Fact-checked by Grok 2 weeks ago

Working fluid

A working fluid is a substance, typically a gas or such as , air, or a , that operates within thermodynamic cycles to facilitate the conversion of into mechanical work or to transfer in systems like heat engines, refrigerators, and heat pumps. In these cycles, the working fluid absorbs from a high-temperature source, undergoes processes involving , , and phase changes, and rejects excess to a low-temperature , enabling net work output while returning to its initial state. The selection and of a working fluid are critical to the and of thermodynamic devices, with key attributes including thermodynamic like , , , , and phase behavior, as well as transport properties such as thermal conductivity and viscosity. For instance, in heat engines operating on the , water serves as the working fluid, evaporating into steam in a to drive a for power generation, then condensing back to liquid in a cooler. Common examples also include air as an in Brayton cycles for gas turbines and refrigerants like R-134a in systems, where the fluid cycles through evaporation to absorb heat and condensation to release it. Applications of working fluids span power generation, , and cooling technologies, with ongoing research focused on developing alternatives to reduce and while maintaining high efficiency. Accurate measurement and modeling of working fluid properties, such as those provided by databases like NIST REFPROP, support the design of safe, reliable, and optimized systems.

Fundamentals

Definition

A working fluid is a substance, typically a or , that serves as the medium in heat engines, heat pumps, and other thermal systems, undergoing thermodynamic cycles to absorb from a high-temperature source, convert a portion of that heat into mechanical work, and reject the remaining to a low-temperature sink. This role is central to energy conversion processes, where the fluid circulates in a closed loop, enabling efficient transfer and transformation of in devices such as turbines and compressors. The basic thermodynamic principles governing working fluids involve cyclic processes of expansion and compression. During expansion, the fluid, heated and pressurized, performs mechanical work on a or blades, increasing volume while decreasing pressure. Compression then returns the fluid to its initial state, requiring input work, with the net output determined by the 's efficiency as dictated by the second law of thermodynamics. These principles allow heat-to-work conversion without net change in the fluid's over a complete . Unlike coolants, which primarily facilitate heat transfer without producing mechanical work, or lubricants, which reduce friction in mechanical components, working fluids actively participate in the full energy conversion cycle, directly interacting with heat sources and sinks to generate output. Common examples include water and steam for high-temperature applications, air for gas cycles, and refrigerants like R-134a for lower-temperature systems.

Historical Context

The concept of a working fluid traces its origins to ancient innovations in harnessing thermal energy. In the 1st century AD, Hero of Alexandria described the aeolipile, a rudimentary device that utilized steam generated from heated water as its working fluid to produce rotational motion through escaping jets, marking one of the earliest documented uses of a vapor-phase fluid for mechanical work. This steam-powered sphere, though not practically applied for power generation, demonstrated the potential of water vapor as a working medium in thermal devices. Centuries later, in 1698, Thomas Savery patented the first commercially viable steam pump, known as the "Miner's Friend," which employed steam as the working fluid to create a vacuum for raising water from mines, initiating the practical application of working fluids in industrial contexts. The accelerated advancements in working fluid technology, particularly with steam. In the 1760s, significantly improved upon earlier steam engines by introducing a separate and rotary motion capabilities, enhancing and versatility while continuing to rely on water as the primary working fluid; his 1769 patent laid the groundwork for widespread adoption in manufacturing and transportation. By the late , engineers shifted toward —vapor heated beyond its saturation point—to reduce condensation losses and improve engine performance, a development that became standard in locomotives and power plants during the early . Concurrently, Sadi Carnot's 1824 theoretical work, "Reflections on the Motive Power of Fire," established the ideal reversible cycle for heat engines, emphasizing the role of the working fluid's properties in maximizing and influencing subsequent selections of fluids for thermodynamic cycles. In the 20th century, diversification of working fluids expanded beyond water to meet specialized needs. The 1930s saw the introduction of chlorofluorocarbons (CFCs), such as dichlorodifluoromethane (R-12), developed by Thomas Midgley Jr. at General Motors as non-toxic, non-flammable alternatives for refrigeration cycles, revolutionizing cooling systems by replacing hazardous ammonia and sulfur dioxide. The organic Rankine cycle (ORC), a modification of the traditional Rankine cycle utilizing organic fluids like refrigerants or hydrocarbons with lower boiling points than water, has roots in the 19th century, with practical developments emerging in the 1930s to recover low-grade waste heat for power generation, including early industrial prototypes deployed in geothermal and industrial applications. Recent developments, up to 2025, reflect a global push toward sustainable working fluids amid environmental concerns. Amendments to the , particularly the 2016 , initiated the phase-out of high (GWP) hydrofluorocarbons (HFCs) that had succeeded CFCs, mandating reductions starting in 2019 for developed nations to curb climate impacts. In response, post-2010 innovations have promoted hydrofluoroolefins (HFOs), such as HFO-1234yf (GWP = 4), as low-GWP alternatives for and , offering comparable thermodynamic performance with minimal and enabling compliance with international regulations.

Properties

Thermodynamic Properties

The thermodynamic properties of working fluids govern their , transfer, and behavior in thermodynamic cycles, directly impacting the and feasibility of engines and systems. Key among these are specific capacities, enthalpies associated with changes, and critical parameters that define the fluid's boundaries. Specific heat capacity at constant (C_p) measures the required to raise the of a mass by one degree without volume restriction, while specific heat at constant volume (C_v) does so under isochoric conditions; for as a common working fluid, C_p is approximately 1.86 kJ/kg·K at 300 K, and C_v is about 1.40 kJ/kg·K, reflecting the fluid's to absorb during compression or expansion. These values vary with and , influencing the input needed for heating processes in cycles. Enthalpy of vaporization (\Delta H_{vap}), or , quantifies the energy absorbed during liquid-to-vapor at constant temperature and ; for at its boiling point of 100°C and 1 atm, \Delta H_{vap} is 2257 kJ/kg, a high value that allows significant per unit mass without temperature rise. This property is essential for fluids in vapor-based systems, where phase change drives much of the cycle's work potential. The critical temperature (T_c) and critical pressure (P_c) mark the conditions above which the fluid cannot be liquefied by pressure alone, eliminating distinct liquid and vapor phases; for , T_c = 647.096 K (374°C) and P_c = 22.064 MPa, setting limits for supercritical in advanced cycles. Fluids with higher T_c enable at elevated temperatures, approaching Carnot efficiency limits. In the gaseous phase, the ideal gas law relates state variables through PV = nRT, where P is , V volume, n moles, R the , and T , providing a foundational model for in turbines or compressors. For sensible heating without phase change, the enthalpy change is calculated as \Delta H = \int C_p \, dT, accounting for temperature-dependent absorption. Thermodynamic state variables—temperature (T), pressure (P), and internal energy (U)—are interrelated and visualized in pressure-volume (P-V) diagrams, where U depends primarily on T for ideal gases and work equals the enclosed area during processes, and in temperature-entropy (T-S) diagrams, where U variations reflect entropy changes (S) and heat transfer is the area under the curve. These representations highlight how P, V, T, and U evolve without depicting full cycle paths. Property tables, such as steam tables for water, tabulate values of specific enthalpy (h), entropy (s), and volume (v) across saturated, superheated, and subcooled states, enabling precise determination of fluid conditions for cycle performance evaluation. Mollier diagrams, plotting enthalpy versus entropy for steam, complement these by simplifying analysis of isentropic expansions and compressions in practical designs. A high enhances cycle efficiency by maximizing the difference between heat addition (during ) and rejection (during ), thereby increasing net work output per unit mass in vapor power cycles like the Rankine. For instance, water's substantial \Delta H_{vap} contributes to thermal efficiencies up to 40% in modern steam plants by optimizing extraction from transitions.

Transport and Chemical Properties

Transport properties of working fluids encompass , , and , which govern , , and momentum transport in thermodynamic cycles. Dynamic (μ) measures a 's resistance to , while kinematic (ν) is defined as ν = μ / ρ, where ρ is ; these properties decrease with increasing for most liquids, affecting pumping and regimes in heat exchangers and turbines. (k) quantifies the 's ability to conduct , typically ranging from 0.07 to 0.15 W/m·K for common liquid organic working fluids like refrigerants, and it influences the efficiency of surfaces in and condensers. (ρ) varies significantly with , often modeled by equations of state; for instance, water's decreases from about 1000 kg/m³ at 20°C to 958 kg/m³ at 100°C, impacting volumetric rates and system design. Chemical properties critical to working fluid performance include and , corrosiveness, and flammability. Thermal stability is characterized by the decomposition temperature, beyond which the fluid breaks down; for example, hydrofluoroolefins like HFO-1336mzz(Z) exhibit decomposition temperatures exceeding 300°C under sealed conditions, enabling high-temperature applications in organic Rankine cycles. prevents unwanted reactions, with many working fluids stable up to 200–400°C in the absence of oxygen or catalysts. Corrosiveness refers to the fluid's tendency to degrade system materials like or ; , a common , is corrosive to copper alloys but compatible with when properly inhibited. Flammability is classified by Standard 34 based on and ; class A1 fluids like R-134a are non-flammable and low-toxicity, while A2L fluids like R-32 have mild flammability for safer handling in systems. In , the () determines flow characteristics for working fluids in pipes and , calculated as \text{Re} = \frac{\rho v D}{\mu} where v is and D is ; flows are typically laminar for < 2300 in pipes, promoting efficient heat transfer but higher pressure drops, and turbulent for > 4000, enhancing mixing in turbine blades. Compatibility with system materials involves assessing interactions such as oxidation resistance and lubrication needs; polyolester lubricants are often required for fluids to prevent wear in compressors, while some organic fluids like offer inherent but may degrade over time. Measurement of these properties follows standardized methods for accuracy. is determined using rotational viscometers per ASTM D445, which measures kinematic viscosity at controlled temperatures for petroleum-derived and synthetic fluids. Thermal conductivity is assessed via transient hot-wire techniques outlined in ASTM D2717, suitable for liquids under . variations are quantified using pycnometers or vibrating-tube densitometers per ASTM D4052, ensuring reliable data for engineering models.

Behavior in Processes

Phase Transitions

Working fluids undergo several key phase transitions during thermodynamic cycles, primarily vaporization (boiling), where liquid converts to vapor by absorbing heat at constant temperature and pressure under saturation conditions, and condensation, the reverse process where vapor releases heat to form liquid. These transitions are essential for heat transfer in cycles like refrigeration and power generation, enabling efficient energy exchange without significant temperature changes. Above the critical point, fluids enter a supercritical state, exhibiting properties intermediate between liquid and gas, with no distinct phase boundary. The vapor pressure curve governing these transitions is described by the Clausius-Clapeyron equation, derived from in two-phase systems: \frac{dP}{dT} = \frac{\Delta H}{T \Delta V} where \Delta H is the , T is the temperature, and \Delta V is the change in between vapor and phases. This relation predicts the slope of the saturation curve, showing how pressure increases with temperature to maintain phase equilibrium, and is approximated for ideal gases as \ln P = -\frac{\Delta H}{R T} + C, allowing estimation of s across temperatures. Phase diagrams for working fluids illustrate these transitions via the liquid-vapor dome, a region bounded by saturated and vapor curves where the two phases coexist; the dome's apex is the critical point, beyond which distinct phases vanish. The marks the intersection where , , and vapor phases equilibrate at a unique and . For (CO₂), used in , the is at -56.6°C and 5.11 atm, while the critical point is 30.98°C and 72.79 atm; (NH₃), common in industrial cooling, has a at -77.7°C and 0.0606 , with a critical point at 132.4°C and 113.5 bar. Hysteresis in phase transitions arises from path-dependent behavior, particularly in , where the curve of versus wall superheat shows different paths for increasing and decreasing heat input due to metastable states. initiates or by forming vapor bubbles at surface imperfections or low-pressure sites; in , this leads to , where discrete bubbles enhance , contrasting with film boiling, where a continuous vapor layer insulates and reduces . , a dynamic process in flowing fluids, occurs when local pressure drops below , forming bubbles that collapse and cause erosion in pumps and turbines handling working fluids. In supercritical states, working fluids like CO₂ behave as a single phase above the critical point, lacking latent heat exchanges but offering liquid-like densities for compact heat transfer and gas-like diffusivity for low viscosity and rapid mixing, as seen in supercritical CO₂ Brayton cycles for power generation. This tunability enhances cycle efficiency by avoiding phase boundaries, with CO₂'s density reaching up to 0.47 g/cm³ near critical conditions while maintaining diffusivities around 10⁻⁸ m²/s, akin to gases.

Work Production

In thermodynamic cycles utilizing working fluids, mechanical work is primarily generated through the expansion of the fluid in turbines, where the work output is calculated as the integral of with respect to , W = \int P \, dV, representing the done by the expanding fluid on the turbine blades. Compression work in pumps, though typically smaller in magnitude, involves the input of to increase the fluid's , often approximated for incompressible liquids but following similar principles of change under . The fundamental operation of these cycles involves addition (Q_{in}) to the working fluid from a high-temperature source, which increases its and enables to produce net work output (W_{net} = Q_{in} - Q_{out}), where Q_{out} is the rejected to a low-temperature . The of the cycle is then defined as \eta = \frac{W_{net}}{Q_{in}}, quantifying the fraction of input converted to useful work. Idealized processes assume isentropic in the and in the , where (S) remains constant, maximizing work extraction by avoiding irreversibilities such as across finite differences. In practice, real processes deviate due to , fluid , and non-equilibrium effects, leading to reduced work output. These losses are quantified by the isentropic , defined for turbines as the ratio of actual work to the work achievable under isentropic conditions, and similarly for pumps as the ratio of isentropic work input to actual input. The of the working fluid during expansion significantly influences work production; in wet expansion, where the fluid enters the two-phase region with liquid droplets present, erosion of turbine blades can occur due to moisture impact, reducing and longevity. Conversely, dry expansion, typically achieved with superheated vapors or dry fluids that remain gaseous post-expansion, minimizes such damage and allows for higher work extraction without risks.

Selection Criteria

Performance Metrics

Performance metrics for working fluids in thermodynamic cycles primarily focus on energetic and exergetic indicators that quantify efficiency and work output under given operating conditions. Thermal efficiency, defined as the ratio of net work output to heat input, serves as a fundamental measure for power cycles like the Rankine cycle, where it typically ranges from 5-15% for low-temperature organic Rankine cycles (ORC) depending on the fluid and temperature span. For refrigeration and heat pump cycles, the coefficient of performance (COP) evaluates cooling or heating effectiveness, expressed as COP = Q_{cold} / W, where Q_{cold} is the heat absorbed from the cold reservoir and W is the compressor work input; typical values range from 2 to 5 for vapor-compression systems using fluids like R134a, reflecting the trade-off between temperature lift and irreversibilities. Exergy efficiency, or second-law efficiency, further refines these by accounting for irreversibilities, calculated as the ratio of actual exergy gain to the maximum available exergy from the heat source, often yielding 40-60% in ORC systems and highlighting fluids that minimize entropy generation during phase changes. Evaluation methods for working fluids incorporate figures of merit (FOM) tailored to specific cycles, particularly , to predict overall performance without full . One widely used FOM for low-temperature is a dimensionless that balances sensible and contributions, defined as FOM = Ja^{0.1} \left( \frac{T_{evap}}{T_{cond}} \right)^{0.8}, where Ja is the Jakob number (∫ C_p dT / ΔH_vap) and T_{cond}, T_{evap} are condensing and evaporating temperatures; this metric ranks fluids like higher than for geothermal applications due to better integration. Additionally, the volumetric , σ_v = V_{out} / V_{in} at expander inlet and outlet, is evaluated alongside FOM to assess expander sizing and , with optimal values around 3-10 for or radial turbines to avoid over- or under-expansion losses in fluids like R245fa. These methods enable comparative ranking of fluids, such as outperforming R123 in second-law for mid-temperature sources by closer alignment with Carnot limits (η_{Carnot} = 1 - T_{low}/T_{high}), where fluids with critical temperatures near the heat source temperature achieve up to 80% of the Carnot . Simulation tools like the NIST REFPROP database facilitate property-based performance prediction by providing accurate thermodynamic data for over 140 fluids, enabling cycle modeling to compute metrics such as isentropic efficiency and destruction; for instance, REFPROP calculations show R141b yielding higher in than hydrocarbons at evaporator temperatures below -20°C. Optimization strategies emphasize matching the fluid's critical point to operating temperatures for maximum work extraction, as fluids with critical temperatures 20-50 K above the heat source peak temperature, like for 150-200°C sources, maximize net power by enabling supercritical operation and reducing pinch point losses. This approach, validated in parametric studies, prioritizes fluids that enhance cycle irreversibility minimization without exceeding material limits.

Environmental and Safety Factors

The selection of working fluids in thermodynamic cycles must account for their environmental impacts, primarily measured by (GWP), (ODP), and atmospheric lifetime. GWP quantifies a fluid's contribution to relative to CO₂ over 100 years, with high-GWP hydrofluorocarbons (HFCs) like R-404A (GWP 3,260) posing significant risks from direct emissions during leaks or disposal. ODP assesses damage, where HFCs and hydrofluoroolefins (HFOs) have zero ODP, unlike earlier hydrochlorofluorocarbons (HCFCs) such as HCFC-22 (ODP >0). Atmospheric lifetime indicates persistence, with HFC-134a lasting 14 years and contributing to prolonged greenhouse effects, whereas low-GWP alternatives like (R-717) have negligible lifetime as a natural substance. International regulations govern these impacts through the , adopted in 1987 to phase out ozone-depleting substances, with subsequent amendments targeting HFCs via the 2016 . The sets global HFC phase-down targets, aiming for an 80-85% reduction by 2047, with developed countries freezing production in 2019 and reducing by 85% by 2036; amendments extend to 2025 compliance milestones, including U.S. allowance reviews for HFC sectors like . As of 2025, the U.S. AIM Act implements these targets with restrictions on high-GWP HFCs in new equipment effective January 1, 2025, aligning with global phase-down efforts. These frameworks promote transitions to low-GWP fluids to mitigate climate forcing. Safety considerations include and flammability, classified by Standard 34 into groups based on occupational exposure limits (OEL) and flame propagation. A (lower toxicity, OEL ≥400 ppm) includes most common fluids like , while Class B (higher toxicity, OEL <400 ppm) requires stricter handling. Flammability ranges from Class 1 (no flame propagation, e.g., A1 group like ) to Class 2L (lower flammability, burning velocity <10 cm/s, e.g., A2L like R-32 with lower flammability limit >10%) and Class 3 (higher flammability, e.g., A3 hydrocarbons like ). requirements under standards like ASME mandate leak testing to design pressures and relief valves to prevent overpressurization, especially for high-pressure fluids. Mitigation strategies address these risks through leak detection systems, such as electronic sensors triggering alarms at 25% of the , and of alternatives like R-1234yf (GWP 4, ODP 0, A2L classification), introduced in the 2010s for as a for HFC-134a. These measures, including adsorbent materials for containment, reduce emission risks. Life-cycle assessments evaluate total environmental footprints, encompassing manufacturing (e.g., 4.26–10.5 kg CO₂-eq per kg for HFCs like ), operational leaks, and end-of-life disposal. Reclamation recycles fluids, emitting 5.7–15.9 kg CO₂-eq less per kg than destruction, while lowering energy use by 82.5–250.6 per kg compared to . Such analyses highlight that disposal phases can account for up to 20% of impacts, favoring reclamation for .
ASHRAE Safety GroupToxicityFlammabilityExampleKey Characteristics
A1Lower (A)None (1)No flame propagation; OEL ≥400
A2LLower (A)Lower (2L)R-32, R-1234yfBurning velocity <10 cm/s; LFL >10%
A3Lower (A)Higher (3) (R-290)Fast propagation; requires

Applications

Heat Engines

In heat engines, working fluids play a central role in converting from sources into through thermodynamic , where the absorbs , expands to produce work, and rejects to complete the . These systems, such as and gas turbines, rely on the fluid's ability to undergo changes or compression-expansion processes efficiently, enabling the extraction of useful work while minimizing energy losses. The choice of fluid influences the operating temperatures, pressures, and overall , with common fluids selected for their thermodynamic and with engine components. A primary example is in the , widely used in steam power plants, where serves as the working fluid due to its high and of , allowing efficient energy transfer at moderate temperatures. In this , the fluid is heated in a to produce high-pressure , which expands through a to generate mechanical work, achieving efficiencies up to 40% in modern supercritical steam plants operating at elevated pressures and temperatures. Another key example is air as the working fluid in the for gas s, where atmospheric air is compressed, heated by combustion, and expanded in the , making it suitable for high-temperature applications like and , with efficiencies typically ranging from 30% to 40% depending on pressure ratios. Water and steam are particularly effective for low-temperature heat sources, such as those below 500°C, due to their and properties that match conventional or combustion. For advanced cycles targeting higher efficiencies with compact designs, is used in closed Brayton cycles for applications owing to its inertness and high thermal conductivity, while (sCO₂) enables recompression cycles with densities closer to liquids, improving performance; sCO₂ concepts were first proposed in the late by E.G. Feher and have seen commercialization in pilot plants during the through U.S. Department of Energy-funded projects, with ongoing advancements including the STEP Demo project's Phase 2 reconfiguration in 2025 and ' involvement as of September 2025. The core components of these heat engines include the (or ), , and , through which the working fluid follows a closed-loop path to ensure continuous operation. In the , from the source vaporizes or superheats the fluid, raising its and ; the high-energy fluid then flows to the , where it expands isentropically, converting into rotational shaft work that drives a ; finally, in the , the fluid rejects to a cooling medium, condensing back to a before being pumped to for return to the , minimizing parasitic losses. This fluid flow path optimizes energy extraction by maintaining steady-state conditions and reducing irreversibilities at each stage. In employing pressurized reactors (PWRs), acts as both and working fluid in the primary , where it is maintained at high pressure (around 15 MPa) to prevent boiling, transferring heat from the reactor core to a secondary that produces steam for the , enabling safe and efficient power output up to 1,000 MW per unit. For geothermal applications, (ORC) systems utilize as the working fluid to harness low-grade heat (80–150°C) from geothermal brines, where the organic fluid's low allows evaporation at lower temperatures than , driving a with efficiencies of 10–20% in like those in the U.S. . Advancements in efficiency include binary cycles that combine multiple fluids to better match variable heat source profiles, such as the , which uses an ammonia-water mixture as the working fluid to enable variable-temperature boiling and condensation, achieving 10–20% higher than traditional Rankine cycles for low-temperature sources like geothermal or recovery. This mixture allows internal heat recovery through and recombination processes, reducing destruction and improving overall system performance in commercial installations since the 1990s.

Refrigeration Systems

In refrigeration systems, working fluids facilitate absorption and rejection to achieve cooling, primarily through the vapor-compression cycle, which consists of four main processes: , where the fluid absorbs from the cooled space; , increasing the fluid's pressure and temperature; , releasing to the surroundings; and , reducing pressure to prepare for re-evaporation. This cycle relies on the fluid's ability to undergo phase changes efficiently at desired temperatures, with the typically operating at low pressures to match the cooling load and the at higher pressures for dissipation. Common working fluids in vary by application scale and environmental requirements. (R-717) is widely used in industrial systems due to its excellent thermodynamic properties and low cost, enabling efficient cooling in large-scale facilities like plants. For domestic and commercial , hydrofluorocarbons (HFCs) such as have been prevalent, but the prohibited new single-split systems containing F-gases with a (GWP) of 750 or higher—like , which has a GWP of 2088—starting January 1, 2025, to reduce climate impacts; similarly, in the United States, the American Innovation and Manufacturing (AIM) Act restricts higher-GWP HFCs in new refrigeration and equipment effective January 1, 2025. Hydrofluoroolefins (HFOs), such as R-1234yf with a GWP under 1, are increasingly adopted as low-GWP alternatives in these settings. Hydrocarbons like (R-290) serve as eco-friendly options in smaller systems, prized for their zero (ODP) and negligible GWP, though they require careful handling due to flammability. System design emphasizes matching the working fluid's boiling point to operational temperatures for optimal heat transfer in evaporators and condensers. In low-temperature freezers targeting -40°C, fluids like R-404A ( approximately -46.5°C at ) are selected to ensure evaporation occurs below the desired cooling level, maximizing efficiency while preventing frost buildup. , often coil-based, allow the fluid to boil and absorb from the refrigerated space, while condensers facilitate phase change back to liquid under elevated temperatures, typically 30–50°C above ambient, to reject heat effectively. Absorption refrigeration systems offer a non-mechanical , using rather than for compression, with serving as the refrigerant and lithium bromide as the absorbent in a common pair. In this , evaporates in a low-pressure generator-absorber setup, absorbing from the cooled medium, while the lithium bromide solution concentrates via heating (often from or solar sources) to release , followed by and dilution steps to maintain the process. This configuration is suited for applications where is scarce, such as remote or large building cooling, achieving coefficients of performance around 0.7–1.2 depending on source . Emerging technologies highlight natural working fluids like (CO2, R-744) in transcritical cycles, particularly for supermarket refrigeration, where the fluid operates above its critical point (31.1°C) in the gas cooler instead of a traditional . These systems gained widespread adoption post-2000s, with the first North American installations around 2010, offering energy savings of 10–20% over HFC-based systems in warm climates due to CO2's high heat transfer coefficients and non-flammability. In supermarkets, setups with CO2 handle both medium- and low-temperature loads, integrating subcritical for display cases and transcritical boosting for efficiency.

References

  1. [1]
    Refrigerants and Working Fluids | NIST
    May 3, 2011 · The conversion of primary energy (e.g. heat generated by the combustion of a fuel) into useful work most often involves a working fluid ...Missing: definition | Show results with:definition
  2. [2]
    None
    ### Summary: Working Fluid in Heat Engines
  3. [3]
    6.1 Heat engine – Introduction to Engineering Thermodynamics
    In a heat engine cycle, a working fluid may remain as a single-phase fluid or experience phase changes. A steam engine is a type of heat engine commonly used in ...
  4. [4]
  5. [5]
    Working Fluid - Wärtsilä
    May 21, 2018 · The working fluid of a heat engine or heat pump is a gas or liquid, usually called a refrigerant, coolant, or working gas, that primarily converts thermal ...
  6. [6]
    Hero's Aeolipile: The "Steam Engine" Built By Ancient Greeks
    Oct 5, 2023 · Hero of Alexandria described a sort of steam engine, used for an unknown purpose. The device is a simple steam turbine, which turns when the water container is ...
  7. [7]
    Steam Power | History of Western Civilization II - Lumen Learning
    A heat engine that performs mechanical work using steam as its working fluid. aeolipile: A simple bladeless radial steam turbine, also known as a Heron's ...
  8. [8]
    Thomas Savery and the Beginning of the Steam Engine - ThoughtCo
    Jul 3, 2019 · He patented the design of his first engine on July 2, 1698. A working model was submitted to the Royal Society of London. The Road ...Missing: fluid | Show results with:fluid
  9. [9]
  10. [10]
    Carnot Heat Engines: 200-Year-Old Technology Holds Key To Net ...
    In his 1824 treatise “Reflections on the Motive Power of Fire,” Carnot laid out a theoretical model for an engine operating between two heat reservoirs. Using ...
  11. [11]
    Chlorofluorocarbons (CFCs) - Global Monitoring Laboratory - NOAA
    CFCs were first synthesized in 1928 by Thomas Midgley, Jr. of General Motors, as safer chemicals for refrigerators used in large commercial appilications1.
  12. [12]
    Story of ORC Organic Rankine Cycle - Kaymacor
    Larjola led the development of high-speed hermetic turbogenerators in the hundreds kWE range, in which the turbine, generator and pump share the same shaft.
  13. [13]
    Hydrofluorocarbons and the Kigali Amendment to the Montreal ...
    Oct. 15, 2016 The international community reached agreement on the Kigali Amendment to the Montreal Protocol to phase out HFCs because of their global warming ...Missing: high fluids HFOs 2010
  14. [14]
    Environmental impact of HFO refrigerants & alternatives for the future
    Jun 11, 2021 · HFOs (Hydrofluoroolefins) are currently being promoted as low-GWP refrigerants by the chemical industry and various refrigeration and air conditioning system ...Missing: rise post-
  15. [15]
    2.2 Thermodynamic properties – Introduction to Engineering ...
    The common properties of a pure substance include pressure, temperature, specific volume, density, specific internal energy, specific enthalpy, and specific ...
  16. [16]
    Thermophysical Properties of Fluid Systems - the NIST WebBook
    Accurate thermophysical properties are available for several fluids. These data include the following:
  17. [17]
    Water Vapor - Specific Heat vs. Temperature
    Online calculators, figures and tables showing specific heat, Cp and Cv, of ... Specific Heat Capacity of Water: Temperature-Dependent Data and Calculator.
  18. [18]
    Water Properties: Vaporization Heat vs. Temperature - Charts and ...
    Online calculator, figures and tables showing heat of vaporization of water, at temperatures from 0 - 370 °C (32 - 700 °F) - SI and Imperial units.
  19. [19]
    Critical Temperatures and Pressures for some Common Substances
    Critical temperature is the temperature above which a substance cannot exist as a liquid. Critical pressure is the pressure needed to liquify a substance at ...
  20. [20]
    [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 ...
  21. [21]
    STEAM TABLES - Thermopedia
    The following tables of the properties of steam are taken directly from Chapter 5.5.3 of the Heat Exchanger Design Handbook, 1986, by CF Beaton.
  22. [22]
    [PDF] Thermodynamics II PCE 320 VAPOR POWER CYCLES
    ✓High enthalpy of vaporization. • Commonly used in coal plants, natural gas ... • The overall effect is an increase in thermal efficiency since the ...
  23. [23]
    [PDF] Transport Properties of Fluids
    They are the diffusion coefficient, the viscosity and the thermal conductivity and are associated with the transport of mass, momentum and energy respectively.
  24. [24]
    Reference Values and Reference Correlations for the Thermal ... - NIH
    Introduction. In this work, we review reference values and correlations for two important fluid transport properties: thermal conductivity and viscosity.
  25. [25]
    Thermal stability and decomposition mechanism of HFO‐1336mzz(Z ...
    May 15, 2019 · The decomposition temperatures of HFO-1336mzz(Z) at different pressure are much higher than its critical temperature (171.3°C), and the ...
  26. [26]
    Review of the Working Fluid Thermal Stability for Organic Rankine ...
    Jul 3, 2019 · This paper presents a review of the working fluid thermal stability for ORCs, including an analysis of the main theoretical method for thermal stability.<|control11|><|separator|>
  27. [27]
    [PDF] MATERIALS COMPATIBILITY AND LUBRICANTS RESEARCH ON ...
    The MCLR program addresses refrigerant and lubricant properties and materials compatibility. The primary elements of the work include data collection and ...Missing: corrosiveness | Show results with:corrosiveness
  28. [28]
    The Reynolds Number: A Journey from Its Origin to Modern ... - MDPI
    Generally, the flow in pipes is laminar when Re < 2000 and turbulent when Re > 2000–2300, with an intermediate transitional range, but varies depending on the ...
  29. [29]
    Material compatibility of ORC working fluids with polymers
    Aug 6, 2025 · This study points out the importance of material compatibility testing especially investigating the difference between hydrofluorocarbons (HFC) ...
  30. [30]
    D117 Standard Guide for Sampling, Test Methods, and ... - ASTM
    Jun 13, 2022 · This guide describes methods of testing and specifications for electrical insulating oils of petroleum origin intended for use in electrical cables.<|control11|><|separator|>
  31. [31]
    [PDF] Computational Fluid Dynamics - Therminol Heat Transfer Fluid
    Dec 18, 2019 · D2717 measures thermal con- ductivity, D1298 or D4052 measures specific gravity, and D445 or D7042 measures viscosity. D2766, the test method ...
  32. [32]
    8.1 Behavior of Two-Phase Systems - MIT
    First, there is a region in which liquid and vapor can coexist, bounded by the liquid saturation curve on the left and the vapor saturation curve on the right.
  33. [33]
    Calculation of self-diffusion coefficients in supercritical carbon ...
    Apr 1, 2021 · This paper presents an application of mean force kinetic theory (MFT) to the calculation of the self-diffusivity of CO2 in the supercritical ...
  34. [34]
    8.4 The Clausius-Clapeyron Equation - MIT
    The working fluid ... We can then derive an important relation known as the Clausius-Clapeyron equation, which gives the slope of the vapor pressure curve.
  35. [35]
    Clausius-Clapeyron Equation - Chemistry LibreTexts
    Mar 21, 2025 · The Clausius-Clapeyron equation allows us to estimate the vapor pressure at another temperature, if the vapor pressure is known at some ...Missing: fluids | Show results with:fluids
  36. [36]
    12.4: Phase Diagrams - Chemistry LibreTexts
    Jul 12, 2023 · The triple point also represents the lowest pressure at which a liquid phase can exist in equilibrium with the solid or vapor. At pressures less ...Missing: dome ammonia
  37. [37]
    (PDF) Flow-driven hysteresis in the transition boiling regime
    Sep 30, 2025 · The transition from nucleate to film boiling exhibits pronounced hysteresis,. with metastable coexistence observed along the rewetting branch, ...Missing: cavitation | Show results with:cavitation
  38. [38]
    Nucleation of boiling and cavitation - ResearchGate
    Aug 9, 2025 · A number of measurements of the conditions required to nucleate boiling and cavitation in stagnant water are described, and used to test ...
  39. [39]
    Immiscible CO2‐H2O fluids in the shallow crust - AGU Journals - Wiley
    Oct 6, 2006 · In addition to liquid-like density and solvent strength, supercritical CO2 possesses gas-like transport properties of viscosity and diffusivity ...
  40. [40]
    [PDF] E Q W ∆ = − 1∫=
    Expansion and compression work a. Transfer of energy to a system by boundary work requires that a force ... 171). 2. 1. ,. V b out. V. W. P dV. = ∫. Remember ...
  41. [41]
    [PDF] Isentropic Efficiency Guide - ME 201 (
    The isentropic is defined differently depending on whether we have a work producing device (i.e., turbine) or a work consuming device (pump or compressor) ...
  42. [42]
    Thermodynamic Foundations – Introduction to Aerospace Flight ...
    Specific Heats. The specific heats of a substance are terms frequently used in thermodynamics; they help measure how much its internal energy or enthalpy ...
  43. [43]
    [PDF] Selection of Working Fluids for Low Enthalpy Geothermal Organic ...
    Feb 12, 2020 · If the expansion exceeds the saturated vapor curve wet steam occurs which may erode the turbine blades due to the impact of water droplets. The ...
  44. [44]
    [PDF] Design of Organic Rankine Cycles for Conversion of Waste Heat in ...
    Organic Rankine cycles provide an alternative to traditional steam Rankine cycles for the conversion of low grade heat sources, where steam cycles are known ...<|control11|><|separator|>
  45. [45]
    Energy and exergy analyses for organic Rankine cycle driven by ...
    Results show that ORC operating with R245fa and R123 as working fluids have the highest thermal efficiencies (7.76% and 7.49 % respectively), and low exergy ...
  46. [46]
    Performance analysis of low temperature organic Rankine cycle with ...
    This paper proposed a dimensionless term, the “Figure of Merit” (FOM), to investigate the thermal performance of a low temperature, organic Rankine cycle ...
  47. [47]
    Geometric Design of Scroll Expanders Optimized for Small Organic ...
    From this data the desired volume expansion ratio RVP of the expander can be inferred from the working fluid properties, assuming 95% volumetric efficiency ...
  48. [48]
    REFPROP | NIST - National Institute of Standards and Technology
    Apr 18, 2013 · NIST Reference Fluid Thermodynamic and Transport Properties Database (REFPROP): Version 10. Download REFPROP 10: $325.00 PLACE ORDER.Ferrophosphorus (powder form) · Free SRD · FAQ
  49. [49]
    [PDF] Maximizing ORC performance with optimal match of working fluid ...
    The working fluids are arranged by decreasing critical temperature with R601 and R143a having highest and lowest critical temperature respectively.
  50. [50]
    [PDF] Energy and Global Warming Impacts of HFC Refrigerants and ...
    from inadvertent releases of the working fluid to the environment, but it can change the indirect ... APPENDIX B: ATMOSPHERIC LIFETIMES AND GWPs FOR REFRIGERANTS ...
  51. [51]
    The Montreal Protocol on Substances That Deplete the Ozone Layer
    In 2016, Parties to the Montreal Protocol adopted the Kigali Amendment to phase down production and consumption of hydrofluorocarbons (HFCs) worldwide. HFCs ...Missing: 2025 | Show results with:2025
  52. [52]
    [PDF] The Montreal Protocol Kigali and HFCs 2024 Transition ... - EPA
    The Montreal Protocol phases out ozone-depleting substances. The Kigali Amendment reduces HFCs, and the AIM Act enables the phasedown of HFCs in the US.Missing: phase targets
  53. [53]
    [PDF] Lower Global Warming Potential Refrigerants - AHRI
    Nov 1, 2019 · The ASHRAE 34 Standard. Committee determines toxicity and flammability classification. • Class A refrigerants have lower toxicity;. Class B ...
  54. [54]
    [PDF] A2l Refrigerants and ASHRAE Standard 15 - Trane
    2 This standard classifies each refrigerant into a “safety group” according to its toxicity (Class A or B) and its flammability. (Class 1, 2L, 2, or 3).
  55. [55]
    46 CFR Part 58 Subpart 58.20 -- Refrigeration Machinery - eCFR
    (a) All pressure vessels, compressors, piping, and direct expansion cooling coils must be leak tested after installation to their design pressures, ...
  56. [56]
    [PDF] New Climate-Friendly Motor Vehicle Air Conditioning Refrigerants
    HFO-1234yf (R-1234yf) HFO-1234yf is a refrigerant being introduced by many automobile manufacturers. There are cars on the road today using this alternative. ...
  57. [57]
    Refrigerant Safety | US EPA
    Mar 12, 2025 · As an example, refrigerant concentrations can be lowered by designing equipment with reduced leakage and promptly repairing leaks that do occur.
  58. [58]
    Mitigation of safety and environmental challenges posed by ...
    Leak Mitigation: Incorporating adsorbent materials within the refrigeration system can function as a secondary containment measure. In the event of a leak, the ...
  59. [59]
    Life-Cycle Assessment of Refrigerants for Air Conditioners ... - MDPI
    The purpose of this study was to evaluate and compare the environmental impact of two treatment methods: reclamation and destruction after refrigerant recovery.Missing: footprint | Show results with:footprint
  60. [60]
    [PDF] advanced rankine and brayton cycle power systems
    Current steam power plants operate at conversion efficiencies of nearly forty percent. By adding heat at higher temperatures, ad- vanced Rankine and Brayton ...
  61. [61]
    Thermodynamic Analysis of Supercritical CO2 Power Cycle with ...
    Jul 24, 2018 · Detailed analysis began in the late 1960s. Feher [4] proposed CO2 as the working fluid for the regenerative supercritical cycle that operated ...
  62. [62]
    [PDF] Pressurized Water Reactor (PWR) Systems
    PWR systems use primary and secondary systems to convert heat from fuel to electricity. The primary system transfers heat to steam generators, and the ...
  63. [63]
    [PDF] Selection of Optimum Working Fluid and Cycle Configuration of ...
    Feb 12, 2020 · Three working fluids are selected: isopentane, isobutane, and n-pentane. Two-cycle configurations are compared: basic ORC and recuperative ORC ...
  64. [64]
    [PDF] EXERGY STUDY OF THE KALINA CYCLE
    The Kalina cycle, which use a ammonia-water mixture, may show 10 to 20% higher exergy efficiency than the conventional Rankine cycle (Kalina, 1984 and El-Sayed ...
  65. [65]
    The Vapor Compression Refrigeration Cycle, Step By Step - ARANER
    The Vapor Compression Refrigeration Cycle involves four components: compressor, condenser, expansion valve/throttle valve and evaporator.
  66. [66]
    [PDF] Chapter 9 - Civil, Environmental and Architectural Engineering
    That is, process 1-2 is isentropic compression, 3-4 is isentropic expansion, and 4-1 is constant-volume heat rejection. The similarity between the two.
  67. [67]
    Advanced Refrigeration Technologies | US EPA
    Mar 26, 2025 · Primary refrigerants can include ammonia (NH3), R-744 (CO2), hydrocarbons, or HFCs and HFC blends such as R-407A and R-404A.
  68. [68]
    Air conditioning - Climate-friendly alternatives to F-gases
    From 1 January 2025, single split AC systems containing less than 3 kg of F-gases with a GWP of 750 or more are prohibited. From 1 January 2027, split air-to- ...
  69. [69]
    Refrigerant Overview in the RACHP Industry - RefIndustry
    Sep 8, 2024 · Common HFCs, such as R-134a, R-410A, and R-404A, became widely used in refrigeration, air conditioning, and automotive applications. However, ...
  70. [70]
    [PDF] Natural Refrigerants Fact Sheet
    Aug 16, 2024 · Natural refrigerants include water, air, ammonia, carbon dioxide, and hydrocarbons. Examples include propane (R-290) and iso-butane (R-600a).
  71. [71]
    Refrigerants - Physical Properties - The Engineering ToolBox
    Physical properties of refrigerants - molecular weight, boiling, freezing and critical points. Physical properties of some common refrigerants.
  72. [72]
    Refrigerants Grouped by Temperature Range - Super Radiator Coils
    Jul 25, 2024 · Ultra Low Temperature (ULT) Working Fluids ; Boiling temp. °F (°C). -320.4 (-195.8). -109 (78.5). -27.9 (-33.3) ; Critical temp °F (°C). -232.5 (- ...
  73. [73]
    Lithium-Bromide Absorption Cycle - an overview - ScienceDirect.com
    A lithium bromide absorption cycle refers to a thermodynamic process that utilizes a dilute lithium bromide aqueous solution as a working fluid for cooling.
  74. [74]
    Lithium Bromide Absorption Refrigeration System - LinkedIn
    Jul 1, 2023 · In a water-lithium bromide vapor absorption refrigeration system, water is used as the refrigerant while lithium bromide (Li Br) is used as the absorbent.
  75. [75]
    Transcritical CO2 Refrigeration: Basics and Benefits | The Super Blog
    Jun 23, 2020 · Around 2010, the first refrigeration systems operating on transcritical CO2 began cropping up in North American supermarkets.
  76. [76]
    [PDF] Transcritical CO2 (R744) Supermarket Refrigeration in Australia By ...
    The first transcritical CO2 supermarket refrigeration system was commissioned in Australia in December 2007 at the Drakes Foodmarkets store at Angle Vale, North.