Fact-checked by Grok 2 weeks ago

Binary cycle

A binary cycle is a type of plant designed to generate from moderate-temperature geothermal resources by using a secondary with a lower than , typically in a closed-loop system. In this process, hot geothermal , often at temperatures between 107°C and 182°C (225°F and 360°F), passes through a to transfer its to the secondary fluid—commonly an like or —which vaporizes at lower temperatures and drives a connected to a . The geothermal , remaining in liquid form, is then reinjected into the , minimizing surface emissions and environmental release of fluids or gases. This technology offers significant advantages over traditional dry steam or flash steam plants, which require higher temperatures above 182°C (360°F) and often involve flashing geothermal water to , potentially releasing non-condensable gases and minerals. Binary cycles enable the exploitation of more abundant lower-temperature resources, operate with higher for such conditions, and produce near-zero emissions of sulfur compounds or compared to plants of similar capacity. They also reduce and issues in the power equipment by keeping the geothermal fluid isolated. Developed in the mid-20th century to address limitations of earlier geothermal technologies, the first began in 1967 at Paratunka in Russia's , using a as the . Since then, have grown in prevalence, accounting for the majority of new geothermal installations worldwide due to the predominance of moderate-temperature reservoirs; as of 2024, they represent a key pathway for expanding geothermal capacity, with global installed power exceeding 16 gigawatts across all types. Today, are integral to strategies, particularly in regions like the , , and , where they support baseload power with high reliability and low operational costs.

Overview

Definition and Applications

A binary cycle is a designed for generating from geothermal sources with temperatures typically below 180°C, where the geothermal itself does not directly drive a . It involves a primary closed cycle using the geothermal brine to transfer via a to a secondary closed cycle employing a low-boiling-point , such as an , which vaporizes to power the . This dual-fluid approach prevents scaling, corrosion, and non-condensable gases from affecting the , making it suitable for moderate-temperature resources that are incompatible with traditional or methods. Geothermal energy originates from the Earth's internal heat, stored in subsurface reservoirs of hot water or steam formed by natural geological processes, accessible through drilling wells to depths of several kilometers. For binary cycles, viable resources generally range from 85°C to 150°C, allowing exploitation of lower-enthalpy fields that constitute a significant portion of global geothermal potential. Key applications include deploying wellhead generator units at individual production wells to capture energy from isolated or marginal flows, often in remote areas for off-grid power or industrial use. Binary cycles also integrate effectively with enhanced geothermal systems (EGS), where engineered reservoirs in hot dry rock are stimulated to create fluid circulation, enabling power generation from deeper, higher-temperature but low-permeability formations. As of , binary cycles represent a growing share of new geothermal installations worldwide, driven by their adaptability to diverse and lower-temperature resources, with variants comprising about 25% of total installed capacity and increasing in recent developments. This versatility supports expanded deployment in regions like the , , and emerging markets in and , where moderate-temperature fields are abundant.

Basic Operating Principles

In a binary cycle system, the primary geothermal fluid, typically hot from low- to moderate-temperature reservoirs (below 200°C), circulates through a where it transfers to a secondary with a lower , such as or . This secondary fluid absorbs the heat, undergoes without direct mixing of the fluids, and the resulting vapor expands to drive a connected to a , producing electrical power. After expansion, the vapor condenses back into liquid form through cooling and is then pumped to return to the heat exchanger, completing the closed loop of the secondary cycle. The heat transfer occurs indirectly via the , ensuring the secondary cycle remains isolated from the corrosive and scaling-prone geothermal , which allows for efficient operation without contamination or material degradation in the power generation components. This closed-loop design for the secondary fluid prevents the release of geothermal fluids into the atmosphere and facilitates reinjection of the cooled primary fluid back into the , promoting . The energy flow in a binary cycle converts low-enthalpy heat sources into mechanical work through an adapted (ORC), optimized for temperatures as low as 90°C, where the secondary fluid's of plays a key role in maximizing efficiency by enabling phase change at lower pressures. This process assumes familiarity with basic thermodynamic principles, such as heat addition and rejection in a vapor power cycle, while emphasizing the utilization of to achieve thermal-to-electrical conversion efficiencies of 9–15% in practical systems. The primary geothermal loop and secondary power cycle interact solely through this heat exchange, with detailed cycle analyses covered in thermodynamic sections.

Thermodynamic Cycles

Primary Cycle

The primary cycle in a binary geothermal power plant functions as an open loop utilizing geothermal as the heat source. Hot is extracted from production wells in the , directed through a where it transfers to the secondary , and subsequently reinjected into the subsurface without undergoing any phase change. This design keeps the geothermal fluid contained in a single pass, preventing direct contact with power generation equipment. The geothermal brine in this cycle is predominantly water-based, exhibiting temperatures typically between 85°C and 180°C to suit moderate-temperature resources. It often features high , which contributes to corrosive properties, along with dissolved minerals and potential non-condensable gases such as CO₂ and H₂S that require management to avoid operational issues. These characteristics make the brine unsuitable for direct use but ideal for indirect applications. Key process steps include initial production from the geothermal via pressurized wells to maintain the in liquid form, followed by delivery to the for controlled cooling and heat release to the secondary loop. The cooled , now at a lower temperature, is then reinjected into injection wells to sustain and promote resource over time. The provides the critical interface for this thermal handover, ensuring separation between the primary and secondary fluids. Advantages of the primary cycle include enhanced resource utilization for lower-temperature fields that cannot support , thereby broadening the applicability of . By eschewing , it minimizes silica risks, as the maintained pressure prevents rapid silica and that occur during phase changes in other systems.

Secondary Cycle

The secondary cycle in a binary power plant operates as a closed-loop akin to the , utilizing a low-boiling-point —such as hydrocarbons like or ammonia-water mixtures—that is heated by the primary geothermal fluid without direct contact. This fluid undergoes vaporization in a , expands to drive a for , condenses in a cooling , and is recirculated via a , ensuring efficient reuse and minimal environmental impact due to the sealed . The process begins with preheating and vaporization of the secondary fluid in the , where it absorbs to reach its and form vapor. This vapor then flows to the for expansion, converting into mechanical work that powers the . Post-expansion, the vapor passes through a regenerator if present to recover heat, followed by in an air- or water-cooled , after which a feed compresses the liquid back to the to complete the cycle. A defining characteristic of the secondary cycle is its ability to function at significantly lower pressures and temperatures than traditional steam-based cycles, enabling effective exploitation of low- to moderate-temperature geothermal resources in the range of 90–175°C that would otherwise be unsuitable for direct generation. This adaptability stems from the secondary fluid's lower , which matches the source profile and enhances overall thermodynamic efficiency in (ORC) or variants. In terms of , the secondary cycle remains isolated from the primary cycle through the intermediary to prevent fluid mixing and , though it can serve as a bottoming cycle in setups to capture residual ; turbine sharing occurs only in specific combined configurations, but separation is standard to maintain fluid integrity.

Historical Development

Early Innovations

The binary cycle concept for generation traces its roots to early 20th-century efforts in low-temperature heat engines, where engineers sought to utilize secondary working fluids to harness geothermal heat more efficiently than direct systems. The first geothermal power plant at Larderello in 1904, developed by Piero Ginori Conti, used direct to generate , but early experiments explored indirect methods to mitigate from geothermal vapors. By the , experimental developments expanded on these ideas, with initial concepts exploring organic fluids for closed cycles to improve in moderate-temperature resources, adapting principles from reciprocating engines to geothermal applications. A key milestone occurred in the early 1940s on the island of , , where Luigi D’Amelio of the University of led the construction of the world's first experimental binary cycle geothermal plant. Operational from 1940 to 1943, this prototype generated 11 kW of mechanical power using monochloroethane as the secondary , heated by geothermal at around 150°C. engineers, including D’Amelio, drew on designs to create compact expansion units suitable for low-enthalpy sources, marking a shift toward variants for geothermal exploitation. A follow-up 250 kW plant was built in 1943 but never entered full operation due to wartime disruptions and technical hurdles. Early prototypes faced significant challenges, particularly with fluid stability and efficiency. Organic working fluids like monochloroethane exhibited risks at elevated temperatures, limiting operational reliability and requiring careful to prevent in heat exchangers. inefficiencies arose from and in prototypes, compounded by the need to scale from models (e.g., a prior 2.6 kW unit) to field applications, which strained and economic feasibility. These issues were partially addressed through improved exchanger configurations, but they highlighted the need for more stable fluids and robust components in subsequent iterations.

Commercial Expansion

The first commercial binary cycle geothermal power plant began operation in 1967 at Paratunka in Russia's , with a capacity of approximately 0.7 MW, using R-12 as the secondary to harness geothermal water at 81°C. This facility served a local village and greenhouses, marking the initial commercial application of binary technology for low-temperature resources. The first commercial binary cycle plant in the United States was commissioned in 1979 at East Mesa in California's Imperial Valley, boasting a capacity of 10 MW and utilizing a secondary organic fluid to harness lower-temperature geothermal resources previously uneconomical for flash steam systems. This facility represented a pivotal shift, enabling broader commercialization of geothermal energy by addressing limitations in resource temperature and reducing issues like corrosion and scaling associated with direct steam extraction. The expansion of binary cycle technology gained momentum through supportive policies, notably the U.S. Geothermal Energy Research, Development, and Demonstration Act of 1974, which allocated federal funding for research into low-temperature conversion systems and spurred early prototyping efforts. This legislative framework, enacted amid the that quadrupled global oil prices and prompted a search for domestic alternatives, catalyzed investments in renewables including . Subsequent incentives, such as tax credits and loan guarantees in the U.S. and similar measures abroad, further accelerated adoption by offsetting high upfront costs for binary installations. Early commercialization focused on regions with accessible moderate-temperature fields, including the U.S. where multiple plants followed East Mesa, New Zealand's Taupo Volcanic Zone with binary units integrated into existing fields like Wairakei, and Iceland's Reykjanes Peninsula where binary elements supplemented high-enthalpy resources. By the early , these areas exemplified policy-driven growth, with the U.S. leading through federal R&D that de-risked deployment in liquid-dominated reservoirs. Technological advancements post-1980s were crucial enablers, particularly refinements in organic working fluids such as mixed hydrocarbons (e.g., isobutane-isopentane blends) that improved thermodynamic matching and for temperatures between 100–200°C. Concurrent innovations in heat exchangers, including direct-contact designs, fluidized-bed systems, and polymer-coated surfaces to mitigate scaling, reduced operational downtime and boosted net output by up to 5% in early installations. These developments, funded largely by U.S. Department of Energy programs, lowered levelized costs and facilitated scaling. By 2014, binary cycle plants had proliferated globally to 203 facilities across 15 countries, accounting for 35% of all stations and generating approximately 2,339 MW, driven by the cumulative effects of needs from the crises and maturing renewable support frameworks.

Design Variations

Dual Pressure Configurations

In dual pressure configurations of binary cycle geothermal power plants, the secondary cycle is divided into high-pressure () and low-pressure () stages to achieve better matching of the 's temperature profile with the geothermal heat source, thereby enhancing overall heat recovery efficiency. This approach addresses the limitations of single-pressure cycles by allowing the , typically an like , to evaporate and expand in staged processes that more closely follow the of the geothermal . Such designs are particularly suited for mid-enthalpy resources where the content of the geothermal fluid can be better utilized without requiring fluid changes. Operationally, vapor generated in the is split or directed through staged expansion, with the stage handling initial high-temperature evaporation and the stage capturing residual from the cooling geothermal fluid. This staged process minimizes temperature mismatches in the , reducing losses while the expanded vapor from the may be reheated or mixed before entering the for further power generation. The configuration often incorporates recuperators to preheat the using exhaust, ensuring closed-loop circulation without direct contact between the geothermal and the secondary fluid. The primary benefits include a 10–20% increase in net output compared to single-pressure systems, achieved through improved utilization of available and lower irreversibilities in . Additionally, dual pressure setups reduce pinch point losses in the by maintaining smaller temperature differences throughout the process, leading to higher efficiencies often exceeding 30% in optimized cases. These enhancements make the configuration economically viable by lowering the unit cost of electricity production, with reported reductions of up to 17% in some analyses. Examples of dual pressure binary cycles are prominently applied in mid-enthalpy geothermal fields operating at 120–150°C, such as the Raft River plant in , , where a subcritical, dual-pressure generates approximately 11–13 MW while optimizing recovery from fluids at around 140°C. Similar implementations have been studied for fields like Velika Ciglena in , demonstrating net power gains and efficiencies up to 34% for comparable temperature ranges. These applications highlight the configuration's role in maximizing output from moderate-temperature resources without the need for advanced fluid mixtures.

Dual Fluid Systems

Dual fluid systems in binary cycle geothermal power plants employ two distinct working fluids within the secondary cycle to enhance efficiency across a wider range. A high-boiling-point fluid is utilized at the hot end of the to capture heat from the higher-temperature portion of the geothermal , while a low-boiling-point fluid handles the colder end, achieving a closer match to the temperature glide of the geothermal source. This configuration minimizes differences during heat exchange, reducing losses and improving overall thermodynamic performance. In operation, the two fluids can be arranged in series or flows through the , allowing sequential or simultaneous tailored to the geothermal fluid's temperature profile. The vaporized fluids then drive , either through separate for each fluid or a shared multi-stage turbine setup, before and recirculation. This staged approach enables better utilization of the available heat, particularly in resources with variable or moderate temperatures. The primary benefits include higher cycle efficiency, with improvements of up to 17% compared to single-fluid basic Rankine cycles, due to optimized thermal matching in variable temperature sources. Additionally, the reduced temperature gradients lower on heat exchanger components, potentially extending equipment lifespan. However, these systems introduce drawbacks such as increased complexity from managing multiple fluids, requiring separate handling, recovery, and containment systems, which elevate capital and operational costs.

Key Components

Turbine and Expansion

In binary cycle geothermal power plants, the facilitates the isentropic of the vaporized secondary , converting its into mechanical shaft work that drives an electrical . This occurs in a closed-loop (ORC), where the secondary fluid—typically an organic compound like or —enters the at elevated and after in the . The drops the fluid to a lower , producing work output while maintaining near-ideal thermodynamic conditions to maximize energy extraction. Turbine design in binary cycles is tailored to the properties of the secondary fluid, particularly its higher molecular weight compared to , which influences flow characteristics and requires compact geometries for efficient operation at low pressures and temperatures. Radial inflow or outflow are commonly employed for smaller (under 5 MW), offering simplicity and suitability for the dense, high-molecular-weight vapors, while axial are preferred for larger installations to handle higher flow rates with multi-stage configurations. Isentropic efficiencies typically range from 80% to 90%, reflecting the ratio of actual to ideal drop during expansion, as specified by manufacturers to optimize performance under varying conditions. Performance is significantly influenced by inlet conditions, with higher temperatures and pressures increasing output and ; for instance, elevating the inlet temperature from 90°C to 150°C can boost net by about 8-10% in single-stage setups. To mitigate risks from wet vapors during expansion, is often applied to ensure the enters as dry or superheated vapor, reducing content below 15% at the exit and protecting blades from liquid droplet impingement. In hybrid configurations, the binary cycle may integrate with cycles as a bottoming unit, utilizing residual heat to enhance overall plant output without dedicated hardware.

Condenser and Cooling

In binary cycle geothermal power plants, the condenser serves as the primary component for heat rejection, where the low-pressure vapor exiting the is condensed back into a state, typically using shell-and-tube or plate to transfer to a cooling medium. This process rejects approximately 70–80% of the input to the ambient environment, enabling the secondary to be recirculated efficiently in the closed loop. Achieving in the —often on the order of 0.1–1°C below the temperature—helps prevent and enhances the for subsequent pumping, thereby improving overall cycle reliability. Cooling methods for the condenser vary based on site conditions, with air-cooled systems (dry cooling towers) preferred in water-scarce regions to minimize freshwater use, while water-cooled evaporative towers are employed where water availability supports higher efficiency. Air-cooled condensers rely on from large finned-tube arrays, but they result in higher at the turbine exhaust—typically 40–50°C condensing temperatures—reducing power output by up to 10–20% compared to water-cooled alternatives under similar ambient conditions. In contrast, water-cooled systems maintain lower condensing temperatures (around 30–40°C) through evaporative cooling, lowering and boosting cycle efficiency, though they require careful design to handle and . Performance of the is critically influenced by the approach , defined as the between the cooling medium outlet and the working fluid's saturation , typically maintained at 5–10°C to optimize without excessive surface area. A smaller approach enhances by allowing closer operation to the ambient sink , but it increases condenser size and cost; deviations can reduce net plant output by 1–2% per °C. Additionally, non-condensable gases, primarily air ingress in the closed secondary loop, accumulate in the and impair , necessitating periodic removal through vents or purging systems to sustain conditions (around 0.05–0.1 ). Condenser sizing is determined by the total heat rejection load, which scales with turbine exhaust conditions and ambient temperatures, often requiring surface areas of 500–1000 m² per MW of gross power to accommodate the non-isothermal condensation of organic fluids. Shell-and-tube designs dominate due to their robustness against pressure differentials and ease of maintenance, featuring horizontal double-pass configurations for counterflow heat exchange, while plate condensers offer compactness with heat transfer coefficients up to 10–20 kW/m²K for space-constrained installations. Proper sizing ensures minimal pressure drop (under 0.01 bar) to avoid efficiency losses, with modular air-cooled units allowing scalability for varying plant capacities.

Feed Pump and Circulation

In binary cycle geothermal power plants, the feed pump serves to pressurize the condensed secondary working fluid, typically an organic liquid such as or , after it exits the and before it enters the for reheating. Centrifugal pumps are commonly employed for this purpose, operating under steady-state adiabatic conditions to increase the fluid pressure from near-vacuum levels at the outlet to the required high pressure at the inlet, ensuring continuous circulation through the closed (ORC). These pumps consume approximately 2-3% of the gross power in higher-temperature ORC configurations, representing a minor but essential parasitic load that impacts net plant efficiency. Design considerations for these feed pumps emphasize reliability and in handling the low-density organic fluids. Multi-stage centrifugal configurations are often utilized to achieve the necessary pressure heads, particularly in systems with significant elevation differences or pressure drops, while maintaining compact footprints suitable for modular plant layouts. To prevent , which can degrade performance and cause mechanical damage, pumps are engineered with adequate (NPSH) margins; this may involve pre-feed booster pumps to elevate inlet pressure, strategic pump placement below the for , or the to reduce . Efficiencies typically range from 70-85% for well-designed centrifugal units, optimized for the specific fluid properties and operating pressures. Circulation in the secondary loop relies on precise control of the feed pump to match fluid flow rates with the available heat input from the primary geothermal brine, ensuring optimal heat transfer without overloading the system. Flow rates are scaled to geothermal fluid throughput, for example, around 40 L/s per production well in mid-enthalpy fields, adjustable via variable speed drives (VSDs) that enable load-following capabilities in response to reservoir variations or demand fluctuations. This dynamic control minimizes energy waste and supports stable operation across partial loads, with VSDs integrated into inverter-driven centrifugal pumps for proportional flow adjustment. On the primary side, reinjection pumps handle the cooled geothermal after heat extraction, returning it to the to sustain pressure and prevent . These pumps are robustly designed to manage , dissolved solids, and corrosive elements in the brine, often using multistage axial-split configurations capable of flows up to 3,200 m³/h and heads to 2,900 m. They operate at elevated temperatures up to 200°C (or higher in specialized models reaching 425°C), incorporating wear-resistant materials like or hard coatings to withstand from silica and abrasives prevalent in geothermal fluids.

Heat Exchanger Design

In binary cycle geothermal power plants, the primary serves as the critical interface for transferring from the hot geothermal to the lower-boiling secondary , enabling efficient power generation without direct contact between the fluids. Common designs include shell-and-tube and plate s, with shell-and-tube configurations often preferred in binary applications due to their robustness against high pressures and temperatures associated with geothermal brines. To enhance coefficients, these exchangers may incorporate surface modifications such as fins or twisted tubes, which promote and increase the effective contact area between fluids. Plate designs, while more compact, are selected for smaller-scale or lower-temperature operations where space constraints apply. Operationally, these heat exchangers typically employ a counterflow to minimize the temperature difference across the unit, thereby maximizing thermodynamic efficiency and reducing losses. Geothermal brines, laden with minerals like silica and salts, promote on the exchanger surfaces, which can degrade performance over time; designs incorporate features such as accessible cleaning ports or removable tube bundles to facilitate periodic mechanical or chemical cleaning. Performance evaluation focuses on the overall (U), which ranges from approximately 500 to 2000 W/m²K depending on fluid properties, flow regimes, and surface enhancements, with higher values achieved in due to their thin boundaries. Sizing of the relies on pinch point analysis, which identifies the minimum temperature approach between the geothermal and secondary fluid (often 5–10°C) to ensure feasible while optimizing surface area requirements. Materials selection emphasizes corrosion resistance on the geothermal side, where are widely used for their immunity to pitting, , and stress cracking induced by acidic or saline brines at temperatures up to 200°C. The secondary fluid side may employ stainless steels or other alloys compatible with organic working fluids like or .

Efficiency and Performance Metrics

First and Second Law Efficiencies

The efficiency of a binary cycle, also known as , quantifies the conversion of input into net mechanical work based on the . It is calculated as \eta_I = \frac{W_{net}}{Q_{in}}, where W_{net} represents the net work output, primarily the work minus the work, and Q_{in} is the absorbed from the geothermal source in the . In practical binary plants, \eta_I typically ranges from 10% to 13%, reflecting losses due to rejection in the and parasitic loads like pumping. This value is derived from an energy balance across the : for steady-state operation, Q_{in} = W_t + W_p + Q_{out}, where W_t is work, W_p is work, and Q_{out} is rejected, leading directly to \eta_I = 1 - \frac{Q_{out}}{Q_{in}}. Irreversibilities in components, such as drops and , reduce the actual W_{net} compared to ideal assumptions. The second law efficiency, or exergy efficiency, provides a more insightful measure by accounting for the quality of , using the concept of —the maximum useful work obtainable from a relative to the environment. It is defined as \eta_{II} = \frac{W_{net}}{Q_{in} (1 - \frac{T_0}{T_{in}})}, where T_0 is the ambient and T_{in} is the geothermal , representing the of actual work to the reversible work potential. In binary cycles, \eta_{II} often achieves 30–50% of the theoretical maximum, highlighting opportunities for improvement beyond mere . This efficiency is derived from exergy balances, incorporating generation: exergy destruction I = T_0 \Delta S_{gen} quantifies irreversibilities, with \eta_{II} approaching 1 only in reversible processes. Component-level analysis reveals that increases due to finite differences and mixing, particularly in the and . Key factors influencing both efficiencies include the geothermal fluid inlet temperature, which elevates Q_{in} and the potential for higher \eta_I and \eta_{II}, and ambient conditions, where elevated T_0 diminishes the and thus the available work. Improvements focus on minimizing destruction, with studies indicating that a significant portion of total losses, often over 50%, occur in the due to thermal irreversibilities; optimizing designs like counterflow configurations or enhanced surfaces can reduce these by 20–30%. Such enhancements elevate \eta_{II} toward 50% without altering the fundamental cycle . The second law serves as a practical relative to ideal reversible cycles, emphasizing recovery over raw energy throughput.

Carnot Efficiency and Limitations

The Carnot efficiency represents the theoretical maximum for any operating between a hot source temperature T_\text{hot} and a cold sink temperature T_\text{cold}, both in , serving as an upper bound for binary cycle systems in generation. It is given by the formula \eta_C = 1 - \frac{T_\text{cold}}{T_\text{hot}}. This efficiency arises from the ideal reversible , which assumes no losses and perfect . For a typical binary cycle with a geothermal source at 180°C (453 ) and an ambient sink at 30°C (303 ), the Carnot efficiency reaches approximately %, highlighting the potential but rarely achieved limit for low-to-medium temperature resources. The derivation of the Carnot efficiency begins with the first and second laws of thermodynamics applied to a reversible heat engine. Consider a cycle with heat input Q_\text{hot} at constant temperature T_\text{hot} during isothermal expansion and heat rejection Q_\text{cold} at constant temperature T_\text{cold} during isothermal compression, connected by adiabatic processes. From the first law, the net work output is W = Q_\text{hot} - |Q_\text{cold}|. The second law requires zero net entropy change for reversibility: \Delta S = \frac{Q_\text{hot}}{T_\text{hot}} + \frac{Q_\text{cold}}{T_\text{cold}} = 0, so \frac{|Q_\text{cold}|}{Q_\text{hot}} = \frac{T_\text{cold}}{T_\text{hot}}. Thus, the efficiency is \eta_C = \frac{W}{Q_\text{hot}} = 1 - \frac{|Q_\text{cold}|}{Q_\text{hot}} = 1 - \frac{T_\text{cold}}{T_\text{hot}}, emphasizing the entropy constraint that bounds all real engines. In binary cycles, actual efficiencies are significantly lower than this ideal, typically 10–13% for plants operating between 85°C and 180°C, due to inherent limitations that introduce irreversibilities. Low source s inherently cap the Carnot efficiency at modest levels (often below 20–30%), while variations in ambient sink further reduce it seasonally. cycles typically achieve 40–60% of the Carnot limit, with advanced designs reaching up to 85% relative to more realistic ideal models like the triangular cycle, primarily because of non-isothermal heat addition in the —unlike the in the Carnot cycle—and finite differences across heat exchangers, which generate through irreversible . Additional losses stem from in turbines and pumps, as well as drops in piping, all deviating from the reversible adiabatic assumptions. The Carnot model thus overestimates performance for systems, where the triangular cycle (accounting for linear decline in the geothermal fluid) provides a more realistic upper bound of about 60–70% of Carnot.

Working Fluids

Selection Criteria

The selection of working fluids for binary cycles, particularly in geothermal applications, hinges on a balance of thermodynamic performance, environmental compatibility, safety, and practical feasibility to ensure efficient from the primary to the power generation cycle. Key thermodynamic criteria include matching the fluid's critical and to the heat source , typically requiring the critical to exceed the source for effective without exceeding the critical point. Additionally, fluids with high of are preferred, as they enable greater heat absorption per unit mass, leading to higher work output and more compact equipment designs. For zeotropic mixtures, the temperature glide during phase change is a critical factor, as it allows better thermal matching with the variable-temperature heat source, reducing losses and maximizing cycle efficiency compared to pure fluids with isothermal . Environmental considerations prioritize fluids with negligible (ODP ≈ 0) and low (typically GWP < 150), excluding chlorofluorocarbons (CFCs) and hydrochlorofluorocarbons (HCFCs) to comply with international regulations. Safety factors emphasize low toxicity and non-flammability to minimize risks in plant operations, particularly for fluids handling high pressures and temperatures. Practical aspects involve ensuring fluid availability and cost-effectiveness, as well as chemical compatibility with system materials like heat exchangers and turbines to prevent corrosion or degradation. Thermal and chemical stability under operating conditions is essential to avoid decomposition, which could reduce performance or require frequent maintenance. Evaluation of candidate fluids typically employs thermodynamic screening tools such as the NIST REFPROP software, which provides accurate property data including equations of state for pure fluids and mixtures. Complementary methods include T-Q (temperature-heat) diagrams, which visualize the temperature profile of the working fluid against heat transfer to identify optimal matches and minimize temperature differences in heat exchangers. These criteria collectively guide the choice toward fluids like hydrocarbons or refrigerants that align with binary cycle requirements.

Common Fluids and Properties

In binary cycle systems, particularly those employing organic Rankine cycles (ORC), isobutane serves as a widely used organic working fluid due to its favorable thermodynamic properties for low-to-medium temperature geothermal sources. With a critical temperature of approximately 135°C and a critical pressure of 36.4 bar, isobutane enables efficient vaporization and expansion in heat sources ranging from 100–150°C, achieving thermal efficiencies around 10-12% in typical configurations. However, its flammability poses safety challenges in plant design and operation. Pentane, including n-pentane and isopentane variants, is another common organic fluid in ORC-based binary cycles, valued for its enhanced thermal stability at higher temperatures. N-pentane has a critical temperature of 196.7°C and a normal boiling point of 36.1°C, making it suitable for geothermal resources up to 180–200°C, where it supports greater heat recovery and cycle efficiencies compared to lower-critical-temperature fluids. Its non-polar nature contributes to good compatibility with turbine materials, though it shares flammability risks with other hydrocarbons. For Kalina cycle implementations in binary systems, ammonia-water mixtures are the standard working fluid, leveraging their non-azeotropic properties for improved matching of temperature profiles during heat exchange. These mixtures exhibit a variable boiling point that spans approximately 0–100°C depending on ammonia concentration (typically 50–70 wt%), enabling closer approach to the heat source temperature curve and up to 20% higher exergy efficiency than single-component ORC fluids in low-temperature applications. The zeotropic behavior enhances heat recovery by reducing irreversibilities, though corrosion from ammonia requires specialized materials. Other fluids employed in binary cycles include R134a, a hydrofluorocarbon refrigerant with a boiling point of -26.1°C, which is effective for very low-temperature sources below 100°C due to its high latent heat and non-flammable nature, yielding system efficiencies around 7–8% in subcritical ORC setups. Supercritical CO2 is gaining traction in advanced binary designs, with its low critical temperature of 31.1°C and pressure of 73.8 bar allowing compact, high-efficiency cycles (up to 15% thermal efficiency) that are environmentally benign and non-toxic, albeit demanding robust high-pressure components. R134a offers operational simplicity but carries a high global warming potential (GWP) of 1430, while CO2 provides near-zero GWP at the cost of elevated operating pressures. Post-2020 regulatory shifts, including the U.S. EPA's hydrofluorocarbon phasedown under the American Innovation and Manufacturing Act, have accelerated the adoption of low-GWP alternatives such as hydrofluoroolefins (HFOs) in ORC binary cycles. Fluids like R1234yf (GWP <1) and HFO-R245fa mixtures demonstrate comparable performance to traditional refrigerants in low-temperature operations, with boiling points around -29°C and improved environmental profiles, supporting net-zero emission goals in geothermal power generation. Emerging research as of 2025 explores water as a working fluid in binary cycles for higher-temperature resources above 300°C, particularly in superhot rock geothermal systems, offering thermal efficiencies of 22-27% and reduced reliance on organic fluids.

Environmental and Economic Aspects

Environmental Impacts

Binary cycle geothermal power plants exhibit minimal direct emissions of greenhouse gases, primarily due to their closed-loop secondary fluid circulation, which prevents the release of carbon dioxide (CO₂) and methane (CH₄) associated with the geothermal brine. Trace amounts of hydrogen sulfide (H₂S) may originate from the primary geothermal fluid, but these are effectively managed through abatement technologies such as liquid redox processes and reinjection, achieving removal efficiencies of 90% to 99.9%. Water consumption in binary cycle operations is relatively low compared to other thermal power technologies, with air-cooled variants using approximately 1 L/kWh (0.27 gallons per kWh) and wet-cooled (evaporative) systems ranging from 5 to 18 L/kWh. Reinjection of geothermal fluids back into the reservoir not only conserves water but also minimizes surface subsidence by maintaining reservoir pressure. These plants require a compact land footprint of 1 to 8 acres per megawatt (MW), enabling potential coexistence with agriculture or other land uses and limiting impacts on biodiversity. However, exploratory and production drilling can disrupt local aquifers through fluid extraction or injection, potentially altering groundwater flow. In enhanced geothermal systems (EGS) integrated with binary cycles, there is a risk of induced seismicity from reservoir stimulation, though events are generally microseismic and manageable through monitoring and controlled injection practices. Life cycle assessments indicate that approximately 98% of the environmental impact from binary cycle plants stems from exergy destruction during heat transfer and conversion processes. Cogeneration applications, such as combined heat and power from binary systems, can substantially reduce this overall footprint by improving resource utilization efficiency.

Economic Considerations

Binary cycle geothermal power plants typically incur higher capital costs compared to flash steam plants, ranging from $2,500 to $5,500 per kilowatt of installed capacity (as of 2017), primarily due to the need for extensive heat exchanger systems to transfer heat from the geothermal fluid to the secondary working fluid without direct contact. Recent estimates as of 2024 indicate averages around $4,350 per kW. Operation and maintenance (O&M) costs are relatively low at approximately $0.01 to $0.02 per kilowatt-hour, benefiting from the technology's high reliability and minimal fuel requirements. The levelized cost of energy (LCOE) for binary cycle plants generally falls between 5 and 10 cents per kilowatt-hour, making it competitive with variable renewables like solar and wind in regions with suitable low- to medium-temperature geothermal resources. Key economic factors influencing viability include resource exploration risks, which can increase upfront uncertainties and costs due to the need for extensive drilling and testing to confirm reservoir productivity. Government incentives, such as the U.S. Production Tax Credit (PTC) providing about $0.0275 per kilowatt-hour for qualified geothermal electricity, help mitigate these risks and improve project economics. Under the 2022 Inflation Reduction Act, the PTC can be multiplied up to five times the base rate for projects meeting prevailing wage, apprenticeship, and domestic content requirements. With a high capacity factor of around 90%, payback periods typically range from 7 to 10 years, assuming stable resource conditions and effective financing. The global market for geothermal power, including binary cycle systems, is valued at $6.95 billion in 2025 and is projected to grow to $10.78 billion by 2034, at a compound annual growth rate (CAGR) of 5-8%, driven by increasing demand for baseload renewable energy and technological advancements in low-temperature resource utilization.

Modern Applications and Future Prospects

Notable Power Plants

Binary cycle power plants represent a significant portion of global geothermal electricity generation, accounting for approximately 20–30% of the total installed capacity of around 16 GW as of 2025. This equates to roughly 3.2–4.8 GW of binary cycle capacity worldwide, distributed across an estimated 250 plants, with the majority being smaller modular units. Leading countries include the United States with about 2.5 GW of binary capacity, followed by Indonesia and Turkey, though the latter two rely more heavily on flash steam systems overall. One of the earliest commercial binary cycle installations is the East Mesa geothermal complex in California's Imperial Valley, USA, which began operations in 1979 with a pioneering 10 MW Magmamax unit and expanded to a total capacity of approximately 45 MW through additional Ormat binary modules by the 1980s. This plant demonstrated the viability of binary cycles for moderate-temperature resources around 165–177°C, achieving reliable operation in a challenging brine environment. In New Zealand, the Ngatamariki power plant, commissioned in 2013, stands as one of the largest dedicated binary facilities globally, with an installed capacity of 100 MW from four Ormat Energy Converter units producing 82 MW net. Designed for a low-enthalpy field with fluids at about 250–300°C, it utilizes air-cooled condensers to minimize water use and has operated with high uptime since startup. For lower-temperature demonstrations, the Chena Hot Springs plant in Alaska, USA, operational since 2006, generates 400 kW using binary cycle technology with geothermal fluids as low as 74°C, making it a benchmark for micro-scale applications in remote, low-resource settings. This air-cooled organic Rankine cycle system powers the resort's facilities and highlights binary cycles' adaptability to temperatures below 100°C. The Salton Sea geothermal field in California, USA, features multiple binary cycle units integrated into its 400 MW total capacity across ten plants, including dual-pressure configurations that enhance efficiency from hypersaline brines exceeding 200°C. These plants exhibit high operational reliability, with average availability rates of 95%, contributing to consistent baseload power in the region.

Organic Rankine and Kalina Cycles

The Organic Rankine Cycle (ORC) functions as a secondary cycle in binary geothermal power systems, employing pure organic working fluids like isobutane or n-pentane that vaporize at low temperatures to facilitate efficient, non-contact heat exchange with the geothermal brine. These systems are noted for their straightforward design, which minimizes complexity and enhances reliability, with implementations often featuring modular units developed by leading manufacturers such as . ORC-based binary plants achieve thermal efficiencies typically ranging from 10% to 12%, making them well-suited for exploiting low- to medium-temperature resources under 150°C, and they constitute the majority of operational binary geothermal facilities worldwide. In contrast, the Kalina cycle utilizes an ammonia-water mixture as its working fluid in binary secondary cycles, enabling a variable boiling temperature that aligns closely with the declining temperature profile of the geothermal heat source for superior heat recovery. This zeotropic mixture exhibits a temperature glide during phase change, where the bubble point temperature (initiation of vaporization) and dew point temperature (completion of vaporization) differ, allowing the fluid to evaporate and condense over a range rather than at a fixed temperature; these points are calculated using correlations such as those by Patek and Klomfar (1995), given by: T_b = 273.15 + \frac{1}{\frac{1}{T_{sat,NH_3}} - A \ln\left(\frac{P}{P_{crit,NH_3}}\right) + B \left(\frac{P}{P_{crit,NH_3}}\right)^{0.5}} and similar forms for the dew point T_d, parameterized by the ammonia mass fraction and pressure, derived from experimental vapor-liquid equilibrium data. Kalina cycles demonstrate 15–20% higher power output than equivalent ORC systems under comparable low-temperature conditions (75–150°C), owing to reduced irreversibilities in heat transfer. Pilot installations include the 2 MW facility in Húsavík, Iceland, operational since 2000, and a 1 MW demonstration in New Mexico, United States. When comparing the two cycles in binary applications, ORC systems provide advantages in maintenance ease due to their simpler piping and component layouts using non-reactive organic fluids, whereas Kalina cycles offer better performance for variable-temperature heat recovery but require corrosion-resistant materials to mitigate the aggressive effects of the ammonia-water mixture on standard alloys. The ammonia component promotes general corrosion in alkaline environments (pH >9), necessitating specialized alloys like with inhibitors or stainless steels in key components such as heat exchangers and piping. ORC adoption dominates in regions with abundant low-temperature resources, such as the (where Ormat operates over 1 GW of capacity) and (with multiple sites under 150°C leveraging ORC for ), while Kalina remains confined to niche, high-efficiency sites prioritizing thermodynamic gains over operational simplicity.

Recent Developments and Innovations

In recent years, binary cycle geothermal systems have seen significant capacity expansions, particularly in regions with established geothermal resources. added 225 MW of geothermal capacity in 2024 through two new installations, including a binary-cycle unit, marking one of the largest single-year increases globally for this technology. By the end of 2024, global geothermal installed capacity reached 15.4 GW, with binary cycles playing a pivotal role in recent additions due to their suitability for lower-temperature resources, representing a substantial growth from approximately 1.25 GW of binary capacity reported in 2014. Innovations in binary cycle integration have focused on enhancing resource accessibility and sustainability. Enhanced Geothermal Systems (EGS) have been increasingly paired with binary cycles to tap deeper, low-permeability reservoirs, enabling efficient heat extraction through closed-loop secondary fluids and hydraulic stimulation. Additionally, 2025 studies have explored hybrid configurations combining binary geothermal plants with (DAC) units, leveraging excess heat from the cycle to drive processes, potentially enabling carbon-negative power generation. Efficiency improvements have been driven by advancements in turbines and working fluids, with optimized designs achieving up to 15% thermal efficiency in binary systems operating at moderate temperatures. A 2024 analysis highlighted the development of climate-resilient binary cycle designs, incorporating adaptive cooling and fluid selections like isobutane to mitigate performance declines from rising ambient temperatures due to global warming. Looking ahead, projections indicate strong growth potential for binary and next-generation geothermal technologies, emphasizing modular units for scalable deployment. The International Energy Agency (IEA) forecasts that supportive policies could drive cumulative investments in geothermal to $1 trillion by 2035, facilitating widespread adoption of advanced systems. The U.S. Department of Energy () projects that enhanced geothermal, including modular configurations, could reach 90 of capacity by 2050, a twentyfold increase from current levels.

References

  1. [1]
    Geothermal power plants - U.S. Energy Information ... - EIA
    Dec 21, 2022 · Binary-cycle power plants transfer the heat from geothermal hot water to another liquid. The heat causes the second liquid to turn to steam, and ...
  2. [2]
    Geothermal Electricity Production Basics - NREL
    Aug 27, 2025 · Binary cycle power plants operate on water at lower temperatures of about 225–360°F (107–182°C). Binary cycle plants use the heat from the hot ...Missing: definition | Show results with:definition
  3. [3]
    Geothermal Energy Factsheet | Center for Sustainable Systems
    Binary cycle plants use separate closed-loop systems for geothermal water and working fluid. Heat exchangers transfer heat to the working fluid, causing it to ...Missing: definition | Show results with:definition
  4. [4]
    Types of Geothermal Power Plants - California Energy Commission
    Geothermal power plants include direct dry steam, flash and double flash cycle, and binary cycle plants. All use steam to turn turbines.
  5. [5]
    Binary Power Plants - Stanford University
    Dec 3, 2011 · Environmentally, binary plants possess key advantages in that they do not release geothermal fluids into the environment. Earth's gases do not ...
  6. [6]
    Electricity Generation | Department of Energy
    Binary-cycle geothermal power plants can use lower temperature geothermal resources, making them an important technology for deploying geothermal electricity ...
  7. [7]
    [PDF] 100 Years of Geothermal Power Production
    Feb 2, 2005 · The power plant began operation in 1967 (Figure 14), and was the first to use an organic binary fluid in the power cycle, R-12 refrigerant, as ...
  8. [8]
    [PDF] Geothermal Binary Power Plants - ESMAP
    The binary technology allows for the production of electricity from resources that could otherwise not be used for such a purpose, typically at reservoir ...
  9. [9]
    Geothermal Basics | Department of Energy
    Geothermal energy is heat from the earth, from reservoirs of hot water at varying temperatures and depths below the earth's surface.
  10. [10]
    Geothermal wellhead technology power plants in grid electricity ...
    Binary cycles use two fluids in a closed loop cycle, one being the geothermal resource fluid and the other an organic fluid [11]. The geothermal fluid is passed ...Missing: EGS | Show results with:EGS
  11. [11]
    Geothermal energy & EGS - Exergy ORC
    Apr 30, 2025 · Binary cycle systems: a key technology for EGS and AGS. EGS and AGS subsurface technologies can be efficiently utilized through binary cycle ...
  12. [12]
    Worldwide geothermal power (2024) - Research Communities
    Jul 20, 2025 · ... binary ORC type units with 25.1% of the installed capacity. The select list of geothermal power countries continues to be headed by the US ...Missing: percentage | Show results with:percentage
  13. [13]
    Geothermal energy - IRENA
    The total installed capacity of geothermal energy reached 15.4 GW globally by the end of 2024, representing a modest increase from around 13.0 GW at the end of ...
  14. [14]
    [PDF] Binary geothermal power plants - Geoelec
    Main features: • Power generation by means of closed thermodynamic cycle. • Geothermal fluid loop and power cycle are completely separated.
  15. [15]
    Binary Cycle - an overview | ScienceDirect Topics
    The binary cycle system is used for geothermal resources at lower temperatures (≤ c. 150°C). The binary cycle is consisted of two cycles, as its name defined.Missing: EGS | Show results with:EGS<|control11|><|separator|>
  16. [16]
    [PDF] Siliceous scaling aspects of geothermal power generation using ...
    Siliceous scaling potential of geothermal brines has long been an obstacle to the efficient energy recovery from some geothermal resources.
  17. [17]
    [PDF] GEOTHERMAL POWER PLANT CYCLES AND MAIN COMPONENTS
    Jan 22, 2011 · A binary cycle uses a secondary working fluid in a closed power generation cycle. A heat exchanger is used to transfer heat from the geothermal ...
  18. [18]
  19. [19]
    [PDF] New Concepts Organic Rankine Cycle Power Systems - POLITesi
    Sep 29, 2014 · island of Ischia in 1940. A second power plant of 250 kWM based on ... binary cycle configura- tion, whereby direct thermal storage is ...
  20. [20]
    [PDF] A History or Geothermal Energy Research and Development in the ...
    This document is a history of the US Department of Energy's geothermal research program from 1976-2006, dedicated to the program's employees.<|control11|><|separator|>
  21. [21]
    GEOTHERMAL SYSTEMS - Thermopedia
    The first binary cycle plant was commissioned in 1979 at East Mesa (Imperial Valley, California), and was rated at 10 MW. A number of other binary cycle plants ...
  22. [22]
    Looking Back on the 1973 Oil Crisis, New Perspectives on Energy ...
    Dec 5, 2023 · The crisis also inspired new science and technology policies and several innovations, including alternative nuclear, solar, wind, and geothermal energy sources.
  23. [23]
    [PDF] Renewable power generation costs in 2023 - IRENA
    Renewable power is increasingly cost-competitive with fossil fuels – 81% of renewable capacity additions in 2023 produce cheaper electricity than fossil fuel ...
  24. [24]
  25. [25]
  26. [26]
    [PDF] Exergetic Sensitivity Analysis of ORC Geothermal Power Plant ...
    They concluded that dual-pressure ORC, works with lower thermal efficiency but higher exergy efficiency and net power. Performance of the combined flash-binary ...
  27. [27]
  28. [28]
    [PDF] ANALYSIS OF BINARY THERMODYNAMIC CYCLES ...
    This report analyzes geothermal binary cycles to find low-cost energy from low-temp resources, using isobutane, pentane, and other fluids, and direct-contact  ...
  29. [29]
    A Review on the Preliminary Design of Axial and Radial Turbines for ...
    Organic working fluids with a higher molecular weight ... Organic Rankine Cycle turbine for electricity production using a geothermal binary power plant.
  30. [30]
    Performance Analysis of Organic Rankine Cycle With Preliminary ...
    Sep 17, 2013 · The radial turbo expanders play an important role in the performance of the organic rankine cycle (ORC) for binary-cycle geothermal plants.
  31. [31]
    [PDF] Water-Based Geothermal Binary Cycles
    Feb 12, 2025 · 1.1 History of Geothermal Power Cycles. The first commercial geothermal power plant, Larderello 1, began operating in Italy in 1914 with a ...
  32. [32]
    [PDF] GEOTHERMAL BINARY PLANT OPERATION AND MAINTENANCE ...
    a) The critical temperature of materials used in plant components must be far below the maximum temperature that can be reached in the cycle; b) Saturation ...
  33. [33]
    None
    ### Summary of Cooling Methods for Binary Geothermal Plants
  34. [34]
    [PDF] Development of a Geothermal Binary Cycle Simulator
    When the turbine exit presure has been finalized, the condenser exit temperature is chosen to represent 0.2° F subcooling (to avoid cavitation in the cycle pump) ...
  35. [35]
    Binary plants - Ormat Technologies
    The organic vapors drive the turbine and then are condensed in a condenser, which is cooled by either air or water. The turbine rotates the generator.
  36. [36]
    Techno-economic survey of Organic Rankine Cycle (ORC) systems
    Pre-feed pump. A first pump with low NPSH is added before the main feed pump (see Fig. 16) to provide the required pressure head. This is the strategy ...
  37. [37]
    Brine re-injection pump for geothermal plants - Sulzer
    The brine re-injection pump re-injects the low enthalpy reject brine back to the geothermal field to renew the available resources.Missing: abrasives | Show results with:abrasives
  38. [38]
    (PDF) Enhancing the Efficiency of the Double-Tube Heat Exchanger ...
    Sep 16, 2023 · This study utilized two double tube-type heat exchangers. The first exchanger employed a smooth inner tube, while the second one utilized a ...
  39. [39]
    Use of plat heat exchanger for binary cycle system in utilization of ...
    Sep 6, 2024 · For this reason, it is necessary to design a practical small-scale binary cycle unit, by designing the type and dimensions of the heat exchanger ...
  40. [40]
    [PDF] THE GEOTHERMAL BINARY FLUID CYCLE; HEAT EXCHANGER ...
    Pumping power is neglected. 0. The heat exchanger will be of the simple counterflow type and the overall heat transfer coefficient.
  41. [41]
    Corrosion and Scaling in Geothermal Heat Exchangers - MDPI
    This review critically examines the challenges posed by corrosion and scaling in geothermal heat exchangers, with a primary focus on three key mitigation ...
  42. [42]
    [PDF] Comparison of shell-and-tube with plate heat exchangers for the use ...
    Jul 9, 2014 · Abstract. Organic Rankine cycles (ORCs) can be used for electricity production from low-temperature heat sources.
  43. [43]
    [PDF] Optimal design of binary cycle power plants for water ... - CORE
    The geothermal fluid inlet temperatures considered are in the 110-160 °C range, while the return temperature of the brine is assumed to be between 70 and 100 ° ...<|control11|><|separator|>
  44. [44]
    [PDF] TITANIUM IN THE GEOTHERMAL INDUSTRY
    Direct binary cycle geothermal systems require that binary heat exchanger tubes be immune to all forms of corrosion. These systems can also include brine.
  45. [45]
    Materials used for geothermal heat exchanger applications.
    Titanium alloys and stainless steels found application in tube and shell heat exchangers for geothermal binary cycle plants.
  46. [46]
    Second Law assessment of binary plants generating power from low ...
    Geothermal binary plants are relatively poor converters of heat into work. First Law or thermal efficiencies typically lie in the range of 8–12%. As a ...
  47. [47]
    Thermodynamic and economic analysis and optimization of power ...
    Maximum first law efficiencies vary between 6.9% and 10.6% while the second law efficiencies vary between 38.5% and 59.3% depending on the cycle considered. ...
  48. [48]
    Exergy analysis of a dual-level binary geothermal power plant
    The study describes an easy-to-follow procedure for exergy analysis of binary geothermal power plants and how to apply this procedure to assess the plant ...
  49. [49]
    [PDF] SMALL GEOTHERMAL POWER PLANT DEVELOPMENTS
    Since the basic binary cycle can have alow second law or utilization ... Note :For thisbinary plant a second law efficiency of 33.5% is achieved,basedonthe.
  50. [50]
    Investigation of waste heat recovery of binary geothermal plants ...
    In the binary geothermal cycle, it has been found out that the highest exergy destruction for all refrigerants occurs in the heat exchanger. And the highest and ...
  51. [51]
    Efficiency of geothermal power plants: A worldwide review
    Aug 9, 2025 · Binary cycle plants operate at temperatures between 85°C and 180°C, achieving efficiencies of 10% to 13%. In contrast, dry steam plants, ...
  52. [52]
    Ideal thermal efficiency for geothermal binary plants - ScienceDirect
    This paper presents a basis for comparing geothermal power plants of the binary type with the ideal thermodynamic cycle appropriate for such plants.<|control11|><|separator|>
  53. [53]
    A Review of Working Fluids for Organic Rankine Cycle (ORC ...
    In this study, comprehensive review of working fluids selection for ORC applications has been carried out. ... A Guide to working fluid selection for Organic ...
  54. [54]
    (PDF) A Review of Working Fluids for Organic Rankine Cycle (ORC ...
    This study helps in identifying the possible most suitable organic fluids for various ORC applications depending on the operating conditions.
  55. [55]
    based binary zeotropic mixtures as working fluids for geothermal ...
    Mar 25, 2017 · The results revealed that mixed working fluids have better performance than pure working fluids under the lower initial temperature of the heat ...
  56. [56]
    Criteria for the Selection of Working Fluids for Geothermal Power ...
    Nov 26, 2018 · The selection of working fluid among the four candidates is made using the criterion of maximum net power. For which a basic subcritical cycle ...
  57. [57]
    [PDF] Tailored Working Fluids for Enhanced Binary Geothermal Power ...
    Objective: Down-select of Working Fluid Selection, System and Component-level Designs. •Costs of reservoir characterization, drilling and pumping resource ...
  58. [58]
    [PDF] REFPROP Documentation - Thermodynamics Research Center
    Jun 4, 2018 · REFPROP, or REFerence fluid PROPerties, calculates thermodynamic and transport properties of fluids and mixtures, developed by NIST.Missing: cycle glide zeotropic
  59. [59]
    Multi-Objective NSGA-II Optimization of Single- and Dual-Fluid ORC ...
    Nov 28, 2024 · Isobutane, with the similar molar mass as butane, has a slightly lower critical temperature (134.66 °C) and pressure (36.29 bar), with a ...
  60. [60]
    [PDF] Selection of Optimum Working Fluid and Cycle Configuration of ...
    Feb 12, 2020 · This research presents the thermo-economic approach for determining the optimal setting of Organic Rankine Cycle (ORC) as the proposed bottoming ...
  61. [61]
    Pentane - Thermophysical Properties - The Engineering ToolBox
    Chemical, physical and thermal properties of pentane, also called n-pentane. Phase diagram included ... Critical temperature, 469.8, K, 196.7, °C, 386.0, °F.
  62. [62]
    [PDF] Comparison of different Organic Rankine Cycle for power ... - IIETA
    Among the five considered organic fluids, the critical temperature of Isopentane. (187.2 °C) is the closest to the waste heat temperature (160 °C). Hence, it ...
  63. [63]
    [PDF] A Review of Kalina Cycle
    Third, ammonia–water has thermophysical property that causes mixed fluid temperatures to increase or decrease without a change in the heat content. The ...
  64. [64]
    Kalina Cycle - an overview | ScienceDirect Topics
    The Kalina cycle is defined as a thermodynamic cycle that generates electricity using a mixture of ammonia and water, enhancing efficiency in various ...
  65. [65]
    [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 ...
  66. [66]
    Thermodynamic analysis of R134a in an Organic Rankine Cycle for ...
    Mar 16, 2024 · The conclusion was that ORC with R134a can function at lower temperatures; it is an excellent candidate for creating usable energy as a working ...
  67. [67]
    Thermodynamic analysis of R134a in an Organic Rankine Cycle for ...
    The results showed a system efficiency of about 7% and 4%, respectively. Other studies that have analyzed the use of R134a as working fluid in the ORC cycles ...
  68. [68]
  69. [69]
    Energy, exergy and exergo-economic analyses of supercritical CO2 ...
    Dec 30, 2024 · ... geothermal heat exchanger, respectively. In the second expression ... binary-cycle geothermal plant. Appl Therm Eng, 125 (2017), pp. 153 ...
  70. [70]
    Technology Transitions HFC Restrictions by Sector | US EPA
    Beginning January 1, 2025, certain technologies may no longer use high global warming potential (GWP) hydrofluorocarbons (HFCs) or HFC blends.
  71. [71]
    On the Viability of Hydrofluoroolefin-R245fa Mixtures as Organic ...
    Aug 7, 2025 · The aim of this study is to investigate the thermo-economic performance of the transcritical organic Rankine cycle (TORC) system using R1234yf/ ...
  72. [72]
    Environmental impact of cogeneration in binary geothermal plants
    In particular, closed-loop circulation of geothermal fluids in so-called binary systems prevents critical on-site emissions of carbon dioxide and methane, which ...
  73. [73]
    [PDF] Life Cycle Environmental Impacts of Geothermal Systems
    Feb 1, 2012 · U.S. geothermal power plants. Their average includes the zero GHG emissions from binary plants, which represent 14% of the surveyed capacity ...
  74. [74]
    (PDF) Review of H2S abatement methods in geothermal plants.
    Feb 10, 2025 · This paper discusses some of the H2S abatement methods available to the industry, such as liquid redox methods, reinjection, Selectox, Dow-Spec ...
  75. [75]
    (PDF) Water Use in the Development and Operation of Geothermal ...
    Jul 14, 2015 · For the EGS scenarios, plant operations consume between 0.29 and 0.72 gal/kWh. The binary plant experiences similar operational consumption, at ...<|separator|>
  76. [76]
    Geothermal FAQs - Department of Energy
    Geothermal heat pumps (GHPs) are cost-effective technologies in the long run, but the costs of ground heat-exchanger loops make them more expensive up front.
  77. [77]
    [PDF] Geothermal Technologies Program - NREL
    Geothermal power plants have small footprints; that is, they require little land compared to coal and nuclear power plants. An entire geothermal field uses 1-8 ...
  78. [78]
    Environmental Impacts of Geothermal Energy
    Mar 5, 2013 · Estimates of global warming emissions for open-loop systems are approximately 0.1 pounds of carbon dioxide equivalent per kilowatt-hour. In ...
  79. [79]
    Environmental assessment of a binary geothermal sourced power ...
    Sep 1, 2019 · ... geothermal power plant is the cause of exergy destruction of the equipment involved. ... heat exchanger, pump, and turbine. Restrepo and ...
  80. [80]
    Geothermal Electric Technology | WBDG
    Binary-cycle, or organic Rankine cycle, power plants use lower temperature geothermal resources around 100°C (212°F), which is a major advantage in areas where ...
  81. [81]
    [PDF] Geothermal power: Technology brief - IRENA
    Typical costs for geothermal power plants range from. USD 1 870 to USD 5 050 per kilowatt. (kW), noting that binary plants are normally more expensive than ...
  82. [82]
    Renewable Electricity Production Tax Credit (PTC) - DSIRE
    Jul 18, 2025 · Note, projects that commence construction on or after January 1, 2025 can receive a tax credit under the new Clean Energy Production Tax Credit ...
  83. [83]
    Geothermal Power Market Size to Hit USD 10.78 Billion by 2034
    Asia Pacific dominated the position sensor market with the largest market share of 34% in 2024. · Market Size in 2025: USD 6.95 Billion · Calpine · Binary Cycle ...Missing: percentage | Show results with:percentage
  84. [84]
    GSR 2025 | Geothermal - REN21
    At the start of the 2024-2025 heating season, geothermal heating capacity was reported to have grown more than 17%, serving an additional 120 million m2 of ...Missing: percentage | Show results with:percentage
  85. [85]
    Geothermal Energy Market Size And Share Report, 2030
    ... share of 30.5% in 2023. Binary cycle power plants utilize lower temperature geothermal resources which are more abundant globally compared to high ...
  86. [86]
    [PDF] 2021 U.S. Geothermal Power Production and District Heating Market ...
    Feb 10, 2020 · In 1974, the Geothermal Energy Research Development and. Demonstration Act (30 U.S.C. § 24) established the first federal loan guarantee ...<|separator|>
  87. [87]
    East Mesa Magmamax Power Process Geothermal Generating Plant ...
    Dec 1, 1980 · During recent months, Magma Power Company has been involved in the shakedown and startup of their 10 MW binary cycle power plant at East ...
  88. [88]
    Magmamax Binary Power Plant, East Mesa, Imperial Valley ...
    The first commercial scale binary power plant operating on geothermal energy was dedicated on 6 December 1985. This 50 MW (gross) power plant is built on a 20- ...
  89. [89]
    Ormat Successfully Completed Ngatamariki Geothermal Plant
    Mighty River Power's modular Ngatamariki geothermal power plant is the largest singular binary power plant ever constructed. The Ormat Energy Converters are ...
  90. [90]
    Ngā Tamariki Geothermal System
    The Ngā Tamariki Power Station was commissioned in 2013. It has four Ormat 20.5 MW(e) binary cycle units, the largest Ormat Binary Cycle units in the world ...
  91. [91]
    Performance analysis of the Chena binary geothermal power plant
    In this paper, the IPSEpro model of the Chena Geothermal ORC Power Plant had been developed and validated using the real data.
  92. [92]
    1000+ MW of geothermal power generation capacity
    Dec 14, 2018 · One of them is POWER Engineers that is proud having worked on geothermal projects of a total of more than 1,000 MW in capacity.
  93. [93]
    [PDF] SUSTAINABILITY OF THE SALTON SEA - Colorado College
    The majority of the Salton Sea geothermal power plants are flash steam plants but there are binary plants too. ... geothermal since it is a reliable and ...<|separator|>
  94. [94]
    Organic Rankine Cycle – ORC - Ormat Technologies
    The Organic Rankine Cycle (ORC) is an evolving energy system for power production utilizing geothermal resources and recovered waste-heat.Missing: 10-12% 80%
  95. [95]
    Thermodynamic performance evaluation of a geothermal ORC ...
    According to the results of the study, thermal efficiency of the geothermal binary plants mostly vary between 8 and 12%; as a result, only 1–2 point increase in ...
  96. [96]
    Kalina Cycle Enjoying Commercial Success - Power Engineering
    Feb 1, 2002 · The Kalina Cycle uses a binary working fluid of an ammonia-water mixture to drive a turbine-generator. By varying the composition of the binary ...Missing: ORC Denmark Israel
  97. [97]
    [PDF] Small-Scale Geothermal Power Plant Field Verification Projects
    The projects include a 1-MW binary-cycle plant in Nevada, a 1-MW Kalina-cycle plant in New Mexico, and a 750-kW flash plant in Utah, with direct heating uses.
  98. [98]
    [PDF] Corrosion in the Kalina cycle - Skemman
    They use a working fluid of ammonia and water, benefiting from the property of mixtures of boiling over a temperature range at constant pressure, to achieve ...
  99. [99]
    Kalina Cycle power systems in waste heat recovery applications
    Aug 6, 2012 · General corrosion​​ In Kalina Cycle environments the presence of ammonia (NH3) ensures strongly alkaline conditions and consequently pH values of ...
  100. [100]
    [PDF] Development of a Low Temperature Geothermal Organic Rankine ...
    Nov 20, 2013 · Low temperature geothermal is an abundant resource in New Zealand with over 260 sites with a resource temperature of 150°C and lower.Missing: Adoption Kalina niche
  101. [101]
    Advancing geothermal energy utilization opportunities: potential and ...
    Jun 18, 2025 · This study explores the potential of geothermal resources to meet the thermal and electrical demands of DAC systems through the development of a geothermal-DAC ...2. Methodology · 3.1. Geothermal-Dac... · 3.2. Geothermal-Dac...
  102. [102]
    Modern geothermal power: Binary cycle geothermal power plants
    Aug 6, 2025 · Binary cycle power plants have an average unit capacity of 6.3 MW, 30.4 MW at single-flash power plants, 37.4 MW at double-flash plants, and 45 ...
  103. [103]
    Impact of Climate Change on the Thermoeconomic Performance of ...
    Sep 3, 2024 · This study focuses on the performance analysis of a binary cycle with isobutane for geothermal power generation under the impact of climate change.
  104. [104]
    Technology breakthroughs are unlocking geothermal energy's vast ...
    Dec 13, 2024 · The report finds that the total investment in geothermal could reach $1 trillion by 2035 and $2.5 trillion by 2050. If next-generation ...
  105. [105]
    U.S. DOE unveils roadmap for next generation of geothermal power
    Mar 25, 2024 · This report shows how advanced geothermal technology could increase the US' geothermal energy production to 90 GW or more by 2050, a twentyfold increase.