Fact-checked by Grok 2 weeks ago

Heat engine

A is a that extracts from a high-temperature source, converts a portion of it into mechanical work, and rejects the remainder to a low-temperature , operating through a repeating . These engines are fundamental to converting into useful work in applications ranging from automotive internal engines to large-scale power generation systems. The operation of a heat engine relies on the second law of , which states that it is impossible to convert all heat from a reservoir into work without some being expelled, limiting the of the process. Key components include a hot reservoir (e.g., or ), a cold reservoir (e.g., atmosphere or cooling water), and a working substance (e.g., gas or ) that undergoes cyclic changes in , , and temperature to produce net work. The \eta of a heat engine is defined as the ratio of work output W to heat input Q_h, given by \eta = \frac{W}{Q_h} = 1 - \frac{Q_c}{Q_h}, where Q_c is the heat rejected to the cold reservoir; real engines achieve efficiencies typically between 20% and 40%, far below theoretical maxima. The theoretical foundation for heat engine efficiency was established by Sadi Carnot in 1824 through his analysis of an idealized reversible cycle, known as the , which operates via two isothermal and two adiabatic processes and sets the upper limit for efficiency as \eta = 1 - \frac{T_c}{T_h}, where T_h and T_c are the absolute temperatures of the hot and cold reservoirs, respectively. Common types include external combustion engines like steam turbines, which powered the , and internal combustion engines such as the in gasoline vehicles or the in heavy machinery. Despite advances, all heat engines are constrained by in irreversible processes, underscoring the second law's role in dictating fundamental limits on energy conversion.

Introduction

Definition and Scope

A heat engine is a device that converts thermal energy extracted from a hot reservoir into mechanical work, while expelling the remaining unusable energy as waste heat to a cold reservoir. This process typically involves a working fluid, such as a gas or vapor, that undergoes changes in state to facilitate the energy transfer./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/04%3A_The_Second_Law_of_Thermodynamics/4.03%3A_Heat_Engines) The scope of heat engines is confined to systems that operate through cyclic thermodynamic processes, where the working fluid returns to its initial state after each cycle, ensuring continuous operation. These processes are fundamentally governed by the second law of , which dictates that not all heat input can be converted to work, as some must be rejected to the cold reservoir to maintain the cycle. Heat engines exclude non-cyclic devices or those that convert energy through non-thermal means, such as electrochemical reactions in fuel cells, which directly transform chemical potential into electrical work without relying on temperature gradients. In contrast to refrigerators and heat pumps, which require net work input to transfer heat from a cold source to a hot sink against the natural flow, heat engines produce a net work output by exploiting the spontaneous flow of heat from hot to cold. This fundamental directional difference underscores their roles: heat engines generate useful mechanical energy, whereas refrigerators and heat pumps achieve cooling or heating effects./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/04%3A_The_Second_Law_of_Thermodynamics/4.04%3A_Refrigerators_Heat_Pumps_and_the_First_Law_of_Thermodynamics) Key terminology includes the heat input from the hot reservoir (Q_h), the heat rejected to the cold reservoir (Q_c), and the net work output (W), with thermal efficiency defined as the ratio W/Q_h. These quantities form the basis for analyzing engine performance within thermodynamic constraints.

Basic Components and Operation

A heat engine fundamentally comprises four core components: a hot reservoir serving as the source of high-temperature heat, a working fluid—typically a gas, liquid, or phase-changing substance like steam—that undergoes thermodynamic changes, a cold reservoir acting as the sink for rejected waste heat, and a mechanical linkage such as a piston in reciprocating engines or blades in turbines that converts the fluid's energy into useful mechanical work. The operational sequence of a follows a cyclic involving , expansion for work extraction, rejection, and compression to restore the initial state. The first absorbs Q_h from the hot , causing it to expand and drive the mechanical linkage to produce work. This is followed by the rejection of lower-grade Q_c to the cold , after which the is compressed, often with minimal work input, to complete the cycle and prepare for renewed . This sequence adheres to the first law of , which states that the change in over a complete cycle is zero (\Delta U = 0), implying that the net work output equals the difference between absorbed and rejected : W_{net} = Q_h - Q_c. The directional flow of operation—from hot to cold reservoir—is enforced by the second law of thermodynamics, which dictates that heat transfers spontaneously only from higher to lower temperatures and prohibits devices that could convert heat entirely into work without such a differential, thereby ruling out perpetual motion machines of the second kind.

Thermodynamic Principles

Fundamental Laws and Cycles

The establishes the concept of , stating that if two systems are each in with a third system, then they are in with each other. This law provides the foundation for defining as a measurable property of systems in equilibrium, which is essential for heat engines to operate by identifying and reservoirs. Without this prerequisite, the consistent transfer of between components in a heat engine would be impossible to quantify or control. The first law of , a statement of , asserts that the change in of a equals the added to the system minus the work done by the system. In the context of heat engines, this law ensures that the work output derives from the conversion of input, with no net creation or destruction of during the process. It sets the basic framework for heat-to-work but does not address the directionality or of such transformations. The second law of introduces the principle of directionality in natural processes, with two equivalent statements relevant to heat engines: the Clausius statement, which prohibits from spontaneously flowing from a colder body to a hotter one without external work, and the Kelvin-Planck statement, which declares that no heat engine can convert all absorbed into work without rejecting some to a colder . These statements imply the existence of , a measure of or unavailable energy, which increases in all irreversible processes, including those in real heat engines due to , leaks, and finite temperature differences. Consequently, complete conversion of to work is impossible, mandating waste expulsion and limiting engine performance. A in a heat engine consists of a closed of processes that returns the working substance to its initial , enabling repeated operation without net change in system properties. Cycles are classified as reversible, where the system and surroundings can be restored to their original states with no net change, or irreversible, where increases due to dissipative effects like or unrestrained . Reversible cycles serve as theoretical ideals for analyzing maximum possible , while irreversible cycles reflect practical operations with inherent losses. Among idealized cycles, the stands as the benchmark for heat engine performance, comprising two reversible isothermal processes—at constant temperature, where heat is absorbed from a hot reservoir and rejected to a cold one—and two reversible adiabatic processes—without , involving expansion and compression. Proposed by Sadi Carnot in , this cycle achieves the highest possible efficiency for given reservoir temperatures but remains unattainable in practice because real processes inevitably involve irreversibilities that increase .

Key Processes in Heat Engines

Heat engines operate through a series of thermodynamic processes that convert thermal energy into mechanical work, typically idealized in cycles like the Carnot cycle. These processes are reversible in the ideal case, ensuring maximum efficiency, and include two isothermal steps where heat transfer occurs at constant temperature and two adiabatic steps where no heat is exchanged. The working fluid, often modeled as an ideal gas, undergoes changes in pressure, volume, temperature, and entropy during these steps, governed by the first and second laws of thermodynamics. The first key process is isothermal heat addition, where the working fluid absorbs Q_h from a high-temperature at constant T_h. During this expansion, the fluid's remains unchanged for an , so the absorbed fully converts to work output, with the volume increasing while decreases. This process increases the of the by \Delta S = Q_h / T_h, as occurs reversibly at constant . Following this is the adiabatic expansion, an where the fluid expands without any heat transfer (Q = 0), converting into additional work. For an , the pressure and follow the P V^{\gamma} = \constant, where \gamma = C_p / C_v is the (e.g., \gamma = 5/3 for monatomic gases). The temperature decreases as the fluid does work, with remaining constant due to the reversibility. This step steepens the pressure- curve compared to isothermal expansion. The third process, isothermal heat rejection, occurs at a lower constant temperature T_c, where the fluid releases heat Q_c to a cold reservoir while contracting. Similar to heat addition, internal energy is unchanged, and the rejected heat equals the work input, decreasing the system's entropy by \Delta S = -Q_c / T_c. Volume decreases as pressure rises, maintaining thermal equilibrium with the reservoir. Finally, adiabatic compression reverses the expansion: the fluid is compressed without , requiring work input to increase its and temperature back toward T_h. Again, for an , P V^{\gamma} = \constant holds, with constant and no heat exchange. This process prepares the fluid for the next by restoring initial conditions. These processes are visualized using pressure-volume (P-V) and temperature- (T-S) diagrams. In a P-V , isothermal processes appear as hyperbolas (P V = \constant), while adiabatics are steeper curves; the enclosed area represents net work. The T-S shows horizontal lines for isothermals (with changes) and vertical lines for adiabatics (constant ), highlighting the cycle's reversibility through equal increases and decreases. In real engines, irreversibilities such as mechanical , fluid turbulence, and unintended losses across finite temperature differences degrade these ideal processes, reducing by generating .

Classification and Examples

Conventional Macroscopic Engines

Conventional macroscopic heat engines encompass traditional large-scale devices that convert into work, primarily through external or internal processes, and are widely employed in and transportation sectors. External combustion engines, where heat is supplied from an external source to a , include engines operating on the and engines. The , fundamental to steam power plants, involves four key components: a where water is heated to produce high-pressure , a that extracts work from the expanding , a that liquefies the exhaust , and a that returns the liquid water to the . In this cycle, plays a crucial role during the phase change in the , where water evaporates into , absorbing significant energy at constant temperature to enable efficient heat addition and subsequent work extraction in the . The , another external combustion type, operates as a closed-cycle regenerative heat engine using a permanently gaseous , such as air or , where heat is transferred through cyclic compression and expansion with internal regeneration to store and reuse , minimizing losses. Internal combustion engines, which burn fuel directly within the working chamber, dominate automotive and heavy-duty applications through cycles like the Otto and Diesel. The Otto cycle models spark-ignition gasoline engines, featuring constant-volume heat addition via spark-induced combustion after isentropic compression, followed by expansion and exhaust, enabling efficient operation in passenger vehicles. In contrast, the Diesel cycle powers compression-ignition engines using diesel fuel, with heat addition occurring at constant pressure during fuel injection and combustion after high compression, which allows for higher compression ratios and better fuel economy in trucks and generators. Gas turbines, operating on the Brayton cycle, provide continuous-flow power through a compressor that pressurizes intake air, a combustor that adds heat at constant pressure by burning fuel, and a turbine that drives both the compressor and an external load, such as a propeller or generator. These engines find broad applications in automotive propulsion via and cycles, stationary power generation using turbines, gas turbines, and reciprocating engines, and primarily through large engines and gas turbines for ships. Typical efficiencies for internal combustion engines range from 20% to 40%, influenced by factors like and load conditions, though real-world performance varies with design and operation.

Specialized and Natural Heat Engines

The Earth's atmosphere operates as a planetary heat engine, powered by that unevenly heats the surface, driving currents, patterns, and systems through the redistribution of . This process converts absorbed into mechanical work, such as , while dissipating excess heat to space via . The overall of this natural heat engine is approximately 1-2%, limited by irreversible processes like friction in air flows and radiative losses, far below theoretical Carnot limits due to the broad temperature range from surface highs to cosmic background lows. Refrigeration cycles function as specialized reverse heat engines, absorbing from a low-temperature reservoir and rejecting it to a higher one, typically using external work or input, with performance measured by the (COP), defined as the ratio of cooling effect to input . The vapor-compression , akin to a reversed , employs four key components: a to raise refrigerant pressure and temperature, a to release , an expansion valve to reduce pressure, and an to absorb , achieving COP values of 3-5 in practical systems depending on operating temperatures. In contrast, absorption cycles replace mechanical compression with thermal absorption using an absorbent-refrigerant pair, such as ammonia-water, driven by from sources like waste streams, yielding lower COPs around 0.7 for applications but enabling operation without . Evaporative heat engines leverage humidity gradients and water to produce cooling or limited mechanical work, exploiting the of to transfer without moving parts. In these systems, dry air passes over water-saturated media, where cools the air stream by absorbing , increasing while lowering by up to 15-20°C in arid conditions, though effectiveness diminishes in high- environments. At mesoscopic and nanoscale regimes, heat engines manipulate electron flow or molecular vibrations to harvest thermal energy, operating under quantum and fluctuation-dominated thermodynamics distinct from macroscopic counterparts. These devices, often fabricated in solid-state systems, convert heat gradients into directed electron currents or mechanical oscillations at the single-molecule level, with prototypes demonstrating work extraction from ambient fluctuations via ratchet-like mechanisms. Magnetic cycles, based on the magnetocaloric effect, enable cooling by cyclically applying and removing magnetic fields to materials like , causing reversible temperature changes of several kelvins near Curie points, achieving COPs up to 10 in prototype refrigerators for near-room-temperature applications. Phase-change and liquid-only heat engines adapt thermodynamic cycles for low-grade heat sources, prioritizing organic or alternative fluids over steam to avoid phase-change challenges at reduced temperatures. The (ORC) uses organic working fluids like refrigerants in a closed loop to generate power from between 80-200°C, with typical thermal efficiencies of 5-15% depending on fluid selection and temperature differential, enabling recovery from or geothermal sources. Thermoelectric engines, grounded in the Seebeck effect where temperature differences across junctions of dissimilar materials induce voltage via charge carrier diffusion, operate without fluids or moving parts, converting heat directly to electricity with efficiencies reaching 10% for materials with figure-of-merit ZT around 1.25, suitable for scavenging in .

Efficiency and Performance

Theoretical Efficiency Limits

The Carnot theorem establishes that no heat engine operating between two thermal reservoirs can exceed the of a reversible Carnot engine operating between the same reservoirs, and that all reversible engines between those reservoirs achieve identical . This theorem, originally articulated by Sadi Carnot in his analysis of ideal heat engines, underscores the second law of thermodynamics by prohibiting any process from converting heat entirely into work without some rejection to a colder reservoir. The maximum efficiency of a reversible heat engine, known as the Carnot efficiency, is derived from the condition of zero net entropy change in a cyclic process. For a reversible cycle, the total entropy change is \Delta S = 0 = \frac{Q_h}{T_h} + \frac{Q_c}{T_c}, where Q_h > 0 is the heat absorbed from the hot reservoir at temperature T_h and Q_c < 0 is the heat rejected to the cold reservoir at T_c (both temperatures in ). Rearranging gives \frac{|Q_c|}{Q_h} = \frac{T_c}{T_h}. The efficiency \eta is then the ratio of net work output to heat input, \eta = \frac{W}{Q_h} = 1 - \frac{|Q_c|}{Q_h} = 1 - \frac{T_c}{T_h}. This formula holds regardless of the working fluid, as the derivation relies solely on thermodynamic reversibility and the temperatures of the reservoirs. The implications of Carnot efficiency are profound: it sets an absolute upper bound on heat engine performance, dependent only on the , which limits practical applications to scenarios with significant differences. For instance, with T_h = 800 K and T_c = 300 K, \eta_{Carnot} \approx 62.5\%, illustrating that even engines cannot approach 100% without an infinite . To address limitations of the infinite-time reversible assumption, endo-reversible models within finite-time thermodynamics provide bounds that assume internal reversibility but incorporate external irreversibilities from finite-rate . In these models, the engine operates between intermediate temperatures due to thermal gradients at the boundaries, yielding a maximum power of \eta = 1 - \sqrt{T_c / T_h}, as derived by Curzon and Ahlborn for an endoreversible Carnot engine. This expression offers a more attainable target for real systems prioritizing power output over ultimate .

Real-World Efficiency and Losses

In practical heat engines, efficiency is invariably lower than theoretical limits due to various irreversibilities that generate and dissipate useful . These losses stem primarily from in , such as bearings and pistons, which converts into ; across finite temperature differences, leading to irreversible conduction; incomplete in engines where fuel is not fully oxidized, resulting in unburned hydrocarbons and loss; and inefficiencies in pumps, turbines, and compressors due to fluid and non-ideal flow. Additionally, second law losses arise from generation during processes like mixing of gases, chemical reactions, and throttling, which reduce the available work potential beyond what reversible models predict. Performance in real-world heat engines is quantified using metrics that account for these losses. The , defined as η = W_net / Q_in, where W_net is the net work output and Q_in is the input, measures the fraction of converted to useful work. Specific fuel consumption (SFC), often expressed as (BSFC) in grams of fuel per , indicates fuel usage per unit power and inversely relates to . analysis provides a more comprehensive assessment by evaluating the maximum available work from energy streams, highlighting destruction due to irreversibilities like those mentioned above, and is particularly useful for identifying loss hotspots in complex systems such as power plants. Typical thermal efficiencies vary by engine type and are constrained by material limits, such as maximum operating temperatures around 1,000–1,500°C for blades to avoid and oxidation. Coal-fired power plants achieve 30–40% efficiency, limited by and losses. Internal engines range from 20–35% for variants, affected by pumping and rejection, to 30–45% for engines with higher ratios. Combined plants, integrating gas and turbines, reach up to 60% by recovering exhaust , though real values often fall to 50–55% due to component mismatches. These figures underscore the gap to Carnot limits, often 10–20 percentage points lower in practice.
Engine TypeTypical Thermal Efficiency (%)Key Limiting Factors
Coal-Fired Steam Plant30–40Heat transfer losses, material temperature limits
Gasoline IC Engine20–35Incomplete ,
Diesel IC Engine30–45Pumping losses, in expansion
Combined Cycle Plant50–60 inefficiencies, heat recovery limits
Techniques like regeneration and intercooling can mitigate some losses by recovering or reducing work, but their implementation is explored in subsequent discussions on enhancements.

Historical Development

Ancient and Early Modern Concepts

The concept of harnessing heat to produce mechanical work dates back to antiquity, with the standing as the earliest documented example of such a device. Invented by the Greek engineer around 10–70 AD, the was a simple reaction consisting of a hollow spherical vessel mounted over a containing water. As the water boiled, escaped through two opposing L-shaped nozzles attached to the sphere, generating reactive thrust that caused the device to rotate rapidly. This demonstration illustrated the potential for heat to drive rotary motion, though the functioned more as a novelty or temple ornament than a utilitarian machine, producing no significant work output beyond its spin. In the medieval and early modern periods, sporadic innovations built on these ancient ideas, particularly in the and , where practical needs like pumping and spurred experimentation. In 1551, the polymath Taqi al-Din Muhammad ibn Ma'ruf described a steam jack in his treatise Al-Turuq al-saniyyah fi al-alat al-ruhaniyyah, an early that directed jets against angled vanes on a to rotate a roasting spit automatically, representing the first known practical steam-powered mechanism. Around the same time, Taqi al-Din also engineered a six-cylinder reciprocal capable of raising , which, while not steam-driven, exemplified advancing technology for fluid displacement in applications like or . In European mining contexts, particularly in Germany's region during the 16th and 17th centuries, deepening shafts exacerbated flooding issues, leading engineers to conceptualize steam-assisted systems for drainage; however, these remained theoretical or rudimentary, relying instead on water wheels and horse-powered gins for actual implementation. Theoretical advancements in the further laid the groundwork for heat engines by exploring and dynamics. , a and , invented the first functional air pump in the 1650s, a -cylinder device that evacuated air from sealed vessels to create partial , famously demonstrated through the experiment where held two hemispheres together against teams of horses. This work illuminated the force of air and the effects of reduced , providing essential insights into principles that would influence later engine designs. Complementing this, French developed the in 1679, a sealed high-pressure vessel used to soften bones with under a weighted lid; observing the steam's expansive force, Papin proposed in 1690 a -cylinder arrangement where steam could lift weights, marking the first explicit concept of a steam-driven engine. Despite these innovations, ancient and early modern heat engine concepts faced profound limitations due to the era's incomplete scientific framework. Without knowledge of —particularly on and the second law limiting heat-to-work conversion—devices like the and steam jack achieved negligible , often wasting as uncontrolled or steam leakage. Lacking , valves, and materials to withstand sustained , these inventions served primarily as scientific curiosities or isolated tools rather than scalable engines for industry or transport, hindering their transition to practical power sources until the 18th century.

Industrial and Contemporary Advances

Building on Papin's ideas, English engineer Thomas Savery patented the first commercially used steam-powered device in 1698, a pump that used steam to create a vacuum and draw water from mines, though it was inefficient, requiring high fuel consumption and limited to low lifts due to steam pressure constraints. This was followed by the Industrial Revolution marking a pivotal era for heat engines, beginning with the atmospheric steam engine developed by Thomas Newcomen in 1712, which was primarily used for pumping water out of mines but suffered from low efficiency due to its integrated cylinder-condenser design. This was significantly improved by James Watt's 1769 patent for a separate condenser, which prevented the cylinder from cooling during each cycle, boosting thermal efficiency from Newcomen's approximately 0.5% to around 2-4% and reducing fuel consumption by up to 75%. Watt's innovations, commercialized in partnership with Matthew Boulton from 1775, enabled broader applications beyond mining, powering factories and laying the groundwork for mechanized industry. In 1824, Sadi Carnot published "Reflections on the Motive Power of Fire," establishing the theoretical foundations of by analyzing the reversible heat engine , which set the upper limit on based on differences between heat source and sink. This work influenced subsequent developments, including the shift toward internal combustion engines in the . Étienne Lenoir's 1860 single-acting gas engine was the first commercially viable internal combustion design, operating on the principle of constant-volume combustion with an of about 4%. Nikolaus Otto's 1876 four-stroke engine improved this to around 12-15% by incorporating intake, compression, power, and exhaust strokes, while Rudolf Diesel's 1892 compression-ignition engine achieved up to 26% through higher compression ratios and . The , utilizing steam in a closed loop with boilers, turbines, and condensers, became dominant in central power plants by the late , enabling large-scale . Gas turbines emerged in the , with practical implementations in and power generation following Frank Whittle's and Hans von Ohain's independent designs in the late 1930s, offering higher power-to-weight ratios than engines. Standardization accelerated through patents and mass manufacturing; for instance, Henry Ford's 1913 moving for the Model T automobile streamlined production, reducing costs and enabling widespread adoption. Key milestones included powering railroads from the , transforming transportation and commerce; automobiles commercialized in the by and others; and propelled by engines from the 1900s, as demonstrated by the ' 1903 flight. By the mid-20th century, these advances yielded efficiency gains to 20-30% for typical and around 30% for steam Rankine plants, reflecting optimized cycles and materials.

Enhancements and Emerging Technologies

Methods to Improve Efficiency

One primary to enhance heat engine involves thermodynamic modifications that maximize the temperature differential between the heat source and sink, as dictated by the Carnot efficiency limit. Increasing the hot-side temperature (T_h) allows engines to approach higher theoretical ; advanced ceramic materials, such as (Si₃N₄) and (SiC), enable turbine inlet temperatures up to 2500°F or more by providing superior high-temperature strength, oxidation resistance, and tolerance compared to traditional metal alloys. These ceramics reduce material degradation and support multi-fuel operations, potentially boosting overall through higher operating temperatures without excessive cooling demands. Conversely, decreasing the cold-side temperature (T_c) via improved cooling strategies, such as intercooling in multi-stage compressors or advanced heat exchangers, minimizes heat rejection and enhances net work output; for instance, intercooling in Brayton cycles cools compressed gas toward ambient levels, reducing compression work and allowing regeneration to operate more effectively. Cycle modifications further optimize by recovering and refining expansion/compression processes. Regeneration, implemented via recuperators in Brayton cycles, preheats using exhaust , significantly reducing fuel input and improving at low to moderate ratios where exhaust temperatures exceed compressor outlet temperatures. Reheat cycles add intermediate heating stages in multi-stage turbines, raising the average temperature of heat addition and increasing specific work output, though they pair best with regeneration to offset added heat requirements and achieve net gains. The exemplifies near-Carnot performance through continuous regeneration and isothermal compression/expansion, theoretically matching Carnot while using practical processes, as demonstrated in configurations approaching 73% for specific temperature ranges. Fluid and process optimizations leverage alternative working fluids for better thermodynamic matching. (sCO₂) cycles employ CO₂ above its critical point for higher fluid density, enabling compact and efficient heat recovery via recuperators that limit heat rejection; these cycles achieve up to 45% with low-temperature heat sinks, surpassing traditional Rankine cycles in recovery applications. The uses an ammonia-water mixture as the , whose variable allows closer temperature gliding to the heat source during evaporation, improving matching and yielding 10-20% higher than conventional Rankine cycles for low-grade heat sources like turbine exhaust. Component-level improvements target frictional and aerodynamic losses to elevate real-world performance. Variable geometry turbines adjust vane angles to optimize flow incidence across operating conditions, maintaining high turbine (up to 60% peak) and broadening the engine's efficient speed range, which enhances overall cycle in variable-load applications like automotive and turbines. Low-friction bearings, such as super-precision ball bearings with advanced coatings, reduce mechanical losses by minimizing viscous drag and heat generation, contributing to 1-5% gains in high-speed rotating components while improving reliability under stresses. Combined cycles integrate multiple engine types to cascade energy recovery, achieving efficiencies over 60% by utilizing exhaust from a topping cycle (e.g., ) to drive a bottoming cycle (e.g., ) via heat recovery steam generators. This approach recovers otherwise lost , with modern systems reaching 64% through high-temperature s and optimized conditions. Efficiency improvements are often quantified using second law metrics like exergy recovery, which measures the fraction of available work potential () converted rather than thermal efficiency alone. In regenerative Brayton cycles, reveals that recuperators can recover up to 80% of exhaust , elevating second law efficiency by identifying and minimizing irreversibilities in and processes. These methods collectively address real-world losses such as and incomplete , enabling practical engines to approach theoretical limits without venturing into experimental designs.

Modern and Exotic Developments

In recent years, advancements in sustainable heat engine technologies have focused on recovering low-temperature , typically below 100°C, using (ORC) systems enhanced by . These systems employ organic working fluids with low boiling points to convert into , achieving efficiencies up to 20% through the integration of nanostructured materials like carbon nanotubes and graphene oxide, which improve and reduce thermal losses. Such nanomaterial enhancements have been demonstrated in post-2020 prototypes, enabling practical applications in and data centers where is abundant. Integration of systems with renewable sources, particularly solar thermal hybrids, has further expanded their viability. Solar-powered engines, for instance, have seen efficiency improvements to over 30% in hybrid configurations combining collectors with ORC bottoming cycles, allowing continuous operation by storing excess solar heat. These hybrids mitigate intermittency in solar input, providing stable power output for off-grid applications and contributing to decarbonization efforts in remote areas. At the nanoscale and mesoscopic levels, molecular heat engines have emerged as experimental platforms harnessing for directed motion. DNA-based heat engines, utilizing programmable DNA nanostructures as working media, operate via cyclic changes to drive conformational switches, achieving synchronized at frequencies up to 1 Hz and converting into mechanical work with near-100% in controlled environments. mechanisms in these systems rectify random fluctuations into net displacement, inspired by biological motors, and have been optimized using to maximize power output in fluctuating thermal baths. Complementing these, electron heat engines in semiconductors exploit single-electron tunneling in quantum dots to manage nanoscale heat flows, with prototypes demonstrating thermoelectric efficiencies exceeding 10% at by leveraging spin-dependent transport. Quantum heat engines represent a , leveraging quantum to surpass classical efficiency bounds in specialized cycles. The quantum , implemented with superconducting qubits or photonic systems, has achieved work extraction with efficiencies up to 25% of the Carnot limit while maintaining times over 100 μs, as shown in trapped-ion experiments where quantum correlations enhance beyond semiclassical predictions. Maser-like quantum devices, operating without , convert heat directly into coherent microwave emission, with recent demonstrations yielding positive work output at efficiencies rivaling classical engines but with tunable quantum advantages from entanglement. from 2021 to 2025 has certified these enhancements through resource-theoretic comparisons, confirming that quantum steady-state operations can outperform classical thermal machines under identical thermodynamic constraints. Exotic heat engine concepts include magnetic refrigeration systems based on room-temperature magnetocaloric materials, which cycle magnetic fields to drive adiabatic demagnetization for cooling without vapor-compression refrigerants. Low-dimensional magnetocalorics, such as gadolinium-based nanostructures, exhibit giant magnetocaloric effects with temperature spans up to 10 K per cycle, enabling efficient heat pumping for cryogenic applications and sustainable air conditioning. Chemical heat engines utilizing thermochemical storage materials, like metal hydrides or salt complexes, store energy via reversible reactions, releasing heat on demand with energy densities over 1 MJ/kg and minimal losses over months. Recent nano-engineered variants incorporate perovskites for faster kinetics, achieving round-trip efficiencies above 90% in solar-driven prototypes. In hypersonic regimes, scramjet engines have advanced with active-cooled designs tested at Mach 6+, incorporating regenerative cooling channels to manage heat fluxes exceeding 10 MW/m², paving the way for reusable hypersonic vehicles. Despite these innovations, challenges in persist, particularly for nanoscale engines where integrating millions of molecular units into macroscopic devices remains limited by fabrication precision and synchronization losses. Materials like offer promise for enhancing thermal conductivity by factors of 10 in nano-engines, but production and cost barriers hinder widespread adoption. Environmentally, these developments enable zero-emission cycles by waste heat and eliminating harmful refrigerants, potentially reducing global by 20% in cooling sectors through magnetocaloric and thermochemical systems. The outlook emphasizes hybrid quantum-classical architectures to bridge lab-scale proofs to industrial viability, fostering conversion.

References

  1. [1]
    4.2 Heat Engines – University Physics Volume 2 - UCF Pressbooks
    A heat engine is a device used to extract heat from a source and then convert it into mechanical work that is used for all sorts of applications.Missing: principles | Show results with:principles
  2. [2]
    Introduction to the Second Law of Thermodynamics: Heat Engines ...
    A device that uses heat to do mechanical work is called a heat engine. These include car engines, jet turbines, and power plant generators. A heat engine ...
  3. [3]
    Heat Engines
    Definition of an engine: A engine works in a closed cycle, returning to its initial state periodically at the end of each cycle. It is an energy transformer ...
  4. [4]
    OpenStax | Free Textbooks Online with No Catch
    **Summary of Heat Engines from OpenStax University Physics Volume 2, Section 4.3:**
  5. [5]
    Comparison of the theoretical performance potential of fuel cells and ...
    It is a common misconception or conclusion that fuel cells have a greater performance potential than heat engines since they are not restricted by the Carnot ...Missing: distinction | Show results with:distinction
  6. [6]
    Heat Engines - HyperPhysics
    A heat engine typically uses energy provided in the form of heat to do work and then exhausts the heat which cannot be used to do work.
  7. [7]
  8. [8]
    6.1 Heat engine – Introduction to Engineering Thermodynamics
    A heat engine is a continuously operating device that produces work by transferring heat from a heat source (high-temperature body) to a heat sink ...
  9. [9]
    15.3 Introduction to the Second Law of Thermodynamics: Heat ...
    Most heat engines, such as reciprocating piston engines and rotating turbines, use cyclical processes. The second law, just stated in its second form ...
  10. [10]
    Second Law of Thermodynamics - HyperPhysics
    Second Law of Thermodynamics: In any cyclic process the entropy will either increase or remain the same. Entropy: a state variable whose change is defined for a ...
  11. [11]
    The Second Law of Thermodynamics - Florida State University
    A heat engine whose thermal efficiency is given by Equation 6-41 is a theoretical heat engine only, an idealization that engineers use to compare with real heat ...
  12. [12]
    Zeroth Law - Thermal Equilibrium | Glenn Research Center - NASA
    May 2, 2024 · The zeroth law of thermodynamics is an observation. When two objects are separately in thermodynamic equilibrium with a third object, they are in equilibrium ...
  13. [13]
    1.3 Changing the State of a System with Heat and Work - MIT
    Zeroth Law: There exists for every thermodynamic system in equilibrium a property called temperature. Equality of temperature is a necessary and sufficient ...
  14. [14]
    12.2 First law of Thermodynamics: Thermal Energy and Work - Physics
    Mar 26, 2020 · The first law of thermodynamics applies the conservation of energy principle to systems where heat and work are the methods of transferring ...
  15. [15]
    The Laws of Thermodynamics and Limits on Engine Efficiency
    The first law of thermodynamics: total energy, including heat energy, is always conserved. He explicitly assumed that heat was just the kinetic energy of the ...
  16. [16]
    6.3 The second law of thermodynamics: Kelvin-Planck and Clausius ...
    Any device that violates the Kelvin–Planck statement must violate the Clausius statement, and vice versa. Let us consider a combined system, which consists of a ...
  17. [17]
    5.1 Concept and Statements of the Second Law - MIT
    The entropy change of any system and its surroundings, considered together, is positive and approaches zero for any process which approaches reversibility.
  18. [18]
    [PDF] Supplementary Notes on Entropy and the Second Law of ...
    3 Entropy Change for an Irreversible Process ... We see that even though we cannot prove the Kelvin-Planck or the Clausius statements of the Second Law, they are.
  19. [19]
    Reversible and irreversible heat engine and refrigerator cycles
    May 1, 2018 · This is a review, clarification, and extension of results and concepts for quasistatic, reversible and irreversible processes and cycles, ...
  20. [20]
    [PDF] Reversible and irreversible heat engine and refrigerator cycles
    A general definition of a heat engine is a cyclic device that takes a working substance through a CW thermodynamic cycle, receiving energy Qin from a range ...
  21. [21]
    3.3 The Carnot Cycle - MIT
    A Carnot cycle is shown in Figure 3.4. It has four processes. There are two adiabatic reversible legs and two isothermal reversible legs.
  22. [22]
    Carnot's Perfect Heat Engine: The Second Law of Thermodynamics ...
    This cycle is composed exclusively of reversible processes: two isothermal processes (constant temperature) and two adiabatic processes (no heat exchange). Real ...
  23. [23]
    13.8: Carnot Cycle, Efficiency, and Entropy - Chemistry LibreTexts
    Jan 2, 2019 · The Carnot cycle consists of the following four processes: A reversible isothermal gas expansion process. In this process, the ideal gas in the ...The Carnot Cycle · P-V Diagram · T-S Diagram · Efficiency
  24. [24]
    Carnot Cycle - HyperPhysics
    The most efficient heat engine cycle is the Carnot cycle, consisting of two isothermal processes and two adiabatic processes.
  25. [25]
    Carnot Cycle
    The first process performed on the gas is an isothermal expansion. · The second process performed on the gas is an adiabatic expansion. · The third process ...
  26. [26]
    3.6 Adiabatic Processes for an Ideal Gas - UCF Pressbooks
    On an adiabatic process of an ideal gas pressure, volume and temperature change such that p V γ is constant with γ = 5 / 3 for monatomic gas such as helium and ...
  27. [27]
    [PDF] • Reversible Adiabatic Expansion (or compression) of an Ideal Gas
    In an adiabatic expansion (V2 > V1), the gas cools (T2 > T1). And in an adiabatic compression (V2 < V1), the gas heats up. For an ideal gas (one mole) γ γ γ.
  28. [28]
    Reversible and irreversible heat engine and refrigerator cycles
    May 1, 2018 · Actual heat engines and refrigerators have unavoidable sources of irreversibility like friction and heat leaks. They typically operate so ...
  29. [29]
    [PDF] Section 4. Technology Characterization – Steam Turbines
    The thermodynamic cycle for the steam turbine is known as the Rankine cycle. This cycle is the basis for conventional power generating stations and consists ...
  30. [30]
    [PDF] Rankine Cycle - MIT OpenCourseWare
    Apr 1, 2020 · Operating in a closed cycle (to recirculate the working fluid), the turbine exhausts into vacuum, the pressure is determine by the cold.Missing: engine components
  31. [31]
    Chapter 3b - The First Law - Closed Systems - Stirling Ebdines ...
    Jul 5, 2014 · The same working gas is used over and over again, making the Stirling engine a sealed, closed cycle system. ... regeneration. Furthermore ...
  32. [32]
    3.5 The Internal combustion engine (Otto Cycle) - MIT
    The heat absorbed occurs during combustion when the spark occurs, roughly at constant volume. The heat absorbed can be related to the temperature change ...Missing: addition | Show results with:addition
  33. [33]
    Design of a Diesel Cycle
    Mar 25, 2002 · Compression-ignition cycles use fuels that begin combustion when they reach a temperature and pressure that occurs naturally at some point ...
  34. [34]
    Turbine Engine Thermodynamic Cycle - Brayton Cycle
    The Brayton cycle is a thermodynamic process used in gas turbine engines, involving compression, combustion, and power turbine, and is used to predict ...
  35. [35]
    How Gas Turbine Power Plants Work - Department of Energy
    Gas turbines have three main sections: compressor, combustion, and turbine. The turbine spins the compressor and a generator to produce electricity.Missing: Brayton | Show results with:Brayton
  36. [36]
    Understand Transportation
    Mar 18, 2025 · A “perfect” internal combustion engine can only achieve ~50% efficiency, and a typical internal combustion engine is only about 20-40% efficient ...Missing: marine | Show results with:marine
  37. [37]
    [PDF] Lectures 7-8 Thurs 23.iv.09 HAS 222d Introduction to energy ...
    The atmosphere/ocean system is a 'heat engine' largely driven by the sun…that is, a contraption in which heated fluid (air or water) expands and, ...<|separator|>
  38. [38]
    Toward Quantifying the Climate Heat Engine: Solar Absorption and ...
    A heat-engine analysis of a climate system requires the determination of the solar absorption temperature and the terrestrial emission temperature.
  39. [39]
    Thermodynamics of the climate system - Physics Today
    Jul 1, 2022 · As a result, the winds and currents driven by the planetary heat engine affect both its efficiency and the amount of heat that it transports.
  40. [40]
    Design of Vapor-Compression Refrigeration Cycles
    Dec 16, 1997 · An ideal refrigeration cycle looks much like a reversed Carnot heat engine or a reversed Rankine cycle heat engine. The primary distinction ...
  41. [41]
    Module 10: Absorption refrigeration - CIBSE Journal
    In practice a typical COP for an absorption cycle in air conditioning would be about 0.7, compared to about 3.5 for a vapour compression system. It appears that ...
  42. [42]
    A comprehensive review on evaporative cooling systems
    The humidity of air increases due to evaporation of water in the air stream. The cooling effect produced by the ECS is good, however the power consumption of ...
  43. [43]
    Evaporative Cooling Systems: How and Why They Work
    Water Evaporation Alters. Both Temperature and Humidity. Water evaporating from your wet, cool cell pads has a cooling effect on the hot air passing through the ...Missing: engines | Show results with:engines
  44. [44]
    Fundamental aspects of steady-state conversion of heat to work at ...
    This review aims to introduce some of the theories used to describe these steady-state flows in a variety of mesoscopic or nanoscale systems.
  45. [45]
    Controlling Motion at the Nanoscale: Rise of the Molecular Machines
    Jul 14, 2015 · In this Review, we outline the conditions that distinguish simple switches and rotors from machines and draw from a variety of fields.
  46. [46]
    Efficient Room-Temperature Cooling with Magnets - ACS Publications
    Jun 6, 2016 · It is based on the magnetocaloric effect, which is associated with the temperature change of a material when placed in a magnetic field. We ...Introduction · Figure 1 · Figure 3Missing: engines | Show results with:engines
  47. [47]
    Performance investigation on Organic Rankine Cycle from a low ...
    The results show that the energy efficiency and exergy efficiency increased from 4.6 % to 10.5 % and 2.4 %–5.6 %, respectively, for the above-specified ...
  48. [48]
    Thermoelectric Generators - Stanford University
    Oct 24, 2010 · A thermoelectric device with a ZT of 1.25 will have an efficiency of around 10% and could be used to generate useable power directly towards the ...<|control11|><|separator|>
  49. [49]
    [PDF] I.D The Second Law - MIT
    • Carnot's Theorem: No engine operating between two reservoirs (at temperatures TH and TC) is more efficient than a Carnot engine operating between them. Proof ...
  50. [50]
    [PDF] reflections on the motive power of fire, and on machines fitted to ...
    Sadi Carnot. 1825. (Excerpts). Page 2. 2. EVERYONE knows that heat can produce motion. That it possesses vast motive‐power none can doubt, in these days when ...
  51. [51]
    [PDF] Chapter 4 Entropy and the second law of thermodynamics
    Efficiency. The efficiency of the Carnot engine is defined as η ≡ performed work absorbed heat. = −W. Q1. = Q1 + Q2. Q1. = Q1 − |Q2|. Q1 . η is 100% if there ...
  52. [52]
    Efficiency of a Carnot engine at maximum power output
    Jan 1, 1975 · It is shown that the efficiency, η, at maximum power output is given by the expression η = 1 − (T2/T1)1/2 where T1 and T2 are the respective ...
  53. [53]
    Engine Efficiency - DieselNet
    These include the chemical energy loss in emissions, heat losses from the engine and through the exhaust gas, and gas pumping and friction losses in the engine.Missing: world | Show results with:world<|control11|><|separator|>
  54. [54]
    [PDF] Sources of Combustion Irreversibility - Penn Engineering
    Combustion irreversibility is due to combined diffusion/fuel oxidation, internal thermal energy exchange, and product constituent mixing. Most exergy ...
  55. [55]
    [PDF] MARTINS-Generation of Entropy in S I Engines - CORE
    From these five main sources of entropy (combustion, free expansion, heat transfer, friction and flow through valves), the free expansion and combustion are the ...
  56. [56]
    Specific Fuel Consumption - an overview | ScienceDirect Topics
    Specific fuel consumption is defined as the fuel flow rate per unit of power output, measuring the efficiency of an engine in using fuel to produce work.
  57. [57]
    Transformative Power Systems - Department of Energy
    The average coal-fired power plant in the United States operates near 33% efficiency. The Transformative Power Systems Research Program aims to increase the ...
  58. [58]
    Power Plant Efficiency: Coal, Natural Gas, Nuclear, and More ...
    Apr 17, 2023 · Coal power plant efficiency is very similar to nuclear, with a typical U.S. coal plant operating at 32% to 33% efficiency. The U.S. Department ...
  59. [59]
    Advanced Combustion Engines - Stanford
    Dec 9, 2011 · For example, advanced internal combustion engines found in modern automobiles have peak thermal efficiencies around 35-40% for gasoline and 40- ...
  60. [60]
    [PDF] Performance and Efficiency of Combined Cycle Power Plants
    The thermal efficiency of a CCGT can exceed 60% depending on various conditions and the design of the cycle's components (Bachmann et al.,1999). The combined ...
  61. [61]
    Combined cycle power plant efficiency: what you need to know
    Combined cycle power plants are known for their high thermal efficiency, typically reaching over 50% and sometimes above 60%.
  62. [62]
    Brief History of Rockets - NASA Glenn Research Center
    About three hundred years after the pigeon, another Greek, Hero of Alexandria, invented a similar rocket-like device called an aeolipile. It, too, used steam as ...
  63. [63]
    Hero Reaction Turbine Engine - Main Page and Introduction
    Hero or Heron of Alexandria is credited with the first written record of a steam engine between 10 and 70 A.D. Hero's Aeolipile is actually a reaction steam ...Missing: historical | Show results with:historical
  64. [64]
    [PDF] A New Look at Heron's "Steam Engine"
    Having noted the force of his liquid-filled squirt and of the heated liquid-filled Aeolus, 15 HERON probably sought to increase the effect by the use of heated ...
  65. [65]
    [PDF] A Bulleted/Pictorial History of Mechanisms and Machines
    A Steam Jack driven by a Steam Turbine (the earliest practical steam-powered machine) was invented in 1551 by Taqi al-Din Muhammad ibn Ma'ruf in Ottoman Egypt, ...
  66. [66]
    The Six-Cylinder Water Pump of Taqi al-Din - Muslim Heritage
    Jan 24, 2009 · The main objective of this study is to investigate into the six-cylinder water raising pump described around 1550 by the Ottoman Muslim scientist Muhammad Ibn ...
  67. [67]
    [PDF] A History of the Schemnitz (Banská Štiavnica) Silver-Gold Mines
    Dec 1, 2014 · These mines led in the early use of gunpowder, and in using horses, waterwheels, dams, reservoirs, ditches, and leats to meet their energy needs ...
  68. [68]
    Vacuum Pump - Physics - Kenyon College
    The piston-type vacuum pump was invented by Otto von Guericke (1602-1686) in the course of a series of experiments on the production and effects of a vacuum ...
  69. [69]
    The Second Generation of British Air Pumps - PubMed
    The air pump originated in the 1650s and 1660s thanks to the work of Otto von Guericke in Magdeburg, Robert Boyle in Oxford and London, and Accademia del ...
  70. [70]
    [PDF] LECTURE NOTES ON THERMODYNAMICS
    May 17, 2025 · ... HISTORICAL MILESTONES. 15. • 1679: Denis Papin develops his steam digester, forerunner to the steam engine. • 1687: Isaac Newton publishes ...
  71. [71]
    [PDF] Statistical Physics (PHY831): Part 1 - The foundations Phillip M ...
    Even before the establishment of the ideal gas law, an associate of Boyle's named Denis Papin in 1679 built a high pressure steam cooker, also called a bone ...
  72. [72]
    A History of Thermodynamics: The Missing Manual - PMC
    We present a history of thermodynamics. Part 1 discusses definitions, a pre-history of heat and temperature, and steam engine efficiency, which motivated ...
  73. [73]
    How Did We Get Here? The Tangled History of the Second Law of ...
    Jan 31, 2023 · It's a fascinating story, that both explains what's been believed about thermodynamics, and provides some powerful examples of the complicated dynamics.
  74. [74]
    [PDF] Heat-to-work
    First modern steam engine: James Watt, 1769 (patent), 1774 (prod.) Higher efficiency than Newcomen by introducing separate condenser. Page 11. First modern ...
  75. [75]
    [PDF] Creator Of The Steam Engine
    James Watt improved the design of the steam engine by adding a separate condenser, which greatly increased its efficiency and made it more practical for ...
  76. [76]
    [PDF] Sadi Carnot, 'Founder of the Second Law of Thermodynamics'
    Thurston's translation was also reprinted in 1960 as part of a collection edited by E. Mendoza entitled Reflections on the Motive Power of Fire by Sadi Carnot ...
  77. [77]
    15.4 Carnot's Perfect Heat Engine: The Second Law of ...
    Any engine using the Carnot cycle, which uses only reversible processes (adiabatic and isothermal), is known as a Carnot engine. Any engine that uses the Carnot ...
  78. [78]
    [PDF] Internal Combustion Engine History - DSpace@MIT
    1860 ... Displacement 1/16 of Otto and Langen engine. Weight 1/3. 1878 Dugald Clerk. Two-stroke cycle engine. Higher output per revolution. 1892 Rudolf Diesel.
  79. [79]
    [PDF] Internal Combustion Engines, Lecture 1 Introduction to SI and DI ...
    1892. Compression Ignition 4-stroke. Rudolf Diesel. ― Key concepts: prevent the very rapid and high pressure heat. 2. Page 3. process via introducing fuel late ...
  80. [80]
    [PDF] Increasing the Efficiency of Existing Coal-Fired Power Plants
    The expansion of water into steam vapor (and condensation back into liquid water) in this manner is called a Rankine Cycle,11 and is the basis for most electric ...
  81. [81]
    Early Gas Turbine History
    1939 World's first gas turbine for power generation by Brown Boveri Company, installed in Neuchâtel, Switzerland
  82. [82]
    [DOC] Chapter Five - Princeton University
    But the key to Ford's success, rigid standardization of parts and procedures, also made Ford an authoritarian manager whose sense of the market and whose ...
  83. [83]
    Internal Combustion Engine - Otto Cycle | Glenn Research Center
    28 Jul 2023 · On this page we will discuss the basic thermodynamic principles and on a separate page we present the thermodynamic analysis that allows you to ...
  84. [84]
    [PDF] liquid-metal binary cycles for stationary power
    The Rankine cycle closely approximates the ideal or Carnot cycle because most of the heat is added or rejected at constant temper- atures. Practical steam ...
  85. [85]
    [PDF] Reliability of Ceramics For Heat Engine Applications
    The potential for use of monolithic ceramics in heat engines presents an attractive possibility for significant improvement in engine durability, efficiency ...
  86. [86]
    Ceramics in Heat Engines 790645 - SAE International
    30-day returnsJan 31, 1979 · The thermal efficiencies of the Otto, Diesel, Brayton and the Stirling cycle can now be improved by higher operating temperatures, reduced heat ...Missing: thermodynamic enhancements increasing
  87. [87]
    Brayton Cycle with Reheat, Regeneration and Inter-cooling
    Reheat, Inter-cooling and Regeneration in Brayton Cycle. As was discussed, reheat and inter-cooling are complementary to heat regeneration.
  88. [88]
    6.4 Carnot cycles – Introduction to Engineering Thermodynamics
    The Carnot heat engine cycle is an ideal cycle consisting of only reversible processes, it produces the maximum power output and has the maximum thermal ...
  89. [89]
    Regenerative Brayton Cycle - an overview | ScienceDirect Topics
    The regenerative Brayton cycle is defined as an improved version of the Brayton cycle that incorporates a regenerator to preheat the pressurized air from the ...
  90. [90]
    [PDF] Ericsson Cycle Gas Turbine Powerplants - RAND
    Both simple and regenerative Brayton cycles can be made to approach the Carnot cycle efficiency (73 percent for the temperatures considered in Figs. 2, 3, and 4) ...
  91. [91]
    Supercritical CO2 power cycles | netl.doe.gov
    This recovery of waste heat in the recuperators limits the heat rejection from the cycle and improves efficiency. Additionally, the ability to modulate the ...
  92. [92]
    [PDF] Performance Improvement Options for the Supercritical Carbon ...
    Overall, it is shown that the S-CO2 Brayton cycle efficiency can potentially be increased to 45 %, if a low temperature heat sink is available and incorporation ...
  93. [93]
    Parametric Analysis of the Kalina Cycle | J. Eng. Gas Turbines Power
    The Kalina Cycle utilizes a mixture of ammonia and water as the working fluid in a vapor power cycle. When the liquid mixture is heated the more volatile.
  94. [94]
    Variable Geometry Turbochargers - DieselNet
    The peak efficiency of a variable geometry turbine occurs at about 60% nozzle opening. It is usually comparable to or a few percent lower than that for a fixed ...Introduction · Applications of Variable... · Variable Geometry Turbine...Missing: heat | Show results with:heat
  95. [95]
    Bearings for thermal and kinetic energy recovery - ScienceDirect.com
    Barden is at the forefront of the latest developments in innovative, energy efficient, low friction, super precision ball bearings for these emerging ...
  96. [96]
    Friction Reduction and Reliability for Engines Bearings - MDPI
    However, for a low friction bearing solution, it is important to design or to validate new solutions with respect to running engine conditions and reliability ...
  97. [97]
    Combined Cycle Power Plants (CCPP) - Siemens Energy
    With over 64% efficiency, our Combined Cycle Power Plants (CCPP) utilize advanced gas and steam turbines to ensure flexible output and lower emissions.
  98. [98]
    Exergy Analysis and Second Law Efficiency of a Regenerative ...
    The effect of two heat additions, rather than one, in a gas turbine engine is analyzed from the second law of thermodynamics point of view.
  99. [99]
    From Fossil Fuels toward Renewable Sources. A Mini Review
    This study reviews different technologies for hydrogen production using renewable and non-renewable resources.
  100. [100]
    Nanoscale thermal transport at metal-semiconductor interfaces
    In metals, electrons are the main heat carriers while in semiconductors thermal transport is governed by phonons [7]. Thermal transport between metal (copper) ...Missing: engines | Show results with:engines
  101. [101]
  102. [102]
    Hybrid solar energy device for simultaneous electric power ...
    Sep 18, 2024 · This paper proposes a hybrid device combining a molecular solar thermal (MOST) energy storage system with PV cell.Missing: engines post-<|separator|>
  103. [103]
    Control and synchronization of rapid nanoscale DNA heat engine by ...
    Jul 30, 2025 · Under different frequencies of square wave temperature changes, we observe that the DNA hinge machine can switch between folded and unfolded ...
  104. [104]
    [PDF] From Brownian to deterministic motor movement in a DNA-based ...
    Jan 28, 2024 · Molecular devices that have an anisotropic, periodic potential landscape can be operated as Brownian motors. When the potential landscape is ...Missing: 2020-2025 | Show results with:2020-2025<|separator|>
  105. [105]
    Single-electron devices could manage heat flow in ... - Phys.org
    Dec 8, 2022 · RIKEN physicists have fabricated a nanoscale "heat engine" that uses a property of electrons known as spin as the effective working medium.
  106. [106]
    Learning the best nanoscale heat engines through evolving network ...
    Mar 8, 2021 · This allows for an alternative strategy to design the best heat engines by optimizing interactions instead of single-electron levels.
  107. [107]
    Autonomous quantum heat engine based on non-Markovian ...
    Apr 24, 2024 · We propose a recipe for demonstrating an autonomous quantum heat engine where the working fluid consists of a harmonic oscillator, the frequency of which is ...<|separator|>
  108. [108]
    Two-level masers as heat-to-work converters - PNAS
    We propose a paradigm of heat-powered maser. In contrast to textbook knowledge, it does not require population inversion or coherent driving.Missing: like | Show results with:like
  109. [109]
    Certifying quantum enhancements in thermal machines beyond the ...
    Oct 7, 2025 · Introducing a comparison with classical machines using the same set of thermodynamic resources, we show that for steady-state quantum thermal ...
  110. [110]
    Low-dimensional magnetocaloric materials for energy-efficient ...
    One of the most promising among them is magnetic refrigeration, which is based on the magnetocaloric effect (MCE) – a phenomenon in which a magnetic material ...Missing: magnetocalorics | Show results with:magnetocalorics
  111. [111]
    Performance optimization of thermochemical heat storage reactor ...
    Sep 15, 2025 · Thermochemical heat storage (TCHS) offers high energy density and long-term storage with minimal heat loss, providing a promising solution ...Missing: engines | Show results with:engines
  112. [112]
    Recent Advances in Nano-Engineered Thermochemical Energy ...
    Thermochemical energy storage (TCES) has gained significant attention as a high-capacity, long-duration solution for renewable energy integration, ...Missing: engines | Show results with:engines
  113. [113]
    Scramjet test another milestone in India's hypersonic weapons ...
    Jun 5, 2025 · India has taken another decisive step in hypersonic weapons development with the successful ground test of an active-cooled, scramjet subscale combustor.
  114. [114]
    Recent developments in technological innovations in scramjet engines
    This paper aims to review various challenges associated with various components used in the scramjet engines along with various research that has been done
  115. [115]
    Graphene for next-generation technologies: Advances in properties ...
    Challenges in graphene's scalability, cost, and integration must be addressed for widespread industrial adoption. •. Advancements in graphene-based composites ...
  116. [116]
    Magnetocaloric Refrigeration in the Context of Sustainability - MDPI
    Jul 21, 2024 · In this review, the physical basis of magnetic refrigeration is analysed, in order to propose this technology for industrial use.Missing: magnetocalorics | Show results with:magnetocalorics