Fact-checked by Grok 2 weeks ago

Thermodynamic cycle

A thermodynamic cycle is a sequence of thermodynamic processes in which the working substance undergoes a series of changes in state, returning to its initial at the completion of the cycle, thereby allowing for the periodic conversion of into mechanical work or the reverse process. This closed-loop nature distinguishes thermodynamic cycles from open processes and forms the basis for analyzing energy transformations in systems. Thermodynamic cycles are fundamental to the operation of heat engines, refrigerators, and heat pumps, where they facilitate efficient energy transfer while adhering to the , particularly the first law () and the second law (limiting due to increase). The of a cycle, often expressed as the ratio of net work output to input, is maximized in ideal reversible cycles like the , which sets the theoretical upper limit for any operating between two temperatures. Real-world cycles approximate these ideals but account for irreversibilities such as and losses, influencing design choices in practical applications. Thermodynamic cycles are broadly classified into power cycles, which produce net work from heat (e.g., Rankine for steam turbines, Brayton for gas turbines, and for internal combustion engines), and refrigeration cycles, which transfer heat from low to high temperatures using work input (e.g., vapor-compression for ). These cycles underpin diverse technologies, from generation and automotive to cooling systems and aerospace , driving advancements in and .

Fundamentals of Thermodynamic Cycles

Definition and Basic Principles

A thermodynamic cycle consists of a series of thermodynamic processes through which a passes, returning it to its initial , thereby forming a closed in the state space. This cyclical path ensures that the working substance experiences no net change in its properties at the completion of the cycle, allowing for repeatable operation. Prerequisite to understanding cycles are the concepts of s and states: a is a defined region of under study, which may be closed (with fixed and no material exchange across boundaries, only ) or open (permitting flow in addition to energy exchange), while an state is one in which the 's properties are uniform and do not change spontaneously over time, encompassing mechanical, thermal, and . In engineering applications, are fundamental to devices that convert into mechanical work or vice versa, enabling continuous operation in systems such as engines, refrigerators, and power plants without requiring a net alteration in the system's internal . By facilitating the absorption of from a high-temperature source, partial conversion to work, and rejection of the remainder to a low-temperature sink, cycles underpin efficient utilization in practical machinery. This repetitive is essential for sustained performance, as it allows the to indefinitely, optimizing resource use in thermal systems. Thermodynamic cycles are typically visualized on diagrams such as pressure-volume (P-V) or temperature-entropy (T-S) plots, where the processes trace a closed path representing the sequence of state changes. On a P-V diagram, the area enclosed by the cycle corresponds to the net work output (or input) of the cycle, providing a graphical measure of energy conversion efficiency. Similarly, the T-S diagram highlights heat transfers, with the enclosed area relating to the net heat interactions, aiding in the analysis of cycle performance and irreversibilities. These representations leverage the fact that state functions like pressure, volume, temperature, and entropy fully describe equilibrium states along the path.

Heat, Work, and Energy Transfer

In thermodynamic cycles, (Q) and work (W) represent distinct modes of energy transfer across the boundary. is the transfer of due to a difference between the and its surroundings, occurring through random molecular motions without macroscopic displacement. In contrast, work involves organized motion, typically arising from the or compression of the , such as in piston-cylinder arrangements where forces act over distances. Both are path-dependent quantities, meaning their values depend on the specific sequence of states traversed during the cycle, unlike state functions such as . The first law of thermodynamics, which expresses the , applies directly to thermodynamic cycles. For a complete cycle, the change (ΔU) returns to zero because the system starts and ends in the same state, leading to the relation Q_net = W_net, where Q_net is the net and W_net is the net work. This equality highlights that the net energy input as must equal the net energy output as work over the cycle. The work for a process within the cycle is calculated as W = ∫ P dV, integrating the over the volume change. For reversible processes, the is given by Q = ∫ T dS, where T is temperature and dS is the infinitesimal change. For heat engines operating on thermodynamic cycles, thermal efficiency (η) quantifies performance as η = W_net / Q_in, where Q_in is the absorbed from the high-temperature source. This metric indicates the fraction of input converted to useful work, with the remainder rejected as Q_out. Sign conventions in these analyses define positive Q as added to the and positive W as work done by the on the surroundings, consistent with formulation ΔU = Q - W. These conventions ensure consistent tracking of energy flows in cycle analyses.

Common Thermodynamic Processes

Thermodynamic processes represent the fundamental changes in state that a undergoes, serving as the building blocks for constructing thermodynamic cycles in engines, refrigerators, and other devices. These processes are characterized by constraints on variables such as , , , or energy transfer, and they are analyzed using the , particularly relating heat, work, and changes. For ideal gases, many processes admit simple analytical expressions derived from the and specific heat capacities. An occurs at constant , where addition or removal leads to changes in and while remains fixed. In such a process, the work done by the is given by W = P \Delta V, where P is the constant and \Delta V is the change in ; for an , this equals P(V_2 - V_1). The transfer is Q = \Delta H = m c_p \Delta T, reflecting the change, with c_p as the specific heat at constant and m as . are common in open systems like combustion chambers, where expansion at fixed converts to work. An maintains constant , often achieved through heat exchange with a , resulting in no change in for an (\Delta U = 0). For an , the pressure-volume relation follows PV = constant, and the work done is W = nRT \ln(V_2 / V_1), where n is the number of moles, R is the , and T is the constant ; the absorbed equals the work, Q = W. This process exemplifies reversible in idealized cycles, balancing compression or expansion without temperature variation. An involves no (Q = 0), so changes in directly equal work done, \Delta U = -W. For a reversible with an , the relation is PV^\gamma = constant, where \gamma = c_p / c_v is the ratio of specific heats; equivalently, TV^{\gamma-1} = constant or T P^{1-\gamma}/\gamma = constant. The work is W = \frac{P_1 V_1 - P_2 V_2}{\gamma - 1} for reversible cases. Adiabatic processes model insulated expansions or compressions in turbines and compressors. An , or constant-volume process, features fixed volume (V = constant), resulting in zero work (W = 0) since no occurs. The equals the internal energy change, Q = \Delta U = m c_v \Delta T, where c_v is the specific heat at constant volume. and vary proportionally via the , making this process relevant for heating in closed vessels without mechanical work. A throttling process is an isenthalpic expansion (h_1 = h_2) through a restriction like a , with no or work, leading to a and potential change. For steady flow, simplifies to constant , and the Joule-Thomson \mu = \left( \frac{\partial T}{\partial P} \right)_h determines the shift, which is zero for gases but nonzero for real gases due to intermolecular forces. This is essential in refrigeration cycles for expanding refrigerants. Thermodynamic processes are classified as reversible or irreversible based on generation. A reversible is quasi-static, maintaining at every stage with no or dissipation, resulting in zero change for the and surroundings (\Delta S = 0); it represents the theoretical maximum , as all changes can be undone without net effects. In contrast, an irreversible involves nonequilibrium phenomena like or unrestrained expansion, generating (\Delta S > 0) and reducing available work, as seen in real devices where losses degrade performance. The distinction implies that cycles composed of reversible processes achieve ideal limits, while irreversibilities in actual s lower .
ProcessConstraintKey Relation (Ideal Gas)Work WHeat Q
IsobaricConstant PP = constP \Delta Vm c_p \Delta T
IsothermalConstant TPV = constnRT \ln(V_2 / V_1)= W (since \Delta U = 0)
AdiabaticQ = 0PV^\gamma = const\frac{P_1 V_1 - P_2 V_2}{\gamma - 1} (rev.)0
IsochoricConstant VV = const0m c_v \Delta T
ThrottlingConstant hh_1 = h_200
These processes combine to form closed thermodynamic cycles, such as those in heat engines, by sequencing state changes that return the system to its .

Classification of Thermodynamic Cycles

Power Cycles

Power cycles are thermodynamic cycles engineered to convert into mechanical work, characterized by a net positive work output that exceeds any required input work, and they form the operational basis of heat engines. These cycles operate by absorbing from a high-temperature source and rejecting a portion to a low-temperature sink, with the difference enabling useful work production. The general structure of power cycles encompasses sequential processes: heat addition at elevated temperatures to increase the of the , followed by that extracts mechanical work; subsequent heat rejection dissipates excess , and prepares the fluid for the next by requiring work input. This framework ensures the system returns to its initial state after one complete , allowing continuous operation. Applications of power cycles span diverse technologies, including internal combustion engines for vehicular propulsion, steam turbines in , and gas turbines for both power production and . Efficiency in power cycles is constrained by fundamental thermodynamic principles, with the Carnot efficiency serving as the theoretical upper limit determined by the temperature ratio of the heat source and , beyond which no real engine can operate. Power cycles are further distinguished as open or closed based on management: closed cycles recirculate a sealed , such as vapor in systems, maintaining constant ; open cycles, like those in gas turbines, involve continuous intake and exhaust of the , often approximated using air-standard assumptions for analysis. Historically, power cycles evolved significantly in the amid the , with early steam engines driving mechanization; the , a cornerstone for vapor power systems, was systematically described by Scottish engineer William John Macquorn Rankine in his 1859 Manual of the Steam Engine and Other Prime Movers, providing the thermodynamic foundation for efficient operations.

Refrigeration and Heat Pump Cycles

Refrigeration and heat pump cycles are thermodynamic cycles that operate as reversed heat engines, using external work input to transfer heat from a low-temperature reservoir to a higher-temperature one, thereby achieving a cooling effect in the low-temperature space. These cycles violate the natural direction of heat flow dictated by the second law of thermodynamics without work input, making them essential for cooling and heating applications. The performance of a refrigeration cycle is quantified by its coefficient of performance (COP), defined as the ratio of heat absorbed from the cold reservoir (Q_c) to the work input (W): \text{COP}_R = \frac{Q_c}{W}. The vapor-compression cycle is the most common type of refrigeration cycle, consisting of four main components: a , a , an (or ), and an . In the cycle, the enters the as a low-pressure vapor and is compressed to and (process 1-2), increasing its ability to release . The hot vapor then flows to the , where it rejects to the surroundings and condenses into a (process 2-3). The passes through the , undergoing a throttling that reduces its and (process 3-4), before entering the . In the , the low-pressure absorbs from the cooled space and evaporates back into vapor (process 4-1), completing the . Absorption cycles differ from vapor-compression by using as the input rather than work, employing a binary mixture of a and an absorbent. A typical pair is as the and as the absorbent, where the process involves and desorption driven by temperature differences. In the absorber, the vapor dissolves into the absorbent, releasing and forming a strong ; this is then heated in a to desorb the vapor, which is condensed and evaporated similarly to vapor-compression but without a . The weak returns to the absorber after , enabling heat-driven operation suitable for or sources. Heat pump cycles utilize the same fundamental processes as cycles but emphasize heat delivery to the high-temperature for heating purposes, such as space or . The COP for a is defined as the ratio of heat rejected to the hot (Q_h) to the work input: \text{COP}_{HP} = \frac{Q_h}{W}. Since Q_h = Q_c + W, the heating COP is always greater than the refrigeration COP by unity for the same cycle. These cycles find widespread applications, including household refrigerators and freezers for , air conditioners for building cooling, and industrial chilling systems for processes like or chemical . are commonly used for residential and commercial heating, leveraging ambient air, ground, or water sources. The maximum possible COP for reversible and cycles is given by the Carnot limits: for , \text{COP}_{R,\text{Carnot}} = \frac{T_c}{T_h - T_c}, and for heating, \text{COP}_{HP,\text{Carnot}} = \frac{T_h}{T_h - T_c}, where T_c and T_h are the absolute temperatures of the cold and hot reservoirs, respectively. Real vapor-compression systems achieve COP values of 3-5 for typical conditions (e.g., T_c = 5^\circC, T_h = 35^\circC), which is approximately 30-55% of the Carnot COP due to irreversibilities like inefficiencies and drops. cycles typically have lower COPs, around 0.5-0.7, reflecting their reliance on lower-grade heat inputs.

Ideal Versus Real Cycles

Assumptions in Ideal Cycles

In the analysis of ideal thermodynamic cycles, several simplifying assumptions are employed to facilitate theoretical calculations of performance metrics such as and net work output. These include the treatment of all processes as quasi-static and reversible, meaning they occur infinitely slowly with no dissipative effects like or throttling, allowing the system to remain in at every stage. Additionally, no heat losses to the surroundings are assumed, so occurs only between the working fluid and designated thermal reservoirs during specified processes. For ideal gas power cycles (e.g., air-standard models of , , and Brayton cycles), the is modeled as an obeying the equation of state PV = nRT, where P is , V is volume, n is the number of moles, R is the universal , and T is absolute ; specific heats at constant (c_p) and constant volume (c_v) are taken as constant, independent of , which simplifies and changes to linear functions of . In contrast, for vapor power cycles like the ideal , real fluid properties from tables (e.g., steam tables) are used, accounting for phase changes and variable properties. A particularly common framework is the air-standard , where air is assumed to be the with constant composition and properties matching those of dry air at conditions (e.g., molecular weight of 29 g/mol, c_p = 1.005 kJ/kg·K, c_v = 0.718 kJ/kg·K). Under these assumptions, is idealized as external addition at constant or , without chemical reactions altering the fluid's composition, and the is closed with the same mass of air recirculating. This model applies broadly to gas power cycles like Otto and Diesel, replacing complex fuel-air interactions with from an infinite-capacity source. These assumptions enable closed-form analytical solutions for cycle performance. For instance, the of an reversible operating between high-temperature T_H and low-temperature T_L is derived as \eta = 1 - \frac{T_L}{T_H}, representing the maximum possible bounded by the second of . Such derivations rely on the reversibility assumption to equate input and output via balances, yielding expressions for net work as W_{net} = Q_H (1 - \frac{T_L}{T_H}), where Q_H is the absorbed at T_H. For air-standard with constant specific heats, further simplifies to functions of or pressure ratio, such as \eta = 1 - \frac{1}{r^{k-1}} for the , where r is the and k = c_p / c_v. While these idealizations provide valuable benchmarks for understanding fundamental limits, they inherently overestimate real-world performance by neglecting irreversibilities like fluid friction, heat leaks, and variable specific heats, which reduce actual efficiencies below theoretical values (e.g., real Rankine cycles achieve 30-42%, Brayton 30-40%, 20-35%).

Deviations and Modeling in Real Systems

In real thermodynamic cycles, deviations from ideal models arise primarily from irreversibilities such as in moving components, losses across finite differences, behavior of non-ideal gases under high s or s, drops due to fluid flow resistances, and the inherent limitations of finite-time processes that prevent quasi-static conditions. These factors introduce generation, reducing the overall reversibility and performance compared to idealized assumptions of isentropic compression/expansion and isothermal . To model these real-world effects, engineers employ approaches like the mean effective pressure (MEP), which quantifies the average pressure exerted on the piston during a cycle to assess net work output relative to displacement volume, aiding comparisons across engine designs. Polytropic processes, described by the relation PV^n = \constant, where n is the polytropic index (typically between 1 and \gamma for gases, accounting for heat transfer and friction), provide a more accurate representation of compression and expansion than purely isentropic paths. Exergy analysis further evaluates losses by tracking the available work potential degraded by irreversibilities, highlighting inefficiencies in heat and work transfers. These deviations significantly lower cycle compared to ideal predictions; for instance, in Rankine cycles, pump work, which is typically about 1% of work (leading to a net reduction of around 0.3-1%), plus additional losses from and drops, consumes a portion of the output. Similarly, in Otto cycles, various irreversibilities, including inefficiencies from incomplete fuel burning and heat losses to cylinder walls, contribute to real thermal efficiencies of typically 20-35%, compared to ideal air-standard values of 52-60% for common compression ratios of 8-12, representing an overall reduction of 25-40%. Correction methods mitigate these losses through regenerative cycles, which recover via internal heat exchangers to preheat working fluids, boosting by 5-20% in applications. Multi-staging, such as intercooling in compression or reheat in , reduces work input and extremes, improving overall in large-scale systems. Advanced simulations using (CFD) model complex fluid behaviors, pressure gradients, and heat flows to optimize component designs and predict real efficiencies with . In general, these real-system deviations typically reduce power output by 20-60% and increase fuel consumption by 30-150% compared to ideal cycles, with the exact figures varying by cycle type (e.g., smaller for Rankine, larger for ), underscoring the need for precise modeling to enhance utilization in applications.

Key Examples of Thermodynamic Cycles

Carnot Cycle

The , proposed by French engineer Sadi Carnot in his 1824 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), represents an idealized, reversible thermodynamic cycle that serves as the theoretical benchmark for the maximum efficiency of heat engines and refrigerators operating between two thermal reservoirs. Carnot's analysis focused on the fundamental limits of converting heat into mechanical work, assuming no dissipative losses such as friction or across finite temperature differences, thereby laying the groundwork for the second law of thermodynamics. The cycle comprises four reversible processes executed with an ideal working fluid, typically modeled as a perfect gas: (1) isothermal expansion at the high temperature T_h, during which heat Q_h is absorbed from the hot reservoir while the gas expands and performs work; (2) adiabatic expansion, where the gas continues to expand without heat transfer, cooling to the low temperature T_c; (3) isothermal compression at T_c, rejecting heat Q_c to the cold reservoir as the gas is compressed; and (4) adiabatic compression, returning the gas to the initial state at T_h without heat exchange. These processes ensure the cycle is fully reversible, with the net work output equal to the difference between heat absorbed and rejected. On the pressure-volume (P-V) diagram, the forms a closed loop consisting of two hyperbolic isotherms—where decreases as increases at constant —and two steeper adiabatic curves connecting them, with the enclosed area quantifying the net work done by the cycle. In contrast, the - (T-S) diagram depicts the cycle as a : horizontal lines represent the isothermal processes, with increasing during addition at T_h and decreasing by the same amount during rejection at T_c, while vertical lines indicate the adiabatic processes where remains constant; the rectangular area corresponds to the net work, interpreted as T_h \Delta S - T_c \Delta S, where \Delta S is the change magnitude. The efficiency \eta of the Carnot cycle, defined as the ratio of net work output to heat input (\eta = W / Q_h), derives from the first and second . By the first law, W = Q_h - |Q_c| for a , so \eta = 1 - |Q_c| / Q_h. The second law, applied to the reversible , requires zero net change (\Delta S = 0), implying Q_h / T_h = |Q_c| / T_c and thus |Q_c| / Q_h = T_c / T_h, yielding \eta = 1 - \frac{T_c}{T_h}, where temperatures are absolute (in ); this expression depends solely on the reservoir temperatures, independent of the . For refrigerators operating in reverse, the provides the maximum , T_c / (T_h - T_c), establishing the reversible limit for heat extraction from a cold . This efficiency formula underscores profound implications: the Carnot cycle achieves the highest possible conversion of heat to work between given temperatures, with all real cycles exhibiting lower due to inherent irreversibilities like and non-equilibrium . Carnot's theorem formalizes this boundary, stating that no operating between two specified temperatures can surpass the of a reversible (Carnot) engine between the same reservoirs, a principle that prohibits machines of the second kind and enforces the directional flow of energy dictated by the second law.

Otto Cycle

The Otto cycle serves as the ideal thermodynamic model for spark-ignition internal combustion engines, such as those used in gasoline-powered automobiles. It was developed based on the patented by Nikolaus August Otto in , which marked the first practical implementation of a cycle featuring , compression, power, and exhaust strokes. This cycle approximates the behavior of the engine by simplifying combustion as heat addition and exhaust as heat rejection, focusing on the closed-system of the during the compression and expansion phases. The Otto cycle comprises four distinct processes in a closed system:
  1. Isentropic compression (1-2): The compresses the air-fuel mixture adiabatically and reversibly, increasing and without or friction.
  2. Constant-volume heat addition (2-3): occurs at fixed near top dead center, modeled as instantaneous input from an external source, raising and sharply.
  3. Isentropic expansion (3-4): The hot gases expand adiabatically and reversibly, performing work on the as increases.
  4. Constant-volume heat rejection (4-1): Residual is expelled at fixed , cooling the gases back to the initial state.
    These processes repeat in a cyclic manner, converting into work.
The air-standard assumptions underpin the Otto cycle analysis, treating the working fluid as air behaving as an throughout the cycle. Key assumptions include constant specific heats at (C_v and C_p), negligible changes in kinetic and , reversible processes without or losses, and replacement of actual with external addition while exhaust is modeled as rejection to a . These simplifications enable analytical tractability but diverge from real engine conditions involving variable composition and properties. On a pressure-volume (P-V) diagram, the Otto cycle appears as a closed : the isentropic (1-2) and expansion (3-4) trace steep curves from low to , while the constant-volume processes (2-3 and 4-1) form vertical lines connecting them. The enclosed area represents the net work output, with the r = V_1 / V_2 determining the loop's shape and . This highlights the cycle's reliance on changes for work extraction, distinct from constant-pressure cycles. The of the ideal depends solely on the and the specific heat ratio, given by \eta = 1 - \frac{1}{r^{\gamma - 1}} where r is the and \gamma = C_p / C_v (typically 1.4 for at standard conditions). To derive this, start with the general expression for a : \eta = 1 - |Q_\text{out}| / Q_\text{in}. For the , addition at constant volume is Q_\text{in} = C_v (T_3 - T_2), and rejection is Q_\text{out} = C_v (T_4 - T_1), so \eta = 1 - (T_4 - T_1) / (T_3 - T_2). Applying the isentropic relations T_2 / T_1 = r^{\gamma - 1} and T_3 / T_4 = r^{\gamma - 1}, it follows that T_4 / T_3 = T_1 / T_2, leading to (T_4 - T_1) / (T_3 - T_2) = T_1 / T_2 = 1 / r^{\gamma - 1}. Thus, increases with r but is independent of input magnitude. For typical automotive ratios of 8 to 12, \eta ranges from about 56% to 60% theoretically. The (MEP) quantifies the cycle's work capacity per unit displacement volume, defined as p_\text{m} = W_\text{net} / (V_1 - V_2), where W_\text{net} = \eta Q_\text{in}. For a naturally aspirated , MEP is approximately 1000 kPa, rising above 1500 kPa with turbocharging, providing a for comparing performance independent of size. This metric emphasizes the benefits of higher r in maximizing work output. Practical limitations constrain the Otto cycle's performance. High compression ratios, while improving efficiency, promote knocking—uncontrolled pre-ignition of the air-fuel mixture due to excessive temperatures and pressures during , which can damage the and typically caps [r](/page/R) at 8-12 for fuels. Additionally, real fuel-air mixtures introduce deviations from air-standard assumptions, as produces variable specific heats, of products at high temperatures, and non-ideal gas behavior, reducing actual efficiencies to 20-30% in engines.

Diesel Cycle

The Diesel cycle models the idealized operation of compression-ignition internal combustion engines, where air is compressed to ignite injected fuel without a spark. In 1893, German engineer received a from the Imperial Patent Office in for a "working method and construction for internal combustion engines operating with ," laying the foundation for this cycle. Unlike spark-ignition cycles, it relies on high compression to achieve auto-ignition, enabling higher thermal efficiencies in practical applications. The cycle consists of four reversible processes assuming an air-standard model with behavior and constant specific heats. First, isentropic (process 1-2) occurs as the moves inward, reducing while increasing and adiabatically, with no or . Second, constant- addition (process 2-3) simulates and , where is supplied at fixed , causing expansion. Third, isentropic expansion (process 3-4) follows, with the driven outward to produce work, again adiabatically. Fourth, constant- rejection (process 4-1) expels to the surroundings at fixed , completing the cycle before of . In the pressure-volume (P-V) diagram, the cycle traces a closed : process 1-2 is a steep adiabatic upward (), 2-3 is a line to the right ( addition at ), 3-4 is a shallower adiabatic downward (), and 4-1 is a vertical line leftward ( rejection at ). The area enclosed represents work output. The temperature- (T-S) diagram shows 1-2 and 3-4 as vertical lines (isentropic, ), 2-3 as a sloped line upward (increasing at ), and 4-1 as a line leftward ( , decreasing temperature). addition and rejection areas are bounded by these paths, highlighting irreversibility in real systems. The thermal efficiency η of the ideal Diesel cycle is derived from the heat input and output, expressed as: \eta = 1 - \frac{1}{r^{\gamma-1}} \cdot \frac{\rho^\gamma - 1}{\gamma (\rho - 1)} where r = V_1 / V_2 is the compression ratio, \rho = V_3 / V_2 is the cutoff ratio, and \gamma = C_p / C_v is the specific heat ratio (typically 1.4 for air). To derive this, start with the first law for a closed system: net work W = Q_{in} - Q_{out}, so \eta = W / Q_{in} = 1 - Q_{out} / Q_{in}. For process 2-3, Q_{in} = C_p (T_3 - T_2); for 4-1, Q_{out} = C_v (T_4 - T_1). Using isentropic relations: T_2 = T_1 r^{\gamma-1}, T_3 = T_2 \rho = T_1 r^{\gamma-1} \rho, and T_4 = T_3 (\rho / r)^{\gamma-1} = T_1 \rho^\gamma. Substitute into Q_{out} / Q_{in} = [C_v (T_4 - T_1)] / [C_p (T_3 - T_2)] = [(T_4 - T_1)] / [\gamma (T_3 - T_2)] = [T_1 (\rho^\gamma - 1)] / [\gamma T_1 r^{\gamma-1} (\rho - 1)] = (\rho^\gamma - 1) / [\gamma r^{\gamma-1} (\rho - 1)], yielding the efficiency formula. This shows efficiency increases with higher r but decreases with larger \rho, as more heat is added late in the cycle. Compared to the in spark-ignition engines, the permits higher compression ratios (typically 14:1 to 25:1 versus 8:1 to 12:1 for ), as fuel is injected after compression, preventing auto-ignition knocking during the compression stroke. This advantage stems from compressing air alone, allowing greater rise without premature . engines power heavy-duty applications like trucks, locomotives, ships, and construction equipment due to their high and , as well as stationary generators for reliable backup power in hospitals and data centers. In real systems, deviations from the ideal cycle include finite rates causing non-constant pressure heat addition, frictional losses, and to cylinder walls, reducing by 10-20% compared to ideal predictions; modern implementations often incorporate turbocharging to elevate intake pressure, boosting but adding and irreversibilities.

Rankine Cycle

The , named after Scottish engineer William John Macquorn Rankine, was first systematically described in his 1859 publication Manual of the Steam Engine and Other Prime Movers, which laid foundational principles for vapor power cycles in . This cycle serves as the thermodynamic model for practical power systems, converting heat into mechanical work through phase changes in the . The ideal Rankine cycle consists of four reversible processes: (1) isentropic compression of liquid in a , increasing its with minimal work input; (2) constant- addition in a , where the fluid evaporates and may be superheated; (3) isentropic expansion in a , extracting work as the drops; and (4) constant- rejection in a , where the vapor condenses back to liquid. These processes utilize as the , leveraging its phase changes—liquid to saturated vapor in the boiler and vapor to liquid in the —to achieve efficient at constant temperatures during and . The cycle's performance is often analyzed using an enthalpy-entropy (h-s) diagram, which illustrates the states of the across the processes, highlighting the area under the expansion curve for turbine work and above the compression curve for pump work. The η of the ideal is defined as the net work output divided by the input in the : \eta = \frac{W_\text{turbine} - W_\text{pump}}{Q_\text{boiler}} where W_\text{turbine} is the work extracted from , W_\text{pump} is the work (typically small), and Q_\text{boiler} is the added at . This efficiency is inherently lower than the due to the average temperature of addition being below the maximum temperature, but modifications can narrow the gap. To improve and reduce moisture in the exhaust, practical Rankine cycles incorporate reheat, where partially expanded is returned to the for additional heating before further expansion, and regeneration, using feedwater heaters to preheat the pumped liquid with extracted , thereby increasing the average addition temperature and approaching Carnot limits. These enhancements can boost from typical values of 30-40% in basic cycles to over 45% in advanced configurations. The underpins most large-scale , particularly in coal-fired and plants, where drives turbines to produce from sources like or . In coal plants, it handles effectively due to the closed-loop operation, while in nuclear plants, it converts to without direct fluid contact.

Brayton Cycle

The , also known as the Joule cycle, is a thermodynamic cycle that models the operation of engines through continuous-flow processes involving , addition, , and rejection. It was originally proposed by American engineer in the 1870s as a constant-pressure engine using oil as , marking an early concept for gas power generation. The cycle gained prominence in modern applications through the development of engines by British engineer , who patented a practical turbojet design in and achieved the first flight test in 1941, revolutionizing aviation propulsion. In the ideal , the working fluid—typically air modeled as an with constant specific heats—undergoes four reversible processes on a temperature-entropy diagram: (1) isentropic compression from state 1 to 2 in a , raising and ; (2) constant- heat addition from state 2 to 3 in a , where is burned to increase ; (3) isentropic expansion from state 3 to 4 in a , extracting work; and (4) constant- heat rejection from state 4 to 1, often to the atmosphere or a . These steady-flow processes distinguish the from reciprocating cycles, enabling high-speed operation in turbines suitable for continuous power output. The thermal efficiency of the ideal Brayton cycle depends solely on the compressor pressure ratio r_p = P_2 / P_1, where P denotes pressure, and is given by \eta = 1 - \frac{1}{r_p^{(\gamma-1)/\gamma}} with \gamma = c_p / c_v as the specific heat ratio of the gas (approximately 1.4 for air). To derive this, start with the isentropic relations for an ideal gas: for compression, T_2 / T_1 = r_p^{(\gamma-1)/\gamma}, so T_2 = T_1 r_p^{(\gamma-1)/\gamma}; for expansion, T_3 / T_4 = r_p^{(\gamma-1)/\gamma}, so T_4 = T_3 / r_p^{(\gamma-1)/\gamma}. The net work is turbine work minus compressor work: w_{net} = c_p (T_3 - T_4) - c_p (T_2 - T_1). Heat input is q_{in} = c_p (T_3 - T_2), so efficiency \eta = w_{net} / q_{in} = 1 - (T_4 - T_1)/(T_3 - T_2). Substituting the temperature relations yields \eta = 1 - (T_1 / T_2) = 1 - 1 / r_p^{(\gamma-1)/\gamma}, independent of heat addition. Brayton cycles operate in open or closed configurations. In the open cycle, atmospheric air is drawn in for and exhausted after , as in jet engines where the working fluid is not recirculated. The closed cycle recirculates a fixed (e.g., or air) through a for heat addition and rejection, avoiding direct atmospheric interaction and enabling use in controlled environments like systems. Modifications enhance Brayton cycle performance beyond the ideal case. Regeneration involves a heat exchanger that transfers exhaust heat from the turbine outlet to preheat air entering the combustor, reducing fuel requirements and boosting efficiency—for instance, an 85% effective regenerator can increase efficiency from about 49% to 63% at typical pressure ratios. Intercooling cools the air between multi-stage compressor sections using an external sink, lowering the work input for compression and improving net output, though it slightly reduces ideal efficiency without other enhancements. The powers aircraft via turbojets and turbofans, where high thrust-to-weight ratios enable efficient high-speed flight, and stationary gas turbines for electrical power generation, often in combined-cycle plants achieving up to 60% . Performance improves with higher pressure ratios, as rises asymptotically toward a limit set by \gamma, but practical constraints like material limits cap ratios at 30-40 for modern engines; for example, increasing from 5 to 20 can raise from 30% to 50%.

Stirling Cycle

The is a thermodynamic cycle that operates as a closed-cycle regenerative , patented by Scottish clergyman Robert Stirling in 1816 for efficient fuel use in air engines. This cycle consists of four distinct processes: isothermal compression of the working fluid at the cold temperature T_c, constant-volume heat addition through a regenerator, isothermal expansion at the hot temperature T_h, and constant-volume heat rejection back through the regenerator. The isothermal compression reduces the volume of the gas while rejecting heat to the cold reservoir, followed by isochoric heating where the regenerator transfers stored thermal energy to the gas, raising its temperature without volume change. The subsequent isothermal expansion performs work by absorbing heat from the hot reservoir as the volume increases, and the final isochoric cooling returns heat to the regenerator, completing the cycle. A defining feature of the Stirling cycle is its use of a thermal regenerator, a porous that stores during the constant-volume rejection and releases it during addition, minimizing external input and enabling reversible operation. This regeneration approaches perfect in conditions, allowing the cycle's thermal to match the Carnot limit, given by \eta = [1](/page/1) - \frac{T_c}{T_h}, where temperatures are in . Under assumptions of complete regeneration and no losses, the Stirling cycle achieves this maximum for given reservoir temperatures, serving as a practical approximation to the theoretical Carnot cycle. Stirling engines implementing this cycle are classified into three main configurations: alpha, , and gamma, differing in and displacer arrangements for gas displacement between hot and cold zones. In the alpha configuration, two separate power s operate in opposed cylinders, one hot and one cold; the beta type uses a single cylinder with a power and a displacer for gas shuttling; while the gamma variant employs two cylinders, separating the displacer and power functions for simpler sealing. These designs facilitate the cycle's external , allowing operation with diverse heat sources such as , , or . Applications of the are prominent in low-power devices and , leveraging its quiet, vibration-free operation and ability to utilize low-grade sources without . In , reversed Stirling cycles achieve cryogenic temperatures below 80 with high efficiency, up to 30% of Carnot, for sensors and space instrumentation, consuming minimal electrical input like 70 W for 2 W cooling at 80 . For low-power generation, Stirling engines power remote sensors or auxiliary systems, offering reliable, emission-free performance in environments where noise and fuel flexibility are critical.

Advanced Thermodynamic Concepts in Cycles

State Functions and Path Independence

In thermodynamics, state functions are thermodynamic properties that depend solely on the current of the system, defined by variables such as , , and volume, and are independent of the path taken to reach that state. Examples include U, H, and S, which allow changes to be calculated directly from initial and final states without regard to intermediate processes. For a complete thermodynamic cycle, where the system returns to its initial , the net change in any state function is zero, such as \Delta U = 0, simplifying the analysis of cyclic processes. In contrast, path functions like heat Q and work W depend on the specific sequence of processes or path traversed between states, resulting in different values for the same initial and final conditions depending on the route taken./19%3A_The_First_Law_of_Thermodynamics/19.03%3A_Work_and_Heat_are_not_State_Functions) This distinction is crucial in cycles, where state functions provide a consistent "bookkeeping" framework—net changes are zero—while path functions account for the actual energy transfers that enable net work output or heat absorption. Enthalpy, defined as H = U + PV, is particularly useful in open systems and steady-flow processes, as it incorporates the flow work PV associated with fluid entry and exit, streamlining energy balance calculations in devices like turbines or compressors. The G = H - TS and A = U - TS play roles in assessing the spontaneity of processes within cycles: a negative change \Delta G < 0 indicates spontaneity at constant temperature and pressure, while \Delta A < 0 does so at constant temperature and volume, guiding the feasibility of cycle steps./22%3A_Helmholtz_and_Gibbs_Energies) For an ideal gas in a thermodynamic cycle, the internal energy U depends only on temperature, so the change in a process is \Delta U = n C_v \Delta T, where n is the number of moles and C_v is the molar heat capacity at constant volume; over the full cycle, with \Delta T = 0, \Delta U = 0, allowing focus on heat and work balances.

Entropy Production and the Second Law

The second law of thermodynamics states that the entropy of an isolated system never decreases; it either remains constant in reversible processes or increases in irreversible ones. In the context of thermodynamic cycles, which return a system to its initial state, the net change in entropy of the working fluid is zero, as entropy is a state function. For a reversible cycle, the integral of reversible heat transfer over temperature, \oint \frac{\delta Q_\text{rev}}{T} = 0, reflects this balance. However, real cycles involve irreversibilities, leading to positive entropy production in the universe (\Delta S_\text{universe} > 0), which limits below the reversible ideal. Entropy change for a system is defined as \Delta S = \int \frac{\delta Q_\text{rev}}{T}, where the integral is along a reversible path connecting initial and final states. For an , this yields specific expressions: \Delta s = c_v \ln \frac{T_2}{T_1} + R \ln \frac{V_2}{V_1} at constant specific heats, or equivalently \Delta s = c_p \ln \frac{T_2}{T_1} - R \ln \frac{p_2}{p_1}. These formulas quantify variations during cycle processes like or expansion, aiding in assessments. In irreversible processes, actual change exceeds the heat transfer term: dS = \frac{\delta Q}{T} + dS_\text{gen}, where dS_\text{gen} > 0 accounts for generation due to internal effects. Irreversibilities in cycles arise from mechanisms such as , fluid mixing, and across finite temperature differences, each producing . dissipates into , increasing without useful work; mixing of dissimilar fluids generates spontaneously toward ; and finite \Delta T violates reversibility, as flows inefficiently from hot to cold reservoirs. These effects ensure \Delta S_\text{universe} \geq 0, with equality only for idealized reversible cycles. The Clausius inequality formalizes this for any cycle: \oint \frac{\delta Q}{T} \leq 0, where equality holds for reversible paths and strict inequality indicates . Temperature-entropy (T-S) diagrams visualize these concepts in cycles, plotting against to depict state changes. The area under a curve represents (Q = \int T \, dS), with isentropic (reversible adiabatic) processes appearing as vertical lines. In real cycles, deviations from vertical lines or enclosed areas quantify generation, highlighting irreversibilities that reduce the net work area compared to reversible ideals. Entropy production links to , or available work, where lost work potential equals T_0 \Delta S_\text{gen}, with T_0 as ambient . This "lost work" represents the portion of energy rendered unavailable due to irreversibilities, providing a second-law metric for cycle optimization beyond first-law efficiencies.

References

  1. [1]
    Thermodynamic cycles
    A thermodynamic cycle is a linked series of processes such that the outlet of the final process is the same state as the input of the first process.
  2. [2]
    5. Thermodynamics — Introduction to Statistical Mechanics
    This is an example of a thermodynamic cycle: a sequence of thermodynamic processes that end with the system (other than the heat baths) in exactly the same ...
  3. [3]
    Thermodynamic cycles and heat engines(VW, S & B: Chapter 9) - MIT
    You will also learn how to model these heat engines as thermodynamic cycles and how to apply the First Law of Thermodynamics to estimate thermal efficiency and ...
  4. [4]
    1.2 Definitions and Fundamental Ideas of Thermodynamics - MIT
    A thermodynamic system is a quantity of matter of fixed identity, around which we can draw a boundary (see Figure 1.3 for an example).
  5. [5]
    7.6. Rankine cycle | EME 812: Utility Solar Electric and Concentration
    We are going to overview the principle of thermodynamic cycle operation using Rankine cycle example, since most of solar power cycles currently operating ...
  6. [6]
    [PDF] Chapter 9 - Civil, Environmental and Architectural Engineering
    Thermo- dynamic cycles can be divided into two general categories: power cycles and refrigeration cycles. The devices or systems used to produce a net power out ...
  7. [7]
    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 ...
  8. [8]
    Thermodynamic Foundations – Introduction to Aerospace Flight ...
    Thermodynamics is the science of energy and the principles governing its transformation from one form to another, such as work and heat, and the production of ...
  9. [9]
    [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 ...<|control11|><|separator|>
  10. [10]
    The First Law of Thermodynamics and Some Simple Processes
    Example 109.1 Total Work Done in a Cyclical Process Equals the Area Inside the Closed Loop on a PV Diagram. Calculate the total work done in the cyclical ...
  11. [11]
    [PDF] 249 Chapter 7 Heat, Work, and the First Law of Thermodynamics ...
    Aug 7, 2021 · Heat is the transfer of energy through random motions of molecules and work is the transfer of energy through organized motion. Consider a gas ...
  12. [12]
    I THE FIRST LAW OF THERMODYNAMICS - MIT
    I THE FIRST LAW OF THERMODYNAMICS ; 1 Heat ; 2 Zeroth Law of Thermodynamics ; 4 Work vs. Heat - which is which?
  13. [13]
    [PDF] dW = Fds = pAds = pdV W = pdV
    The work W done by the system during a transformation from an initial state to a final state depends on the path taken. The heat Q absorbed by the system ...
  14. [14]
    [PDF] T dq ds ≡ T dq ds ≥
    d) heating of an ideal gas at constant pressure. For a reversible process ds = dqrev/T = cpdT/T = cpdlnT. e) entropy changes during phase transitions. For a ...
  15. [15]
    [PDF] Heat and 1st Law of Thermodynamics
    The first law of thermodynamics relates the change of the internal energy of a system to the heat absorbed/released and the work done on/by a system. It is ...
  16. [16]
    Joule-Thompson Throttling - Richard Fitzpatrick
    Joule-Thompson throttling is a continuous-flow process where a gas is constricted through a porous plug, maintaining a pressure difference, and its enthalpy ...
  17. [17]
    6.7 Examples of Lost Work in Engineering Processes - MIT
    The throttling process is a representation of flow through inlets, nozzles, stationary turbomachinery blades, and the use of stagnation pressure as a measure ...
  18. [18]
    Engine Power Cycle - an overview | ScienceDirect Topics
    An engine power cycle is defined as a thermodynamic cycle that converts thermal energy into work, exemplified by various types such as the Otto, Diesel, ...
  19. [19]
    8 Analysis of Thermodynamic Cycles – Thermo 101
    Thermodynamic power cycles are used in heat engines and produce mechanical power by converting thermal energy into work (examples pictured above).
  20. [20]
    (PDF) Chapter 6 Thermodynamics of Cycles - ResearchGate
    Mar 7, 2018 · steam is exhausted into the atmosphere. The basic processes of the cycle, either in open or closed, are heat addition, heat. rejection, ...
  21. [21]
    Gas Power Cycle - an overview | ScienceDirect Topics
    Commonly, steam and gas power cycles are employed in power generation. Steam power cycles have been applied for commercial purposes in various sectors.
  22. [22]
    [PDF] 1 Thermodynamic Cycles - Wiley-VCH
    Power cycles are classified according to the type of heat engine they model, e.g. we have the Otto cycle and the Diesel cycle for modelling internal combustion ...
  23. [23]
    Carnot Limit - an overview | ScienceDirect Topics
    The Carnot limit refers to the maximum efficiency achievable by heat engines, which is determined by the temperatures of the heat source (T1) and heat sink ...
  24. [24]
    Thermodynamic Cycle - an overview | ScienceDirect Topics
    In closed cycles, the working fluid is returning to the initial state at the end of the cycle and is recirculated. By the same token, in open cycles, the ...
  25. [25]
    3.7 Brayton Cycle - MIT
    Figure 3.16(a) shows an ``open'' cycle, where the working fluid enters and then exits the device. ... closed cycle, which recirculates the working fluid.
  26. [26]
    [PDF] The Life and Legacy of William Rankine - Purdue e-Pubs
    Professor William Rankine, who was born on July 5 1820, made a substantial contribution to the science of heat and power and his influence in refrigeration ...
  27. [27]
    3.4 Refrigerators and Heat Pumps - MIT
    Typically the thermodynamic system in a refrigerator analysis will be a working fluid, a refrigerant, that circulates around a loop, as shown in Figure 3.7. The ...
  28. [28]
    [PDF] 3.8.3. Vapor Compression Refrigeration and Heat Pump Cycles
    Dec 15, 2021 · The objective of a vapor compression refrigeration cycle is to remove energy from a cold reservoir and move it to the hot reservoir.
  29. [29]
    Design of Vapor-Compression Refrigeration Cycles
    Dec 16, 1997 · The design is to be based upon the ideal vapor-compression refrigeration cycle, with four components: a cooler (where we reject the heat), a ...
  30. [30]
    [PDF] ABSORPTION REFRIGERATION - OSTI.GOV
    The most common type of absorption cycle employed for these applications is the water/ammonia cycle. In this case, water is the absorbent and ammonia is the ...
  31. [31]
    Absorption Heat Pumps: An Emerging Technology for Large Homes
    Absorption heat pumps use an ammonia-water absorption cycle to provide heating and cooling. Similar to a standard heat pump, the refrigerant (ammonia) ...
  32. [32]
    Vapor Compression Refrigeration System - School of Engineering
    The cycle starts with the refrigerant being pressurized in the compressor, the super heated vapor then moves through the condenser which releases the heat ...
  33. [33]
    Heat Pump Systems - Department of Energy
    Like your refrigerator, heat pumps use electricity to transfer heat from a cool space to a warm space, making the cool space cooler and the warm space warmer.Geothermal Heat · How Ductless Minisplit Heat... · Residential Cold Climate Heat...<|control11|><|separator|>
  34. [34]
    [PDF] Fundamentals of engineering thermodynamics
    Typeset in 10/12 pt Times by Techbooks. Printed and bound in Great Britain by Scotprint, East Lothian. This book is printed on acid-free paper responsibly ...
  35. [35]
    [PDF] Unit I:Air Standard Cycles - Sathyabama
    The analysis of all air standard cycles is based upon the following assumptions: 1. The gas in the engine cylinder is a perfect gas i.e., it obeys the gas laws ...
  36. [36]
    [PDF] Outline Thermodynamic cycle analysis Idealized cycles
    P-v diagram represents net work/heat for internally reversible cycles ... circulates in a closed loop and always behaves as an ideal gas. 2. All the ...
  37. [37]
    Brayton Cycle Efficiency - an overview | ScienceDirect Topics
    In the real cycle, irreversibilities occur and entropy is produced due to frictions and heat losses, so the real processes deviate from the ideal process as ...
  38. [38]
    [PDF] Chapter23 - Rose-Hulman
    Two important areas of application for thermodynamics are power generation and refrigeration. Both power generation and refrigeration are usually ...<|control11|><|separator|>
  39. [39]
    Simulation and Exergy Analysis of Energy Conversion Processes ...
    Sep 30, 2018 · Relevant basics of a thermodynamic analysis with exergy-based methods and necessary fluid property models are explained. Thermodynamic models ...Missing: effective | Show results with:effective
  40. [40]
    Rankine Cycle Analysis and Efficiency - Thermodynamics II - Fiveable
    The isentropic efficiencies of the turbine and pump affect the overall efficiency of the Rankine cycle · The presence of irreversibilities, such as friction and ...Rankine Cycle Components And... · Basic Rankine Cycle... · Factors Affecting Rankine...
  41. [41]
    Effect of several irreversibilities on the thermo-economic ...
    Aug 8, 2025 · Regeneration is one of these modifications which increases thermal efficiency for the same power output and provides less fuel consumption.
  42. [42]
    Advanced energy conversion strategies using multistage radial ...
    This study presents a novel approach by designing a multistage subsonic radial inflow turbine as an alternative, mitigating inherent losses in supersonic ...
  43. [43]
    CFD Simulation of Stirling Engines: A Review - MDPI
    An SE operates on the Stirling cycle, a highly efficient thermodynamic cycle with a theoretical efficiency equivalent to the Carnot cycle. The SE elements ...
  44. [44]
    Thermodynamic Cycle Enhancement → Term
    Feb 10, 2025 · Reduced Fuel Consumption → Improved efficiency translates to lower fuel consumption in power plants and engines, saving resources and reducing ...
  45. [45]
    [PDF] Reflections on the motive power of heat and on machines fitted to ...
    -SADI CARNOT. Bg Mons. H. Carnoe, •. III. REFJ,ECTIONS ON THE MOTIVE POWER OF ... It was before this period (in 1824) tbat Sadi had published his Rljlexion.
  46. [46]
    Heat Engines: the Carnot Cycle - Galileo
    Efficiency = TH−TCTH. This was an amazing result, because it was exactly correct, despite being based on a complete misunderstanding of the nature ...<|control11|><|separator|>
  47. [47]
    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.
  48. [48]
  49. [49]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics 3.8 ...
    Dec 15, 2021 · Carnot cycles include power, refrigeration, and heat pump cycles. This section, however, focuses specifically on Carnot power cycles.
  50. [50]
    Otto (1832) - Energy Kids - EIA
    Born in 1832 in Germany, Nicolaus August Otto invented the first practical alternative to the steam engine - the first successful four-stroke cycle engine.
  51. [51]
    Otto or Not, Here it Comes - ASME Digital Collection
    Ever since Nicolaus Otto demonstrated the first working four-stroke engine in 1876, engineers have been struggling to come up with ways to sidestep a ...Missing: Nikolaus | Show results with:Nikolaus
  52. [52]
    Model of Basic Otto Cycle
    The air standard Otto cycle is modeled as a closed system with a fixed air/fuel mass, and uses the following process assumptions: 1-2 Compression which is ...
  53. [53]
    3.5 The Internal combustion engine (Otto Cycle) - MIT
    1 Efficiency of an ideal Otto cycle. The starting point is the general expression for the thermal efficiency of a cycle: $\displaystyle \eta = \frac{\textrm ...
  54. [54]
    Mean Effective Pressure
    A naturally aspirated Otto cycle engine has a Pmean,b ~ 1000 kPa. If turbo charged, the engine Pmean,b can increase to above 1500 kPa.
  55. [55]
    [PDF] Otto and Diesel Cycles
    Another approach to improve efficiency is to reduce the back-work ratio through compressor cooling or intercooling. Regeneration. Page 34. Gas Power Cycles - ...
  56. [56]
    [PDF] LECTURE NOTES ON INTERMEDIATE THERMODYNAMICS
    Mar 28, 2025 · The temperature at the beginning of the compression process of an air-standard Otto cycle with a compression ratio of 8 is 540 ◦R, the ...<|control11|><|separator|>
  57. [57]
    Diesel Engine: Patent (February 23, 1893)
    On February 23, 1893, German engineer Rudolf Diesel (1858-1913) was granted a patent by the Imperial Patent Office in Berlin for “working methods and design ...
  58. [58]
    The Diesel Engine - HyperPhysics Concepts
    The input and output energies and the efficiency can be calculated from the temperatures and specific heats: It is convenient to express this efficiency in ...
  59. [59]
    Diesel Cycle - Definition, Process, PV Diagram and TS Diagram
    It is the cycle used in the Diesel (compression-ignition) engine. The heat is transferred to the working fluid at constant pressure.
  60. [60]
    Use of diesel - U.S. Energy Information Administration (EIA)
    Diesel-engine-powered machinery can do demanding construction work, such as lifting steel beams, digging foundations and trenches, drilling wells, paving roads ...
  61. [61]
    History of diesel engines | Cummins Inc.
    Apr 4, 2023 · In 1897, after patenting the first compression ignition engine design in 1892, the German inventor and mechanical engineer Rudolf Diesel ...
  62. [62]
    Introduction 1
    To a Scottish engineer, William Rankine, goes the credit of writing the first textbook on engineering thermodynamics. Entitled Manual of the Steam Engine.
  63. [63]
    [PDF] Rankine Cycle - MIT OpenCourseWare
    Apr 1, 2020 · Large heat transfer (latent heat). Both work and efficiency increase monotonically because of small pumping work. © Ahmed F. Ghoniem. 4 ...<|control11|><|separator|>
  64. [64]
    [PDF] Chapter 8 - WPI
    ▻ While pump work input is much less than turbine work output, irreversibilities in the pump affect net power output of the vapor plant. ▻ Isentropic pump ...
  65. [65]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Dec 15, 2021 · The standard. Rankine cycle consists of the following four processes (refer to Figure 3.41):. • Process 1 – 2: Expansion of the working fluid ...
  66. [66]
    8.5 Rankine Power Cycles - MIT
    For the Rankine cycle, $ T_{m1} \approx T_1$ , $ T_{m2} <T_2$ . From this equation we see not only the reason that the cycle efficiency is less than that of ...
  67. [67]
    Rankine Cycle with Regeneration
    Dec 15, 1997 · Using CyclePad, we will modify a Rankine cycle and examine the effects of regeneration on the cycle's thermal efficiency.
  68. [68]
    8.6 Enhancements of Rankine Cycles - MIT
    The main practical advantage of reheat (and of superheating) is the decrease in moisture content in the turbine because most of the heat addition in the cycle ...Missing: regeneration | Show results with:regeneration
  69. [69]
    [PDF] The Development of a Steady State Rankine Cycle Optimization ...
    Large base load thermal power plants, such as coal and nuclear, use the Rankine steam cycle to generate electricity [14]. Many natural gas power plants use ...
  70. [70]
    The Brayton Cycle - Intro
    Introduction to the Brayton Cycle. The Brayton cycle is gas power cycle originally devised by George Brayton in 1870 as an oil burning engine.
  71. [71]
    Whittle W.1X Engine | National Air and Space Museum
    This is one of the first turbojet engines. British engineer Sir Frank Whittle patented his pioneering design in 1932. The engine first flew on the E. 28/39 in ...
  72. [72]
    The Development of the Whittle Turbojet | J. Eng. Gas Turbines Power
    His work in developing the turbojet can truly be said to represent one of the greatest mechanical engineering achievements in the last 70 years. The development ...
  73. [73]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Dec 15, 2021 · The Brayton cycle is a thermodynamic model for gas turbine engines. In a simple gas turbine engine. (components shown schematically in ...
  74. [74]
    [PDF] Aircraft Engines - Federal Aviation Administration
    The Brayton cycle is the name given to the thermodynamic cycle of a gas turbine engine to produce thrust. This is a variable volume constant-pressure cycle ...Missing: generation | Show results with:generation
  75. [75]
    [PDF] Combustion Turbines - U.S. Environmental Protection Agency
    Gas turbines are ideally suited for CHP applications because their high-temperature exhaust can be used to generate process steam at conditions as high as 1,200 ...
  76. [76]
    History of Stirling engines
    The original Stirling engine was invented and patented by Robert Stirling on September 27, 1816 and first used in 1818 as a pumping device to push water into a ...
  77. [77]
    Thermodynamic Theory of the Ideal Stirling Engine
    Therefore, the Carnot efficiency at a given hot section and cold section temperature is equal to the Stirling efficiency between the same hot and cold sections.
  78. [78]
    Stirling Engine Configurations - updated 3/30/2013 - Ohio University
    Mar 30, 2013 · The mechanical configurations of Stirling engines are generally divided into three groups known as the Alpha, Beta, and Gamma arrangements.Missing: variants | Show results with:variants
  79. [79]
    Development of a Low-Power Stirling Cycle Cryocooler for Space ...
    Stirling cycle cryocoolers have successfully been used to provide thermal control for many space applications, particularly Earth observation instruments ...Missing: devices | Show results with:devices
  80. [80]
    [PDF] High-Capacity and Efficiency Stirling Cycle Cryocooler
    The cryocooler performance responds well with demonstrated efficiencies of up to. 30+% of Carnot based on net useful cooling capacity and net electrical power ...
  81. [81]
    The Core of Cryogenic Cooling Systems - Stirling Cryogenerators
    Technical Features of Stirling Cryocoolers · Highest efficiency possible (through reversed Stirling Cycle technology) leading to low electricity consumption ...
  82. [82]
    State Functions: Enthalpy, Entropy, Energy & Internal Energy - EMBIBE
    A state function can be defined as a function whose value depends only upon the initial and final states of the system and not upon the path through which this ...
  83. [83]
  84. [84]
    5.3 Enthalpy - Chemistry 2e | OpenStax
    Feb 14, 2019 · ... enthalpy (H) to describe the thermodynamics of chemical and physical processes. Enthalpy is defined as the sum of a system's internal energy ...
  85. [85]
    Steady Flow Energy Equation - MIT
    In an open flow system, enthalpy is the amount of energy that is transferred across a system boundary by a moving flow. This energy is composed of two parts ...
  86. [86]
    4.1 Internal energy in a system
    They can be used to calculate the changes of specific internal energy, Δ u , and specific enthalpy, Δ h , respectively, in a process involving ideal gases, ...
  87. [87]
    II THE SECOND LAW OF THERMODYNAMICS - MIT
    Entropy on the Microscopic Scale · 7.1 Entropy Change in Mixing ... 7 Combined Cycles for Power Production · 8.8 Some Overall Comments on Thermodynamic Cycles ...
  88. [88]
    None
    ### Summary of Clausius Inequality, Entropy as a State Function, and Application to Cycles
  89. [89]
    [PDF] Supplementary Notes on Entropy and the Second Law of ...
    Since the Carnot cycle is reversible and all the heat transfer interac- tions are reversible, all processes in the universe are reversible. So the universe ...Missing: ∫ | Show results with:∫
  90. [90]
    Entropy of a Gas
    This slide shows math derivations for the evaluation of the change of entropy for a gas ... For an ideal gas, the equation of state is written: p * V = R ...Missing: cycles | Show results with:cycles
  91. [91]
    P-V and T-S Diagrams
    P-V and T-s diagrams are often used to visualize the processes in a thermodynamic cycle and help us better understand the thermodynamics of engines.
  92. [92]
    6.5 Irreversibility, Entropy Changes, and ``Lost Work'' - MIT
    The entropy of a system can be altered in two ways: (i) through heat exchange and (ii) through irreversibilities.Missing: exergy | Show results with:exergy