From well before recorded human history, man has quested for different sources of energy for survival and comfort. Today, the need for useful energy plays a role in almost all aspects of society. Certainly, there is a benefit to having an efficient source of mechanical energy. When designing an engine, heat pump, or other thermodynamic cycle, one can not get around the laws of thermodynamics. Prevalent is the first law, which stipulates the conservation of energy; no energy can be created or destroyed. The second law is a result of the fact that heat can only flow from hot to cold, and not cold to hot; as a result, heat transfer processes ultimately result in thermodynamic disorder known as entropy throughout the universe. These two natural limitations have to be recognized in the design of a thermodynamic machine to achieve a net mechanical work output.
Under dense, pressurized conditions, a fluid ceases to become an ideal gas, and becomes a real gas following its equation of state. At a certain point, the intermolecular attractive forces of the fluid causes the gas to condense to a liquid, where these forces are too much for the kinetic energy of the fluid molecules to overcome, and the particles converge into a more ordered liquid state. During condensation, the fluid exists at two distinct phases at a constant temperature and pressure until it is a single consistent phase. As the pressure is constant with reduced volume during condensation, the intermolecular forces will reduce the work input during condensation from a saturated gas to a mixed-phase fluid.
The inventor proposes a closed-loop, internally reversible, piston-cylinder heat engine, not dissimilar to the Stirling cycle. Rather than use an ideal gas, this cycle uses a real fluid that partially condenses during the isothermal compression stage of the cycle. The isothermal compression phase starts off as a saturated gas, and compresses isothermally at the cool temperature until a percentage of the gas has condensed. It then is heated to the hot temperature isochorically, at a temperature greater than the critical temperature. Afterwards, it expands isothermally back to the original saturated gas volume, recovering energy in the process. Finally, the gas is cooled isochorically back to the original stage pressure and temperature, where it is a saturated gas.
The engine takes advantage of the fluid's intermolecular attractive forces that enable the fluid to condense into a liquid. The impact of these forces is profound during condensation when the fluid is stable as two distinct phases of liquid and gas, as described by Maxwell's Construction. These forces keep the pressure consistent throughout condensation, rather than increasing with reduced volume as would be described during the equation of state; this ultimately results in less work input to compressed the gas isothermally, and thus greater efficiency of the heat engine.
List of labeled components in
This heat engine is a modification of the Stirling cycle, a heat engine cycle of isothermal compression at the cold temperature sink, followed by isochoric heating up to the high temperature source, followed by isothermal expansion at the high temperature back to the original volume, and ending with isochoric cooling back to the original temperature and pressure. The original Stirling cycle operated under the assumption that the working fluid was constantly an ideal gas, where the equation of state is
Pv=RT, (1)
where P (Pa) is the pressure, v (m3/kg) is the specific volume, T (K) is the absolute temperature, and R (J/kg·K) is the specific gas constant, where
where Mm (kg/M) is the molar mass, and Ru is the universal gas constant (8.314 J/M·K) defined as
Ru=A·κ, (3)
where A is Avogadro's Number 6.02214.1023, and κ is Boltzman's Constant 1.38.10−23 (J/K). The number of moles M is defined as the total number of particles over Avogadro's Number
The novel aspect of this engine is that it does not use an ideal gas as the working fluid, but a real gas that is subjected to condensation and evaporation. The hot temperature of the engine is above the critical temperature Tc (K), whereas the cold temperature of the engine is below the critical temperature, but above the triple point temperature Ttp (K). The working fluid is a saturated gas at the initial, low temperature, high volume stage of the engine cycle. The working fluid partially condenses during the isothermal compression, which ends when the working fluid is a liquid-gas mixture. The working fluid is then heated isochorically to the hot temperature, upon which there is isothermal expansion back to the original stage volume, and where mechanical work is recovered. Finally, the working fluid undergoes isochoric cooling back to a saturated gas at the cool temperature, and the cycle repeats itself.
There famous Van der Waals (VDW) equation of state for a real gas is defined as
where a and b are the gas specifics VDW constants, where
where Pc (Pa), Tc (K), and vc (m3/kg) are the critical pressure, temperature, and specific volume, where the first and second derivative of the pressure over volume are zero,
and at temperatures greater than Tc, gas is the only possible phase of the fluid. If the specific volume is significantly greater than the critical specific volume (v>>vc), then
and thus the VDW equation 5 becomes the ideal gas equation 1.
The critical pressure, temperature, and volume are material-specific, and are determined experimentally. The dimensionless reduced pressure PR, temperature TR, and volume vR are dimensionless ratios of the pressure, temperature, and volume over the critical values
The VDW equation of state can be reduced to its dimensionless state, defined as
and equation 8 can be used for an arbitrary fluid.
One limitation of the VDW equation of state is that they cannot be used to represent the change in the fluid from liquid to gas. Following the VDW equation of state, for a constant temperature (
u=(1−χ)·uliquid+χ·ugas,
h=(1−χ)·hliquid+χ·hgas,
s=(1−χ)·sliquid+χ·sgas,
v=(1−χ)·vliquid+χ·vgas, (9)
where χ is the mass ratio present in the stable liquid-gas state.
This sudden change in the equation of state at the point of phase change from liquid to gas is explained with Maxwell's Construction (
where a (J/kg) is the Helmholtz free energy. Another feature of Maxwell's Construction is that the total work applied
W=∫P·dv, (11)
from the liquid to gas phase equals the value of the VDW equation of state,
∫v
where PVDW (Pa) is the pressure found with the VDW equation of state
and the reduced pressure following the VDW equation of state is simply
The values of PR, vgas, and vliquid are determined numerically, and some reduced examples are given in Table 1.
The reduced pressure-volume diagram for this heat engine has been generated in
Many of the derivations of traditionally used thermodynamic equations are operating under the assumption that the fluid is an ideal gas. An ideal gas was used to derive the efficiency of the Carnot engine, and the entropy increase during heat transfer
as well as the derivation of the specific internal energy
where f is the number of degrees of freedom of the gas particles (f=1 for monatomic gases, f=2 for diatomic gases). Additionally, the assumption of equation 14 is used to predict the total internal energy change
which can be used when the equation of state is known. It can be easily derived from equation 16 that for isothermal ideal gas compression or expansion, there is no change in internal energy or enthalpy δu=δh=0.
For this real gas bounded by the VDW equation of state and Maxwell's Construction, these ideal-gas assumptions are not valid; attempts to apply them result in an imbalance in the internal energy after completion of the internally reversible cycle, especially when there is partial condensation. Due to the kinetic theory of gas, for a monatomic gas (f=1), the pressure P (Pa) of a gas is proportional to the average velocity of each gas particle
where N is the total number of particles, Ekinetic (J) is the average kinetic energy of each gas particle, V (m3) and v (m3/kg) is the volume and specific volume, and u (J/kg) is the specific internal energy. The internal energy U (J), by definition, is related to the average kinetic energy of the gas
U=N·E
kinetic,
and the specific internal energy u (J/kg) is simply the total internal energy U divided by the mass. To derive equation 15 to find the specific internal energy of an ideal gas, equation 17 is plugged into the ideal gas equation 1. As this heat engine does not deal with ideal gases, but with real gases that follow VDW equation of state, the specific internal energy is derived by plugging in the definition of P from equation 5 into equation 17,
The specific heat at a constant volume can thus be easily found as
If one wants to work in terms of dimensionless reduced values, the reduced internal energy, defined as
can be found with a reduced version of equation 18
The reduced specific heat at a constant volume is simply the reduced temperature derivative of equation 21
One observed phenomenon of Maxwell's Construction is the fact that when there are two phases in stable equilibrium, the internal energy initially decreases more for a given reduction in volume, than what would be calculated with equation 18 and 21 for a single phase of the same temperature and specific volume. This phenomenon is demonstrated in
A demonstration was conducted of the condensing Stirling cycle heat engine demonstrated in
The condensing Stirling cycle heat engine is a moving boundary cycle, as seen in a piston-cylinder system. At Stage 1 and Stage 4, the piston is at Bottom Dead Center (BDC), and the reduced volumes are the saturated gas reduced volume (vR=4.1724); whereas Stage 2 and Stage 3, the piston is at Top Dead Center (TDC), and the reduced volume is the equivalent volume when the VDW pressure equals the reduced pressure vR=1.2083. The reduced temperatures at Stage 1 and 2 are low (TR=0.8), whereas at Stage 3 and 4 the reduced temperatures are high (TR=1.1). The reduced pressures PR are found with equation 8, whereas the reduced internal energy uR was found with equation 21. The results of the cycle are in Table 2.
With the change of each stage in this cycle, there is some heat exchanged with the ambient universe, as well as a work applied when there is a moving boundary. The first law of thermodynamics states that energy can not be created or destroyed, and that the change in internal energy equals the heat and work input into the working fluid,
δuij=Qij−Wij, (23)
where δuij (J/kg) is the change in internal energy, Qij (J/kg) is the heat transferred, and Wij (J/kg) is the work applied across the boundary, from stage i to j. As the pressure is constant during the isothermal compression with partial condensation, the reduced work input from Stage 1 to 2 is simply
where reduced work is defined as the work (J/kg) divided by the product of the critical pressure and critical temperature, similar to equation 20
The reduced work applied across the boundary can be found by integrating the VDW pressure, defined in equation 13, plugged into equation 11,
The change in reduced internal energy is found by taking the difference in internal energy at each stage, determined with equation 21. Finally, the value of the heat transfered during each stage is found with the first law of thermodynamics equation 23. These results are tabulated in Table 3. It can be noted that the summation of the heat and work changes is equal to zero, as this is an internally reversible cycle.
What is interesting about this engine cycle is the entropy change of the universe (Table 4) for each phase of the cycle, when entropy is determined with equation 14, which was determined for the ideal Carnot cycle, which assumes an ideal gas equation of state (equation
1). For the isothermal compression and expansion stages, these are easily determined,
The reduced constant specific heat of a constant volume is determined with equation 22
and Cv,41 can be used to find the equivalent entropy change out of the universe during stage 41,
Because of Maxwell's Construction, the entropy change during stage 2-3 had to be determined numerically until the fluid was a single-phase super-heated gas. At each subsequent reduced temperature increment, the saturated liquid and gas reduced volumes are found numerically with Maxwell's Construction, the quality is determined as the volume is held constant during the heating, and then the cumulative internal energy is found as the summation of the liquid and gas reduced internal energies (equation 9). The gas becomes super-heated after TR=0.9930, and then the reduced internal energy increase is found the same way as δs41. The change in entropy is determined by finding the change in internal energy for each temperature increment, and dividing by the reduced temperature. When heating a two-phase fluid from TR=0.8 to a super-critical gas at TR=1.1 at a constant reduced volume of vR=1.2083, the entropy increase is demonstrated in
In ideal heat transfer, where the difference in temperature is kept to a minimum, the summation of the entropy changes out of the known universe
−(δs12+δs23δs34δs41)R=δsnet
3.5526−1.7487−3.5686+1.3844=−0.3803.
is observed to be negative. This phenomenon is observed for real gases; when the specific volume is expanded significantly (reducing the influence of intermolecular attractive forces) to simulate ideal gases, the net-total entropy goes to zero. This phenomenon can be observed by the fact that the heat engine efficiency,
exceeds the ideal-gas Carnot efficiency,
This reduction in ideal-gas entropy is increased due to Maxwell's Construction and mixed-phase condensation; the reduced pressure and work input to compress the gas results in less heat transfer out and thus less entropy generated to the surrounding universe. Of course, this does not encompass the real losses, as heat transfer has to have a temperature gradient, and there is some irreversible loss from friction. Nevertheless, under ideal conditions, the condensing Stirling cycle heat engine demonstrated in
The first law of thermodynamics, described in equation 23, is consistently observed, as everything, including energy, comes from somewhere. The second law of thermodynamics can be ascribed as the fact that heat can only flow from hot to cold, thus increasing the overall disorder during heat transfer. The fact that in a natural process heat transfer only flows from hot to cold has consistently been observed, and is therefore a law of nature.
How can this condensing Stirling cycle heat engine be reconciled with the second law of thermodynamics? This can be explained by the fact that the reduction in overall entropy is observed near the point of condensation, when the intermolecular attractive forces are profound due to the fluid molecules being in close proximity. During the isothermal compression, these intermolecular forces seek to pull the gas molecules together, in effect generating order with less work input by the boundary piston. By removing the intermolecular force component a from the equation of state, which effectively happens when the specific volume is increased and the fluid becomes an ideal gas, there is no reduction in net entropy. For this reason, this condensing Stirling cycle heat engine can reduce the net overall entropy in the universe without violating the first and second law of thermodynamics.
The condensing Stirling cycle heat engine described so far has been a theoretical cycle following a reduced VDW equation of state. The real engine that the inventor claims is a piston-cylinder system with the monatomic fluid argon; the engine cycle can work with any monatomic fluid if sized and designed accordingly. Argon was selected because helium and neon have extremely low critical temperatures of 5 K and 44 K; this cycle utilizes a cold temperature sink colder than the critical temperature. The heavier monatomic fluids of Krypton, Xenon, and Radon have higher critical temperatures of 209 K, 289 K, and 377 K, but their expense and rarity would make them infeasible to be a practical working fluid in this engine. For this reason, argon was selected as the best practical working fluid.
In addition, while the VDW equation of state is often a reasonable representation of molecular behavior, it is still fairly inaccurate when compared to experimental measurements. There are numerous equations of states for different molecules, and they are constantly evolving to better fit new experimental data. For the purpose of this design, the tables in Thermodynamic Properties of Argon from the Triple Point to 1200 K with Pressures to 1000 MPa by Stewart and Jacobsen 1989 (DOI: 10.1063/1.555829) will be used.
To best represent the theoretical condensing Stirling cycle heat engine demonstrated in
At Stage 1 of this cycle, the fluid is a saturated gas at the low temperature of 120 K; according to the referenced tables, the pressure is 1.2139 MPa, and the saturated liquid and gas densities are 29.1230 mol/dm3 and 1.5090 mol/dm3. The densities can easily be converted to the specific volumes, which are 0.8595.10−3 m3/kg and 16.5888.10−3 m3/kg for saturated liquid and gas argon at 120 K. This engine will compress the fluid to a quality χ of 10%, and therefore the volume is
This corresponds to a density of 10.2910 mol/dm3.
The hot, super-critical portion of the engine cycle will occur at a consistent temperature of 166 K, as the specific volume expands isothermally from 2.4325.10−3 m3/kg to the 120 K saturated gas specific volume of 16.5888.10−3 m3/kg. Referencing Table 5, the pressures and densities at 166 K can be determined, and the work output during isothermal expansion is calculated with the numerical summation of equation 11
The work input during isothermal compression with condensation is more easily calculated, as due to Maxwell's Construction, the pressure remains constant,
and thus the net mechanical work out of this engine per unit mass of working fluid for each cycle is -31.6919 kJ/kg.
It is now possible to characterize the pressure, temperature, specific volume, internal energy, and enthalpy of the condensing Stirling cycle heat engine with argon. The pressures are determined from the referenced tables; the pressure of condensation for T=120K of P1−P2−1.2139 MPa, and the interpolated super-critical pressures of P3−6.9007 MPa and P4=1.8689 MPa. The temperatures are by design, with T1=T2=120 K and T3−T4−166 K. The specific volumes are designed by the piston and cylinder, with the Top Dead Center volume of v2=v3=2.4325.10−3 (m3/kg), and the Bottom Dead Center volume of v1=v4=16.5888.10−3 (m3/kg). The internal energy u and enthalpy h are determined from the kinetic gas theory (equation 18), which for a monatomic fluid such as argon,
and thus the results can be found in Table 6.
Next, the first law of thermodynamics is used to determine the heat input and output at each stage. The work applied during isothermal compression and expansion has been determined, and the heat input is simply the summation of the change in internal energy minus the work applied by the fluid (equation 23)
Q
ij
=δu
ij
W
ij,
and thus using the internal energies in Table 6, the net heat inputs and outputs can be determined and included in Table 7. The summation of the heat and work in Table 7 is zero,
E
ij(Q+W)ij=−42.961+20.7496+70.1532−16.2498+17.1844−48.8764=0,
showing that this cycle is an internally reversible cycle that complies with the first law of thermodynamics.
Finally, the efficiency η of this engine
can be determined from the values in Table 7
If there is perfect regeneration of the heat output from the isochoric cooling (41) into the heat input from the isochoric heating (23), the efficiency is improved
Remarkably, both of these values are greater than the Carnot efficiency
and this efficiency that exceeds the Carnot efficiency is evidence that the intermolecular attractive forces are reducing the disorder of the molecules during the isothermal compression with condensation.
An example of this engine cycle being practically implemented is represented in
The pressure vessel volume can expand and contract by an isentropic piston; this piston recovers mechanical energy during expansion and inputs mechanical energy during compression. During the isochoric heating of the argon, the volume of the surrounding ideal gas will compress slowly so that the ideal gas will heat up slowly, and the temperature difference during heat transfer will be minimized, reducing the overall entropy of heat transfer of the engine cycle. A mechanical work input will be used during this compression; this work will be recovered when the piston expands while the argon is undergoing isochoric cooling.
For the practical implementation of the argon engine described, 100 kg of air will be used as the surrounding heat transfer fluid; air has a specific heat ratio of 1.4 and a gas constant of 287 J/kg·K. The pressure vessel can be of an arbitrary volume; for the given mass, decreasing the volume will result in an increase in pressures, but not affecting the work inputs and outputs. For 100 kg of air, 0.7575 kg of argon, and a temperature range between 120 K and 166 K, the ideal gas volume decreases by a factor of 3.892, and the work input for each compression stroke would be 3.3162 megajoules. This compression will serve to raise the temperature from 120 K to 166 K, and allow for sufficient heat loss to heat the argon simultaneously. This energy is recovered during the argon cooling stage, where the piston expands and recovers this energy. By using this method, reducing the temperature difference significantly during heat transfer, the ideal engine efficiency (excluding friction and irreversible losses) can even exceed the minimum predicted 34.86% and get closer to the 42.45% possible with this engine cycle.
The pistons are synchronized, so that the ideal gas piston is fixed when the argon engine piston is in motion, and vice versa. During the isothermal compression of the argon, the heat input into the ideal gas is removed by the heat exchanger fluid (at 120 K), and the ideal gas piston remains fixed at Bottom Dead Center. During the isochoric heating, the heat exchanger fluid ceases to flow, the argon piston is held fixed, and the ideal gas piston compresses the gas to Top Dead Center. For the isothermal expansion of the argon, the ideal gas piston remains fixed at Top Dead Center, and the heat exchanger fluid flowing provides a source of heat at 166 K. Finally, the argon gas is held fixed by the piston, while the gas cools to saturation; during this time the ideal gas piston is expanding back to Bottom Dead Center and recovering mechanical energy.
To synchronize these two pistons, each piston is controlled by a gear, which is operated by a mutilated gear. These two mutilated gears have teeth on half of the circumference, divided into four 90° sections of gear-teeth and no-gear-teeth. These gears are connected to a cam-shaft, that operates a brake that holds the piston fixed in place during the no-gear-teeth angles; without this feature, the pressurized ideal and argon gas will expand against the piston prematurely.
This cycle can run at varying speed so long as it is slow enough to ensure sufficient heat transfer at each stage. The greater and more consistent the heat transfer, the less entropy will generate and thus the efficiency of the heat engine will increase. With sufficient heat transfer, and a temperature source and sink of 120 K and 166 K, heat engine efficiencies in excess of the 27.71% Carnot efficiency can be achieved by taking advantage of the attractive intermolecular forces during condensation.