Fact-checked by Grok 2 weeks ago

Carnot cycle

The Carnot cycle is a theoretical thermodynamic cycle that achieves the highest possible efficiency for converting heat into work in a heat engine operating between two fixed temperatures, consisting of four reversible processes: two isothermal and two adiabatic. Proposed by French engineer and physicist Nicolas Léonard Sadi Carnot in 1824, it was detailed in his seminal work Réflexions sur la puissance motrice du feu et sur les machines propres à développer cette puissance (Reflections on the Motive Power of Fire and on Machines Fitted to Develop That Power), which analyzed the efficiency of steam engines and laid the foundation for the second law of thermodynamics, based on the then-prevailing caloric theory of heat. In the cycle, an ideal working substance—such as a perfect gas—undergoes isothermal at the hot reservoir temperature T_h, absorbing Q_h; followed by adiabatic to the cold reservoir temperature T_c; then isothermal at T_c, rejecting Q_c; and finally adiabatic back to the initial . These processes are represented on a pressure-volume as a closed loop, with the area enclosed equaling the net work output. The efficiency \eta of the Carnot cycle is \eta = 1 - \frac{T_c}{T_h}, where temperatures are in , providing an upper bound for any reversible and demonstrating that depends solely on temperatures, not the . This formula implies no engine can convert all heat to work, as some must be rejected to the cold , a principle central to the second law of . Carnot's cycle, later formalized with concepts by and William Thomson () in the 1850s, underpins the analysis of real engines like internal combustion and systems, highlighting irreversibilities that reduce practical efficiencies below the Carnot limit. Its reversible nature makes it a for thermodynamic performance, influencing fields from power generation to climate modeling.

History

Sadi Carnot's Contributions

(1796–1832) was a French military engineer whose brief career bridged the and the early . Born in on June 1, 1796, as the son of , a renowned and revolutionary leader, Sadi entered the in 1812 and graduated in 1814 amid the turmoil of Napoleon's final campaigns. He then pursued advanced studies in military engineering at the École d'Application in , serving as an officer in the until his resignation in 1828, when he turned to independent engineering research. Disillusioned by France's technological lag behind in development, Carnot shifted focus to engineering research, driven by a patriotic desire to enhance national industrial competitiveness through improved efficiency. In 1824, at age 28, Carnot published his seminal work, Réflexions sur la puissance motrice du feu et sur les machines propres à développer cette puissance (Reflections on the Motive Power of Fire and on Machines Fitted to Develop that Power), a slim volume of 118 pages issued at his own expense in Paris, with 600 copies printed. Drawing on observations of steam engines, Carnot proposed an idealized theoretical cycle for a heat engine operating reversibly between two heat reservoirs at different temperatures, positing that this configuration yields the maximum possible motive power from heat. He envisioned the engine as a closed system where a working fluid undergoes a sequence of transformations, absorbing heat from a hot reservoir and rejecting it to a cold one, without specifying the fluid's nature to emphasize universality. Carnot's original framework centered on the principle that an engine's is bounded by the disparity between the reservoirs, independent of the processes or substances involved, and achievable only through reversible operations that avoid dissipative losses. He described the cycle's changes as isothermal expansions and compressions—where the interacts with the reservoirs at constant —interlinked by adiabatic transformations, in which no is exchanged. This , rooted in the then-prevailing yet presciently abstract, highlighted reversible processes as the cornerstone for attaining peak performance, laying groundwork for thermodynamic ideals without resolving heat's nature. Despite its insight, Carnot's treatise garnered little attention during his lifetime; he died of in 1832, with only a few copies sold during his lifetime. The work's significance emerged in the 1850s through independent rediscoveries: William Thomson (later ) encountered it in 1848 while studying heat engines and incorporated its principles into his absolute temperature scale by 1851, while Rudolf Clausius referenced it in his 1850 memoir on heat efficiency, adapting the cycle to articulate foundational thermodynamic principles. These efforts propelled Carnot's ideas into prominence, transforming his overlooked analysis into a pillar of .

Historical Context and Influences

The Carnot cycle emerged during the height of the , when were transforming economies in and . James Watt's 1769 patent for an improved , featuring a separate , boosted from the Newcomen engine's roughly 1% to about 2-3%, enabling broader applications in , , and transportation. By the early 19th century, the demand for even greater efficiency was acute in , where engineers sought to rival British industrial dominance amid resource constraints and military needs, prompting theoretical analyses of heat engines to optimize performance without excessive fuel consumption. Preceding Carnot's work, the dominated understandings of , positing it as an invisible, self-repellent that flows from hotter to colder bodies without creation or destruction, a concept advanced by in the late 18th century through experiments like the ice calorimeter co-developed with . Laplace extended this by treating as a conserved in , influencing early thermodynamic models, while Joseph Fourier's 1822 Théorie Analytique de la Chaleur established the mathematical framework for conduction via partial differential equations, independent of caloric assumptions yet providing tools that indirectly shaped Carnot's critique of inefficient engines and fallacies. These ideas framed as a transferable substance, allowing Carnot to model engine cycles without invoking mechanical equivalents initially. Following the 1824 publication of Réflexions sur la puissance motrice du feu, Carnot's ideas faced initial neglect and critique from contemporaries focused on practical , though they gained traction through Émile Clapeyron's 1834 graphical reformulation. By the mid-19th century, William Thomson (later ) in 1848 adapted Carnot's principle to define an absolute temperature scale, reconciling it with emerging laws, while in 1850 reformulated the cycle within a mechanical theory of , introducing the concept of and fully abandoning caloric in favor of heat-work equivalence. This integration marked the foundation of classical , emphasizing reversible processes over caloric flow. In a revealing 1830 on in engines, unpublished until 1878, Carnot began shifting toward these mechanical views, analyzing variable steam volumes and foreshadowing the first law of through detailed calculations.

Fundamental Concepts

Reversible Processes

A reversible thermodynamic process is one in which both the system and its surroundings can be returned to their exact initial states without any net change in the universe, achieved through a series of infinitesimal steps where the system remains in thermodynamic equilibrium at every stage. This contrasts with irreversible processes, where dissipation such as friction or uncontrolled heat transfer leaves permanent changes. Key characteristics of reversible processes include their quasi-static nature, meaning they proceed infinitely slowly to maintain , with no dissipative effects like or that would generate . Additionally, for state functions such as or , the integrals along the forward and reverse paths are equal, ensuring no net work or is lost to irreversibilities. Examples of reversible processes include the isothermal or of an , where heat exchange with a keeps constant while the system follows the equilibrium path defined by the . Another is a reversible , such as the frictionless of a gas in an insulated container with no heat loss to the surroundings, preserving constant . The mathematical foundation for reversible work in a process depicted on a pressure-volume (PV) diagram is given by the infinitesimal work element: \delta W = -P \, dV where P is the and dV is the volume change, integrated along the path to yield the maximum possible work for a given change. In contrast, an like free expansion into a performs no work (W = 0) despite a volume increase, as there is no opposing . Reversible processes serve as the ideal prerequisite for the Carnot cycle, which assumes them to achieve the theoretical maximum by eliminating all irreversibilities that would reduce useful work output.

Heat Reservoirs and the Second Law

Thermal reservoirs, also known as reservoirs, are idealized bodies with infinite that maintain a constant regardless of the amount of transferred to or from them. In the context of thermodynamic cycles like the Carnot cycle, two such reservoirs are typically involved: a hot reservoir at T_h serving as a heat source, and a cold reservoir at T_c (where T_h > T_c) acting as a sink. These reservoirs enable isothermal heat transfers, where the working fluid absorbs Q_h from the hot reservoir and rejects Q_c to the cold reservoir without altering the temperatures of the reservoirs themselves. The second law of thermodynamics imposes fundamental constraints on heat engines interacting with these reservoirs, establishing limits on the conversion of into work. The Kelvin-Planck statement asserts that it is impossible for any device operating in a to absorb from a single reservoir and convert it entirely into work without producing other effects. Similarly, the Clausius statement declares that it is impossible to construct a cyclic device whose sole effect is to transfer from a colder body to a hotter body without external work input. Together, these statements prohibit machines of the second kind, which would violate the directional flow of energy dictated by temperature differences. Sadi Carnot's 1824 analysis of engines operating between two reservoirs anticipated these limits, demonstrating that depends solely on the reservoir temperatures and setting an upper bound unachievable by real engines. This insight was later formalized as the second law by in the 1850s, who reconciled Carnot's with James Joule's work on as motion, integrating the concept into a broader framework of and irreversibility. A key concept emerging from these constraints is , a that quantifies the degree of irreversibility in thermodynamic processes. For reversible processes, the change in \Delta S is defined as: \Delta S = \int \frac{dQ_\text{rev}}{T} where dQ_\text{rev} is the infinitesimal reversible and T is the absolute . This definition highlights how increases in irreversible processes involving heat reservoirs, underscoring the second law's prohibition on reversing natural flows without work.

Cycle Description

The Four Stages

The Carnot cycle consists of four reversible processes executed with an as the in a , returning the gas to its initial state after one complete cycle and producing net positive work given by the W = \oint P \, dV > 0. These stages are enabled by reversible processes and interaction with two heat reservoirs at constant temperatures T_h (hot) and T_c (cold). Each stage is reversible, resulting in no net over the cycle. Stage 1: Isothermal Expansion. In the first stage, the undergoes reversible isothermal at the hot temperature T_h, absorbing Q_h from the while the increases from V_1 to V_2. The system performs work on the surroundings during this , with the remaining constant due to the isothermal condition for an . Stage 2: Adiabatic Expansion. The second stage involves reversible adiabatic expansion, during which no is transferred to or from the , and the gas decreases from T_h to T_c as the volume expands further to V_3. This cooling occurs through a decrease in the gas's , which is converted into additional work done by the . Stage 3: Isothermal Compression. In the third stage, the gas experiences reversible isothermal compression at the cold reservoir temperature T_c, rejecting heat Q_c to the cold reservoir as the volume decreases from V_3 to V_4. Work is done on the system by the surroundings during this compression, maintaining constant internal energy for the ideal gas. Stage 4: Adiabatic Compression. The final stage is reversible adiabatic compression, with no , raising the gas back from T_c to T_h as reduces to the initial value V_1. The increase in during this process is achieved through work input from the surroundings, completing the .

Thermodynamic Diagrams

The Carnot cycle is commonly represented graphically using thermodynamic diagrams that plot key state variables, aiding in the visualization of the cycle's processes and quantities such as work and heat transfer. The two primary diagrams are the pressure-volume (PV) diagram and the temperature-entropy (TS) diagram, each offering distinct insights into the cycle's behavior for an ideal working fluid like a gas. In the PV diagram, the Carnot cycle appears as a closed consisting of two isothermal processes depicted as curves and two adiabatic processes shown as steeper curves connecting them. The isothermal expansion occurs along a at the higher T_h, where decreases as increases while is absorbed; this is followed by an adiabatic expansion along a curve where no heat is exchanged, leading to a drop to T_c. The isothermal compression at T_c traces another with decreasing and heat rejected, and the adiabatic compression returns to the initial state along a steeper curve. For an , the adiabatic processes follow the PV^\gamma = \text{[constant](/page/Constant)}, where \gamma = C_p / C_v is the ratio of specific heats, resulting in curves that are steeper than the isotherms due to the absence of . The area enclosed by this represents the net work output of the cycle per unit mass or mole, providing a direct graphical measure of the engine's performance. The TS diagram presents the Carnot cycle as a rectangle, with temperature plotted vertically and entropy horizontally, offering a clearer view of heat transfers and entropy changes. The top horizontal line at T_h represents the isothermal expansion, spanning an entropy increase of \Delta S = Q_h / T_h, where Q_h is the heat absorbed; the right vertical line is the adiabatic expansion, where entropy remains constant (\Delta S = 0) as temperature decreases to T_c. The bottom horizontal line at T_c shows the isothermal compression with an entropy decrease of \Delta S = Q_c / T_c (where Q_c is the magnitude of heat rejected), and the left vertical adiabat returns to the starting entropy with no change. The heights of the rectangle correspond to T_h and T_c, while the widths reflect the normalized heat transfers Q_h / T_h and |Q_c| / T_c, which are equal in magnitude for the cycle. The enclosed area of the rectangle equals the net work, (T_h - T_c) \Delta S, but more importantly, the closed loop with vertical adiabats visually confirms the cycle's reversibility, as the net entropy change over the cycle is zero (\Delta S_\text{cycle} = 0), distinguishing it from irreversible processes that would show a net entropy production. These diagrams complement each other in analysis: the diagram is particularly useful for calculating work through the enclosed area and understanding volume and pressure variations during the four stages, while the TS diagram excels in illustrating balances and heat interactions, emphasizing the theoretical limits imposed by the second law. Both confirm the cycle's closure, returning to the initial after completing the processes.

Thermodynamic Analysis

Heat and Work Calculations

The analysis of the Carnot cycle often assumes an as the working substance, for which the U depends solely on , with the change given by \Delta U = n C_v \Delta T, where n is the number of moles, C_v is the at constant volume, and \Delta T is the change. For the isothermal processes in the cycle, \Delta T = 0, so \Delta U = 0. In the first stage, reversible isothermal expansion at the hot reservoir temperature T_h, the system absorbs heat Q_h from the reservoir while expanding from volume V_1 to V_2 > V_1. The heat transfer is Q_h = n R T_h \ln(V_2 / V_1), where R is the universal gas constant. Using the first law of (\Delta U = Q - W, with W as work done by the system), \Delta U = 0 implies the work done by the system W_1 = Q_h > 0. The second stage is reversible adiabatic expansion, with no heat transfer (Q_2 = 0). The temperature drops from T_h to the cold reservoir temperature T_c < T_h, so \Delta U_2 = n C_v (T_c - T_h) < 0. By the first law, the work done by the system is W_2 = -\Delta U_2 = n C_v (T_h - T_c) > 0. During the third stage, reversible isothermal compression at T_c, the system rejects heat Q_c to the cold reservoir while the volume decreases from V_3 to V_4 < V_3. The heat transfer is Q_c = n R T_c \ln(V_4 / V_3), which is negative due to the volume ratio being less than unity, and the magnitude |Q_c| < Q_h. With \Delta U_3 = 0, the work done by the system is W_3 = Q_c < 0, corresponding to work input to the system of magnitude |W_3| = |Q_c|. The fourth stage is reversible adiabatic compression, again with Q_4 = 0. The temperature rises from T_c to T_h, yielding \Delta U_4 = n C_v (T_h - T_c) > 0. The work done by the system is W_4 = -\Delta U_4 = n C_v (T_c - T_h) < 0, corresponding to work input to the system of magnitude |W_4| = n C_v (T_h - T_c). Over the complete cycle, the net work output is the sum of the stage works: W_\text{net} = W_1 + W_2 + W_3 + W_4 = Q_h + Q_c (noting Q_c < 0), or equivalently W_\text{net} = Q_h - |Q_c|. The total change in internal energy is zero (\Delta U_\text{cycle} = 0), as expected for a cyclic process, confirming the first law balance \oint \delta Q = \oint \delta W. Due to the reversibility of all processes, the cycle satisfies Q_h / T_h = - Q_c / T_c.

Efficiency Derivation

The thermal efficiency \eta of a heat engine is defined as the net work output divided by the heat input from the hot reservoir, \eta = W_\text{net} / Q_h. For a cyclic process, the first law of thermodynamics gives W_\text{net} = Q_h - |Q_c|, where |Q_c| is the magnitude of the heat rejected to the cold reservoir, yielding \eta = 1 - |Q_c| / Q_h. In the Carnot cycle, heat transfers occur isothermally at the hot reservoir temperature T_h and cold reservoir temperature T_c. For reversible processes, the second law implies that the total entropy change over the cycle is zero, \oint dQ_\text{rev} / T = 0. Since no heat is exchanged during the adiabatic stages, this reduces to Q_h / T_h + Q_c / T_c = 0. With Q_c negative, |Q_c| / T_c = Q_h / T_h, so |Q_c| / Q_h = T_c / T_h. Thus, the efficiency is \eta = 1 - T_c / T_h, with temperatures in Kelvin. This result assumes full reversibility in all stages; any irreversibility, such as friction or non-quasistatic processes, reduces the efficiency below this value. The formula depends solely on the reservoir temperatures, independent of the working fluid, establishing it as the theoretical maximum for heat engines between T_h and T_c. This derivation forms the basis for Carnot's theorem on efficiency limits. For an ideal gas working substance, the ratio can be confirmed via stage-specific calculations. During isothermal expansion at T_h, Q_h = n R T_h \ln(V_2 / V_1); during isothermal compression at T_c, |Q_c| = n R T_c \ln(V_3 / V_4). The adiabatic stages relate volumes such that \ln(V_2 / V_1) = \ln(V_3 / V_4), derived from the adiabatic condition T V^{\gamma - 1} = \text{constant}, giving T_h / T_c = (V_3 / V_2)^{\gamma - 1} = (V_4 / V_1)^{\gamma - 1}. This equality ensures |Q_c| / Q_h = T_c / T_h.

Significance

Carnot's Theorem

Carnot's theorem asserts that no heat engine operating between two thermal reservoirs at temperatures T_h (hot) and T_c (cold) can achieve an efficiency greater than that of a reversible engine, such as the , and that all reversible engines between the same reservoirs possess identical efficiency given by \eta = 1 - \frac{T_c}{T_h}, where temperatures are in absolute scale. This establishes the as the universal upper bound for heat-to-work conversion, independent of the working substance or cycle details. The theorem was originally conceived by Sadi Carnot in 1824 but formally proved and popularized by William Thomson (Lord Kelvin) in 1851, who built upon Carnot's principles to align them with the emerging second law of thermodynamics. Kelvin's contribution emphasized the theorem's implications for the impossibility of perpetual motion machines of the second kind. A classical proof proceeds by contradiction: suppose an engine A operates irreversibly between the same reservoirs with efficiency \eta_A > \eta (Carnot efficiency). Coupling A with a reversible Carnot engine B operated in reverse as a —absorbing from the cold reservoir and rejecting it to the hot one—yields a composite system where the net work input to B is less than the net work output from A, resulting in overall positive work from a single flow without increase, violating Clausius's statement of the second law that cannot spontaneously flow from cold to hot without work. Thus, no such engine A can exist, confirming the Carnot limit. An alternative proof invokes : in any irreversible engine, \Delta S > 0 within the system or surroundings reduces the available work compared to a reversible , where \Delta S = 0, thereby lowering the effective below the Carnot value. This entropic perspective, formalized post-Kelvin by Clausius, underscores that irreversibilities universally diminish performance. The theorem implies that the temperatures T_h and T_c alone dictate the absolute ceiling for any , rendering material choices and cycle configurations secondary to achieving reversibility.

Reversed Carnot Cycle

The reversed Carnot cycle functions as an ideal model for a or by executing the Carnot cycle processes in reverse order. It consists of four reversible stages: adiabatic compression of the , isothermal heat rejection to the hot at T_h, adiabatic , and isothermal heat absorption from the cold at T_c. In operation, the cycle absorbs heat |Q_c| from the cold during the isothermal at T_c and rejects heat Q_h to the hot during the isothermal compression at T_h. This process transfers heat from the lower- reservoir to the higher- one, opposing the natural thermal gradient, and requires a net work input W = Q_h - |Q_c| to drive the cycle. The performance of the reversed Carnot cycle is quantified by the (). For , the is defined as the ratio of absorbed from the cold reservoir to the work input: \text{COP}_R = \frac{|Q_c|}{W} For a reversible cycle, this yields \text{COP}_R = \frac{T_c}{T_h - T_c}. For pumping, the measures the delivered to the hot reservoir per unit work: \text{COP}_H = \frac{Q_h}{W} = \frac{T_h}{T_h - T_c}. These COP expressions derive from the second law of thermodynamics applied to reversible processes, where the change is zero over the cycle, implying \frac{Q_h}{T_h} = \frac{|Q_c|}{T_c}. Substituting into the work relation W = Q_h - |Q_c| gives W = |Q_c| \left( \frac{T_h}{T_c} - 1 \right), leading directly to the temperature-based formulas. The reversed Carnot cycle establishes the theoretical maximum for any or heat pumping device operating between two fixed temperatures; irreversible real-world systems can approach but never surpass this bound due to generation. This limit extends Carnot's to reversed operations.

Real-World Applications

Ideal versus Actual Engines

The Carnot cycle represents an idealized that operates under strict assumptions of reversibility, wherein all processes occur quasi-statically over infinite time to maintain , with no mechanical , perfect to eliminate leaks, and no dissipative losses. These conditions ensure that the cycle achieves the theoretical maximum for given reservoir temperatures, as dictated by the second law of thermodynamics. In contrast, actual heat engines encounter fundamental limitations from irreversibilities that deviate from these ideals. Mechanical friction in pistons, bearings, and turbines dissipates energy as , while unintended leaks occur across gradients due to imperfect . Operations at finite speeds introduce non-equilibrium effects, such as and gradients within the , and material constraints—such as strength limits at high temperatures or —restrict achievable conditions. These factors collectively reduce the conversion of to work, making real engines inherently less efficient than their Carnot counterparts. This efficiency gap is evident in practical applications: Otto cycle engines in gasoline vehicles typically attain 20-30% thermal efficiency, and Diesel engines achieve 30-40%, far below the Carnot limit of 50-70% for comparable hot reservoir temperatures around 800-1000 and cold reservoirs near 300 . For instance, steam turbines in conventional power plants operate at 30-40% , compared to a Carnot of about 60% under similar temperature conditions. The second law of thermodynamics remains valid for real engines, mandating that their is strictly less than the Carnot efficiency because irreversibilities generate (S_gen > 0), increasing the total entropy of the universe and limiting available work output. loss, representing the destruction of useful energy potential, and availability analysis further quantify these inefficiencies by assessing how much work could have been extracted under reversible conditions but is lost due to such .

Practical Approximations and Examples

Engineers approximate the Carnot cycle in practical heat engines through techniques such as multi-stage and to better mimic isentropic processes, regenerative heating to recover internally, and the use of high-temperature materials like nickel-based superalloys for blades to approach higher hot-reservoir temperatures. These methods reduce irreversibilities and elevate the average temperature during addition, thereby increasing closer to the Carnot limit without achieving perfect reversibility. In steam power plants, the Rankine cycle serves as a key approximation to the Carnot cycle, particularly through superheating the steam to raise the temperature of heat addition and avoid wet steam erosion in turbines, which enhances efficiency by making the process more akin to the isothermal expansion stage. The Stirling engine provides another near-reversible approximation via its internal regenerator, which stores and reuses heat during cyclic compression and expansion of a fixed gas volume, theoretically matching Carnot efficiency under ideal conditions. In automotive applications, the Otto cycle partially approximates the Carnot cycle through its isentropic compression and expansion phases, though constant-volume heat addition limits its efficiency compared to the ideal isothermal processes. Historically, the , patented in 1849 and widely adopted in the 1850s, improved efficiency by about 30% over conventional engines through advanced that enabled variable cutoff and better distribution, advancing practical performance toward theoretical limits like those of the Carnot cycle. In modern power generation, combined-cycle s integrate Brayton and Rankine cycles to achieve efficiencies exceeding 60%, approaching Carnot bounds by utilizing exhaust from the to drive a , with overall performance enhanced by high inlet temperatures. For , the in jet engines relies on isentropic approximations for compressor and stages, while cooling techniques—such as internal air channels and coatings—allow operation near material limits to sustain high temperatures and boost efficiency. Future trends in heat engines emphasize , including ceramics and nickel superalloys with improved resistance, to enable higher operating temperatures and further narrow the gap to Carnot ; for example, as of 2025, Carnot Engines has developed a hydrogen-fueled that achieves approximately 70% of the theoretical Carnot , demonstrated in a 40-day . However, the second law of imposes fundamental caps, preventing efficiencies from exceeding the theoretical maximum for given reservoir temperatures.

Theoretical Interpretations

Macroscopic Construct

The Carnot cycle, from a , models the working substance as a characterized by bulk thermodynamic properties such as P, V, T, and S. This approach treats the system as a continuous medium without regard to its underlying molecular structure, focusing instead on aggregate behaviors. The cycle is conceptualized as a closed sequence of four reversible processes—two isothermal and two adiabatic—that connect a series of states, ensuring that the system returns to its initial condition after one full loop. This macroscopic framework offers significant advantages for analysis at scales, where it provides reliable predictions of system performance by emphasizing average properties over microscopic fluctuations. By disregarding atomic-level details, the model simplifies computations and enables the design of practical engines that approximate behavior under assumptions. The formulation of the Carnot cycle relies on the foundational laws of classical : establishing equilibrium, conserving through \Delta U = [Q](/page/Q) - [W](/page/W), and the second law dictating the directionality of flow and increase. State functions like U, H = U + PV (relevant for considering potential open-system extensions), and S are employed to describe path-independent changes across the cycle's equilibrium states. As a , the Carnot cycle serves to empirically validate the second law by demonstrating that no can exceed its derived bound, \eta = 1 - \frac{T_C}{T_H}, thereby establishing an upper limit on heat-to-work conversion without invoking microscopic mechanisms. This macroscopic output underscores the cycle's role in setting theoretical benchmarks for real engines. However, the macroscopic construct of the Carnot cycle has limitations at small scales, where quantum coherence and finite-size statistical effects disrupt the assumptions, allowing deviations from classical bounds and second-law predictions.

Microscopic and Statistical Views

In , the Carnot cycle is interpreted as the collective behavior of an ensemble of particles, where macroscopic thermodynamic quantities arise from averages over microscopic configurations. The system is modeled as a large number of interacting particles following dynamics, with emerging as T = \frac{2}{3k_B} \left\langle \frac{p^2}{2m} \right\rangle, linking it directly to the average per degree of freedom, where k_B is Boltzmann's constant, p is , and m is mass. This probabilistic framework, developed in the late , shows how the cycle's isothermal and adiabatic processes correspond to controlled changes in the volume accessible to the particles, maintaining distributions like the Maxwell-Boltzmann at each stage. Microscopic reversibility underpins this view, as the equations of motion for individual particles are time-symmetric, yet the Carnot cycle exhibits apparent irreversibility on macroscopic scales, raising : how can irreversible thermodynamic behavior derive from reversible microscopic laws? The resolution lies in the dependence on initial conditions; typical starting microstates in lead to trajectories that rapidly explore high- regions, making reverse evolution statistically improbable without precise preparation. Thus, the Carnot cycle traces reversible paths in , but its directionality emerges from the overwhelming probability of forward evolution due to entropy maximization. Entropy provides the key probabilistic justification, defined microscopically by Boltzmann as S = k_B \ln \Omega, where \Omega counts the microstates consistent with a macrostate. In the Carnot cycle, the net entropy change \Delta S = 0 over the full loop implies equal \Omega for the hot and cold reservoir interactions in forward and reverse directions, preserving reversibility as a balance of microstate multiplicities. The fluctuation-dissipation theorem bridges microscopic fluctuations to macroscopic dissipation, revealing how small-scale random motions in the working fluid generate the irreversible entropy production that upholds the second law in the Carnot cycle, even though individual particle paths are reversible. Molecular dynamics simulations, treating the system as interacting atoms, approximate these cycles by evolving trajectories under controlled potentials, confirming that efficiency bounds hold amid thermal noise. Recent advances extend this to quantum regimes, where Carnot-like cycles operate in nanoscale devices such as engines, with work and heat defined via quantum expectation values; however, the remains capped at $1 - T_c / T_h, as quantum cannot evade the without violating the second . Studies from the 2020s, including implementations with superconducting qubits, demonstrate these engines achieving near-Carnot performance under weak coupling. For instance, in 2025, experimental realizations using superconducting qubits have demonstrated cyclic quantum heat engines and absorption refrigerators operating near the Carnot limit under controlled thermal environments. This highlights ' universality across scales.

References

  1. [1]
    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.
  2. [2]
    Carnot's Reflection on the Motive of Fire & Power - ASME
    Nicholas Sadi Carnot's Reflections on the Motive Power of Fire and on Machines Fitted to Develop that Power, published in France in 1824, was among the ...
  3. [3]
    Carnot Cycle - HyperPhysics
    The most efficient heat engine cycle is the Carnot cycle, consisting of two isothermal processes and two adiabatic processes.
  4. [4]
    4.5 The Carnot Cycle – University Physics Volume 2
    A hypothetical working cycle with the highest possible efficiency between the same two reservoirs, known now as the Carnot cycle.<|control11|><|separator|>
  5. [5]
    PHYS 200 - The Second Law of Thermodynamics and Carnot's Engine
    The Carnot heat engine is discussed in detail to show how there is an upper limit to the efficiency of heat engines and how the concept of entropy arises.
  6. [6]
    Lecture 5: Second Law of Thermodynamics - ESM Intranet Site
    The Carnot cycle is a reversible cycle that consists of an isotherm, an adiabat, a second isotherm, and a second adiabatic process to complete the cycle. The ...
  7. [7]
    Heat Engines: the Carnot Cycle - Galileo
    This simplest heat engine is called the Carnot engine, for which one complete heating/cooling, expanding/contracting cycle back to the original gas volume and ...
  8. [8]
    Sadi Carnot (1796 - 1832) - Biography - MacTutor
    He showed that the efficiency of the "Carnot engine" depends only on the temperature difference within the engine and not on the substance such as steam that ...
  9. [9]
    Nicolas Léonard Sadi Carnot, Father of Thermodynamics - ASME
    Apr 10, 2012 · Nicolas Léonard Sadi Carnot, the son of high-ranking military leader Lazare Nicholas Marguerite Carnot, was born in Paris in 1796.Missing: Revolution | Show results with:Revolution
  10. [10]
    [PDF] Sadi Carnot, 'Founder of the Second Law of Thermodynamics'
    Sadi Carnot had one of the shortest lifetimes, just 36 years, of any of the giants in the history of physics. Carnot's period was that of the Industrial ...
  11. [11]
    [PDF] Reflections on the motive power of heat and on machines fitted to ...
    These until lately unpublished notes of Carnot contain equally well- constructed arguments in favor of the now accepted theory of heat as energy. While ...Missing: summary | Show results with:summary
  12. [12]
    [PDF] REFLECTIONS ON THE MOTIVE POWER OF FIRE AND ON ... - ASME
    The thermal machines that existed in Carnot's time were simple steam engines in which heat was supplied to a boiler by burning coal or wood, converting water in ...Missing: reservoirs | Show results with:reservoirs
  13. [13]
    2024 'Key Reflections' on Sadi Carnot's 1824 'Réflexions' and ... - NIH
    Carnot's reasoning proves that a reversible cycle cannot have a lower efficiency (power output relative to heat input) than any other cycle, thus all reversible ...Missing: summary | Show results with:summary
  14. [14]
    (PDF) 2024 Key Reflections on the 1824 Sadi Carnot's “Réflexions”
    Feb 13, 2024 · 1. “Reversible and Reverse” Processes and Cycles Dissected · 2. Maximum E ciency and “Reversible Equivalency” Scrutinized · 3. Reversible ...Missing: summary | Show results with:summary
  15. [15]
    Reflections of Sadi Carnot - arXiv
    Aug 23, 2025 · Motive power is Carnot's term for the mechanical work performed by a heat engine. The term fire refers to the heat involved in producing ...Missing: reversible | Show results with:reversible
  16. [16]
    How Did We Get Here? The Tangled History of the Second Law of ...
    Jan 31, 2023 · Kelvin recognized, though, that Carnot had chosen to look at the particular (“equilibrium thermodynamics”) case of processes that occur ...Missing: overlooked | Show results with:overlooked
  17. [17]
    How Kelvin and Clausius discovered Carnot's ideas - carnotcycle
    Aug 4, 2012 · ... Carnot's seminal work – in which lay the seeds of the second law of thermodynamics – into the spotlight of publicity. This popular account ...
  18. [18]
    Sadi Carnot and the second law of thermodynamics - ResearchGate
    Aug 10, 2025 · Carnot, as a pioneer of thermodynamics [3,5,6], did not know yet either the law of conservation of energy or the I law of thermodynamics [7] , ...
  19. [19]
    [PDF] 2. The British Industrial Revolution, 1760-1860
    Watt's first engine improved that to about 2.7% efficiency. ... Efficiency improvements continued in the nineteenth century, especially in France which had few.
  20. [20]
    [PDF] Lecture 6 History of Energy Use II
    Gradual improvements in engine efficiency. 0.6%. 6%. Page 18. First Watt steam engine: James Wac, 1769 patent. (1774 producCon model). Higher efficiency than ...
  21. [21]
    A History of Thermodynamics: The Missing Manual - PMC
    In 1783, Lavoisier and Laplace invented the ice calorimeter [35]. ... Sadi Carnot's father was “hero of the Republic” Lazare Carnot, the general ...
  22. [22]
    [PDF] Theory of Heat
    THE AIM of this book is to exhibit the scientific connexion of the various steps by which our know- ledge of the phenomena of heat has been extended.
  23. [23]
    [PDF] Sadi Carnot, 'Founder of the Second Law of Thermodynamics'
    After the publication of his. Réflexions he continued his research, and his Notes [2, 3] indicate that he abandoned the caloric theory for the mechanical theory ...Missing: post- Coriolis 1850 1848
  24. [24]
    [PDF] LECTURE NOTES ON THERMODYNAMICS
    May 17, 2025 · These are lecture notes for AME 20231, Thermodynamics, a sophomore-level undergraduate course taught in the Department of Aerospace and ...Missing: critiques | Show results with:critiques
  25. [25]
    Nicolas Leonard Sadi Carnot | Encyclopedia.com
    May 11, 2018 · These notes anticipated nearly all of the groundwork for the first law of thermodynamics. They remained undiscovered and unpublished until 1878, ...
  26. [26]
    4.1 Reversible and Irreversible Processes - UCF Pressbooks
    A reversible process is a process in which the system and environment can be restored to exactly the same initial states that they were in before the process ...
  27. [27]
    The Second Law of Thermodynamics - Florida State University
    A reversible process is a process that can be reversed in direction without leaving any trace on the surroundings. A simple example of a reversible process is a ...
  28. [28]
    4.3 Features of reversible processes - MIT
    Reversible processes are idealizations or models of real processes. One familiar and widely used example is Bernoulli's equation, which you saw in Unified. They ...
  29. [29]
    [PDF] Thermodynamic reversibility - University of Pittsburgh
    Sep 3, 2016 · This is the notion of thermodynamically reversible or quasi-static processes. They are used in the definition of entropy and are distinguished ...
  30. [30]
    [PDF] Lecture452_4_14_reversible-work.pdf - University of Washington
    Jun 30, 2014 · Reversible changes involve passage through a series of equilibrium states. • Example: Isothermal expansion of an ideal gas reversiblework w. P ...
  31. [31]
    [PDF] T dq ds ≡ T dq ds ≥
    A system process is defined as reversible if a system, after having experienced several transformations, can be returned to its original state without ...
  32. [32]
    [PDF] Entropy - Ken Intriligator
    /dW ≤ pdV Equality iff process is reversible. /dQ ≤ T dS Equality iff process is reversible. dE = /dQ − /dW = T dS − pdV. These equalities are whether or ...
  33. [33]
    reversible and irreversible processes, entropy and introduction ... - MIT
    Maximum work is achieved during a reversible expansion (or compression). For example, suppose we have a thermally insulated cylinder that holds an ideal gas.
  34. [34]
    Carnot Engine
    The Carnot engine -- or the Carnot cycle -- is important because it describes a heat engine that uses reversible processes that can be handled theoretically.
  35. [35]
    6.3 The second law of thermodynamics: Kelvin-Planck and Clausius ...
    Clausius statement: it is impossible to construct a device that operates in a cycle and produces no effect other than the transfer of heat from a lower- ...
  36. [36]
    The Second Law: From Carnot to Thomson-Clausius, to the Theory ...
    Contribution by Lazare and Sadi Carnot to the caloric theory of heat and its inspirational role in alternative thermodynamics. J. Therm. Anal. Calor. 2009 ...Missing: post- Coriolis
  37. [37]
  38. [38]
    Carnot Cycle
    The cycle begins with a gas ... During the process from State 3 to State 4 heat is transferred from the gas to heat source to maintain the temperature.Missing: four | Show results with:four
  39. [39]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics 3.8 ...
    Dec 15, 2021 · (3) Since each of the processes in a Carnot cycle is internally reversible, the entire Carnot cycle is also internally reversible.
  40. [40]
    [PDF] Efficiency of a Carnot engine
    The Carnot cycle makes an engine. The p-V diagram below sketches the operation of a Carnot engine, where the “working fluid” that expands and contracts within ...
  41. [41]
    [PDF] The Carnot cycle Handout We saw in A gas heat engine, continued ...
    We saw in A gas heat engine, continued that our rectangular cycle in a pV diagram had an efficiency of 15%, where the Carnot efficiency was 75%.Missing: explanation | Show results with:explanation
  42. [42]
    4.5 The Carnot Cycle – General Physics Using Calculus I
    The figure shows four steps of Carnot cycle, namely isothermal expansion, adiabatic expansion, Figure 4.11 The four processes of the Carnot cycle. The working ...Missing: stages | Show results with:stages
  43. [43]
    [PDF] V. The Second Law of Thermodynamics
    If the Carnot heat engine is operated in reverse, we call it a. Carnot refrigeration cycle. •1-2 rev., isothermal expansion. •2-3 rev., adiabatic expansion. •3- ...
  44. [44]
    [PDF] Thermodynamics equation
    CARNOT REFIGERATION CYCLE. COPR,REV = 1. TL/TH − 1. CARNOT REFRIGERATOR. CARNOT HEAT PUMP: COPHP,REV = 1. 1 − TL/TH. QH. Q. L REV. = TL. TH. CARNOT HEAT ENGINE:.
  45. [45]
    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 ...
  46. [46]
    [PDF] Thermal efficiency - McGraw Commons
    Mar 27, 2013 · There is an overall theoretical limit to the efficiency of any heat engine due to temperature, called the Carnot efficiency. Second, specific ...
  47. [47]
    Gasoline Engine - General Physics II
    of 1.4 gives a theoretical efficiency of 56%. That is, the efficiency of an Otto cycle engine would be 56%. Measured efficiencies of real internal ...<|separator|>
  48. [48]
    Steam Generator Efficiency - Power Engineering
    Jun 1, 2007 · The efficiency calculates to 44.5 percent, which is 2 percent higher than the non-reheat example. The increased fuel requirement is ...
  49. [49]
    Entropy and Efficiency of Heat Engines | Physics Van | Illinois
    If everything works ideally, you get the Carnot efficiency: W/Q = 1- TC/TH. However, things don't work ideally, there's always some net production of entropy, S ...
  50. [50]
    Exergy Loss - an overview | ScienceDirect Topics
    The exergy losses are a thermodynamic metric to detect the inefficient energy use in any process due to system irreversibility.
  51. [51]
    Turbine Engine Thermodynamic Cycle - Brayton Cycle
    The Brayton cycle analysis is used to predict the thermodynamic performance of gas turbine engines. The EngineSim computer program, which is available at this ...
  52. [52]
    [PDF] Improving Free-Piston Stirling Engine Power Density
    the same efficiency as the ideal Stirling engine, which is equal to the Carnot efficiency. ... Schmidt cycle while achieving the Carnot efficiency. Ideal motion ...
  53. [53]
    [PDF] HIGH-TEMPERATURE HIGH-STRENGTH NICKEL-BASE ALLOYS
    The material presented in this publication has been prepared for the general information of the reader and should not be used or relied on for specific ...
  54. [54]
    8.6 Enhancements of Rankine Cycles - MIT
    A comparison of the Carnot cycle and the Rankine cycle with superheating is given in Figure 8.14. The maximum and minimum temperatures are the same, but the ...Missing: approximation via
  55. [55]
    Theory of Otto Cycle - Gasoline Engine - Nuclear Power
    In contrast to the Carnot cycle, the Otto cycle does not execute isothermal processes because these must be performed very slowly. In an ideal Otto cycle, the ...
  56. [56]
    A General-Purpose Technology at Work: The Corliss Steam Engine ...
    Mar 29, 2004 · This article examines the role that a particular design improvement in steam power, embodied in the Corliss engine, played in the growth of the ...
  57. [57]
    MODERN GAS TURBINE COMBINED CYCLE
    Nov 1, 2013 · At 60+% net thermal efficiency (officially clocked in a commercial installation in 2011), it is ten percentage points ahead of its nearest ...
  58. [58]
    Models for Predicting the Performance of Brayton-cycle Engines
    The performance of gas-turbine engines is the result of choices of type of cycle for the application, cycle temperature ratio, pressure ra- tio, cooling flows ...
  59. [59]
    What are some recent advances in high-temperature materials for jet ...
    Good impact resistance and stability at high operating temperatures make the silicon carbide (SiC)/SiC ceramic matrix composite system a desirable option for ...
  60. [60]
    The Laws of Thermodynamics and Limits on Engine Efficiency
    Unlike the first law, there are many ways of writing this Second Law: our wording above is a paraphrase of Clausius' original formulation. ... Carnot refrigerator ...
  61. [61]
  62. [62]
    The Carnot Cycle, Reversibility and Entropy - PMC - PubMed Central
    As the piston moves, the system maps out a trajectory through the equilibrium states of thermodynamic phase space. The change in entropy of the states ...Missing: continuum | Show results with:continuum
  63. [63]
    4 Classical Thermodynamics‣ Statistical Physics by David Tong
    In this section, we're going to take a step back and look at classical thermodynamics. This is a theory that cares nothing for atoms and microscopics.Missing: limitations | Show results with:limitations
  64. [64]
    Thermodynamic Foundations – Introduction to Aerospace Flight ...
    Q_H/T_H , while rejected heat to the cold reservoir carries entropy Q_C ... The Carnot cycle establishes the maximum possible efficiency of any heat ...
  65. [65]
    5. Thermodynamics — Introduction to Statistical Mechanics
    In this chapter, I will discuss some of the most important results of classical thermodynamics as seen from a modern statistical viewpoint. 5.1. Thermodynamic ...
  66. [66]
    [PDF] LECTURE NOTES ON INTERMEDIATE THERMODYNAMICS
    Mar 28, 2025 · (Adapted from BS) Consider an ideal gas Carnot cycle with air in a piston cylinder with a high temperature of 1200 K and a heat rejection at ...
  67. [67]
    Carnot Cycle - an overview | ScienceDirect Topics
    The Carnot cycle is a Gedankenexperiment that represents an idealized reversible process to rationalize the function of a steam or combustion engine.Missing: continuum | Show results with:continuum
  68. [68]
    Revisiting Thermodynamic Efficiency - Physical Review Link Manager
    Feb 11, 2013 · This Carnot limit is simply a statement of the second law of thermodynamics. On a practical level, knowing the ideal efficiency of engines ...Missing: thought | Show results with:thought
  69. [69]
    Fundamental limitations for quantum and nanoscale thermodynamics
    Jun 26, 2013 · We find that there are fundamental limitations on work extraction from non-equilibrium states, owing to finite size effects and quantum coherences.
  70. [70]
    [PDF] Statistical Mechanics I - MIT OpenCourseWare
    A Carnot Engine is any engine that is reversible, runs in a cycle, with all of its heat exchanges taking place at a source temperature TH, and a sink ...
  71. [71]
    Macroscopic irreversibility and microscopic paradox: A Constructal ...
    Oct 20, 2016 · Loschmidt objected that this result is inconsistent, because any irreversible process cannot be obtained by using a time-symmetric dynamics1. ...
  72. [72]
    [PDF] THE EVOLUTION OF CARNOT'S PRINCIPLE*
    We have followed the evolution of Carnot's principle, via Kelvin's perception that it de nes a universal temperature scale, Clausius' discovery that it implied ...
  73. [73]
    Finite-time thermodynamics of fluctuations in microscopic heat engines
    Jan 31, 2022 · Applying our framework to the Carnot cycle, it has been found that both the average and the fluctuation of the dissipation can be minimized ...
  74. [74]
    Efficiency at Maximum Power of a Carnot Quantum Information Engine
    Jun 16, 2023 · We concretely introduce a generalized finite-time Carnot cycle for a quantum information engine and optimize its power output in the regime of low dissipation.Missing: 2020s | Show results with:2020s
  75. [75]
    Entropy-based analysis of single-qubit Otto and Carnot heat engines
    May 2, 2025 · We investigate the performance of Otto-like and Carnot-like engines for a single-qubit working medium.Missing: 2020s | Show results with:2020s