Fact-checked by Grok 2 weeks ago

Rankine cycle

The Rankine cycle is a thermodynamic cycle that serves as the fundamental model for vapor power systems, converting heat energy into mechanical work in a closed-loop process typically using water as the working fluid. It comprises four primary reversible processes: isentropic compression of liquid in a pump to increase pressure, isobaric heat addition in a boiler to produce high-pressure vapor, isentropic expansion of the vapor in a turbine to generate work, and isobaric heat rejection in a condenser to return the fluid to its liquid state. Developed by Scottish engineer and physicist William John Macquorn Rankine (1820–1872) in the mid-19th century as an idealized representation of steam engine operation, the cycle provides a practical framework for analyzing the efficiency of heat engines operating between two temperature reservoirs, though its thermal efficiency is inherently lower than that of the Carnot cycle due to the lower average temperature of heat addition. This cycle underpins the majority of conventional steam power plants, including those fueled by , , , and concentrated , where it enables large-scale by driving turbines connected to electrical generators. Key performance metrics, such as (typically 30–45% in practical implementations), depend on factors like pressure, maximum temperature, and temperature, with higher values achieved through modifications including the vapor to avoid wet in the , reheating after partial to boost work output, and regeneration via feedwater heaters to preheat the pumped liquid using extracted . These enhancements address real-world irreversibilities like and losses, making the Rankine cycle a cornerstone of modern energy production while variants like the adapt it for lower-temperature sources using organic fluids for applications in recovery and .

Fundamentals

Overview

The Rankine cycle is an idealized used in engines to convert into mechanical work, serving as the model for vapor systems such as steam turbines. It operates by circulating a , typically , through a closed where is absorbed to generate high-pressure vapor, which then drives mechanical components to produce . This cycle forms the basis for efficient energy conversion in large-scale systems, enabling the practical harnessing of from various sources. The Rankine cycle finds primary application in steam power plants for , where it powers the majority of global power infrastructure. sources include fossil fuels such as and in conventional plants, reactors in power stations, and renewable options like concentrating systems and geothermal flash plants. In these settings, the cycle's adaptability to high-temperature heat inputs makes it essential for reliable, scalable power production across diverse energy portfolios. Compared to the , which offers the highest theoretical efficiency between two temperature limits but proves impractical for vapor-based systems due to the difficulties in managing isothermal processes amid phase changes, the Rankine cycle provides a more feasible alternative through its use of constant pressure addition and rejection. This design avoids the challenges of compressing two-phase mixtures and enables straightforward and , aligning better with real-world constraints. The basic components include a for vaporizing the via input, a for work extraction during expansion, a for cooling and liquefying the exhaust vapor, and a for returning the to . Overall , measured as the net work output divided by input, generally achieves 30-40% in operational power plants, reflecting practical limitations like irreversibilities and heat losses.

Historical Development

The Rankine cycle originated from advancements in steam engine technology during the 18th century, particularly through the improvements made by , who enhanced the efficiency of Newcomen engines by introducing a separate and other mechanisms that reduced energy loss, laying the groundwork for more practical heat engines. Building on these foundations, Scottish engineer William John Macquorn Rankine formalized the cycle in 1859 through his seminal work Manual of the Steam Engine and Other Prime Movers, providing the first systematic thermodynamic analysis of steam power processes and establishing it as a theoretical model for heat-to-work conversion. During the from the late 1700s to the mid-1800s, the principles underlying the Rankine cycle were adopted in reciprocating engines, which powered factories, mines, and transportation, driving widespread industrialization despite their relatively low compared to modern standards. The late 19th century marked a pivotal transition from reciprocating engines to , with Swedish engineer Carl Gustaf Patrick de Laval developing the first impulse in 1883 and British engineer inventing the multi-stage turbine in , enabling higher speeds and outputs. Key milestones included Parsons' demonstration of the first commercial plant in , which generated 7.5 kW and showcased the potential for electrical integration. By the early , turbines were routinely coupled with electrical generators, revolutionizing as seen in General Electric's 500-kW Curtis turbine in 1901 and widespread adoption in utility plants. In the mid-20th century, the Rankine cycle evolved toward high-pressure and supercritical designs to boost in plants, with the first commercial supercritical unit operational at the Philo Power Plant in in 1957, operating above water's critical point of 22.1 MPa and 374°C for improved performance. As of 2025, the cycle remains central to plants, where it converts to , and renewable systems like concentrated facilities, with ongoing incremental gains through advanced materials and controls rather than fundamental shifts.

Cycle Processes

The Four Processes

The Rankine cycle operates as a in a , typically using as the , and involves four sequential processes that convert into mechanical work under steady-flow conditions. These processes occur in components such as the , , , and , assuming reversible behavior in the ideal case and familiarity with transitions between and vapor states. In the first process (1-2), isentropic compression takes place in the , where subcooled liquid exiting the is pressurized to the boiler's . The mechanism involves adiabatic work input to the incompressible liquid, resulting in a small increase in and with negligible . The purpose is to elevate the fluid's efficiently, enabling subsequent addition at elevated levels while minimizing expenditure due to the liquid's low . The second process (2-3) is isobaric heat addition in the , where the pressurized absorbs at from an external . This involves three stages: preheating the subcooled to the temperature, evaporative to produce saturated vapor, and optional to raise the vapor above its point, increasing its . The mechanism relies on steady to drive the phase change and thermal excitation, producing high-energy ready for . The purpose is to maximize the availability of for conversion into work in the subsequent process. During the third process (3-4), isentropic expansion occurs in the , where the high-pressure, high-temperature flows through and expands adiabatically. The mechanism entails the 's driving turbine blades, leading to a decrease in , , and while generating mechanical work output, often resulting in a wet vapor mixture at the . The purpose is to extract useful work from the fluid's stored , powering generators or other machinery. The fourth process (4-1) involves isobaric heat rejection in the condenser, where the low-pressure exhaust steam from the turbine is cooled at constant pressure. The mechanism includes latent heat removal to condense the vapor back into saturated or subcooled liquid, with heat transferred to a cooling medium such as water or air. The purpose is to restore the working fluid to its initial liquid state, facilitating efficient recirculation and completing the cycle while rejecting waste heat to the environment.

Thermodynamic Representation

The thermodynamic representation of the Rankine cycle is typically illustrated using temperature-entropy (T-s) and pressure-volume (P-v) diagrams, which provide visual insights into the state changes of the working fluid, usually water, during the cycle's processes. On the T-s diagram, the cycle is plotted with entropy (s) on the horizontal axis and temperature (T) on the vertical axis, featuring the saturation dome that separates the liquid, vapor, and two-phase regions. State 1 represents the saturated liquid at the condenser pressure, located on the liquid saturation line under the dome. From state 1 to 2, isentropic compression in the pump occurs nearly vertically upward along a constant entropy line, moving the fluid to a compressed liquid state at the boiler pressure, with minimal entropy change due to the low specific volume of the liquid. The process from 2 to 3 involves constant-pressure heat addition in the boiler, following the constant-pressure line that rises from the compressed liquid state 2 to the saturated liquid line, proceeds horizontally through the two-phase region under the dome, and then rises into the superheated vapor region to state 3 as superheated vapor. Isentropic expansion from 3 to 4 in the turbine follows another vertical line downward, ending in the wet vapor region under the dome at state 4. Finally, constant-pressure heat rejection from 4 to 1 is a horizontal line leftward through the two-phase region back to the saturated liquid at state 1. The area under the curve from 2 to 3 represents the heat input (q_in), while the area under the curve from 4 to 1 represents the heat rejected (q_out), allowing for visual estimation of thermal efficiency as the ratio of the difference in these areas to q_in. Constant-pressure lines are horizontal in the two-phase region due to constant saturation temperature but curve upward in the single-phase regions. The P-v diagram, with pressure (P) on the vertical axis and specific volume (v) on the horizontal axis, highlights the phase boundaries and volume changes more prominently than the T-s diagram. State 1 is again the saturated liquid at low pressure, near the left side of the dome. The pump compression from 1 to 2 is a nearly vertical line with a small increase in v, reflecting the incompressible nature of the liquid and resulting in state 2 as compressed liquid at high pressure. Heat addition from 2 to 3 at constant high pressure appears as a horizontal line extending far to the right into the superheated vapor region, with state 3 having a large specific volume. Isentropic expansion from 3 to 4 slopes downward to the right initially then leftward as it enters the two-phase dome, reaching state 4 as wet vapor at low pressure with reduced v. Heat rejection from 4 to 1 at constant low pressure is a horizontal line leftward to the saturated liquid dome boundary. Unlike gas power cycles such as the Otto or Diesel, which operate entirely in the vapor phase with closed loops avoiding phase changes, the Rankine cycle's P-v diagram crosses the saturation dome, illustrating the benefits of latent heat utilization for higher work output. The enclosed area within the cycle loop directly represents the net work output (w_net), as it quantifies the difference between expansion and compression work. These diagrams facilitate the interpretation of cycle performance by visually delineating the state points—1 (saturated liquid), 2 (compressed liquid), 3 (superheated vapor), and 4 (wet vapor)—and revealing key thermodynamic relationships without numerical computation. They are particularly advantageous for identifying irreversibilities, such as non-isentropic processes that cause deviations from vertical lines on the T-s diagram or slanted lines on the P-v diagram, and for spotting opportunities for improvements like increasing superheat or reducing pressure to enlarge the work area.

Mathematical Model

Key Variables

The key variables in the Rankine cycle analysis encompass thermodynamic state properties at the four cycle points and overarching parameters that define operating conditions, enabling the evaluation of energy transfers and cycle performance. These properties are primarily for water as the working fluid and are determined using thermodynamic tables or charts due to the fluid's behavior across phase boundaries. At state point 1 (condenser exit and pump inlet), the fluid is saturated liquid under condenser pressure P_1 = P_\text{low}, with temperature T_1 equal to the saturation temperature at P_\text{low}, specific volume v_1 as the saturated liquid volume v_f, enthalpy h_1 as the saturated liquid enthalpy h_f, and entropy s_1 as the saturated liquid entropy s_f. State point 2 (pump exit and boiler inlet) features compressed liquid at boiler pressure P_2 = P_\text{high}, where entropy s_2 \approx s_1 under ideal isentropic compression, temperature T_2 slightly exceeds T_1, specific volume v_2 \approx v_1, and enthalpy h_2 accounts for the small work input during compression. At point 3 (boiler exit and turbine inlet), the fluid is superheated vapor at P_3 = P_\text{high} and maximum cycle temperature T_3 = T_\text{max}, with enthalpy h_3 and entropy s_3 obtained from superheated vapor tables. Point 4 (turbine exit and condenser inlet) is typically a two-phase mixture at P_4 = P_\text{low}, defined by vapor quality x_4 (the mass fraction of vapor), where h_4 = h_f + x_4 (h_g - h_f), s_4 = s_f + x_4 (s_g - s_f), v_4 = v_f + x_4 (v_g - v_f), and T_4 = T_\text{sat}(P_\text{low}). Cycle parameters include P_\text{high} (typically 10–100 in conventional steam power plants), P_\text{low} (often around 0.1 under saturated conditions), maximum T_\text{max} (ranging from 500–600°C), and exit quality x_4 (ideally near 0.9–1.0 to minimize ). The \dot{m} (in kg/s) scales the for , while \dot{Q}_\text{in} to the and rejection \dot{Q}_\text{out} from the (both in kW), along with pump work \dot{W}_\text{pump} and work \dot{W}_\text{turbine} (in kW), quantify energy interactions; these are often analyzed on a per-unit-mass basis before applying \dot{m}. include in or kPa, in °C or , in m³/kg, enthalpy in kJ/kg, and entropy in kJ/(kg·). These variables facilitate analysis by encapsulating phase transitions—from subcooled liquid at point 1, to compressed liquid at point 2, superheated vapor at point 3, and wet vapor at point 4—allowing precise property retrieval from steam tables to model non-ideal behaviors like latent heat absorption without relying on ideal gas approximations.

Governing Equations

The governing equations for the ideal Rankine cycle are derived by applying the first law of thermodynamics to each component as a steady-flow open system, assuming steady-state operation, negligible changes in kinetic and potential energies, and no heat or work losses other than those specified for each process. The first law for a steady-flow process per unit mass simplifies to q - w = h_2 - h_1, where q is heat transfer, w is shaft work (positive when done by the system), and h is specific enthalpy. For the ideal cycle, processes 1-2 (pump) and 3-4 (turbine) are isentropic, so s_2 = s_1 and s_4 = s_3, enabling determination of outlet states from inlet conditions using thermodynamic property relations. For the pump (process 1-2), the process is adiabatic (q = 0) and reversible. From the first law, w = h_1 - h_2 < 0 (negative, indicating work input to the system). Approximating the working fluid as an incompressible liquid, the enthalpy change is h_2 - h_1 = v_1 (P_2 - P_1), so the magnitude of the pump work input per unit mass is w_\text{pump} = h_2 - h_1 \approx v_1 (P_2 - P_1) > 0. For mass flow rate \dot{m}, the total pump work input is \dot{W}_\text{pump} = \dot{m} v_1 (P_2 - P_1). For the (process 2-3), there is no shaft work (w = 0), so q_\text{in} = h_3 - h_2. The input is therefore \dot{Q}_\text{in} = \dot{m} (h_3 - h_2). For the (process 3-4), the process is adiabatic (q = 0) and reversible, yielding w_\text{turbine} = h_3 - h_4 (work output). The work is \dot{W}_\text{turbine} = \dot{m} (h_3 - h_4). For the condenser (process 4-1), there is no shaft work (w = 0), so q = h_1 - h_4 < 0 (heat leaving the system). The magnitude of the heat rejected per unit mass is thus q_\text{out} = h_4 - h_1 > 0. The total heat rejection is \dot{Q}_\text{out} = \dot{m} (h_4 - h_1). The net work output per unit mass is w_\text{net} = w_\text{turbine} - w_\text{pump} = (h_3 - h_4) - v_1 (P_2 - P_1), and the total net work is \dot{W}_\text{net} = \dot{W}_\text{turbine} - \dot{W}_\text{pump}. The is \eta = \frac{w_\text{net}}{q_\text{in}} = 1 - \frac{q_\text{out}}{q_\text{in}} = 1 - \frac{h_4 - h_1}{h_3 - h_2}.

Ideal and Real Cycles

Ideal Rankine Cycle

The of the ideal Rankine cycle is calculated using the formula \eta = 1 - \frac{h_4 - h_1}{h_3 - h_2}, where h_1, h_2, h_3, and h_4 are the specific enthalpies at the inlet, outlet, inlet, and outlet, respectively. These enthalpies are determined from steam tables based on the cycle conditions, assuming isentropic processes in the and . For typical operating conditions of a pressure of 100 and temperature of 500°C, with a pressure of 0.1 , the enthalpies are approximately h_1 = 192 kJ/kg, h_2 = 202 kJ/kg, h_3 = 3375 kJ/kg, and h_4 = 2090 kJ/kg, yielding an efficiency of about 40%. The of the ideal Rankine cycle improves with higher or greater superheat , as these parameters raise the average at which is supplied to the , thereby increasing the net work output relative to input. Lowering the also enhances by reducing the average of rejection, which widens the overall span of the cycle. These effects are evident in performance analyses using steam property data, where, for instance, increasing superheat from saturation to 500°C at a fixed can boost by 5-10 percentage points. Compared to the operating between the same maximum and minimum s, the ideal Rankine cycle achieves lower because addition occurs over a range of s from the saturation point to the superheat , rather than isothermally at the maximum . Similarly, rejection is isothermal at the T_L, but the effective mean for addition T_{H,\text{mean}} is lower than the peak T_H, resulting in \eta = 1 - T_L / T_{H,\text{mean}} < 1 - T_L / T_H = \eta_{\text{Carnot}}. For example, under conditions yielding 40% Rankine , the corresponding Carnot might exceed 60%. The Rankine cycle assumes perfect isentropic and , with no irreversibilities, and includes the work in calculations, though this work is small—typically 1-3% of the work—and sometimes approximated as negligible in preliminary estimates. These assumptions establish theoretical benchmarks but highlight limits, as real cycles incorporate additional losses.

Deviations in Real Cycles

In real Rankine cycles, deviations from the ideal model arise primarily due to irreversibilities in the components, leading to reduced compared to the theoretical predictions of 40-50% for ideal cases. These non-idealities include , losses, drops, and property effects, which increase generation and alter the work and heat transfers. Pump inefficiencies stem from non-isentropic compression caused by mechanical , , and hydraulic losses, resulting in higher actual work input than the ideal isentropic value. The actual pump work is given by W_{\text{pump, real}} = \frac{W_{\text{pump, ideal}}}{\eta_{\text{pump}}}, where the isentropic efficiency \eta_{\text{pump}} typically ranges from 70% to 90% in practical steam power plants. This inefficiency elevates the at the pump outlet, thereby decreasing the net cycle work output. Boiler losses occur due to drops across the complex circuits, including fittings, headers, and tubing, as well as incomplete and imperfect from the combustion gases to the . These effects reduce the effective heat input Q_{\text{in}} by 2-5% or more, depending on the type and burner design, as incomplete combustion leads to unburned and excess air requirements that lower the overall energy transfer . drops can amount to 5-10% of the boiler inlet , further diminishing the cycle's performance. Turbine inefficiencies arise from non-isentropic expansion due to aerodynamic , steam leakage past blade tips, finite stage expansions, and partial admission losses, causing the actual exhaust to be higher than the isentropic value. The isentropic is defined as \eta_t = \frac{h_3 - h_{4,\text{real}}}{h_3 - h_{4,s}}, where h_3 is the inlet , h_{4,\text{real}} is the actual outlet , and h_{4,s} is the isentropic outlet ; typical values for turbines range from 80% to 90%. In the low-pressure stages, in the expanding exacerbates losses through , reduced aerodynamic , and wet , potentially dropping stage efficiencies by an additional 5-10%. Condenser deviations include pressure drops in the and of the below the , which increases the actual rejection Q_{\text{out}} beyond the ideal saturated state. , often by 5-10°C to ensure proper and prevent in the feedwater system, is constrained by the cooling water inlet , typically limiting the minimum condenser pressure to 5-10 kPa and raising Q_{\text{out}} by 1-3%. These factors, combined with air inleakage and , contribute to higher turbine backpressure and reduced net work. Overall, these deviations result in real Rankine cycle thermal of 30-40% in conventional steam power plants, significantly lower than ideal benchmarks due to the cumulative effects of component losses and generation. Blade losses and in low-pressure stages alone can account for 5-15% of the work reduction. To account for these in , steam tables are used to determine real enthalpies by incorporating irreversibilities, such as through isentropic corrections and balances, enabling more accurate performance predictions.

Advanced Variations

Reheat Rankine Cycle

The reheat Rankine cycle enhances the basic Rankine cycle by introducing a reheating to improve performance in high-pressure steam power plants. In this configuration, from the at and maximum expands isentropically in a to an intermediate . The partially expanded steam is then returned to the or a separate reheater, where it is heated at constant back to the original maximum before entering a low-pressure for further isentropic expansion to the condenser . This setup effectively divides the turbine work into two while elevating the temperature profile during heat addition. The primary processes in the reheat cycle modify the expansion step of the ideal Rankine cycle: the isentropic expansion from state 3 (boiler exit) to state 4 is split into expansion from state 3 to 4' in the high-pressure turbine, followed by constant-pressure reheating from 4' to 5, and then expansion from 5 to 6 in the low-pressure turbine. The reheat heat input is calculated as Q_{\text{reheat}} = \dot{m} (h_5 - h_{4'}), where \dot{m} is the of and h denotes specific . The work and other processes ( and ) remain similar to the basic cycle. The is given by \eta_{\text{reheat}} = \frac{ (h_3 - h_{4'}) + (h_5 - h_6) - W_{\text{pump}} }{ (h_3 - h_2) + (h_5 - h_{4'}) }, where W_{\text{pump}} is the pump work input, typically small compared to turbine work. This equation reflects the increased net work output relative to total heat input, with the reheat term contributing to higher average heat addition temperature. Key benefits of the reheat cycle include an efficiency improvement of 4-5% over the simple Rankine cycle, achieved by raising the average temperature at which heat is added, which approaches the Carnot limit more closely. Additionally, reheating reduces moisture content in the low-pressure turbine exhaust steam—typically limiting it to under 10% wetness fraction—thereby minimizing erosion of turbine blades caused by liquid droplets and extending equipment life. These advantages are particularly valuable in cycles operating at high boiler pressures, where excessive moisture would otherwise occur during deep expansion. Despite these gains, the reheat cycle involves drawbacks such as increased from additional reheater components, , and systems, which add to the plant design. It is thus economically justified mainly in large-scale applications, such as coal-fired or plants with capacities exceeding 500 MW, where the efficiency benefits offset the higher upfront . The reheat Rankine cycle was introduced in the mid-1920s as pressures rose in power generation, enabling reliable operation at elevated conditions.

Regenerative Rankine Cycle

The regenerative Rankine cycle modifies the basic cycle by extracting steam from the turbine at intermediate pressure stages to preheat the feedwater before it enters the boiler, thereby reducing the heat input required from the external source and improving overall efficiency. This preheating is accomplished using feedwater heaters, which transfer heat from the extracted steam to the subcooled liquid, minimizing the temperature difference during heat addition in the boiler and approaching the ideal of reversible heat transfer. In the configuration, steam is bled from the after partial expansion and directed to one or more feedwater heaters arranged in series or parallel along the feedwater line. The preheated feedwater is then returned to the at a higher , while the condensed steam from the heaters is either mixed back into the or pumped separately. Large power plants typically employ 4 to 8 stages of regeneration to optimize the profile, with extraction points selected at pressures that match the conditions suitable for the heaters. Feedwater heaters are classified into open and closed types. Open feedwater heaters operate on direct mixing, where extracted steam bubbles into the feedwater stream, achieving excellent due to intimate contact, but requiring the streams to enter at the same ; the outlet is a single saturated liquid stream at the heater . In contrast, closed feedwater heaters use a shell where condenses on one side without mixing, and the feedwater is heated on the other side before being pumped to the next stage; this allows operation at different pressures but introduces additional pumping requirements for the drain. Open heaters are simpler and common for lower stages, while closed heaters predominate in higher-pressure sections to maintain cycle integrity. The processes in a regenerative involve multiple points during expansion, typically between the high-pressure and low-pressure stages (corresponding to points 3 and 4 in the basic T-s diagram). For an open , the balance ensures : (1 - y) h_6 + y h_4 = h_7, where y is the fraction of steam , h_4 is the at the point, h_6 is the incoming feedwater , and h_7 is the preheated outlet (saturated at heater ). This yields y = (h_7 - h_6) / (h_4 - h_6). For closed heaters, the balance is across the : \dot{m}{\text{}} (h{\text{}} - h_{\text{condensate}}) = \dot{m}{\text{feed}} (h{\text{preheated}} - h_{\text{incoming}}), with no mass mixing. These balances determine the fractions needed to achieve desired preheating temperatures. The primary benefits include raising the average temperature of heat addition, which enhances the Carnot efficiency factor and reduces exergy losses from irreversible mixing in the boiler; this typically improves thermal efficiency by 5-10% over the basic cycle, depending on the number of heaters and operating conditions. Additionally, preheating reduces thermal stresses on boiler tubes by minimizing the temperature gradient between feedwater and combustion gases, extending equipment life. The modified efficiency calculation accounts for reduced turbine work output due to extractions but offsets it with lower boiler heat input: \eta_{\text{regen}} = \frac{W_{\text{turb,net}} - W_{\text{pump}}}{Q_{\text{in,reduced}}}, often approximated as \eta_{\text{basic}} + \Delta T_{\text{avg}} / T_{\text{high}} \times \eta_{\text{Carnot}}, where \Delta T_{\text{avg}} reflects the temperature rise from regeneration. Limitations arise from the increased system complexity, including additional , valves, and pumps, which elevate and costs; optimal pressures are determined through economic analyses balancing gains against these expenses. In , regeneration is most viable in large-scale where the improvements justify the added hardware.

Organic Rankine Cycle

The (ORC) is a variation of the Rankine cycle designed for low-temperature recovery, employing working fluids with high molecular weights instead of to enable efficient power generation from sources typically below 400°C. Unlike steam-based cycles, which struggle with at low temperatures due to the need for operation, the ORC maintains positive pressures throughout, facilitating compact system design and operation with sources as low as 80°C. This adaptation preserves the core four processes—pumping, , , and —but shifts to subcritical conditions with reduced pressures (often below 20 ) and temperatures, making it ideal for low-grade where traditional steam cycles yield efficiencies under 5%. Fluid selection in ORC systems prioritizes organic compounds with low boiling points and favorable thermodynamic properties to match low-temperature sources, such as refrigerants like R134a or hydrocarbons like n-pentane, which exhibit dry vapor expansion curves that avoid liquid droplet formation during expansion and subsequent erosion. These fluids, often with molecular weights exceeding 100 g/mol, enable at temperatures 50–100°C lower than , enhancing and cycle performance while minimizing isentropic losses. Selection criteria also consider critical temperature alignment with the heat source to maximize vapor density and expander efficiency. ORC applications focus on waste heat recovery from industrial processes, such as exhaust gases at 100–300°C, alongside geothermal and solar thermal systems, where net efficiencies range from 10% to 20%—significantly lower than the 30–40% of high-temperature steam cycles but valuable for otherwise unused energy. These systems convert low-grade heat into electricity at scales from kilowatts to megawatts, with modular designs supporting distributed generation in biomass plants or automotive exhaust recovery. Advantages include compact turbines suited to organic vapors' higher densities and the absence of vacuum condensers, reducing complexity and enabling partial-load operation above 50% efficiency. Post-2000 advancements, driven by regulatory incentives for renewables, have scaled modular ORC units up to 10 MW for biomass and industrial cogeneration, enhancing reliability through non-corrosive fluids. Despite these benefits, systems face challenges including of organic fluids above 300–350°C, which limits maximum operating temperatures and requires careful fluid matching, alongside higher specific costs (often 2–3 times that of cycles per kW due to specialized components). Environmental concerns arise from some fluids' (e.g., older HFCs) or high , prompting shifts to low-ODP alternatives like hydrofluoroolefins. Performance optimization centers on the approximate given by the Carnot-like formula: \eta_{\text{ORC}} = 1 - \frac{T_{\text{low}}}{T_{\text{high, mean}}} where T_{\text{low}} is the condenser temperature and T_{\text{high, mean}} is the mean evaporator temperature (both in Kelvin), further refined by expander isentropic efficiencies of 70–85%, which directly impact net power output.

Supercritical Rankine Cycle

The supercritical Rankine cycle operates above the critical point of water (374°C and 221 bar), where there is no distinct boiling phase transition; instead, the working fluid transitions continuously from a liquid-like state to a dense, gas-like supercritical fluid during heat addition, enabling a single-phase process that eliminates the two-phase evaporation stage typical of subcritical cycles. This configuration allows for more efficient heat transfer in the boiler, as the fluid's properties change gradually without the formation of bubbles or steam drums. In practice, supercritical cycles typically employ pressures exceeding 250 and maximum temperatures above 550°C, while ultra-supercritical variants push these limits further to around 300 and 600–700°C, often incorporating double reheat stages to manage thermal stresses and enhance performance. These plants utilize once-through boilers, which feedwater directly through tubes without recirculation drums, simplifying the but requiring precise control to achieve uniform heating. Components must withstand extreme conditions, necessitating specialized materials such as nickel-based alloys (e.g., ) to resist , oxidation, and creep at high temperatures. Efficiency in supercritical cycles typically achieves 38–42%, while ultra-supercritical variants reach 42–48%, compared to 33–38% in subcritical plants, primarily due to a closer approximation to the Carnot limit through higher average addition temperatures and reduced irreversibilities in the single-phase process. This improvement translates to lower specific fuel consumption and reduced CO₂ emissions per in coal-fired applications, potentially cutting emissions by 10–15% relative to subcritical units. The first commercial supercritical plant, the 125 MW Philo Unit 6 in , , entered operation in 1957, marking the transition from experimental to practical deployment. By 2025, supercritical and ultra-supercritical technologies dominate new coal-fired capacity in , particularly in and , where they support "clean coal" initiatives by maximizing from domestic resources; many recent installations utilize ultra-supercritical conditions exceeding 600°C. Ongoing research focuses on and cycles targeting 700°C+ for efficiencies above 50%, driven by international collaborations including the Department of Energy's Advanced Ultra-Supercritical (A-USC) program. However, drawbacks include significantly higher —up to 20–30% more than subcritical plants—due to exotic materials and complex controls, as well as heightened sensitivity to feedwater impurities, which can cause rapid or tube failures. In the temperature-entropy (T-s) , the heat addition process features a characteristic "slide region" near the pseudo-critical line, where peaks, mimicking pseudo-boiling but introducing challenges in flow stability and distribution.

References

  1. [1]
    [PDF] Chapter 9 - Civil, Environmental and Architectural Engineering
    In the Rankine cycle, which is the ideal cycle for steam power plants, the working fluid flows through a series of steady-flow devices such as the turbine and ...
  2. [2]
    Thermodynamic Foundations – Introduction to Aerospace Flight ...
    William John Macquorn Rankine develops the Rankine cycle (ideal vapor power cycle) and publishes a foundational textbook on thermodynamics. Together with ...
  3. [3]
    8.5 Rankine Power Cycles - MIT
    In the Rankine cycle, the mean temperature at which heat is supplied is less than the maximum temperature, $ T_2$ , so that the efficiency is less than that of ...Missing: explanation | Show results with:explanation
  4. [4]
    7.6. Rankine cycle | EME 812: Utility Solar Electric and Concentration
    The Rankine cycle system consists of a pump, boiler, turbine, and condenser. The pump delivers liquid water to the boiler. The boiler heated by the solar heat ...
  5. [5]
    [PDF] Chapter 8 - WPI
    ► Rankine cycle modifications known as superheat and reheat permit advantageous operating pressures in the boiler and condenser while avoiding low-quality ...Missing: explanation | Show results with:explanation
  6. [6]
    [PDF] ASTER - UNT Digital Library
    The efficiency of a Rankine cycle is primarily determined by the temperatures of heat addition and rejection. However, no working fluid has been identified ...<|control11|><|separator|>
  7. [7]
    RANKINE CYCLE - Thermopedia
    The Rankine cycle is the fundamental operating cycle of all power plants where an operating fluid is continuously evaporated and condensed.
  8. [8]
    Steam Rankine Cycle - an overview | ScienceDirect Topics
    The steam Rankine cycle is one of the most important ways to transform on large scale thermal energy into power. Examples are nuclear and coal-fired power ...
  9. [9]
    Thermodynamic cycles for solar thermal power plants: A review
    Oct 17, 2021 · Currently, the steam Rankine cycle is the most widespread and commercially available option, usually coupled to a parabolic trough solar field.Abstract · INTRODUCTION · CONCENTRATING SOLAR... · PROPOSALS OF...
  10. [10]
    None
    ### Summary of Rankine Cycle from the Document
  11. [11]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Dec 15, 2021 · (d) The typical thermal efficiency of a standard Rankine cycle is on the order of 20 - 40%. (e) The ratio of the power required by the pump to ...
  12. [12]
    Watt steam engine | Definition, History, & Facts - Britannica
    Sep 29, 2025 · He added a pump to remove condensed air and steam and added components that ensured that the piston stayed hot and the condensing tank stayed ...
  13. [13]
    Rankine cycle | Thermodynamics, Heat Engines, Efficiency | Britannica
    It is used as a thermodynamic standard for rating the performance of steam power plants. The cycle was described in 1859 by the Scottish engineer William J.M. ...
  14. [14]
    William John Macquorn Rankine | Thermodynamics, Heat Engines ...
    Sep 27, 2025 · His classic Manual of the Steam Engine and Other Prime Movers (1859) was the first attempt at a systematic treatment of steam-engine theory.
  15. [15]
    Steam Engines and the Industrial Revolution - ThoughtCo
    May 1, 2025 · At first, steam was used to pump water back up into reservoirs, but this soon developed and steam was used to power bigger and better blast ...Key Takeaways · Industrial Power Pre 1750 · The Development Of SteamMissing: Rankine reciprocating
  16. [16]
    Steam Engines - an overview | ScienceDirect Topics
    The principle of operation of the steam engines is the same as for the conventional Rankine cycle where an engine is used instead of the turbine. ... Early ...
  17. [17]
    Turbines - Visual Encyclopedia of Chemical Engineering Equipment
    Apr 13, 2022 · The first steam turbine, an impulsive-type steam turbine, was developed by Carl Gustav de Laval in 1883. In 1884, Charles Parsons altered the ...
  18. [18]
    THE HISTORICAL EVOLUTION OF TURBOMACHINERY
    ... 1883) • Sir Charles Parsons (Patent date, 1884) • Auguste Rateau ... Steam Turbine Developments of Gustav De Laval Carl Gustav De Laval was born ...
  19. [19]
    [PDF] Section 4. Technology Characterization – Steam Turbines
    The first steam turbine used for power generation was invented in 1884. Following this initial introduction, steam turbines rapidly replaced reciprocating ...
  20. [20]
    History of Power: The Evolution of the Electric Generation Industry
    Oct 1, 2022 · Sir Charles Parsons, who built the first steam turbine generator (with a thermal efficiency of just 1.6%) in 1884, improved its efficiency two ...
  21. [21]
    Power to the People - ASME Digital Collection
    In the early 20th century, electric power was generated by cogeneration ... steam turbines began to replace steam engines to drive 1,200-rpm electric generators.<|control11|><|separator|>
  22. [22]
    Philo 6 Steam-Electric Generating Unit - ASME
    1957. The world's first commercial supercritical steam-electric generating unit · Landmark Location. American Electric Power · Owner or plaque location, if ...
  23. [23]
    Supercritical Cycle - an overview | ScienceDirect Topics
    In a supercritical Rankine cycle, the working fluid is heated directly from the liquid state into the supercritical state (b and c), bypassing the two phase ...
  24. [24]
    Optimizing Efficiency and Performance in a Rankine Cycle Power ...
    Mar 31, 2025 · This study aims to increase Rankine cycle efficiency by adjusting boiler and condenser pressures, achieving 47.1% efficiency and 85% steam ...
  25. [25]
    Exploring Real-World Applications of the Rankine Cycle - Medium
    Jan 1, 2024 · The Rankine Cycle enables the efficient conversion of geothermal heat into electricity, making it an important technology in the transition to ...Introduction · The role of the Rankine Cycle... · Applying the Rankine Cycle in...
  26. [26]
    World Nuclear Performance Report 2025
    Sep 1, 2025 · World Nuclear Performance Report 2025 charts this record milestone and demonstrates nuclear's endurance as part of a clean and resilient global ...
  27. [27]
    [PDF] Chapter 5 Engines, Power, and the Anthropocene
    Figure 5.12: Ideal Rankine Cycle schematic and T-S diagram. From state 2 to 3, we can again write the first law to relate the heat transferred, ˙Qin = ˙Q2→3 ...
  28. [28]
    [PDF] LECTURE NOTES ON INTERMEDIATE THERMODYNAMICS
    Mar 28, 2025 · This does represent an increase of efficiency over a comparable Rankine cycle without regeneration, which happens to be 0.369. 2.2.4 Losses.
  29. [29]
    Setting Up a Rankine Cycle Using CyclePad
    Sep 11, 2001 · CyclePad uses the equation wpump = -vDP to calculate reversable pump work because the fluid (liquid water) is very close to incompressible.Missing: derivation | Show results with:derivation
  30. [30]
    Theory of Rankine Cycle - Equations and Calculation - Nuclear Power
    The Rankine cycle is an idealized thermodynamic cycle of a constant pressure heat engine that converts part of heat into mechanical work.
  31. [31]
    Investigation of the thermodynamic performance of an existing ...
    Jan 9, 2021 · The analysis shows that the boiler is the major source of exergy destruction, because of the incomplete combustion process, and inappropriate ...
  32. [32]
    Maintaining Maximum Efficiency in Power Generation Units
    Sep 14, 2015 · (In actuality, this would be excessive moisture that could damage low-pressure turbine blades. Techniques such as steam reheating are common for ...
  33. [33]
    [PDF] IBR/_. - NASA Technical Reports Server
    efficiency. In an actual steam Rankine cycle, pressure losses occur in the boiler and condenser and various losses and irreversabili- ties associated with ...
  34. [34]
    [PDF] Chapter 10
    The ideal reheat Rankine cycle. Page 11. 11. The average temperature at which heat is transferred during reheating increases as the number of reheat stages is.
  35. [35]
    Rankine Cycle - an overview | ScienceDirect Topics
    Mathematical expressions for the efficiency of the Rankine cycle are more cumbersome than those for the Carnot efficiency, but the strong reliance on peak ...<|control11|><|separator|>
  36. [36]
    Performance Comparison of 500 MW Coal-Fired Thermal Power ...
    Aug 9, 2024 · A coal-fired power plant is based on the regenerative-reheat Rankine cycle. High-temperature and high-pressure steam with thermal energy is ...
  37. [37]
    [PDF] LECTURE SC06 Introduction to Thermodynamics - ICTP
    The reheat cycle was introduced in the mid-1920s, but was abandoned in the 1930s, because of operational difficulties. The steady increase in boiler pressures ...
  38. [38]
    [PDF] Rankine Cycle - MIT OpenCourseWare
    Apr 1, 2020 · Rankine Cycle: two phase region Superheat and Ultra-superheat Cycles. Reheating. Recuperation.Missing: four | Show results with:four
  39. [39]
    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.
  40. [40]
    [PDF] Expt.2 Experiments on Rankine Cycle Steam Power system
    From the steady state run data plot identify the steady state system parameters for the analysis. Record these into the data sheet provided. Calculate the heat ...Missing: entropy explanation
  41. [41]
    [PDF] Week 13 Chapter 10 Vapor & Combined Power Cycles - UOW
    An open (or direct-contact) feedwater heater is basically a mixing chamber, where the steam extracted from the turbine mixes with the feedwater exiting the pump ...<|control11|><|separator|>
  42. [42]
    Rankine Cycle for Closed Feed Water Heaters and ... - Electrical4U
    May 22, 2024 · Differentiate Between Open and Closed Feed Water Heater ; Open and simple, More complex in design ; Good heat transfer characteristics, Less ...
  43. [43]
    Working fluids for low-temperature organic Rankine cycles
    The ORC cycles operate between 100 and 30 °C typical for geothermal power plants at pressures mostly limited to 20 bar, but in some cases supercritical ...
  44. [44]
    Organic Rankine Cycle Power Systems: From the Concept to ...
    The development of medium-capacity (10–50 MWe) supercritical CO2 CBC power plants is now actively pursued for next generation concentrated solar power (CSP) ...
  45. [45]
    Brief Review of Working Fluid Selection for Organic Rankine Cycle
    In this study, the working fluids selection criteria for the ORC system have been reviewed. This study helps in identifying the potential organic fluids for.
  46. [46]
    (PDF) Working fluid selection of Organic Rankine Cycles
    Organic Rankine Cycles (ORCs) are identified as one of the best candidates to generate electricity from low-grade heat sources. ORCs operate on low ...
  47. [47]
    [PDF] Waste Heat Recovery: Technology and Opportunities in U.S. Industry
    It is estimated that somewhere between 20 to 50% of industrial energy input is lost as waste heat in the form of hot exhaust gases, cooling water, and heat lost ...
  48. [48]
    ORC System | Organic Rankine Cycle manufacturer - Turboden
    The absence of water and vapor in the ORC process eliminates the need for complex steam turbine systems, making the technology more compact, reliable, and ...
  49. [49]
    [PDF] Organic Rankine Cycle and Its Working Fluid Selection-A Review
    Mar 15, 2016 · Many of the organic fluids have a high negative impact on the greenhouse effect and ozone depletion problems. At the same time these organic ...
  50. [50]
    Experimental Investigation of the Performance of a Hermetic Screw ...
    Oct 16, 2025 · The test results show a maximum expander isentropic efficiency of 72.4% and a relative cycle efficiency of 10.5% at an evaporation temperature ...
  51. [51]
    [PDF] Thermodynamic Comparison of Organic Rankine Cycles Employing ...
    In the present work, the thermodynamic performance of two novel modifications to a traditional Organic Rankine Cycle is investigated: Organic Rankine Cycle ...
  52. [52]
    Ultra supercritical thermal power plant material advancements
    Thermal power plant efficiency can be raised by increasing the MS pressure and temperature and reducing the exhaust heat which is mainly in LP turbine exhaust.Ultra Supercritical Thermal... · 1. Introduction · 1.1. Thermal Power Plant...<|separator|>
  53. [53]
    Fundamentals of Industrial Boilers and Steam Generation Systems
    Another aspect is that water and steam become a single phase at supercritical pressure/temperature (3,200 psi/705°F) and above. There can be no steam/liquid ...
  54. [54]
    [PDF] CLEANER ENERGY - Nickel Institute
    For this purpose, a 725 °C test track with thick-walled pipelines and pipe fittings was set up in the Grosskraftwerk Mannheim GKM Power Plant in Germany, and.
  55. [55]
    Subcritical and supercritical Rankine steam cycles, under elevated ...
    Jan 12, 2024 · This study considers the impact of the temperature and pressure, independently, on the performance of a steam Rankine cycle.Missing: mid- | Show results with:mid-
  56. [56]
    Thermodynamic Analysis of Supercritical and Subcritical Rankine ...
    Sep 17, 2019 · Additionally, the supercritical Rankine cycles have better thermal efficiency than subcritical cycles, increases in average 6%, and consequently ...
  57. [57]
    [PDF] Increasing the Efficiency of Existing Coal-Fired Power Plants
    Carbon dioxide emissions from coal-fired power plants could thus be reduced by burning a better grade of coal, or by increasing the efficiency of the power ...<|separator|>
  58. [58]
    Ultrasupercritical Plant - an overview | ScienceDirect Topics
    Philo Unit 6 of American Electric Power (AEP), with 120 MWe of capacity, was considered to be the first supercritical steam generator and power plant. It was ...
  59. [59]
    China Doubles Down on Coal-Fired Power, With Record Plant ...
    “China is deploying ultra-supercritical and supercritical coal technology, which can operate at higher temperatures and pressures (above 600°C and 30 MPa) to ...