Fact-checked by Grok 2 weeks ago

Ericsson cycle

The Ericsson cycle is a consisting of two isothermal processes ( and ) and two isobaric processes (regeneration), designed for use in regenerative heat engines to achieve high efficiency through the reuse of . It operates as a closed cycle, typically with air or another gas as the , where is added and rejected at constant via a regenerator that transfers between the working fluid streams, minimizing external input requirements. Named after Swedish-American inventor , the cycle emerged from his efforts in the early to develop efficient alternatives to steam engines using "caloric" (hot air) principles. In 1833, Ericsson patented a caloric engine incorporating an innovative regenerator—a using tubes or wire gauze to recover heat from exhaust gases—marking the first practical application of regeneration in heat engines. His designs, including experimental engines built between 1840 and 1850, aimed to exploit constant-pressure heat exchange, though challenges like material oxidation at high temperatures (up to 450°F) and mechanical complexity limited commercial success during his lifetime. Theoretical analysis by engineers like W.J.M. Rankine in the 1850s further refined the cycle's principles, distinguishing it from contemporary cycles like the (which uses constant-volume regeneration). In its ideal form, the Ericsson cycle achieves the Carnot efficiency—the theoretical maximum for any operating between two temperatures—given by η = 1 - (T_C / T_H), where T_C and T_H are the absolute temperatures of the cold and hot reservoirs, respectively, assuming perfect regeneration and no irreversibilities. This efficiency surpasses that of the (a common cycle with adiabatic processes) under similar conditions; for example, at a pressure ratio where the Brayton efficiency is 56.5%, the Ericsson cycle reaches 63.7% with near-isothermal and factors of 1.043. On P-V and T-S diagrams, the cycle appears as a , with lines for isothermal processes and vertical lines for isobaric regeneration, highlighting its reversible nature. While primarily theoretical due to practical difficulties in achieving true isothermality and perfect regeneration, the Ericsson cycle influences modern applications in advanced gas turbines, cryogenic refrigeration, and solar thermal engines, where multi-stage intercooling, reheating, and regeneration enhance performance. Its principles also extend to reverse-cycle heat pumps using natural refrigerants like air, offering environmental advantages over traditional vapor-compression systems.

Fundamentals of the Ericsson Cycle

Definition and Processes

The Ericsson cycle is a reversible comprising two isothermal processes and two isobaric regeneration processes, designed for use in heat engines. It is named after the inventor , who developed early hot air engines incorporating regenerative principles. The cycle enables efficient heat-to-work conversion by approximating constant-temperature heat addition and rejection, potentially achieving performance near that of the . The four core processes of the ideal Ericsson cycle are as follows: Process 1-2 involves reversible isothermal compression of the at the low temperature T_L, during which mechanical work is input and heat is rejected to the cold surroundings to maintain constant temperature. Process 2-3 is reversible isobaric heat addition via the regenerator, where the fluid absorbs stored heat internally to raise its temperature to the high value T_H at constant pressure. Process 3-4 entails reversible isothermal expansion at T_H, with mechanical work output and heat absorption from the hot source to sustain the temperature. Finally, process 4-1 is reversible isobaric heat rejection to the regenerator, cooling the fluid back to T_L at constant pressure while transferring heat for later reuse. This cycle operates in a closed configuration, recirculating a gaseous such as air or another , with heat supplied externally through outside the working fluid path. In the pressure-volume (P-V) diagram, the isothermal processes trace curves (PV = constant), while the isobaric processes appear as horizontal lines (constant P), forming a closed loop that highlights the regenerative heat exchange. The regenerator, briefly, facilitates the isobaric processes by enabling near-perfect internal heat recovery, minimizing external heat requirements beyond the isothermal steps.

Thermodynamic Analysis

The Ericsson cycle operates under the assumption of an as the and perfect regeneration, where the regenerator transfers between the isobaric processes without losses, ensuring that the heat added during the constant-pressure heating equals the heat rejected during constant-pressure cooling, both given by Q_{\text{regen}} = C_p (T_H - T_L). Heat is supplied externally only during the isothermal expansion at the high temperature T_H, calculated as Q_{\text{in}} = R T_H \ln \left( \frac{V_4}{V_3} \right), where V_4 > V_3 is the volume ratio during expansion and R is the . Heat is rejected externally only during the isothermal compression at the low temperature T_L, with Q_{\text{out}} = R T_L \ln \left( \frac{V_1}{V_2} \right), where V_1 > V_2 and the magnitude |Q_{\text{out}}| represents the heat leaving the system. The isobaric regeneration processes contribute zero net work, as the work done during constant-pressure expansion equals the work absorbed during constant-pressure compression. The net work output is thus W_{\text{net}} = Q_{\text{in}} - |Q_{\text{out}}|. In the ideal cycle, the pressure ratio across the isothermals ensures the volume equals the , \frac{V_4}{V_3} = \frac{V_1}{V_2} = r > 1, yielding W_{\text{net}} = [R](/page/R) (T_H - T_L) \ln r. To derive the , start with the definition \eta = \frac{W_{\text{net}}}{Q_{\text{in}}} = 1 - \frac{|Q_{\text{out}}|}{Q_{\text{in}}}. Substituting the expressions gives \eta = 1 - \frac{[R](/page/R) T_L \ln r}{[R](/page/R) T_H \ln r} = 1 - \frac{T_L}{T_H}. This matches the Carnot efficiency for the same limits, as the reversibility of all processes and perfect regeneration eliminate irreversible losses, allowing the cycle to approach the theoretical maximum. In the temperature-entropy (T-S) diagram, the cycle appears as two horizontal isothermal lines—at T_H for (entropy increasing) and at T_L for (entropy decreasing)—connected by two sloped isobaric lines representing the regeneration processes, where changes as \Delta S = C_p \ln \left( \frac{T_H}{T_L} \right) but shifted due to differing s. In practice, real Ericsson cycles deviate from this ideal due to imperfect regeneration (finite rates leading to temperature differences), pressure drops in the regenerator, and non-ideal gas behavior, reducing below the Carnot limit.

Comparisons with Other Thermodynamic Cycles

Similarities and Differences with Carnot and Stirling Cycles

The Ericsson cycle shares fundamental similarities with the Carnot cycle in its theoretical reversibility and maximum achievable efficiency, both operating between two thermal reservoirs at temperatures T_H (high) and T_L (low) to yield an efficiency of \eta = 1 - \frac{T_L}{T_H} under ideal conditions with perfect regeneration. Like the Carnot cycle, the Ericsson cycle consists of reversible processes that minimize entropy generation, ensuring no net entropy increase over a complete cycle. However, the Ericsson cycle replaces the Carnot cycle's two adiabatic (isentropic) processes with two isobaric regeneration steps, paired with isothermal compression and expansion, which facilitates practical external combustion while approximating the same efficiency bounds. This substitution allows the Ericsson cycle to bridge the Carnot ideal—unattainable in practice due to the need for infinite heat transfer surfaces during adiabatic steps—with more feasible implementations, as the isobaric regeneration enables heat recovery without the constraints of perfect insulation. In comparison to the Stirling cycle, the Ericsson cycle exhibits strong parallels as both are reversible, external combustion cycles featuring isothermal compression and expansion processes, along with regeneration to achieve near-Carnot efficiency by recycling heat internally and eliminating entropy production from imperfect heat transfer. Both cycles rely on a regenerator to store and release heat during the non-isothermal steps, enabling the working fluid to undergo quasi-isothermal heat addition and rejection, which theoretically matches the Carnot efficiency for the same temperature limits. The primary distinction lies in the regeneration process: the Ericsson cycle employs isobaric (constant-pressure) regeneration, whereas the Stirling cycle uses isochoric (constant-volume) regeneration. This constant-pressure approach in the Ericsson cycle reduces dead volume associated with displacer mechanisms in Stirling engines and supports continuous fluid flow, making it particularly suitable for gaseous working fluids in steady-flow configurations. Overall, the Ericsson cycle positions itself as a practical extension of the Carnot ideal, akin to the Stirling cycle but optimized for pressure-based heat exchange that enhances applicability in gas turbine-like systems.

Comparison with Brayton, Otto, and Diesel Cycles

The Ericsson cycle shares isobaric heat addition and rejection processes with the Brayton cycle, commonly used in gas turbines, but differs fundamentally in its compression and expansion stages: the Ericsson employs isothermal processes, while the Brayton uses adiabatic ones. This isothermal approach in the Ericsson cycle significantly reduces compression work requirements compared to the Brayton cycle (to about 46% for a pressure ratio of 8), leading to higher net work output—up to 180% greater in specific implementations at a pressure ratio of 8—and thermal efficiencies closer to the Carnot limit. Without regeneration, the Ericsson cycle closely resembles the closed Brayton cycle; however, its incorporation of regeneration recovers a substantial portion of exhaust heat, enabling efficiencies of 69–74% under conditions where the Brayton achieves 58–63%. In contrast to the , which models spark-ignition internal engines with constant-volume heat addition, the Ericsson cycle operates via external and isothermal processes supported by regeneration, avoiding the irreversible losses associated with rapid constant-volume . This design yields a higher theoretical potential for the Ericsson—approaching Carnot values—compared to the Otto's typical range of 30–35% in practical engines with ratios of 8–10. However, the Ericsson's external and lower make it less suitable for high-speed mobile applications where the Otto excels. The Ericsson cycle also outperforms the , the ideal model for compression-ignition engines featuring constant-pressure heat addition, by eliminating inefficiencies from high-temperature internal through its isothermal expansion and external heat supply. While engines achieve practical efficiencies of 40–50% with compression ratios of 12–24, the Ericsson's regenerative isothermal processes enable superior performance, particularly with low-grade heat sources, as heat addition occurs externally without limitations. This positions the Ericsson as more versatile for stationary or heat-recovery applications, though its complexity contrasts with the 's robustness in heavy-duty uses.
CycleKey ProcessesTypical Efficiency (%)Example Net Work (kJ/kg at r_p=8)
Isothermal comp/exp, isobaric regen69–74 (theoretical)369
BraytonAdiabatic comp/exp, isobaric heat58–63 (regenerative)131
OttoIsentropic comp/exp, const-vol heat30–35 (practical)N/A (closed cycle, variable)
Isentropic comp, const-press heat/exp40–50 (practical)N/A (closed cycle, variable)

Key Components

The Regenerator

The regenerator serves as a critical counterflow in the Ericsson cycle, capturing and storing released by the during the isobaric cooling process (from state 4 to 1) and subsequently releasing it to preheat the fluid during the isobaric heating process (from state 2 to 3). This internal heat recovery minimizes the need for external heat addition, allowing the cycle to approach the of a operating between the same temperature limits in ideal conditions. With perfect regeneration, the effectiveness approaches 100%, meaning the temperature of the fluid exiting the regenerator toward the hot source equals the temperature after expansion, and vice versa for the cold side. Designs for the regenerator vary to optimize while accommodating the cycle's continuous flow. Fixed-matrix regenerators employ porous media, such as packed beds or screens, where the working fluid passes through a stationary structure to heat. Rotary regenerators, by , use a rotating or drum filled with heat-storing material to provide continuous counterflow between hot and cold streams without interrupting the cycle. John Ericsson's original implementations featured mixed-flow configurations using coiled wires or layered plates to facilitate storage and transfer between descending hot air and ascending cold air streams. The heat transfer in the regenerator can be quantified by the energy recovered, expressed as Q_{\text{regen}} = \int C_p \, dT across the temperature range of the processes, where C_p is the specific heat at constant pressure. Effectiveness \varepsilon measures performance relative to an ideal case, defined as \varepsilon = \frac{T_{\text{in,hot}} - T_{\text{out,hot}}}{T_{\text{in,hot}} - T_{\text{in,cold}}} for the hot stream (assuming equal heat capacities), indicating the fraction of available temperature difference utilized. In practice, imperfections such as finite rates and pressure drops lead to incomplete recovery, introducing axial temperature gradients that can reduce overall cycle thermal efficiency by 10-20% compared to the ideal. Although the regenerator concept was first credited to Robert Stirling in for his hot air engine, Ericsson popularized and refined it in the mid-19th century through extensive experimentation and application in large-scale engines. Materials selection emphasizes high and durability under cyclic , with options like for metallic matrices or ceramics for operation at high temperatures up to 1000°C, ensuring minimal degradation over repeated cycles.

Ericsson Engine Mechanics

The Ericsson engine features a basic mechanical design centered on reciprocating pistons that execute the isothermal and processes, supported by extensive heat exchangers to enable during these phases. A dedicated blower or manages the isobaric addition and rejection, ensuring continuous fluid flow. Configurations exist in both open and closed cycle variants, with the closed cycle recirculating the internally and the open cycle drawing in and exhausting atmospheric air. In operation, is supplied externally through of or concentration, warming the —typically air or —which then circulates sequentially through an integrated regenerator for , expansion cylinders, and units. The fluid's expansion drives the , converting into mechanical work without direct internal . Notable early configurations include John Ericsson's 1833 closed-cycle hot air engine, equipped with a 14-inch diameter working and a separate water-jacketed connected via an organ-pipe regenerator, delivering approximately 5 horsepower through -driven motion. In contrast, the 1851 open-cycle design utilized massive double- assemblies with 14-foot diameter working and 11.5-foot supply , each with a 6-foot , where compressed air from the supply fed the working before atmospheric exhaust, powering a for output. Key mechanical challenges arise in approximating isothermal conditions, which demand slow reciprocation or near-frictionless to permit adequate heat exchange time, often limiting typical operating speeds to 10-60 RPM in early historical prototypes and up to 100-200 RPM in later small-scale and commercial designs. Unlike internal combustion engines, these designs eliminate valves for delivery due to external heating, instead employing actuated valves for and timing to and .

Historical Development

John Ericsson's Inventions

(1803–1889) was a Swedish-born and inventor who emigrated to the in 1839, where he pursued a wide range of mechanical innovations, including early contributions to and ironclad warships. His work on hot air engines stemmed from a desire to create safer, more efficient alternatives to steam power, driven by concerns over explosions. In 1833, while based in , developed and patented his first closed-cycle hot air engine, known as the caloric engine, which incorporated a novel heat-recovery device he termed the regenerator. This engine operated by compressing air, heating it externally, and expanding it to produce mechanical work, marking an early conceptualization of what would later be termed the Ericsson cycle. Ericsson's inventions evolved through several key milestones, reflecting iterative improvements in design and efficiency. He secured multiple for his engines, including British Patent No. 6409 in for the initial closed-cycle caloric engine and U.S. Patent No. 8,481 in 1851 for an open-cycle variant that compressed atmospheric air directly into the system. The regenerator, consisting of coiled tubes to transfer from exhaust to incoming air, was integral from the outset in , enabling partial heat recovery and distinguishing his design from contemporary steam engines; it later evolved to in 1840s prototypes for better performance. By the 1850s, Ericsson had refined these concepts amid growing thermodynamic insights, though his adherence to —a now-obsolete view positing as a conserved, indestructible —shaped his emphasis on heat as a perpetual motive power without conversion losses. During the 1830s, Ericsson conducted extensive experiments in , constructing and testing prototypes that demonstrated the caloric engine's potential for steady, explosion-free operation. These efforts built on his earlier designs from the but incorporated advanced features like double-acting cylinders for continuous power. Collaborations with contemporaries, such as engineer John Braithwaite, facilitated demonstrations and refinements, though largely worked independently from leading thermodynamic theorists like Sadi Carnot. Despite these advances, early engines suffered from material limitations, operating at relatively low hot-side temperatures around 230°C (450°F) due to the fragility of available metals and seals under sustained heat, which constrained efficiency and power output. Over his career, Ericsson constructed several prototypes of caloric engines, including at least eight major experimental units in from 1840 to 1850, along with earlier models from the 1820s and 1830s, ranging from small laboratory models to larger experimental units. These efforts, including variations in cylinder configurations and regenerator materials like , influenced subsequent hot air engine developments, particularly the , by popularizing the regenerator principle and sparking debates on heat recovery in external combustion systems.

Notable Implementations: The Caloric Ship

The Caloric Ship Ericsson, launched in , represented a bold attempt to apply the Ericsson cycle to on a grand scale. This wooden displaced approximately 2,000 tons and measured about 260 feet in length, with a of 40 feet. It was powered by four massive open-cycle air engines, each featuring a working cylinder 14 feet in diameter and a 6-foot , connected to supply cylinders of 11.5 feet in diameter; these delivered a total output of around 300 horsepower. The engines operated by drawing in atmospheric air, compressing it, heating it in spherical furnaces burning anthracite coal for isothermal expansion—achieved through exposure of the finned working cylinders to the stream—and then passing the exhaust through regenerators filled with coiled wire netting disks to recover . Unlike contemporary steamships, the required no for boilers, relying solely on air as the , which promised greater safety and reduced logistical demands for long voyages. Performance during initial trials underscored both the cycle's efficiencies and its practical challenges. On its first major excursion in early 1853, the ship completed a journey from to via the , running flawlessly for 73 hours of effective sailing time at an average speed of about 6 knots, while consuming only around 6 tons of per day—far less than the 50+ tons required by comparable steamers like the Baltic. The trial demonstrated reliable operation without the explosion risks of steam boilers and highlighted the engine's potential for crossings, as envisioned the vessel sustaining extended ocean voyages with minimal fuel and no freshwater needs for cooling. However, the low limited top speeds to 8 knots in smooth conditions, even with paddle wheels turning at 9–12 RPM, revealing the cycle's struggle to match steam's rapid and overall . Despite the promising demonstration, the ship's 1853 ocean-oriented trial exposed critical limitations, ultimately dooming the project. While the engines performed steadily, the vessel's underpowered propulsion proved inadequate against adverse weather and currents, preventing viable transatlantic speeds and leading to skepticism about scalability for naval use. In 1854, during a subsequent bay trial, the Ericsson sank in a storm off after open ports flooded, though the machinery was later salvaged and repurposed for a conversion. Constructed at a cost of approximately $500,000, this implementation influenced early naval engineering by validating the regenerator's heat-recovery efficacy and the Ericsson cycle's fuel economy, yet it was ultimately overshadowed by steam engines' superior and established .

Modern Applications and Future Potential

Challenges and Limitations

One of the primary technical challenges in implementing the Ericsson cycle lies in achieving true and processes, which theoretically require infinite time or perfect to maintain constant temperature during heat addition and rejection. In practice, finite rates and limitations lead to non-isothermal behavior, introducing irreversibilities that degrade performance. Additionally, the regenerator, essential for recovering between processes, increases system complexity and cost due to its large size and the need for high-effectiveness heat exchange surfaces. Real-world efficiencies of Ericsson cycle systems are typically below 40%, falling short of the theoretical Carnot limit due to factors such as pressure drops in the regenerator and imperfect regeneration. These engines also exhibit lower compared to , limiting their suitability for applications requiring compact designs. Modern finite-time thermodynamic analyses further reveal efficiency penalties of 20-30% arising from finite-speed operations, which prevent ideal quasi-static conditions. Material constraints pose another hurdle, particularly the need for components durable at high temperatures exceeding 800°C during addition, where thermal stresses and can compromise . Piston-cylinder assemblies in reciprocating implementations are susceptible to leakage, further reducing . Economically, the high initial costs of advanced heat exchangers and regenerators, combined with slower startup times relative to gas turbines, have historically deterred widespread adoption. A notable historical illustration of these limitations is the 1852 caloric ship , powered by an engine, which achieved only about 8 knots during trials—underpowered compared to contemporary steamships exceeding 10 knots—despite smooth operation, ultimately contributing to the cycle's early commercial setbacks.

Recent Advancements and

Recent has advanced the Ericsson cycle's application in micro-combined and (micro-CHP) systems, particularly through innovative free liquid engine designs. In 2023, experimental testing of a novel free liquid Ericsson engine (FLPEE) configuration demonstrated stable operation and net work output for micro-CHP applications in conditions. This design integrates a free liquid system with a micro-expander, enabling efficient recovery and suitable for residential or small-scale distributed systems. Turbine-based implementations of the Ericsson cycle have seen notable progress, especially in enhancing isothermal processes for improved efficiency. A 2022 U.S. patent (US11530644B1) describes an turbine developed by U.S. researchers, incorporating gas-liquid mixtures to facilitate near-isothermal and . This configuration uses a centrifugal gas , heat exchangers, and a , potentially offering higher thermal efficiencies compared to traditional gas by approximating the cycle's ideal isothermal steps more closely. At the nanoscale, stochastic analyses have explored the Ericsson cycle's feasibility for ultra-small engines. A 2024 study examined an Ericsson nano-engine using a charged quantum oscillator as the working substance in a coupled to a bath, deriving bounds from fluctuations. This approach highlights the cycle's potential at quantum scales. Emerging CO2-based systems are leveraging the Ericsson cycle for integration and carbon management. Research published in 2025 proposes a gas-CO2 hybrid combined cycle power generation system approximating the Ericsson cycle, incorporating utilization and carbon capture. The design achieves a power generation of approximately 30%, facilitating the storage and conversion of sources like or into dispatchable power while mitigating CO2 emissions. Additional developments include rotary variants and pairings. A UKRI-funded project from 2016-2017 developed a revolutionary rotary Ericsson for residential heating, utilizing fluids and a rotary to closely follow the cycle's for high coefficient-of-performance values. Pairings with concentrators are also gaining traction for off-grid , where Ericsson engines convert concentrated into , supporting remote applications with efficiencies enhanced by advanced regenerators. U.S. Navy initiatives from the through the have focused on high-efficiency Ericsson cycle turbines for naval and power. Simulations of these designs, incorporating such as ceramics, indicate potential efficiencies exceeding 60%, addressing challenges like high-temperature operation while enabling compact, fuel-flexible systems.

References

  1. [1]
    None
    ### Summary of Ericsson Cycle from the Document
  2. [2]
    Ericsson's Caloric Engine 1833 - Hot Air Engines
    His own application of a " regenerator " was first made in 1833, when he invented and patented in England, France, the United States, and other countries a " ...
  3. [3]
    The Regenerator Principle in the Stirling and Ericsson Hot Air Engines
    Jan 5, 2009 · During the first half of the nineteenth century the Stirling brothers and John Ericsson made significant attempts to design hot air engines ...
  4. [4]
    Ericsson Cycle - Theory and Efficiency - Nuclear Power
    An ideal Ericsson cycle consists of isothermal compression and expansion processes, combined with isobaric heat regeneration between them.
  5. [5]
    [PDF] 4.3 Ericsson Cycle: - NPTEL Archive
    The processes are: Process 1-2: Reversible isothermal compression. Process 2-3: Constant pressure heat addition. Process 3-4: Reversible isothermal expansion.
  6. [6]
    [PDF] LECTURE NOTES ON INTERMEDIATE THERMODYNAMICS
    Mar 28, 2025 · of isentropic processes in the Brayton cycle, one would obtain the Ericsson cycle. Diagrams for P − v and T − s for the Ericsson cycle are ...
  7. [7]
    Comparison of the Net Work Output between Stirling and Ericsson ...
    Stirling and Ericsson cycles are ideal thermodynamic cycles for external heat engines with regeneration, and both are considered to have the Carnot efficiency ...
  8. [8]
    [PDF] A Comprehensive Review on Stirling Engines - DergiPark
    Ericsson cycle efficiency is on par with Carnot and Stirling cycles with regenerator. However, the amount of heat transferred and converted into work at the ...
  9. [9]
    None
    ### Summary of Ericsson, Carnot, and Stirling Cycles
  10. [10]
    [PDF] Thermodynamic cycle
    A thermodynamic cycle consists of a series of thermodynamic processes transferring heat and work, while varying pressure, temperature, and other state ...
  11. [11]
    None
    ### Summary of Stirling, Ericsson, and Carnot Cycles
  12. [12]
    Parametric analysis on the performance of a revolutionary rotary ...
    Dec 21, 2018 · The Ericsson cycle has significant advantages over the Stirling cycle. In particular, it avoids the 'dead' volume of the Stirling system by ...
  13. [13]
    Brayton Cycle vs Ericsson Cycle - Nuclear Power
    The second Ericsson cycle is similar to the Brayton cycle but uses external heat and incorporates the multiple uses of an inter-cooling and reheat.
  14. [14]
    Ericsson cycle – Knowledge and References - Taylor & Francis
    This cycle combines the features of regeneration, reheating, and intercooling along with steam/fuel reformation. In the ideal Ericsson cycle, both isothermal ...
  15. [15]
  16. [16]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Dec 15, 2021 · (1) Typical values for the compression ratio for a spark ignition IC engine are r = 8 to 10 with engine thermal efficiencies of 7 = 30 to 35% ( ...
  17. [17]
    Analysis of an irreversible Ericsson engine with a realistic regenerator
    The effects of internal irreversibilities on the power output and thermal efficiency of the cycle have been determined using the finite-time thermodynamics.
  18. [18]
    [PDF] An efficient optimization of an irreversible Ericsson refrigeration ...
    Jun 20, 2022 · The efficiency of the Ericsson cycle is in-line with Carnot cycle and Sterling cycle and hence it can be adopted for low temperature production.Missing: derivation | Show results with:derivation
  19. [19]
    [PDF] Heat exchanger design for hot air ericsson-brayton piston engine
    Regenerator functions as heater and cooler while in Ericssonovho engine cooler and heater are separated. The air is compressed in the compressor, flows through ...
  20. [20]
    The Ericsson Regenerator - Hot Air Engines
    An inquiry on hot air engines with a detailled presentation of most popular engines: Stirling engines, Ericsson caloric engine, Rider, Lehmann, Laubereau, ...Missing: design coiled plates mixed- flow
  21. [21]
  22. [22]
    Composite-Matrix Regenerators for Stirling Engines - Tech Briefs
    Aug 31, 2021 · Regenerator matrices are subjected to hot, oscillating gas flows and high temperature gradients. ... steel wool, steel felt, stacked ...Missing: Ericsson cycle
  23. [23]
    [PDF] Dynamic simulation of an original Joule cycle liquid pistons hot air ...
    Jul 21, 2022 · An Ericsson engine is an external heat supply engine working according to a Joule thermodynamic cycle. It is based on reciprocating piston- ...
  24. [24]
    Dynamic modeling of a small open Joule cycle reciprocating ...
    Aug 28, 2013 · The dynamic model of an open Joule cycle reciprocating Ericsson engine is developed. The simulation results predict a stable operation of ...
  25. [25]
    Ericsson's Caloric Engine of 1851
    When the pistons have reached their highest point, a valve is opened by the machine, which establishes a free communication between the compressed and heated ...
  26. [26]
    Ericsson I (Destroyer No. 56) - Naval History and Heritage Command
    Jun 4, 2019 · Ericsson saw caloric power as a cleaner, safer, and more efficient means of propulsion than steam. In 1852, he launched Ericsson as the first ...
  27. [27]
    Ericsson's hot air engines patents
    1833, Ericsson patent, Patent not found, External combustion, closed cycle ; 1851, Ericsson patent #8,481, The famous Ericsson Caloric Engine used on the ...
  28. [28]
    John Ericsson - Hot Air Engines
    The story of Captain John Ericsson's attempt to supplant the marine steam engine by his caloric or hot air engine.The first Hot Air Engines · The Ericsson Caloric Engine of...
  29. [29]
    John Ericsson - Spartacus Educational
    Ericsson lived in London where he formed a partnership with John Braithwaite. In 1829 the two men produced Novelty, one of the entries for the Rainhill Trials.Missing: thermodynamics | Show results with:thermodynamics<|separator|>
  30. [30]
    Seeing The Light - John Ericssson - Maritime design genius
    Dec 2, 2007 · John Ericsson was born in Vermland, Sweden, and immigrated to England in 1826 with a working model of an engine design, which he called his ...
  31. [31]
    john ericsson (1803-1889) - Stirling Engines
    Ericsson settled in New York where he built, between 1840 to 1850, eight experimental engines using wire gauze regenerators. These engines worked on an open ...Missing: collaboration thermodynamics
  32. [32]
    Ericsson's Caloric Engine of 1852
    In view of the fact that the engines consisted of four working-cylinders of 168 inches diameter, 6 feet and four air-compressing cylinders of 137 inches ...Missing: prototypes | Show results with:prototypes
  33. [33]
    Reports and Retorts on the Status of the Air-Engine as Suc
    John Ericsson's caloric engine of 1833 incorporating the controversial 'regenerator' (seen here below the two cylinders). From William Conant Church, The ...
  34. [34]
  35. [35]
    The Hot Air Ship Erricsson | Scientific American
    In the engine room, instead of two close cylinders as in the steam engine, there are four large under cylinders of 22,300 square inches piston area each. These ...Missing: horsepower | Show results with:horsepower
  36. [36]
    Scientific Papers—Dr. Lardner and Newspapers
    Here we see a steamship moving at the rate of 16 miles per hour (the Baltic passing the Ericsson), and a new ship, driven by hot air, moving at the rate of six ...
  37. [37]
    The Story of Ericsson's Caloric Ship - Part 2 - Hot Air Engines
    The Ericsson was finally made ready for another trial, and took a trip down New York Bay on March 15, 1854. A second trip followed on April 27th, and the next ...
  38. [38]
    [PDF] Liquid-Flooded Ericsson Cycle Cooler: Part 1 - Purdue e-Pubs
    A novel implementation of a gas Ericsson cycle heat pump is presented. The concept uses liquid flooding of the compressor and expander to approach ...
  39. [39]
    [PDF] Ericsson Cycle Gas Turbine Powerplants - RAND
    Such plants have a thermodynamic feature of fundamental importance: They can very closely approximate a constant temperature expansion. The theoretical ...
  40. [40]
    [PDF] First Experimental Results of a New Free Liquid Piston Ericsson ...
    Jun 30, 2023 · isentropic dead volume recompression. As mentioned earlier, the ... or an Ericsson cycle, Energy (2013), vol. 49, no. 1, pp. 229–239 ...
  41. [41]
    US11530644B1 - Ericsson cycle turbine engine - Google Patents
    An Ericsson cycle turbine engine. The Ericsson cycle turbine may comprise: a centrifugal gas compressor, shaft, at least one heat exchanger, and a reaction ...Missing: 11530644B1 | Show results with:11530644B1
  42. [42]
    Stochastic thermodynamics and the Ericsson nano engine - arXiv
    Nov 5, 2024 · In this work, we study an Ericsson cycle whose working substance is a charged (quantum) oscillator in a magnetic field that is coupled to a heat ...
  43. [43]
    Research on power generation and waste heat utilization ...
    Apr 15, 2025 · This paper proposes a gas-CO 2 combined cycle power generation system based on the principle of the Ericsson cycle.
  44. [44]
    A Revolutionary Rotary Ericsson Heat Pump/Engine - GtR
    The proposed project is aimed to investigate an innovative heat pump utilizing the Ericsson thermodynamic cycle in early concept stage.Missing: 2020s | Show results with:2020s
  45. [45]
    Ericsson cycle turbine engine - TechLink
    A Navy researcher has developed a turbine engine that operates on the Ericsson cycle, which consists of two isothermal and two constant pressure processes.