -\, \delta W\, +\, \sum_{j=1}^n \mu _j \, \mathrm d N_j\, \,\,\,\,\,\, \text {(suitably defined surrounding subsystems, quasi-static transfers)}\,. \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, (3)</math>
 
The reference <ref name="Aston Fritz account"/> does not actually write equation (3), but what it does write is fully compatible with it.
 
There are several other accounts of this, in apparent mutual conflict.<ref name="Smith 1980"/><ref>Buchdahl, H.A. (1966), Section 66, pp. 121–125.</ref><ref>Callen, J.B. (1960/1985), Section 2-1, pp. 35–37.</ref>
 
===Non-equilibrium transfers===
 
The transfer of energy between an open system and a single contiguous subsystem of its surroundings is considered also in non-equilibrium thermodynamics. The problem of definition arises also in this case. It may be allowed that the wall between the system and the subsystem is not only permeable to matter and to internal energy, but also may be movable so as to allow work to be done when the two systems have different pressures. In this case, the transfer of energy as heat is not defined.
 
Methods for study of non-equilibrium processes mostly deal with spatially continuous flow systems. In this case, the open connection between system and surroundings is usually taken to fully surround the system, so that there are no separate connections impermeable to matter but permeable to heat. Except for the special case mentioned above when there is no actual transfer of matter, which can be treated as if for a closed system, in strictly defined thermodynamic terms, it follows that transfer of energy as heat is not defined. In this sense, there is no such thing as 'heat flow' for a continuous-flow open system. Properly, for closed systems, one speaks of transfer of internal energy as heat, but in general, for open systems, one can speak safely only of transfer of internal energy. A factor here is that there are often cross-effects between distinct transfers, for example that transfer of one substance may cause transfer of another even when the latter has zero chemical potential gradient.
 
Usually transfer between a system and its surroundings applies to transfer of a state variable, and obeys a balance law, that the amount lost by the donor system is equal to the amount gained by the receptor system. Heat is not a state variable. For his 1947 definition of "heat transfer" for discrete open systems, the author Prigogine carefully explains at some length that his definition of it does not obey a balance law. He describes this as paradoxical.<ref>Prigogine, I., (1947), pp. 48–49.</ref>
 
The situation is clarified by Gyarmati, who shows that his definition of "heat transfer", for continuous-flow systems, really refers not specifically to heat, but rather to transfer of internal energy, as follows. He considers a conceptual small cell in a situation of continuous-flow as a system defined in the so-called Lagrangian way, moving with the local center of mass. The flow of matter across the boundary is zero when considered as a flow of total mass. Nevertheless if the material constitution is of several chemically distinct components that can diffuse with respect to one another, the system is considered to be open, the diffusive flows of the components being defined with respect to the center of mass of the system, and balancing one another as to mass transfer. Still there can be a distinction between bulk flow of internal energy and diffusive flow of internal energy in this case, because the internal energy density does not have to be constant per unit mass of material, and allowing for non-conservation of internal energy because of local conversion of kinetic energy of bulk flow to internal energy by viscosity.
 
Gyarmati shows that his definition of "the heat flow vector" is strictly speaking a definition of flow of internal energy, not specifically of heat, and so it turns out that his use here of the word heat is contrary to the strict thermodynamic definition of heat, though it is more or less compatible with historical custom, that often enough did not clearly distinguish between heat and internal energy; he writes "that this relation must be considered to be the exact definition of the concept of heat flow, fairly loosely used in experimental physics and heat technics."<ref>Gyarmati, I. (1970), p. 68.</ref> Apparently in a different frame of thinking from that of the above-mentioned paradoxical usage in the earlier sections of the historic 1947 work by Prigogine, about discrete systems, this usage of Gyarmati is consistent with the later sections of the same 1947 work by Prigogine, about continuous-flow systems, which use the term "heat flux" in just this way. This usage is also followed by Glansdorff and Prigogine in their 1971 text about continuous-flow systems. They write: "Again the flow of internal energy may be split into a convection flow {{math|''ρu'''''v'''}} and a conduction flow. This conduction flow is by definition the heat flow {{math|'''W'''}}. Therefore: {{math|'''j'''[''U''] {{=}} ''ρu'''''v''' + '''W'''}} where {{math|''u''}} denotes the [internal] energy per unit mass. [These authors actually use the symbols {{math|''E''}} and {{math|''e''}} to denote internal energy but their notation has been changed here to accord with the notation of the present article. These authors actually use the symbol {{math|''U''}} to refer to total energy, including kinetic energy of bulk flow.]"<ref>Glansdorff, P, Prigogine, I, (1971), p. 9.</ref> This usage is followed also by other writers on non-equilibrium thermodynamics such as Lebon, Jou, and Casas-Vásquez,<ref>Lebon, G., Jou, D., Casas-Vázquez, J. (2008), p. 45.</ref> and de Groot and Mazur.<ref>de Groot, S.R., Mazur, P. (1962), p. 18.</ref> This usage is described by Bailyn as stating the non-convective flow of internal energy, and is listed as his definition number 1, according to the first law of thermodynamics.<ref name="Bailyn, M. 1994, p. 308"/> This usage is also followed by workers in the kinetic theory of gases.<ref>de Groot, S.R., Mazur, P. (1962), p. 169.</ref><ref>Truesdell, C., Muncaster, R.G. (1980), p. 3.</ref><ref>Balescu, R. (1997), p. 9.</ref> This is not the ''ad hoc'' definition of "reduced heat flux" of Haase.<ref>Haase, R. (1963/1969), p. 18.</ref>
 
In the case of a flowing system of only one chemical constituent, in the Lagrangian representation, there is no distinction between bulk flow and diffusion of matter. Moreover, the flow of matter is zero into or out of the cell that moves with the local center of mass. In effect, in this description, one is dealing with a system effectively closed to the transfer of matter. But still one can validly talk of a distinction between bulk flow and diffusive flow of internal energy, the latter driven by a temperature gradient within the flowing material, and being defined with respect to the local center of mass of the bulk flow. In this case of a virtually closed system, because of the zero matter transfer, as noted above, one can safely distinguish between transfer of energy as work, and transfer of internal energy as heat.<ref>Eckart, C. (1940).</ref>
 
== See also ==
* [[Laws of thermodynamics]]
* [[Entropy production]]
* [[Perpetual motion]]
* [[Relativistic heat conduction]]
 
==References==
{{reflist|colwidth=35em}}
 
===Cited sources===
*Adkins, C.J. (1968/1983). ''Equilibrium Thermodynamics'', (first edition 1968), third edition 1983, Cambridge University Press, ISBN 0-521-25445-0.
*Aston, J.G., Fritz, J.J. (1959). ''Thermodynamics and Statistical Thermodynamics'', John Wiley & Sons, New York.
*Bailyn, M. (1994). ''A Survey of Thermodynamics'', American Institute of Physics Press, New York, ISBN 0-88318-797-3.
*[[Max Born|Born, M.]] (1949). ''Natural Philosophy of Cause and Chance'', Oxford University Press, London. 
*[[George H. Bryan|Bryan, G.H.]] (1907). [https://ia700208.us.archive.org/6/items/Thermodynamics/Thermodynamics.tif ''Thermodynamics. An Introductory Treatise dealing mainly with First Principles and their Direct Applications'', B.G. Teubner, Leipzig].
*[[Radu Bălescu|Balescu, R.]] (1997). ''Statistical Dynamics; Matter out of Equilibrium'', Imperial College Press, London, ISBN 978-1-86094-045-3.
*Buchdahl, H.A. (1966), ''The Concepts of Classical Thermodynamics'', Cambridge University Press, London.
*[[Herbert Callen|Callen, H.B.]] (1960/1985), ''Thermodynamics and an Introduction to Thermostatistics'', (first edition 1960), second edition 1985, John Wiley & Sons, New York, ISBN 0–471–86256–8.
*[[Constantin Carathéodory|Carathéodory, C.]] (1909). Untersuchungen über die Grundlagen der Thermodynamik, ''Mathematische Annalen'', '''67''': 355–386, {{doi|10.1007/BF01450409}}. A translation may be found [http://neo-classical-physics.info/uploads/3/0/6/5/3065888/caratheodory_-_thermodynamics.pdf here].  Also a mostly reliable [http://books.google.com.au/books?id=xwBRAAAAMAAJ&q=Investigation+into+the+foundations translation is to be found] at Kestin, J. (1976). ''The Second Law of Thermodynamics'', Dowden, Hutchinson & Ross, Stroudsburg PA.
*Crawford, F.H. (1963). ''Heat, Thermodynamics, and Statistical Physics'', Rupert Hart-Davis, London, Harcourt, Brace & World, Inc..
*de Groot, S.R., Mazur, P. (1962). ''Non-equilibrium Thermodynamics'', North-Holland, Amsterdam. Reprinted (1984), Dover Publications Inc., New York, ISBN 0486647412.
*Denbigh, K.G. (1951). [http://books.google.com.au/books/about/The_thermodynamics_of_the_steady_state.html?id=uoJGAAAAYAAJ&redir_esc=y ''The Thermodynamics of the Steady State''], Methuen, London, Wiley, New York.
*Denbigh, K. (1954/1971). ''The Principles of Chemical Equilibrium. With Applications in Chemistry and Chemical Engineering'', third edition, Cambridge University Press, Cambridge UK.
*Eckart, C. (1940). The thermodynamics of irreversible processes. {{math|I.}} The simple fluid, ''Phys. Rev.'' '''58''': 267–269.
*Fitts, D.D. (1962). ''Nonequilibrium Thermodynamics. Phenomenological Theory of Irreversible Processes in Fluid Systems'', McGraw-Hill, New York.
*Glansdorff, P., [[Ilya Prigogine|Prigogine, I.]], (1971). ''Thermodynamic Theory of Structure, Stability and Fluctuations'', Wiley, London, ISBN 0-471-30280-5.
*Gyarmati, I. (1967/1970). ''Non-equilibrium Thermodynamics. Field Theory and Variational Principles'', translated from the 1967 Hungarian by E. Gyarmati and W.F. Heinz, Springer-Verlag, New York.
*Haase, R. (1963/1969). ''Thermodynamics of Irreversible Processes'', English translation, Addison-Wesley Publishing, Reading MA.
*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.
*Helmholtz, H. (1847). ''Ueber die Erhaltung der Kraft. Eine physikalische Abhandlung'', G. Reimer (publisher), Berlin, read on 23 July in a session of the Physikalischen Gesellschaft zu Berlin. Reprinted in Helmholtz, H. von (1882), [http://archive.org/details/wissenschaftlic00helmgoog ''Wissenschaftliche Abhandlungen''], Band 1, J.A. Barth, Leipzig. Translated and edited by J. Tyndall, in ''Scientific Memoirs, Selected from the Transactions of Foreign Academies of Science and from Foreign Journals. Natural Philosophy'' (1853), volume 7, edited by J. Tyndall, W. Francis, published by Taylor and Francis, London, pp. 114–162, reprinted as volume 7 of Series 7, ''The Sources of Science'', edited by H. Woolf, (1966), Johnson Reprint Corporation, New York, and again in Brush, S.G., ''The Kinetic Theory of Gases. An Anthology of Classic Papers with Historical Commentary'', volume 1 of ''History of Modern Physical Sciences'', edited by N.S. Hall, Imperial College Press, London, ISBN 1-86094-347-0, pp. 89–110.
*Kestin, J. (1961). On intersecting isentropics, ''Am. J. Phys.'', '''29''': 329–331.
*[[John Gamble Kirkwood|Kirkwood, J.G.]], Oppenheim, I. (1961). ''Chemical Thermodynamics'', McGraw-Hill Book Company, New York.
*Landsberg, P.T. (1961). ''Thermodynamics with Quantum Statistical Illustrations'', Interscience, New York.
*Landsberg, P.T. (1978). ''Thermodynamics and Statistical Mechanics'', Oxford University Press, Oxford  UK, ISBN 0-19-851142-6.
*Lebon, G., Jou, D., Casas-Vázquez, J. (2008). ''Understanding Non-equilibrium Thermodynamics'', Springer, Berlin, ISBN 978-3-540-74251-7.
*Münster, A. (1970), ''Classical Thermodynamics'', translated by E.S. Halberstadt, Wiley–Interscience, London, ISBN 0-471-62430-6.
*[[Brian Pippard|Pippard, A.B.]] (1957/1966). ''Elements of Classical Thermodynamics for Advanced Students of Physics'', original publication 1957, reprint 1966, Cambridge University Press, Cambridge UK.
*[[Max Planck|Planck, M.]](1897/1903). [https://ia700200.us.archive.org/15/items/treatiseonthermo00planrich/treatiseonthermo00planrich.pdf ''Treatise on Thermodynamics'', translated by A. Ogg, Longmans, Green & Co., London.]
*[[Ilya Prigogine|Prigogine, I.]] (1947). ''Étude Thermodynamique des Phénomènes irréversibles'', Dunod, Paris, and Desoers, Liège.
*[[Ilya Prigogine|Prigogine, I.]], (1955/1967). ''Introduction to Thermodynamics of Irreversible Processes'', third edition, Interscience Publishers, New York.
*Reif, F. (1965). ''Fundamentals of Statistical and Thermal Physics'', McGraw-Hill Book Company, New York.
*[[László Tisza|Tisza, L.]] (1966). ''Generalized Thermodynamics'', M.I.T. Press, Cambridge MA.
*[[Clifford Truesdell|Truesdell, C.A.]] (1980). ''The Tragicomical History of Thermodynamics, 1822-1854'', Springer, New York, ISBN 0-387-90403-4.
*[[Clifford Truesdell|Truesdell, C.A.]], Muncaster, R.G. (1980). ''Fundamentals of Maxwell's Kinetic Theory of a Simple Monatomic Gas, Treated as a branch of Rational Mechanics'', Academic Press, New York, ISBN 0-12-701350-4.
*Tschoegl, N.W. (2000). ''Fundamentals of Equilibrium and Steady-State Thermodynamics'', Elsevier, Amsterdam, ISBN 0-444-50426-5.
 
==Further reading==
* {{cite book | author=Goldstein, Martin, and Inge F. | title=The Refrigerator and the Universe | publisher=Harvard University Press | year=1993 | isbn=0-674-75325-9 | oclc=32826343}} Chpts. 2 and 3 contain a nontechnical treatment of the first law.
 
* {{cite book| author=Çengel Y.A. and Boles M.| title=Thermodynamics: an engineering approach|publisher=McGraw-Hill Higher Education|year=2007|isbn=0-07-125771-3}} Chapter 2.
 
* {{cite book| author=Atkins P.| title=Four Laws that drive the Universe|publisher=OUP Oxford |year=2007|isbn=0-19-923236-9}}
 
==External links==
* [http://35.9.69.219/home/modules/pdf_modules/m158.pdf MISN-0-158, ''The First Law of Thermodynamics''] ([[Portable Document Format|PDF file]]) by Jerzy Borysowicz for [http://www.physnet.org Project PHYSNET].
* [http://web.mit.edu/16.unified/www/FALL/thermodynamics/notes/node8.html ''First law of thermodynamics''] in the MIT Course [http://web.mit.edu/16.unified/www/FALL/thermodynamics/notes/notes.html ''Unified Thermodynamics and Propulsion''] from Prof. Z. S. Spakovszky
 
{{DEFAULTSORT:First Law Of Thermodynamics}}
[[Category:Concepts in physics]]
[[Category:Laws of thermodynamics|1]]
 
[[de:Thermodynamik#Erster Hauptsatz]]

Latest revision as of 12:42, 10 January 2015

Msvcr71.dll is an significant file which helps help Windows procedure different components of the system including important files. Specifically, the file is selected to aid run corresponding files inside the "Virtual C Runtime Library". These files are significant inside accessing any settings that help the different applications and programs in the system. The msvcr71.dll file fulfills various significant functions; though it's not spared from getting damaged or corrupted. Once the file gets corrupted or damaged, the computer may have a hard time processing plus reading components of the system. However, consumers require not panic considering this problem could be solved by following several procedures. And I usually show we several strategies about Msvcr71.dll.

You are able to reformat the computer to make it run faster. This will reset a computer to whenever you first used it. Always remember to back up all files plus programs before carrying this out since this usually remove a files from the database. Remember before we do this you need all drivers plus installation files and this should be a last resort in the event you are hunting for slow computer tips.

With RegCure to improve the begin up plus shut down of your computer. The program shows the scan progress and you shouldn't worry where it is actually functioning at which time. It shows you exactly what occurs. Dynamic link library section of the registry will cause severe application failures. RegCure restores plus repairs the registry and keeps you from DLL. RegCure can make individual corrections, so it may functions for your requires.

Always see to it which you have installed antivirus, anti-spyware plus anti-adware programs plus have them updated regularly. This can help stop windows XP running slow.

Many registry reviver s let you to download their product for free, to scan the computer yourself. That way we can see how countless mistakes it finds, where it finds them, and how it can fix them. A ideal registry cleaner will remove the registry difficulties, and optimize plus accelerate your PC, with little effort on the piece.

Reinstall Windows 7 - If nothing appears to work, reinstall Windows 7 with the installation disc which came with all the pack. Kindly backup or restore all your data to a flash drive or another hard drive/CD etc. before operating the reinstallation.

Maybe you are asking why these windows XP error messages appear. Well, for we to be able to know the fix, you need to initially recognize where those errors come from. There is this software called registry. A registry is software that shops everything on your PC from a usual configuration, setting, info, plus logs of escapades from installing to UN-installing, saving to deleting, along with a lot more alterations we do in a system pass from it and gets 'tagged' plus saved as a simple file for recovery reasons. Imagine it because a big recorder, a registrar, of all a records in your PC.

So, the number one thing to do when your computer runs slow is to buy an authentic plus legal registry repair tool that would help you eliminate all problems associated to registry plus aid we enjoy a smooth running computer.