Non ideal compressible fluid dynamics: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
m Fixed a typo found with Wikipedia:Typo_Team/moss and deleted jargon abbreviations.
Expansion of the article for the course Wikipedia:GLAM/PoliMi/2023
Line 1: Line 1:
[[File:Non-ideal nozzle 01.png|thumb|324x324px|Non-[[Monotonic function|monotone]] evolution of the [[Mach number]] ''M'' in the divergent section of a [[De Laval nozzle|supersonic nozzle]]. The fluid is [[siloxane]] MM ([[hexamethyldisiloxane]], <chem>C6H18OSi2
{{Orphan|date=August 2016}}
</chem>) evolving in the non-ideal gasdynamic regime.]]
{{unreferenced|date=November 2015}}
'''Non ideal compressible fluid dynamics''' ('''NICFD'''), or non ideal [[Compressible flow|gasdynamics]], is a branch of [[fluid mechanics]] studying the dynamic behavior of [[Fluid|fluids]] not obeying [[Ideal gas|ideal-gas]] [[thermodynamics]]. It is for example the case of dense [[Vapor|vapors]], [[Supercritical fluid|supercritical flows]] and compressible [[Two-phase flow|two-phase flows]]. With the term dense vapors, we indicate all fluids in the [[Gas|gaseous state]] characterized by thermodynamic conditions close to [[Vapor–liquid equilibrium|saturation]] and the [[Critical point (thermodynamics)|critical point]]<ref>{{Cite book |last=Callen |first=Herbert B. |title=Thermodynamics and an introduction to thermostatistics |date=1985 |publisher=J. Wiley & Sons |isbn=978-0-471-86256-7 |edition=2nd |location=New York |pages=255-261}}</ref>. Supercritical fluids feature instead values of [[pressure]] and [[temperature]] larger than their critical values<ref>{{Cite journal |last=Schlosky |first=Kevin M. |date=1989 |title=Supercritical phase transitions at very high pressure |url=https://pubs.acs.org/doi/abs/10.1021/ed066p989 |journal=Journal of Chemical Education |language=en |volume=66 |issue=12 |pages=989 |doi=10.1021/ed066p989 |issn=0021-9584}}</ref>, wheras two-phase flows are characterized by the simultaneous presence of both [[liquid]] and gas states.<ref>{{Citation |last=Faghri |first=Amir |title=11 - TWO-PHASE FLOW AND HEAT TRANSFER |date=2006-01-01 |url=https://www.sciencedirect.com/science/article/pii/B9780123706102500167 |work=Transport Phenomena in Multiphase Systems |pages=853–949 |editor-last=Faghri |editor-first=Amir |access-date=2023-07-06 |place=Boston |publisher=Academic Press |language=en |doi=10.1016/b978-0-12-370610-2.50016-7 |isbn=978-0-12-370610-2 |last2=Zhang |first2=Yuwen |editor2-last=Zhang |editor2-first=Yuwen}}</ref>


In all these cases, the fluid requires to be modelled as a [[real gas]], since its thermodynamic behavior considerably differs from that of an ideal gas, which by contrast appears for dilute thermodynamic conditions. The ideal-gas law, in fact, can be employed in general as a reasonable approximation of the fluid thermodynamics for low pressures and high temperatures. Otherwise, [[Intermolecular force|intermolecular forces]] and dimension of fluid particles, which are neglected in the ideal-gas approximation, become relevant and can significantly affect the fluid behavior<ref name=":0" />. This is extremely valid for gases made of complex and heavy [[Molecule|molecules]], which tend to deviate more from the ideal model.<ref name=":3">{{Cite journal |last=Colonna |first=P. |last2=Guardone |first2=A. |date=2006 |title=Molecular interpretation of nonclassical gas dynamics of dense vapors under the van der Waals model |url=http://dx.doi.org/10.1063/1.2196095 |journal=Physics of Fluids |volume=18 |issue=5 |pages=056101 |doi=10.1063/1.2196095 |issn=1070-6631}}</ref>
'''Non ideal compressible fluid dynamics''' is a branch of [[fluid mechanics]] studying the actual characteristics of dense vapors, supercritical flows and compressible two-phase flows, namely whereby the thermodynamic behavior of the fluid differs considerably from that of a [[perfect gas]]. At high [[reduced pressure]] and [[Reduced temperature|temperature]], close to the [[Saturation vapor curve|saturation curve]] the [[speed of sound]] is largely sensitive to [[density]] variations along [[Isentropic process|isentropes]]. Consequently, the fluid flow departs from the ideality assumption and under particular conditions may even exhibit non classical gas dynamic phenomena, whose nature is governed by the value of the fundamental derivative of gas-dynamics Γ. A non-monotonic [[Mach number]] trend along an expansion is typical for 0 < Γ < 1, while for Γ < 0 values admit the occurrence of inverse gas-dynamics phenomena such as [[rarefaction]] [[shock wave]]s , splitting waves or even composite waves. Inverse gas-dynamics behavior has been theoretically predicted for heavy complex molecules in the vapor region, and a recent study discovered that two-phase rarefaction shock waves are physically realizable close to the [[Critical point (thermodynamics)|critical point]].


While the fluid dynamics of compressible flows in ideal conditions is well-established and is characterized by several analytical results<ref name=":4" />, when non-ideal thermodynamic conditions are considered, peculiar phenomena possibly occur. This is particularly valid in [[Supersonic speed|supersonic]] conditions, namely for flow velocities larger than the [[speed of sound]] in the fluid considered. All typical features of supersonic flows, in fact, are affected by non-ideal thermodynamics, resulting in both quantitative and qualitative differences with respect to the ideal gasdynamics.<ref name=":1" />
==Applications==
== Non-ideal thermodynamics ==
Application of non-ideal compressible fluid dynamics flows can be found in numerous fields. They are encountered in [[Organic Rankine cycle|Organic Rankine Cycle]] turbogenerators, refrigerator, [[supercritical carbon dioxide]] power system, pharmaceutical processes, transportation of [[fuel]]s at high-speed, and in [[transonic]] and [[Supersonic speed|supersonic]] [[wind tunnel]]s.
{{Main|Real gas}}
[[File:Compressibility factor generalized diagram.png|thumb|336x336px|Compressibility factor ''Z'' for different values of reduced pressure and temperature.]]
For dilute thermodynamic conditions, the ideal-gas [[equation of state]] (EoS) provides sufficiently accurate results in modelling the fluid thermodynamics. This occurs in general for low values of reduced pressure and high values of reduced temperature, where the term [[Reduced properties|reduced]] refers to the ratio of a certain thermodynamic quantity and its critical value. For most applications including [[air]], for example, the assumption of considering ideal conditions is perfectly reasonable and it is widely used<ref name=":4">{{Cite book |last=Thompson |first=Philip A. |title=Compressible-fluid dynamics |date=1972 |publisher=McGraw-Hill |isbn=978-0-07-064405-2 |series=Advanced engineering series |location=New York |pages=76-99}}</ref>.


On the other hand, when thermodynamic conditions approach [[condensation]] and the critical point or when high pressures are involved, real-gas models are needed in order to capture the real fluid behavior. In these conditions, in fact, intermolecular forces and compressibility effects come into play<ref name=":0" />.
==References==

A measure of the fluid non-ideality is given by the [[compressibility factor]] <math> Z </math><ref>{{Cite book |last=Zucker |first=Robert D. |title=Fundamentals of gas dynamics |last2=Biblarz |first2=Oscar |date=2002 |publisher=Wiley |isbn=978-0-471-05967-7 |edition=2nd |location=Hoboken, NJ |pages=327}}</ref>, defined as

:<math> Z = \frac{Pv}{RT} </math>

where
* <math> P </math> is the pressure [Pa];
* <math> v </math> is the [[specific volume]] [m<sup>3</sup>/kg];
* <math> R </math> is the [[gas constant]] [J/kgK];
* <math> T </math> is the absolute temperature [K].

The compressibility factor is a [[dimensionless quantity]] which is equal to 1 for ideal gases and deviates from unity for increasing levels of non-ideality.<ref>{{Cite book |last=Thompson |first=Philip A. |title=Compressible-fluid dynamics |date=1972 |publisher=McGraw-Hill |isbn=978-0-07-064405-2 |series=Advanced engineering series |location=New York |pages=99-101}}</ref>

Several non-ideal models exist, from the simplest [[cubic equations of state]] (such as the [[Van der Waals equation|Van der Waals]]<ref>{{Cite book |last=Parsegian |first=V. Adrian |url=https://www.cambridge.org/core/books/van-der-waals-forces/EFAB2ADFEF0B97F2AEB112AA6F3A7862 |title=Van der Waals Forces: A Handbook for Biologists, Chemists, Engineers, and Physicists |date=2005 |publisher=Cambridge University Press |isbn=978-0-521-83906-8 |location=Cambridge |doi=10.1017/cbo9780511614606}}</ref><ref name=":0">{{Cite book |last=Waals |first=J. D. van der |title=On the continuity of the gaseous and liquid states |last2=Rowlinson |first2=John Shipley |date=1988 |publisher=North-Holland |isbn=978-0-444-87077-3 |series=Studies in statistical mechanics |location=Amsterdam}}</ref> and the [[Peng-Robinson equation|Peng-Robinson]]<ref>{{Cite journal |last=Peng |first=Ding-Yu |last2=Robinson |first2=Donald B. |date=February 1976 |title=A New Two-Constant Equation of State |url=https://pubs.acs.org/doi/abs/10.1021/i160057a011 |journal=Industrial & Engineering Chemistry Fundamentals |language=en |volume=15 |issue=1 |pages=59–64 |doi=10.1021/i160057a011 |issn=0196-4313}}</ref> models) up to complex multi-parameter ones, including the Span-Wagner equation of state<ref>{{Cite journal |last=Span |first=R. |last2=Wagner |first2=W. |date=2003-01-01 |title=Equations of State for Technical Applications. I. Simultaneously Optimized Functional Forms for Nonpolar and Polar Fluids |url=https://doi.org/10.1023/A:1022390430888 |journal=International Journal of Thermophysics |language=en |volume=24 |issue=1 |pages=1–39 |doi=10.1023/A:1022390430888 |issn=1572-9567}}</ref><ref>{{Citation |last=Span |first=Roland |title=Describing Mixtures with Multiparameter Equations of State |date=2000 |url=http://dx.doi.org/10.1007/978-3-662-04092-8_8 |work=Multiparameter Equations of State |pages=319–340 |access-date=2023-07-05 |place=Berlin, Heidelberg |publisher=Springer Berlin Heidelberg |isbn=978-3-642-08671-7}}</ref>.

State-of-the-art equations of state are easily accessible through thermodynamic libraries, such as FluidProp or the open-source software CoolProp<ref name=":5">{{Cite journal |last=Bell |first=Ian H. |last2=Wronski |first2=Jorrit |last3=Quoilin |first3=Sylvain |last4=Lemort |first4=Vincent |date=2014-01-27 |title=Pure and Pseudo-pure Fluid Thermophysical Property Evaluation and the Open-Source Thermophysical Property Library CoolProp |url=http://dx.doi.org/10.1021/ie4033999 |journal=Industrial &amp; Engineering Chemistry Research |volume=53 |issue=6 |pages=2498–2508 |doi=10.1021/ie4033999 |issn=0888-5885}}</ref>.

== Non-ideal gasdynamic regimes ==
The dynamic behavior of compressible flows is governed by the dimensionless thermodynamic quantity <math> \Gamma </math>, which is known as the fundamental derivative of gasdynamics<ref name=":6">{{Cite journal |last=Thompson |first=Philip A. |date=1971 |title=A Fundamental Derivative in Gasdynamics |url=http://dx.doi.org/10.1063/1.1693693 |journal=Physics of Fluids |volume=14 |issue=9 |pages=1843 |doi=10.1063/1.1693693 |issn=0031-9171}}</ref> and is defined as
:<math> \Gamma = \frac{v^2}{2c^2} \left(\frac{\partial^2 P}{\partial v^2} \right)_s = 1+ \frac{c}{v} \left(\frac{\partial c}{\partial P} \right)_s </math>

where

* <math> c </math> is the speed of sound [m/s];
* <math> s </math> is the specific [[entropy]] per unit mass [J/kgK].

From a mathematical point of view, the fundamental derivative is a non-dimensional measure of the curvature of [[Isentropic process|isentropes]] in the [[Pressure–volume diagram|pressure-volume thermodynamic plane]]. From a physical point of view, the definition of <math> \Gamma </math> tells that the speed of sound increases with pressure in isentropic transformations for values of <math> \Gamma > 1 </math>, while, by contrast, it decreases with pressure for <math> \Gamma < 1 </math>.

Based on the value of <math> \Gamma </math>, three gasdynamic regimes can be defined<ref name=":6" />:

* ideal gasdynamic regime for <math> \Gamma > 1 </math>;
* non-ideal classical gasdynamic regime for <math> 0 < \Gamma < 1 </math>;
* non-classical gasdynamic regime for <math> \Gamma < 0 </math>.

=== Ideal gasdynamic regime ===
In the ideal regime, the usual ideal-gas behavior is qualitatively recovered. For an ideal gas, in fact, the value of the fundamental derivative reduces to the constant value <math> \Gamma = \frac{\gamma + 1}{2} </math>, where <math> \gamma </math> is the [[heat capacity ratio]]. By definition, <math> \gamma </math> is the ratio between the constant pressure and the constant volume [[Specific heat capacity|specific heats]], so it is larger than 1, leading to a value of <math> \Gamma </math> larger than 1 too.<ref name=":4" />

In this regime, only quantitative differences with respect to the ideal model are encountered. The flow evolution in fact depends on total, or [[Stagnation point|stagnation]], thermodynamic conditions. For example, the [[Mach number]] evolution of an ideal gas in a [[De Laval nozzle|supersonic nozzle]] depends only on the heat capacity ratio (namely on the fluid) and on the exhaust-to-stagnation pressure ratio<ref name=":4" />. Considering real-gas effects, instead, even fixing the fluid and the pressure ratio, different total states yield different Mach profiles.<ref>{{Cite journal |last=Tsien |first=Hsue-Shen |date=1946 |title=One-Dimensional Flows of a Gas Characterized by Vander Waal's Equation of State |url=http://dx.doi.org/10.1002/sapm1946251301 |journal=Journal of Mathematics and Physics |volume=25 |issue=1-4 |pages=301–324 |doi=10.1002/sapm1946251301 |issn=0097-1421}}</ref>

Typically, for single-phase fluids made of simple molecules, only the ideal gasdynamic regime can be reached, even for thermodynamic conditions very close to saturation. It is for example the case of [[Diatomic molecule|diatomic]] or [[Triatomic molecule|triatomic]] molecules, such as [[nitrogen]] or [[carbon dioxide]], which can only experience small departure from the ideal behavior<ref name=":3" />.

=== Non-ideal classical gasdynamic regime ===
[[File:MM pv plane 03.png|thumb|313x313px|Reduced pressure-volume thermodynamic diagram for siloxane fluid MM ([[hexamethyldisiloxane]], <chem>C6H18OSi2
</chem>), including the liquid-vapor saturation curve, some isentropes and some isolines of the fundamental derivative of gas dynamics <math> \Gamma </math>. The non-ideal gas region (<math> 0 < \Gamma < 1 </math>) is shown close to the saturation curve.]]
For fluids with high molecular complexity, state-of-the-art thermodynamic models predict values of <math> 0 < \Gamma < 1 </math> in the single-phase region close to the saturaton curve, where the speed of sound is largely sensitive to density variations along isentropes<ref name=":7" />. Such fluids belong to different classes of [[Chemical compound|chemical compounds]], including [[Hydrocarbon|hydrocarbons]], [[Siloxane|siloxanes]] and [[Refrigerant|refrigerants]]<ref name=":3" /><ref name=":7">{{Cite journal |last=Kluwick |first=Alfred |date=2004-05-01 |title=Internal flows of dense gases |url=http://dx.doi.org/10.1007/s00707-004-0096-z |journal=Acta Mechanica |volume=169 |issue=1-4 |pages=123–143 |doi=10.1007/s00707-004-0096-z |issn=0001-5970}}</ref>.

In the non-ideal regime, even qualitative differences with respect to ideal gasdynamics can be found, meaning that the flow evolution can be strongly different for varying total conditions. The most peculiar phenomenon of the non-ideal regime is the decrease of the Mach number in isentropic [[Rarefaction|expansions]] occurring in the supersonic regime, namely processes in which the fluid [[density]] decreases<ref name=":8" />. Indeed, for an ideal gas expanding isentropically in a converging-diverging nozzle, the Mach number increases monotonically as the density decreases<ref name=":4" />. By contrast, for flows evolving in the non-ideal regime, a non-monotone Mach number evolution is possible in the divergent section, whereas the density reduction remains monotonic (see figure in the lead section). This particular phenomenon is governed by the quantity <math> J </math>, which is a non-dimensional measure of the Mach number [[derivative]] with respect to density in isentropic processes<ref name=":8">{{Cite journal |last=Cramer |first=M. S. |last2=Best |first2=L. M. |date=1991 |title=Steady, isentropic flows of dense gases |url=http://dx.doi.org/10.1063/1.857855 |journal=Physics of Fluids A: Fluid Dynamics |volume=3 |issue=1 |pages=219–226 |doi=10.1063/1.857855 |issn=0899-8213}}</ref>:
:<math> J = \frac{\rho}{M} \frac{dM}{d \rho} = 1-\Gamma - \frac{1}{M^2} </math>
where

* <math> M </math> is the Mach number;
* <math> \rho </math> is the density [kg/m<sup>3</sup>].

From the definition of <math> J </math>, the Mach number increases with density for flow conditions featuring values of <math> J>0 </math>. Indeed, this is possible only for values of <math> \Gamma < 1 </math>, that is in the non-ideal regime. However, this is not a sufficient condition for the non-monotone Mach number to appear, since a sufficiently large value of <math> M </math> is also required. In particular, supersonic conditions (<math> M > 1 </math>) are necessary<ref name=":8" />.

An analogous effect is encountered in the expansion around [[Prandtl–Meyer expansion fan|rarefactive ramps]]: for suitable thermodynamic conditions, the Mach number downstream of the ramp can be lower than the one upstream<ref>{{Cite journal |last=Cramer |first=M. S. |last2=Crickenberger |first2=A. B. |date=1992 |title=Prandtl-Meyer function for dense gases |url=http://dx.doi.org/10.2514/3.10956 |journal=AIAA Journal |volume=30 |issue=2 |pages=561–564 |doi=10.2514/3.10956 |issn=0001-1452}}</ref>. By contrast, in [[Oblique shock|oblique shock waves]], the post-shock Mach number can be larger than the pre-shock one<ref>{{Cite journal |last=Vimercati |first=Davide |last2=Gori |first2=Giulio |last3=Guardone |first3=Alberto |date=2018-05-21 |title=Non-ideal oblique shock waves |url=http://dx.doi.org/10.1017/jfm.2018.328 |journal=Journal of Fluid Mechanics |volume=847 |pages=266–285 |doi=10.1017/jfm.2018.328 |issn=0022-1120}}</ref>.

=== Non-classical gasdynamic regime ===
Finally, fluids with an even higher molecular complexity can exhibit non-classical behavior in the single-phase vapor region near saturation. They are called Bethe-Zel’dovich-Thompson (BZT) fluids, from the name of physicists [[Hans Bethe]]<ref>{{Citation |last=Bethe |first=H. A. |title=On the Theory of Shock Waves for an Arbitrary Equation of State |date=1998 |url=http://dx.doi.org/10.1007/978-1-4612-2218-7_11 |work=Classic Papers in Shock Compression Science |pages=421–495 |access-date=2023-07-05 |place=New York, NY |publisher=Springer New York |isbn=978-1-4612-7461-2}}</ref>, [[Yakov Zeldovich|Yakov Zel'dovich]]<ref>{{Citation |title=14. On the Possibility of Rarefaction Shock Waves |date=1992-12-31 |url=http://dx.doi.org/10.1515/9781400862979.152 |work=Selected Works of Yakov Borisovich Zeldovich, Volume I |pages=152–154 |access-date=2023-07-05 |publisher=Princeton University Press}}</ref> and Philip Thompson<ref name=":2">{{Cite journal |last=Thompson |first=P. A. |last2=Lambrakis |first2=K. C. |date=1973-08-21 |title=Negative shock waves |url=http://dx.doi.org/10.1017/s002211207300011x |journal=Journal of Fluid Mechanics |volume=60 |issue=1 |pages=187–208 |doi=10.1017/s002211207300011x |issn=0022-1120}}</ref><ref name=":9">{{Cite journal |last=Lambrakis |first=Konstantine C. |date=1972 |title=Existence of Real Fluids with a Negative Fundamental Derivative Γ |url=http://dx.doi.org/10.1063/1.1694004 |journal=Physics of Fluids |volume=15 |issue=5 |pages=933 |doi=10.1063/1.1694004 |issn=0031-9171}}</ref>, who first worked on these kinds of fluids.

For thermodynamic conditions lying in the non-classical regime, the non-monotone evolution of the Mach number in isentropic expansions can be found even in subsonic conditions. In fact, for values of <math> \Gamma < 0 </math>, positive values of <math> J </math> can be reached also in subsonic flows (<math> M < 1 </math>). In other words, the non-monotone Mach number evolution is also possible in the convergent section of an isentropic nozzle<ref name=":9" />.

Moreover, a peculiar phenomenon of the non-classical regime is the so-called ''inverted gasdynamics''. In the classical regime, expansions are smooth isentropic processes, while compressions occur through [[Shock wave|shock waves]], that are discontinuities in the flow. If gasdynamics is inverted, the opposite occurs, namely rarefaction shock waves are physically admissible and compressions occur through smooth isentropic processes<ref name=":2" />.

As a consequence of the negative value of <math> \Gamma </math>, two other peculiar phenomena can occur for BZT fluids: shock splitting and composite waves. Shock splitting occurs when an inadmissible pressure discontinuity evolves in time by generating two weaker shock waves<ref>{{Cite journal |last=Cramer |first=M. S. |date=February 1989 |title=Shock splitting in single-phase gases |url=http://dx.doi.org/10.1017/s0022112089000388 |journal=Journal of Fluid Mechanics |volume=199 |pages=281–296 |doi=10.1017/s0022112089000388 |issn=0022-1120}}</ref><ref>{{Citation |last=Cramer |first=M. S. |title=Nonclassical Dynamics of Classical Gases |date=1991 |url=http://dx.doi.org/10.1007/978-3-7091-2608-0_5 |work=Nonlinear Waves in Real Fluids |pages=91–145 |access-date=2023-07-05 |place=Vienna |publisher=Springer Vienna |isbn=978-3-211-82277-7}}</ref>. Composite waves, instead, are referred to as phenomena in which two elementary waves propagate as a single entity<ref name=":1">{{Cite journal |last=Menikoff |first=Ralph |last2=Plohr |first2=Bradley J. |date=1989-01-01 |title=The Riemann problem for fluid flow of real materials |url=http://dx.doi.org/10.1103/revmodphys.61.75 |journal=Reviews of Modern Physics |volume=61 |issue=1 |pages=75–130 |doi=10.1103/revmodphys.61.75 |issn=0034-6861}}</ref><ref>{{Citation |last=KLUWICK |first=ALFRED |title=Rarefaction Shocks |date=2001 |url=http://dx.doi.org/10.1016/b978-012086430-0/50008-7 |work=Handbook of Shock Waves |pages=339–411 |access-date= |publisher=Elsevier}}</ref>.

Experimental evidence of non-classical gasdynamic regime is not available yet. The main reasons are the complexity of performing experiments in such challenging thermodynamic conditions and the [[Thermal decomposition|thermal stability]] of these very complex molecules<ref>{{Cite journal |last=Borisov |first=A. A. |last2=Borisov |first2=Al. A. |last3=Kutateladze |first3=S. S. |last4=Nakoryakov |first4=V. E. |date=January 1983 |title=Rarefaction shock wave near the critical liquid–vapour point |url=http://dx.doi.org/10.1017/s002211208300004x |journal=Journal of Fluid Mechanics |volume=126 |pages=59–73 |doi=10.1017/s002211208300004x |issn=0022-1120}}</ref>.

== Applications ==
Compressible flows in non-ideal conditions are encountered in several industrial and [[aerospace]] applications. They are employed for example in [[Organic Rankine cycle|Organic Rankine Cycles]] (ORC)<ref>{{Citation |last=Angelino |first=G. |title=Organic Working Fluid Optimization for Space Power Cycles |date=1991 |url=http://dx.doi.org/10.1007/978-1-4612-0945-4_16 |work=Modern Research Topics in Aerospace Propulsion |pages=297–326 |access-date=2023-07-05 |place=New York, NY |publisher=Springer New York |isbn=978-1-4612-6956-4 |last2=Invernizzi |first2=C. |last3=Macchi |first3=E.}}</ref> and [[supercritical carbon dioxide]] (sCO<sub>2</sub>) systems<ref>{{Cite journal |last=Feher |first=E.G. |date=1968 |title=The supercritical thermodynamic power cycle |url=http://dx.doi.org/10.1016/0013-7480(68)90105-8 |journal=Energy Conversion |volume=8 |issue=2 |pages=85–90 |doi=10.1016/0013-7480(68)90105-8 |issn=0013-7480}}</ref> for [[Electricity generation|power production]]. In the aerospace field, fluids in conditions close to saturation can be used as [[Oxide|oxiders]] in [[Hybrid-propellant rocket|hybrid rocket motors]] or for surface cooling of [[Rocket engine nozzle|rocket nozzles]]<ref>{{Cite journal |date=1957 |title=Rocket Propulsion Elements. 2nd edition. George P. Sutton. J. Wiley and Sons Inc., New York, 1956. 483 pp. Illustrated. 82s. |url=http://dx.doi.org/10.1017/s0368393100128512 |journal=The Journal of the Royal Aeronautical Society |volume=61 |issue=559 |pages=503–503 |doi=10.1017/s0368393100128512 |issn=0368-3931}}</ref>. Gases made of molecules of high molecular mass can be used in [[Supersonic wind tunnel|supersonic wind tunnels]] instead of air to obtain higher [[Reynolds number|Reynolds numbers]]<ref>{{Cite journal |last=Sagnier |first=P. |last2=Vérant |first2=J.-L. |date=1998 |title=On the validation of high enthalpy wind tunnel simulations |url=http://dx.doi.org/10.1016/s1270-9638(99)80002-9 |journal=Aerospace Science and Technology |volume=2 |issue=7 |pages=425–437 |doi=10.1016/s1270-9638(99)80002-9 |issn=1270-9638}}</ref>. Finally, non-ideal flows find application in [[Fuel|fuels]] transportation at high-speed and in Rapid Expansion of Supercritical Solutions (RESS) of CO<sub>2</sub> for particles generation or extraction of chemicals<ref>{{Cite journal |last=Helfgen |first=B. |last2=Türk |first2=M. |last3=Schaber |first3=K. |date=2003 |title=Hydrodynamic and aerosol modelling of the rapid expansion of supercritical solutions (RESS-process) |url=http://dx.doi.org/10.1016/s0896-8446(02)00159-6 |journal=The Journal of Supercritical Fluids |volume=26 |issue=3 |pages=225–242 |doi=10.1016/s0896-8446(02)00159-6 |issn=0896-8446}}</ref>.

=== Organic Rankine cycles ===
{{Main|Organic Rankine cycle}}
[[File:ORC power plant.JPG|thumb|ORC [[Turbo generator|turbogenerator]] at the [[LUT University]] in [[Lappeenranta|Lappenranta]], [[Finland]].<ref>{{Cite journal |last=Turunen-Saaresti |first=Teemu |last2=Uusitalo |first2=Antti |last3=Honkatukia |first3=Juha |date=2017 |title=Design and testing of high temperature micro-ORC test stand using Siloxane as working fluid |url=http://dx.doi.org/10.1088/1742-6596/821/1/012024 |journal=Journal of Physics: Conference Series |volume=821 |pages=012024 |doi=10.1088/1742-6596/821/1/012024 |issn=1742-6588}}</ref>]]Usual [[Rankine cycle|Rankine cycles]] are [[Thermodynamic cycle|thermodynamic cycles]] which employ [[water]] as [[working fluid]] to produce [[electric power]] from thermal sources<ref>{{Cite journal |last=Saitoh |first=T.S. |last2=Hoshi |first2=A. |title=Proposed solar Rankine cycle system with phase change steam accumulator and CPC solar collector |url=http://dx.doi.org/10.1109/iecec.2002.1392137 |journal=IECEC '02. 2002 37th Intersociety Energy Conversion Engineering Conference, 2002. |publisher=IEEE |doi=10.1109/iecec.2002.1392137}}</ref>. In Organic Rankine cycles, by contrast, water is substituted by molecularly complex [[Organic compound|organic compounds]]. Since the [[vaporization]] temperature of these kinds of fluids is lower than that of water, low-to-medium temperature sources can be exploited allowing for [[Heat recovery steam generator|heat recovery]], for example, from [[Biomass (energy)|biomass]] [[combustion]], industrial [[waste heat]] or [[Geothermal energy|geothermal heat]].<ref>{{Cite journal |last=Quoilin |first=Sylvain |last2=Broek |first2=Martijn Van Den |last3=Declaye |first3=Sébastien |last4=Dewallef |first4=Pierre |last5=Lemort |first5=Vincent |date=2013 |title=Techno-economic survey of Organic Rankine Cycle (ORC) systems |url=http://dx.doi.org/10.1016/j.rser.2013.01.028 |journal=Renewable and Sustainable Energy Reviews |volume=22 |pages=168–186 |doi=10.1016/j.rser.2013.01.028 |issn=1364-0321}}</ref> For these reasons, the ORC technology belongs to the class of [[Renewable energy|renewable energies]].

For the design of mechanical components, such as [[Turbine|turbines]], working in ORC plants, it is fundamental to take into account typical non-ideal gasdynamic phenomena. In fact, the single-phase vapor at the inlet of an ORC turbine [[stator]] usually evolves in the non-ideal thermodynamic region close to liquid-vapor saturation curve and critical point. Moreover, due to the high molecular mass of the complex organic compounds employed, the speed of sound in these fluids is low compared to that of air and other simple gases. Therefore, turbine stators are very likely to involve supersonic flows even if rather low flow velocities are reached<ref>{{Cite journal |last=Brown |first=Brady |last2=Argrow |first2=Brian |date=1999-01-11 |title=Application of Bethe-Zel'dovich-Thompson fluids in organic Rankine cycle engines |url=http://dx.doi.org/10.2514/6.1999-462 |journal=37th Aerospace Sciences Meeting and Exhibit |location=Reston, Virigina |publisher=American Institute of Aeronautics and Astronautics |doi=10.2514/6.1999-462}}</ref>. High supersonic flows can produce large losses and mechanical [[Stress (mechanics)|stresses]] in the [[Turbine blade|turbine blades]] due to occurrence of shock waves, which cause a strong pressure raise<ref>{{Cite journal |last=Denton |first=J. D. |last2=Xu |first2=L. |date=1989-06-04 |title=The Trailing Edge Loss of Transonic Turbine Blades |url=http://dx.doi.org/10.1115/89-gt-278 |journal=Volume 1: Turbomachinery |publisher=American Society of Mechanical Engineers |doi=10.1115/89-gt-278}}</ref>. However, when working fluids of the BZT class are employed, expander performances could be improved by exploiting some non-classical phenomena.<ref>{{Citation |last=Kluwick |first=A. |title=Interacting laminar boundary layers of dense gases |date=1994 |url=http://dx.doi.org/10.1007/978-3-7091-9310-5_37 |work=Fluid- and Gasdynamics |pages=335–349 |access-date=2023-07-06 |place=Vienna |publisher=Springer Vienna |isbn=978-3-211-82495-5}}</ref><ref>{{Cite journal |last=CRAMER |first=M. S. |last2=PARK |first2=S. |date=1999 |title=On the suppression of shock-induced separation in Bethe–Zel'dovich–Thompson fluids |url=http://dx.doi.org/10.1017/s0022112099005479 |journal=Journal of Fluid Mechanics |volume=393 |pages=1–21 |doi=10.1017/s0022112099005479 |issn=0022-1120}}</ref>

=== Supercritical carbon dioxide cycles ===
{{Main|Supercritical carbon dioxide#Applications}}

When carbon dioxide is held above its critical pressure (73.773 bar)<ref name=":10">{{Cite journal |last=Span |first=Roland |last2=Wagner |first2=Wolfgang |date=1996-11-01 |title=A New Equation of State for Carbon Dioxide Covering the Fluid Region from the Triple-Point Temperature to 1100 K at Pressures up to 800 MPa |url=http://dx.doi.org/10.1063/1.555991 |journal=Journal of Physical and Chemical Reference Data |volume=25 |issue=6 |pages=1509–1596 |doi=10.1063/1.555991 |issn=0047-2689}}</ref> and temperature (30.9780 °C)<ref name=":10" />, it can behave both as a gas and as a liquid, that is it expands to fill entirely its container like a gas but has a density similar to that of a liquid.

Supercritical CO<sub>2</sub> is [[Chemical stability|chemically stable]], very cheap and [[Combustibility and flammability|non-flammable]], making it suitable as a working fluid for [[Transcritical cycle|transcritical cycles]]<ref name=":11">{{Cite journal |last=Ma |first=Yitai |last2=Liu |first2=Zhongyan |last3=Tian |first3=Hua |date=2013 |title=A review of transcritical carbon dioxide heat pump and refrigeration cycles |url=http://dx.doi.org/10.1016/j.energy.2013.03.030 |journal=Energy |volume=55 |pages=156–172 |doi=10.1016/j.energy.2013.03.030 |issn=0360-5442}}</ref>. For example, it is employed in domestic water [[Heat pump|heat pumps]], which can reach high [[Efficiency|efficiencies]]<ref name=":11" />.

Moreover, when used in power generation plants which employ [[Brayton cycle|Brayton]] and Rankine cycles, it can improve efficiency and power output. In fact, its high density enables a strong reduction in turbomachines dimensions, still ensuring high efficiency of these components. Simpler designs can therefore be adopted, while [[Steam turbine|steam turbines]] require multiple turbine stages, which necessarily yield larger dimensions and costs.<ref>{{Cite journal |last=Fleming |first=Darryn |last2=Pasch |first2=Jim |last3=Conboy |first3=Thomas |last4=Carlson |first4=Matt |date=2013-06-03 |title=Testing Platform and Commercialization Plan for Heat Exchanging Systems for SCO2 Power Cycles |url=http://dx.doi.org/10.1115/gt2013-95125 |journal=Volume 8: Supercritical CO2 Power Cycles; Wind Energy; Honors and Awards |publisher=American Society of Mechanical Engineers |doi=10.1115/gt2013-95125}}</ref>

By contrast, mechanical components within sCO<sub>2</sub> Brayton cycles, especially turbomachinery and heat exchangers, suffer from [[corrosion]].<ref>{{Cite journal |last=Fleming |first=Darryn |last2=Kruizenga |first2=Alan |last3=Pasch |first3=James |last4=Conboy |first4=Tom |last5=Carlson |first5=Matt |date=2014-06-16 |title=Corrosion and Erosion Behavior in Supercritical CO2 Power Cycles |url=http://dx.doi.org/10.1115/gt2014-25136 |journal=Volume 3B: Oil and Gas Applications; Organic Rankine Cycle Power Systems; Supercritical CO2 Power Cycles; Wind Energy |publisher=American Society of Mechanical Engineers |doi=10.1115/gt2014-25136}}</ref>

== See also ==

* [[Compressible flow]]
* [[Real gas]]
* [[Equation of state]]
* [[Mach number]]
* [[De Laval nozzle|Supersonic nozzle flow]]
* [[Shock wave]]
* [[Prandtl–Meyer expansion fan]]
* [[Organic Rankine cycle]]
* [[Supercritical carbon dioxide]]

== References ==
<references />
<references />

== Further readings ==

* {{Cite book |last=Anderson |first=John David |title=Modern Compressible Flow: With Historical Perspective |publisher=McGraw-Hill |year=2003 |language=en}}
* {{Cite book |last=di Mare |first=Francesca |title=Non-Ideal Compressible Fluid Dynamics for Propulsion and Power |last2=Spinelli |first2=Andrea |last3=Pini |first3=Matteo |publisher=Springer Link |year=2018 |language=en}}
* {{Cite journal |last=Feher |first=E.G. |date=1968 |title=The supercritical thermodynamic power cycle |url=https://www.sciencedirect.com/science/article/abs/pii/0013748068901058 |journal=Energy Conversion |volume=8 |issue=2 |pages=85-90}}
* {{Cite journal |last=Kluwick |first=Alfred |date=2017 |title=Non-Ideal Compressible Fluid Dynamics: A Challenge for Theory |url=https://iopscience.iop.org/article/10.1088/1742-6596/821/1/012001/pdf |journal=Journal of Physics}}
*{{Cite book |last=Macchi |first=Ennio |title=Organic Rankine Cycle (ORC) Power Systems |last2=Astolfi |first2=Marco |publisher=Elsevier |year=2016 |isbn=9780081005101 |edition=1st |language=en}}

== External links ==

* [http://www.coolprop.org/ Open-source thermodynamic library CoolProp]
* [https://asimptote.com/fluidprop/ Thermodynamic library FluidProp]
* [https://separeco.com/co2-atomization-process/rapid-expansion-of-supercritical-solution/#:~:text=RESS%20(Rapid%20Expansion%20of%20Supercritical,solution%20through%20an%20expansion%20nozzle. Rapid Expansion of Supercritical Solutions (RESS)]



[[Category:Fluid mechanics]]
[[Category:Fluid mechanics]]
[[Category:Thermodynamics]]

Revision as of 10:12, 7 July 2023

Non-monotone evolution of the Mach number M in the divergent section of a supersonic nozzle. The fluid is siloxane MM (hexamethyldisiloxane, ) evolving in the non-ideal gasdynamic regime.

Non ideal compressible fluid dynamics (NICFD), or non ideal gasdynamics, is a branch of fluid mechanics studying the dynamic behavior of fluids not obeying ideal-gas thermodynamics. It is for example the case of dense vapors, supercritical flows and compressible two-phase flows. With the term dense vapors, we indicate all fluids in the gaseous state characterized by thermodynamic conditions close to saturation and the critical point[1]. Supercritical fluids feature instead values of pressure and temperature larger than their critical values[2], wheras two-phase flows are characterized by the simultaneous presence of both liquid and gas states.[3]

In all these cases, the fluid requires to be modelled as a real gas, since its thermodynamic behavior considerably differs from that of an ideal gas, which by contrast appears for dilute thermodynamic conditions. The ideal-gas law, in fact, can be employed in general as a reasonable approximation of the fluid thermodynamics for low pressures and high temperatures. Otherwise, intermolecular forces and dimension of fluid particles, which are neglected in the ideal-gas approximation, become relevant and can significantly affect the fluid behavior[4]. This is extremely valid for gases made of complex and heavy molecules, which tend to deviate more from the ideal model.[5]

While the fluid dynamics of compressible flows in ideal conditions is well-established and is characterized by several analytical results[6], when non-ideal thermodynamic conditions are considered, peculiar phenomena possibly occur. This is particularly valid in supersonic conditions, namely for flow velocities larger than the speed of sound in the fluid considered. All typical features of supersonic flows, in fact, are affected by non-ideal thermodynamics, resulting in both quantitative and qualitative differences with respect to the ideal gasdynamics.[7]

Non-ideal thermodynamics

Compressibility factor Z for different values of reduced pressure and temperature.

For dilute thermodynamic conditions, the ideal-gas equation of state (EoS) provides sufficiently accurate results in modelling the fluid thermodynamics. This occurs in general for low values of reduced pressure and high values of reduced temperature, where the term reduced refers to the ratio of a certain thermodynamic quantity and its critical value. For most applications including air, for example, the assumption of considering ideal conditions is perfectly reasonable and it is widely used[6].

On the other hand, when thermodynamic conditions approach condensation and the critical point or when high pressures are involved, real-gas models are needed in order to capture the real fluid behavior. In these conditions, in fact, intermolecular forces and compressibility effects come into play[4].

A measure of the fluid non-ideality is given by the compressibility factor [8], defined as

where

  • is the pressure [Pa];
  • is the specific volume [m3/kg];
  • is the gas constant [J/kgK];
  • is the absolute temperature [K].

The compressibility factor is a dimensionless quantity which is equal to 1 for ideal gases and deviates from unity for increasing levels of non-ideality.[9]

Several non-ideal models exist, from the simplest cubic equations of state (such as the Van der Waals[10][4] and the Peng-Robinson[11] models) up to complex multi-parameter ones, including the Span-Wagner equation of state[12][13].

State-of-the-art equations of state are easily accessible through thermodynamic libraries, such as FluidProp or the open-source software CoolProp[14].

Non-ideal gasdynamic regimes

The dynamic behavior of compressible flows is governed by the dimensionless thermodynamic quantity , which is known as the fundamental derivative of gasdynamics[15] and is defined as

where

  • is the speed of sound [m/s];
  • is the specific entropy per unit mass [J/kgK].

From a mathematical point of view, the fundamental derivative is a non-dimensional measure of the curvature of isentropes in the pressure-volume thermodynamic plane. From a physical point of view, the definition of tells that the speed of sound increases with pressure in isentropic transformations for values of , while, by contrast, it decreases with pressure for .

Based on the value of , three gasdynamic regimes can be defined[15]:

  • ideal gasdynamic regime for ;
  • non-ideal classical gasdynamic regime for ;
  • non-classical gasdynamic regime for .

Ideal gasdynamic regime

In the ideal regime, the usual ideal-gas behavior is qualitatively recovered. For an ideal gas, in fact, the value of the fundamental derivative reduces to the constant value , where is the heat capacity ratio. By definition, is the ratio between the constant pressure and the constant volume specific heats, so it is larger than 1, leading to a value of larger than 1 too.[6]

In this regime, only quantitative differences with respect to the ideal model are encountered. The flow evolution in fact depends on total, or stagnation, thermodynamic conditions. For example, the Mach number evolution of an ideal gas in a supersonic nozzle depends only on the heat capacity ratio (namely on the fluid) and on the exhaust-to-stagnation pressure ratio[6]. Considering real-gas effects, instead, even fixing the fluid and the pressure ratio, different total states yield different Mach profiles.[16]

Typically, for single-phase fluids made of simple molecules, only the ideal gasdynamic regime can be reached, even for thermodynamic conditions very close to saturation. It is for example the case of diatomic or triatomic molecules, such as nitrogen or carbon dioxide, which can only experience small departure from the ideal behavior[5].

Non-ideal classical gasdynamic regime

Reduced pressure-volume thermodynamic diagram for siloxane fluid MM (hexamethyldisiloxane, ), including the liquid-vapor saturation curve, some isentropes and some isolines of the fundamental derivative of gas dynamics . The non-ideal gas region () is shown close to the saturation curve.

For fluids with high molecular complexity, state-of-the-art thermodynamic models predict values of in the single-phase region close to the saturaton curve, where the speed of sound is largely sensitive to density variations along isentropes[17]. Such fluids belong to different classes of chemical compounds, including hydrocarbons, siloxanes and refrigerants[5][17].

In the non-ideal regime, even qualitative differences with respect to ideal gasdynamics can be found, meaning that the flow evolution can be strongly different for varying total conditions. The most peculiar phenomenon of the non-ideal regime is the decrease of the Mach number in isentropic expansions occurring in the supersonic regime, namely processes in which the fluid density decreases[18]. Indeed, for an ideal gas expanding isentropically in a converging-diverging nozzle, the Mach number increases monotonically as the density decreases[6]. By contrast, for flows evolving in the non-ideal regime, a non-monotone Mach number evolution is possible in the divergent section, whereas the density reduction remains monotonic (see figure in the lead section). This particular phenomenon is governed by the quantity , which is a non-dimensional measure of the Mach number derivative with respect to density in isentropic processes[18]:

where

  • is the Mach number;
  • is the density [kg/m3].

From the definition of , the Mach number increases with density for flow conditions featuring values of . Indeed, this is possible only for values of , that is in the non-ideal regime. However, this is not a sufficient condition for the non-monotone Mach number to appear, since a sufficiently large value of is also required. In particular, supersonic conditions () are necessary[18].

An analogous effect is encountered in the expansion around rarefactive ramps: for suitable thermodynamic conditions, the Mach number downstream of the ramp can be lower than the one upstream[19]. By contrast, in oblique shock waves, the post-shock Mach number can be larger than the pre-shock one[20].

Non-classical gasdynamic regime

Finally, fluids with an even higher molecular complexity can exhibit non-classical behavior in the single-phase vapor region near saturation. They are called Bethe-Zel’dovich-Thompson (BZT) fluids, from the name of physicists Hans Bethe[21], Yakov Zel'dovich[22] and Philip Thompson[23][24], who first worked on these kinds of fluids.

For thermodynamic conditions lying in the non-classical regime, the non-monotone evolution of the Mach number in isentropic expansions can be found even in subsonic conditions. In fact, for values of , positive values of can be reached also in subsonic flows (). In other words, the non-monotone Mach number evolution is also possible in the convergent section of an isentropic nozzle[24].

Moreover, a peculiar phenomenon of the non-classical regime is the so-called inverted gasdynamics. In the classical regime, expansions are smooth isentropic processes, while compressions occur through shock waves, that are discontinuities in the flow. If gasdynamics is inverted, the opposite occurs, namely rarefaction shock waves are physically admissible and compressions occur through smooth isentropic processes[23].

As a consequence of the negative value of , two other peculiar phenomena can occur for BZT fluids: shock splitting and composite waves. Shock splitting occurs when an inadmissible pressure discontinuity evolves in time by generating two weaker shock waves[25][26]. Composite waves, instead, are referred to as phenomena in which two elementary waves propagate as a single entity[7][27].

Experimental evidence of non-classical gasdynamic regime is not available yet. The main reasons are the complexity of performing experiments in such challenging thermodynamic conditions and the thermal stability of these very complex molecules[28].

Applications

Compressible flows in non-ideal conditions are encountered in several industrial and aerospace applications. They are employed for example in Organic Rankine Cycles (ORC)[29] and supercritical carbon dioxide (sCO2) systems[30] for power production. In the aerospace field, fluids in conditions close to saturation can be used as oxiders in hybrid rocket motors or for surface cooling of rocket nozzles[31]. Gases made of molecules of high molecular mass can be used in supersonic wind tunnels instead of air to obtain higher Reynolds numbers[32]. Finally, non-ideal flows find application in fuels transportation at high-speed and in Rapid Expansion of Supercritical Solutions (RESS) of CO2 for particles generation or extraction of chemicals[33].

Organic Rankine cycles

ORC turbogenerator at the LUT University in Lappenranta, Finland.[34]

Usual Rankine cycles are thermodynamic cycles which employ water as working fluid to produce electric power from thermal sources[35]. In Organic Rankine cycles, by contrast, water is substituted by molecularly complex organic compounds. Since the vaporization temperature of these kinds of fluids is lower than that of water, low-to-medium temperature sources can be exploited allowing for heat recovery, for example, from biomass combustion, industrial waste heat or geothermal heat.[36] For these reasons, the ORC technology belongs to the class of renewable energies.

For the design of mechanical components, such as turbines, working in ORC plants, it is fundamental to take into account typical non-ideal gasdynamic phenomena. In fact, the single-phase vapor at the inlet of an ORC turbine stator usually evolves in the non-ideal thermodynamic region close to liquid-vapor saturation curve and critical point. Moreover, due to the high molecular mass of the complex organic compounds employed, the speed of sound in these fluids is low compared to that of air and other simple gases. Therefore, turbine stators are very likely to involve supersonic flows even if rather low flow velocities are reached[37]. High supersonic flows can produce large losses and mechanical stresses in the turbine blades due to occurrence of shock waves, which cause a strong pressure raise[38]. However, when working fluids of the BZT class are employed, expander performances could be improved by exploiting some non-classical phenomena.[39][40]

Supercritical carbon dioxide cycles

When carbon dioxide is held above its critical pressure (73.773 bar)[41] and temperature (30.9780 °C)[41], it can behave both as a gas and as a liquid, that is it expands to fill entirely its container like a gas but has a density similar to that of a liquid.

Supercritical CO2 is chemically stable, very cheap and non-flammable, making it suitable as a working fluid for transcritical cycles[42]. For example, it is employed in domestic water heat pumps, which can reach high efficiencies[42].

Moreover, when used in power generation plants which employ Brayton and Rankine cycles, it can improve efficiency and power output. In fact, its high density enables a strong reduction in turbomachines dimensions, still ensuring high efficiency of these components. Simpler designs can therefore be adopted, while steam turbines require multiple turbine stages, which necessarily yield larger dimensions and costs.[43]

By contrast, mechanical components within sCO2 Brayton cycles, especially turbomachinery and heat exchangers, suffer from corrosion.[44]

See also

References

  1. ^ Callen, Herbert B. (1985). Thermodynamics and an introduction to thermostatistics (2nd ed.). New York: J. Wiley & Sons. pp. 255–261. ISBN 978-0-471-86256-7.
  2. ^ Schlosky, Kevin M. (1989). "Supercritical phase transitions at very high pressure". Journal of Chemical Education. 66 (12): 989. doi:10.1021/ed066p989. ISSN 0021-9584.
  3. ^ Faghri, Amir; Zhang, Yuwen (2006-01-01), Faghri, Amir; Zhang, Yuwen (eds.), "11 - TWO-PHASE FLOW AND HEAT TRANSFER", Transport Phenomena in Multiphase Systems, Boston: Academic Press, pp. 853–949, doi:10.1016/b978-0-12-370610-2.50016-7, ISBN 978-0-12-370610-2, retrieved 2023-07-06
  4. ^ a b c Waals, J. D. van der; Rowlinson, John Shipley (1988). On the continuity of the gaseous and liquid states. Studies in statistical mechanics. Amsterdam: North-Holland. ISBN 978-0-444-87077-3.
  5. ^ a b c Colonna, P.; Guardone, A. (2006). "Molecular interpretation of nonclassical gas dynamics of dense vapors under the van der Waals model". Physics of Fluids. 18 (5): 056101. doi:10.1063/1.2196095. ISSN 1070-6631.
  6. ^ a b c d e Thompson, Philip A. (1972). Compressible-fluid dynamics. Advanced engineering series. New York: McGraw-Hill. pp. 76–99. ISBN 978-0-07-064405-2.
  7. ^ a b Menikoff, Ralph; Plohr, Bradley J. (1989-01-01). "The Riemann problem for fluid flow of real materials". Reviews of Modern Physics. 61 (1): 75–130. doi:10.1103/revmodphys.61.75. ISSN 0034-6861.
  8. ^ Zucker, Robert D.; Biblarz, Oscar (2002). Fundamentals of gas dynamics (2nd ed.). Hoboken, NJ: Wiley. p. 327. ISBN 978-0-471-05967-7.
  9. ^ Thompson, Philip A. (1972). Compressible-fluid dynamics. Advanced engineering series. New York: McGraw-Hill. pp. 99–101. ISBN 978-0-07-064405-2.
  10. ^ Parsegian, V. Adrian (2005). Van der Waals Forces: A Handbook for Biologists, Chemists, Engineers, and Physicists. Cambridge: Cambridge University Press. doi:10.1017/cbo9780511614606. ISBN 978-0-521-83906-8.
  11. ^ Peng, Ding-Yu; Robinson, Donald B. (February 1976). "A New Two-Constant Equation of State". Industrial & Engineering Chemistry Fundamentals. 15 (1): 59–64. doi:10.1021/i160057a011. ISSN 0196-4313.
  12. ^ Span, R.; Wagner, W. (2003-01-01). "Equations of State for Technical Applications. I. Simultaneously Optimized Functional Forms for Nonpolar and Polar Fluids". International Journal of Thermophysics. 24 (1): 1–39. doi:10.1023/A:1022390430888. ISSN 1572-9567.
  13. ^ Span, Roland (2000), "Describing Mixtures with Multiparameter Equations of State", Multiparameter Equations of State, Berlin, Heidelberg: Springer Berlin Heidelberg, pp. 319–340, ISBN 978-3-642-08671-7, retrieved 2023-07-05
  14. ^ Bell, Ian H.; Wronski, Jorrit; Quoilin, Sylvain; Lemort, Vincent (2014-01-27). "Pure and Pseudo-pure Fluid Thermophysical Property Evaluation and the Open-Source Thermophysical Property Library CoolProp". Industrial & Engineering Chemistry Research. 53 (6): 2498–2508. doi:10.1021/ie4033999. ISSN 0888-5885.
  15. ^ a b Thompson, Philip A. (1971). "A Fundamental Derivative in Gasdynamics". Physics of Fluids. 14 (9): 1843. doi:10.1063/1.1693693. ISSN 0031-9171.
  16. ^ Tsien, Hsue-Shen (1946). "One-Dimensional Flows of a Gas Characterized by Vander Waal's Equation of State". Journal of Mathematics and Physics. 25 (1–4): 301–324. doi:10.1002/sapm1946251301. ISSN 0097-1421.
  17. ^ a b Kluwick, Alfred (2004-05-01). "Internal flows of dense gases". Acta Mechanica. 169 (1–4): 123–143. doi:10.1007/s00707-004-0096-z. ISSN 0001-5970.
  18. ^ a b c Cramer, M. S.; Best, L. M. (1991). "Steady, isentropic flows of dense gases". Physics of Fluids A: Fluid Dynamics. 3 (1): 219–226. doi:10.1063/1.857855. ISSN 0899-8213.
  19. ^ Cramer, M. S.; Crickenberger, A. B. (1992). "Prandtl-Meyer function for dense gases". AIAA Journal. 30 (2): 561–564. doi:10.2514/3.10956. ISSN 0001-1452.
  20. ^ Vimercati, Davide; Gori, Giulio; Guardone, Alberto (2018-05-21). "Non-ideal oblique shock waves". Journal of Fluid Mechanics. 847: 266–285. doi:10.1017/jfm.2018.328. ISSN 0022-1120.
  21. ^ Bethe, H. A. (1998), "On the Theory of Shock Waves for an Arbitrary Equation of State", Classic Papers in Shock Compression Science, New York, NY: Springer New York, pp. 421–495, ISBN 978-1-4612-7461-2, retrieved 2023-07-05
  22. ^ "14. On the Possibility of Rarefaction Shock Waves", Selected Works of Yakov Borisovich Zeldovich, Volume I, Princeton University Press, pp. 152–154, 1992-12-31, retrieved 2023-07-05
  23. ^ a b Thompson, P. A.; Lambrakis, K. C. (1973-08-21). "Negative shock waves". Journal of Fluid Mechanics. 60 (1): 187–208. doi:10.1017/s002211207300011x. ISSN 0022-1120.
  24. ^ a b Lambrakis, Konstantine C. (1972). "Existence of Real Fluids with a Negative Fundamental Derivative Γ". Physics of Fluids. 15 (5): 933. doi:10.1063/1.1694004. ISSN 0031-9171.
  25. ^ Cramer, M. S. (February 1989). "Shock splitting in single-phase gases". Journal of Fluid Mechanics. 199: 281–296. doi:10.1017/s0022112089000388. ISSN 0022-1120.
  26. ^ Cramer, M. S. (1991), "Nonclassical Dynamics of Classical Gases", Nonlinear Waves in Real Fluids, Vienna: Springer Vienna, pp. 91–145, ISBN 978-3-211-82277-7, retrieved 2023-07-05
  27. ^ KLUWICK, ALFRED (2001), "Rarefaction Shocks", Handbook of Shock Waves, Elsevier, pp. 339–411
  28. ^ Borisov, A. A.; Borisov, Al. A.; Kutateladze, S. S.; Nakoryakov, V. E. (January 1983). "Rarefaction shock wave near the critical liquid–vapour point". Journal of Fluid Mechanics. 126: 59–73. doi:10.1017/s002211208300004x. ISSN 0022-1120.
  29. ^ Angelino, G.; Invernizzi, C.; Macchi, E. (1991), "Organic Working Fluid Optimization for Space Power Cycles", Modern Research Topics in Aerospace Propulsion, New York, NY: Springer New York, pp. 297–326, ISBN 978-1-4612-6956-4, retrieved 2023-07-05
  30. ^ Feher, E.G. (1968). "The supercritical thermodynamic power cycle". Energy Conversion. 8 (2): 85–90. doi:10.1016/0013-7480(68)90105-8. ISSN 0013-7480.
  31. ^ "Rocket Propulsion Elements. 2nd edition. George P. Sutton. J. Wiley and Sons Inc., New York, 1956. 483 pp. Illustrated. 82s". The Journal of the Royal Aeronautical Society. 61 (559): 503–503. 1957. doi:10.1017/s0368393100128512. ISSN 0368-3931.
  32. ^ Sagnier, P.; Vérant, J.-L. (1998). "On the validation of high enthalpy wind tunnel simulations". Aerospace Science and Technology. 2 (7): 425–437. doi:10.1016/s1270-9638(99)80002-9. ISSN 1270-9638.
  33. ^ Helfgen, B.; Türk, M.; Schaber, K. (2003). "Hydrodynamic and aerosol modelling of the rapid expansion of supercritical solutions (RESS-process)". The Journal of Supercritical Fluids. 26 (3): 225–242. doi:10.1016/s0896-8446(02)00159-6. ISSN 0896-8446.
  34. ^ Turunen-Saaresti, Teemu; Uusitalo, Antti; Honkatukia, Juha (2017). "Design and testing of high temperature micro-ORC test stand using Siloxane as working fluid". Journal of Physics: Conference Series. 821: 012024. doi:10.1088/1742-6596/821/1/012024. ISSN 1742-6588.
  35. ^ Saitoh, T.S.; Hoshi, A. "Proposed solar Rankine cycle system with phase change steam accumulator and CPC solar collector". IECEC '02. 2002 37th Intersociety Energy Conversion Engineering Conference, 2002. IEEE. doi:10.1109/iecec.2002.1392137.
  36. ^ Quoilin, Sylvain; Broek, Martijn Van Den; Declaye, Sébastien; Dewallef, Pierre; Lemort, Vincent (2013). "Techno-economic survey of Organic Rankine Cycle (ORC) systems". Renewable and Sustainable Energy Reviews. 22: 168–186. doi:10.1016/j.rser.2013.01.028. ISSN 1364-0321.
  37. ^ Brown, Brady; Argrow, Brian (1999-01-11). "Application of Bethe-Zel'dovich-Thompson fluids in organic Rankine cycle engines". 37th Aerospace Sciences Meeting and Exhibit. Reston, Virigina: American Institute of Aeronautics and Astronautics. doi:10.2514/6.1999-462.
  38. ^ Denton, J. D.; Xu, L. (1989-06-04). "The Trailing Edge Loss of Transonic Turbine Blades". Volume 1: Turbomachinery. American Society of Mechanical Engineers. doi:10.1115/89-gt-278.
  39. ^ Kluwick, A. (1994), "Interacting laminar boundary layers of dense gases", Fluid- and Gasdynamics, Vienna: Springer Vienna, pp. 335–349, ISBN 978-3-211-82495-5, retrieved 2023-07-06
  40. ^ CRAMER, M. S.; PARK, S. (1999). "On the suppression of shock-induced separation in Bethe–Zel'dovich–Thompson fluids". Journal of Fluid Mechanics. 393: 1–21. doi:10.1017/s0022112099005479. ISSN 0022-1120.
  41. ^ a b Span, Roland; Wagner, Wolfgang (1996-11-01). "A New Equation of State for Carbon Dioxide Covering the Fluid Region from the Triple-Point Temperature to 1100 K at Pressures up to 800 MPa". Journal of Physical and Chemical Reference Data. 25 (6): 1509–1596. doi:10.1063/1.555991. ISSN 0047-2689.
  42. ^ a b Ma, Yitai; Liu, Zhongyan; Tian, Hua (2013). "A review of transcritical carbon dioxide heat pump and refrigeration cycles". Energy. 55: 156–172. doi:10.1016/j.energy.2013.03.030. ISSN 0360-5442.
  43. ^ Fleming, Darryn; Pasch, Jim; Conboy, Thomas; Carlson, Matt (2013-06-03). "Testing Platform and Commercialization Plan for Heat Exchanging Systems for SCO2 Power Cycles". Volume 8: Supercritical CO2 Power Cycles; Wind Energy; Honors and Awards. American Society of Mechanical Engineers. doi:10.1115/gt2013-95125.
  44. ^ Fleming, Darryn; Kruizenga, Alan; Pasch, James; Conboy, Tom; Carlson, Matt (2014-06-16). "Corrosion and Erosion Behavior in Supercritical CO2 Power Cycles". Volume 3B: Oil and Gas Applications; Organic Rankine Cycle Power Systems; Supercritical CO2 Power Cycles; Wind Energy. American Society of Mechanical Engineers. doi:10.1115/gt2014-25136.

Further readings

External links