Fact-checked by Grok 2 weeks ago

Ocean thermal energy conversion

Ocean thermal energy conversion (OTEC) is a technology that exploits the temperature difference between warm surface seawater, heated by solar radiation, and colder —typically around 20–25°C—to drive a for production. The concept relies on vast ocean thermal resources in tropical regions, where surface temperatures average 25–30°C and deep waters below 1,000 meters remain near 5°C, enabling continuous baseload power without inputs. Proposed in 1881 by French physicist Jacques-Arsène d'Arsonval and first experimentally demonstrated in the 1930s by with a small open-cycle plant off that produced 22 kW of electricity but suffered pipe breakage, OTEC has seen intermittent pilot-scale development rather than widespread commercialization. Systems operate in closed-cycle configurations using a low-boiling-point like as the working medium, or open-cycle designs that vaporize directly for power and ; hybrid variants combine both for enhanced efficiency. Theoretical Carnot efficiency is capped at 6–8% due to the modest , with practical net efficiencies often below 3% after accounting for parasitic pumping losses, limiting economic viability without scale or co-products like cold water for or . Notable achievements include the U.S. Department of Energy-funded 100 kW closed-cycle plant operational since 2011 on Hawaii's Big Island by Makai Ocean Engineering, which has delivered over 1 GWh to the grid, demonstrating reliability but highlighting persistent challenges such as , high upfront costs exceeding $10,000 per kW installed, and the need for robust cold-water pipes in dynamic ocean environments. Recent modeling suggests global OTEC potential could expand by 46% by 2100 under high-emission scenarios due to surface warming, yet deployment lags owing to capital risks and competition from cheaper and , with no utility-scale plants beyond pilots as of 2025. Environmental impacts appear minimal, including nutrient that may boost local productivity but risks altering ecosystems if scaled massively.

Basic Principles

Operational Mechanism

Ocean thermal energy conversion (OTEC) harnesses the in ocean waters, typically a difference of 20°C or more between warm surface and cold deep , to drive a for . In tropical regions, surface waters average 24–29°C, while deep waters at 800–1,000 meters depth maintain 4–5°C year-round due to limited solar penetration and vertical mixing. This gradient serves as the high-temperature heat source and low-temperature in a analogous to a Rankine steam cycle. The predominant closed-cycle OTEC system employs a low-boiling-point , such as (boiling point -33°C at ), to amplify the small difference into usable differentials. Warm surface is pumped at rates of several cubic meters per second through an evaporator , transferring heat to boil the into high-pressure vapor without mixing fluids. This vapor expands through a , coupled to a , producing mechanical and electrical power; turbine efficiencies reach 80–90% in prototypes. The vapor then enters a condenser , where cold deep —piped via a flexible, anchored cold water pipe of 1–2 meter diameter and up to 1,000 meters length—absorbs heat to condense the fluid back to liquid. The is repressurized by a feed and returned to the , closing the loop. Pumping requirements for flows, often 10–20 times the mass flow, consume 20–30% of gross output, yielding net efficiencies of 2–4%. must withstand and , typically using plates with enhanced surfaces for coefficients up to 3,000 W/m²K. Open-cycle variants directly evaporate surface seawater under to produce for drive, yielding condensate as a , but require larger vacuum pumps and face scalability issues with non-condensable gases. Hybrid systems integrate elements of both for augmentation. All configurations demand sites with persistent thermoclines and minimal seasonal variation for reliable operation.

Thermodynamic Fundamentals

Ocean thermal energy conversion (OTEC) functions as a low-grade , exploiting the persistent in tropical oceans where surface waters average 25–29°C and deep waters below the maintain 4–7°C, yielding a practical of 18–25°C. This drives a analogous to a but constrained by the second law of thermodynamics, converting into mechanical work via across finite temperature differences. The process requires massive flows—typically 10–20 times the mass in closed systems—to achieve viable power outputs, introducing significant parasitic pumping losses that dominate net efficiency. The upper bound on efficiency is the Carnot limit, η_C = 1 - (T_c / T_h), with temperatures in ; for T_h = 302 K (29°C) and T_c = 277 K (4°C), η_C ≈ 8.3%, though site-specific ΔT often limits this to 6–7%. Actual cycles fall short due to finite-rate , pressure drops, and non-ideal components, with second-law analyses revealing exergy destruction primarily in evaporators and condensers from log-mean temperature differences of 2–5°C. Net power is W_net = W_turbine - ΣW_pumps, where turbine work derives from isentropic (η_t ≈ 80–85%) and pumps consume 20–30% of gross output for circulation at Reynolds numbers exceeding 10^5 in pipes. Closed-cycle OTEC employs a with a volatile fluid like (boiling point -33°C at 1 atm), where warm vaporizes the fluid at 5–10 bar, driving a before at 0.5–1 bar by cold water; the cycle's η_th = W_net / Q_h typically ranges 2–4%, optimized via multi-staging or Kalina variants to approach 50% of Carnot. Open-cycle systems use of under vacuum (pressure ~2.3 kPa at 25°C), producing low-pressure for , but face penalties from vacuum maintenance and non-condensable gases, yielding η_th < 2% without desalination co-production. Hybrid cycles integrate both for enhanced performance, balancing working fluid purity with direct seawater use. Thermodynamic modeling incorporates finite-time analysis to maximize power under realistic constraints, such as evaporator conductance UA ≈ 100–500 kW/K/m², revealing optimal intermediate temperatures that boost output by 10–20% over reversible assumptions. Piping losses and biofouling further degrade performance, necessitating alloys resistant to corrosion at these low ΔT. Overall, OTEC's viability hinges on minimizing irreversibilities, with recent simulations projecting scalable efficiencies up to 5% for 100 MW plants via advanced heat exchangers.

Historical Development

Early Concepts and Patents

The concept of ocean thermal energy conversion (OTEC) originated with French physicist , who proposed in 1881 harnessing the temperature gradient between warm surface seawater and colder deep water to drive a heat engine. D'Arsonval envisioned a closed-cycle system employing a low-boiling-point working fluid, such as , evaporated by surface water heat and condensed by deep water cold, with the resulting vapor pressure difference powering a turbine. This theoretical framework anticipated exploiting typical tropical ocean gradients of 20–25°C between surface temperatures around 25–30°C and deep water below 5°C. D'Arsonval's student, engineer Georges Claude, advanced the idea into practical experiments, favoring an open-cycle approach that used seawater itself as the working fluid via low-pressure evaporation. In 1928, Claude constructed a small-scale test facility in Ougrée, Belgium, utilizing warm effluent from a coastal power plant to simulate surface conditions, demonstrating proof-of-concept power generation. Two years later, in 1930, he erected the first OTEC prototype on land in Matanzas, Cuba, featuring a 60-meter tower and a 1.4-kilometer-long cold-water pipe extending seaward; the system produced 22 kilowatts of electricity intermittently but ceased operation after several months due to biofouling and structural failure of the pipe. Claude's innovations culminated in U.S. Patent 2,006,985, granted on July 2, 1935, to and Paul Boucherot for a "method and apparatus for obtaining power from sea water." The patent detailed an open-cycle process involving vacuum evaporation of warm seawater, turbine expansion of the steam, and direct condensation using cold deep water, addressing engineering challenges like maintaining vacuum and pipe integrity observed in the Cuban trial. These early efforts highlighted fundamental technical hurdles, including parasitic pumping losses and biological encrustation, yet established the feasibility of principles despite limited commercial success at the time.

Prototype Testing and Milestones

In 1930, French engineer constructed the first operational OTEC prototype at Matanzas Bay, Cuba, employing an open-cycle system that utilized seawater as the working fluid to drive a low-pressure turbine, producing approximately 22 kilowatts of electricity but operating for only about 11 days due to challenges with vacuum maintenance and biofouling on pipes. The plant featured a 60-meter evaporator pipe and a 1,000-meter condenser pipe drawing cold deep water, demonstrating proof-of-concept for vacuum-based evaporation but highlighting inefficiencies from non-optimized heat exchangers and pipe losses, with no net power achieved after parasitic loads. Advancing to the late 1970s amid U.S. energy crises, the Mini-OTEC experiment, deployed by off Keahole Point, Hawaii, in 1979, marked the first closed-cycle OTEC system to produce net electricity at sea, generating 50 kilowatts gross and 15-18 kilowatts net using ammonia as the working fluid in a barge-mounted platform at 900 meters depth. This prototype tested heat exchanger performance, cold water pipe dynamics, and platform stability in 3,000-foot waters, validating thermodynamic efficiencies around 3% while identifying issues like biofouling and pipe vibrations, with operations lasting three months before decommissioning. Subsequent U.S. efforts included the OTEC-1 facility in 1980 at the Natural Energy Laboratory of Hawaii Authority (NELHA), a 210-kilowatt gross shore-based plant that operated intermittently to evaluate titanium heat exchangers and desalination integration, achieving milestones in scaling closed-cycle components but facing corrosion and economic hurdles that limited long-term viability. Internationally, Japan's Nauru plant in 1981 produced 30 kilowatts net briefly, testing floating platforms, while the Republic of Korea's KRISO barge in 2019 achieved the largest net output to date at 338 kilowatts from a 1-megawatt system, focusing on hybrid cycle efficiencies and cold water pipe materials in tropical waters. These prototypes underscored persistent challenges in capital costs and scaling, with net efficiencies rarely exceeding 3-4% under real ocean conditions.

Recent Progress and Demonstrations

In 2023, Mitsui O.S.K. Lines, Ltd. (MOL) initiated an ocean thermal energy conversion (OTEC) demonstration project in Hawaii, partnering with local entities to manufacture and test large-scale cold water pipes and heat exchangers, with the goal of achieving 1 MW commercialization by 2026. This effort builds on prior small-scale tests by focusing on component scalability to address deployment challenges in tropical waters. Global OTEC advanced multiple prototypes during 2023–2025, including the PLOTEC initiative—a EU-funded project developing a storm-resistant floating structure, with offshore testing commencing off in October 2025 to validate durability under extreme wave conditions up to 15 meters. In March 2025, the company launched its largest onshore OTEC pilot plant to date, incorporating modular power units for efficient heat exchange and electricity generation from seawater gradients. These developments aim to reduce capital costs, which have historically limited OTEC viability, through optimized designs yielding net power outputs in the tens of kilowatts during initial runs. By September 2025, joined the —comprising , , and —to evaluate integration for powering deepwater oil and gas operations, targeting hybrid systems that leverage existing offshore infrastructure for cold water access. This collaboration highlights potential non-utility applications, with simulations indicating feasible power densities of 1–2 MW per platform in equatorial regions. Concurrently, 's "Dominique" 1.5 MW floating platform progressed toward deployment in in late 2025, marking an anticipated milestone for grid-connected baseload renewable output exceeding prior demonstrations.

System Configurations

Closed-Cycle Systems

Closed-cycle ocean thermal energy conversion systems employ a secondary working fluid with a low boiling point, such as , circulating in a sealed loop isolated from seawater. Warm surface seawater, typically at 25–30°C, flows through an evaporator heat exchanger, vaporizing the fluid to produce high-pressure vapor that drives a turbine connected to a generator for electricity production. The vapor then enters a condenser cooled by cold deep seawater at 4–10°C, liquefying the fluid before it is pumped back to the evaporator, completing the cycle without direct seawater involvement in power generation. Ammonia serves as the primary working fluid due to its favorable thermodynamic properties, including high latent heat of vaporization and compatibility with the low-temperature differentials of OTEC operations, enabling efficient phase changes at oceanic temperatures. Alternative fluids, including propane, R-134a, and mixed refrigerants like R717 blends, have been tested to potentially enhance performance, though ammonia predominates in prototypes for its balance of efficiency, availability, and manageable toxicity risks. Closed cycles, including variants like the , offer improved thermal resource utilization over open-cycle systems by allowing precise fluid selection and reduced scaling issues in heat exchangers. Net power efficiencies in closed-cycle OTEC typically range from 1.5% to 3%, constrained by the Carnot efficiency limit for a 20°C temperature difference and substantial parasitic pumping loads for large seawater volumes—often 10–20 times the working fluid flow. Demonstrations include a 105 kW facility at Hawaii's Natural Energy Laboratory, the first closed-cycle plant grid-connected in the U.S. in 2013, which operated to validate heat exchanger performance and biofouling mitigation. Earlier tests, such as the 1 MW OTEC-1 loop in 1980–1981, confirmed scalability of components like radial-flow turbines suited to low-pressure drops. Ongoing research focuses on optimizing evaporator and condenser designs to minimize exergy losses and boost output.

Open-Cycle Systems

Open-cycle ocean thermal energy conversion (OTEC) systems employ warm surface seawater directly as the working fluid, distinguishing them from closed-cycle configurations that use secondary fluids like ammonia. In operation, warm seawater, typically at 25–30°C, is pumped into a low-pressure evaporator chamber maintained under vacuum, where partial evaporation occurs, producing low-pressure steam. This steam expands through a specialized low-speed turbine designed to handle wet, low-density vapor, converting thermal energy into mechanical work that drives a generator. The exhaust steam then enters a condenser cooled by deep ocean water at 4–10°C, where it condenses into fresh water, yielding desalination as a byproduct. To sustain the vacuum essential for low boiling points, non-condensable gases extracted from the seawater must be continuously vented, often via a vacuum pump or deaerator, preventing accumulation that would reduce efficiency. The unevaporated brine from the evaporator and the cooling water discharge require management to minimize environmental impacts, such as nutrient upwelling from cold water that could alter local ecosystems. System design necessitates corrosion-resistant materials like for heat exchangers and turbines due to direct seawater contact, alongside measures against biofouling on intake pipes and surfaces. Large pipe diameters—up to several meters—are required for steam transport owing to its low density at operating pressures around 2–5 kPa, increasing structural demands. Thermodynamic efficiency in open-cycle OTEC is constrained by the Carnot limit and practical losses, typically achieving 1–3% net efficiency, lower than closed-cycle systems due to irreversibilities in flashing and condensation processes. For instance, a 1993 demonstration plant in Hawaii produced 210 kW gross power using open-cycle principles, highlighting feasibility but underscoring scaling challenges for commercial viability. Advantages include integrated desalination, with fresh water output potentially exceeding 1 m³ per kWh generated, beneficial for water-scarce tropical regions, and simpler fluid handling without secondary refrigerants. However, disadvantages encompass higher pumping power needs for voluminous flows—cold water rates of 2.5–3 m³/s per net MW—and vulnerability to vacuum leaks or gas ingress, which degrade performance. Recent analyses emphasize hybrid integrations, such as combining open-cycle with solar enhancement, to boost output, though pure open-cycle remains challenged by capital costs for vacuum systems and large infrastructure. Overall, while open-cycle OTEC offers dual energy and water production, its deployment lags behind closed-cycle due to technical hurdles in efficiency and materials durability.

Hybrid Systems

Hybrid ocean thermal energy conversion (OTEC) systems integrate open-cycle and closed-cycle processes, leveraging flash evaporation of seawater for desalination alongside a secondary working fluid for enhanced power generation. In this configuration, warm surface seawater enters a vacuum chamber, where partial flash evaporation produces low-pressure steam that expands through a turbine to generate electricity, with the unevaporated seawater serving as the primary working fluid. The resulting steam, after partial expansion, condenses on the surface of a low-boiling-point fluid (such as ammonia or propane) in a heat exchanger, vaporizing it to drive an additional turbine for supplementary power output, thereby combining desalination with higher overall electricity production compared to standalone open-cycle systems. This approach avoids the need for indirect heat exchangers in the evaporation stage, reducing thermal losses and material corrosion risks associated with seawater handling. The hybrid design optimizes resource use by co-producing fresh water—estimated at up to 0.5–1.0 liters per kWh of electricity in conceptual models—and electricity, with net efficiency potentially reaching 3–5% under typical tropical temperature gradients of 20–25°C between surface (25–30°C) and deep (4–10°C) waters. Conceptual analyses from the late 1980s indicated that hybrid plants could flexibly adjust outputs to prioritize either power or desalination, with capital costs projected at $200–400 million for a 100 MW facility, though scalability depends on site-specific ocean conditions like thermocline stability. Unlike pure closed-cycle systems, hybrids utilize direct seawater vapor for initial power, minimizing auxiliary pumping energy, but they require precise vacuum control (typically 2–5 kPa) to sustain flash evaporation and prevent biofouling in the open components. Challenges include higher system complexity, leading to elevated maintenance needs for dual-cycle components, and sensitivity to seawater quality, where impurities can degrade turbine performance or increase scaling in the flash chamber. Despite these, hybrid systems offer economic advantages in regions requiring both energy and water, such as small island states, with studies suggesting payback periods of 10–15 years at electricity prices above $0.15/kWh when desalination revenue is factored in. No commercial-scale hybrid OTEC plants have been deployed as of 2025, with development limited to simulations and small prototypes, reflecting persistent hurdles in cold deep-water pipe deployment and overall cost competitiveness against solar or wind alternatives.

Deployment Strategies

Land-Based Installations

Land-based (OTEC) installations involve onshore power plants that pump warm surface seawater and cold deep ocean water through pipelines to the facility for heat exchange in closed-, open-, or hybrid-cycle systems. This approach facilitates direct grid connection and maintenance access but demands substantial pumping power to lift cold water from depths of 600–1,000 meters over seabed topography, often resulting in net efficiency penalties of 20–30% due to hydrostatic head losses. The earliest documented land-based OTEC prototype was constructed by French engineer in 1930 at Matanzas Bay, Cuba, employing a vacuum flash evaporation method with a 2.2-megawatt thermal capacity evaporator drawing on seawater temperature differentials. The system generated approximately 22 kilowatts gross but achieved negligible net power after pumping losses and operated only briefly before destruction by a hurricane in 1933, highlighting early challenges with biofouling and structural durability. In Hawaii, significant advancements occurred at the Natural Energy Laboratory of Hawaii Authority (NELHA) in Keahole Point. An open-cycle plant operated from 1993 to 1998, producing up to 255 kilowatts gross power and demonstrating net electricity generation using cold seawater for desalination and aquaculture integration. More recently, Makai Ocean Engineering commissioned a 105-kilowatt closed-cycle demonstration plant in 2015 at NELHA, which became the first OTEC system connected to the U.S. grid, supplying continuous baseload power equivalent to 120 homes annually through ammonia as the working fluid and titanium heat exchangers. Japan has also pursued land-based pilots, including a 100-kilowatt facility on Kumejima Island, Okinawa Prefecture, operational since 2013 and grid-connected by 2016, serving as a testbed for hybrid applications amid regional resource constraints. These installations remain at pilot scale, with no commercial-scale land-based plants deployed globally as of 2023, primarily due to high capital costs for deepwater piping exceeding $10 million per kilometer and environmental permitting hurdles for seawater discharge.

Shelf-Based Systems

Shelf-based ocean thermal energy conversion (OTEC) systems are deployed on the continental shelf in water depths typically ranging from 50 to 600 meters, utilizing fixed structures anchored directly to the seabed rather than floating platforms or land-based facilities. This configuration positions the plant closer to shore than deep-sea floating systems, enabling shorter power cables to transmit electricity to the grid and reducing associated transmission losses and infrastructure expenses. Such setups require access to deep cold seawater via vertical or inclined cold water pipes that extend beyond the shelf break to depths of 800–1,000 meters, where temperatures drop to 4–7°C, maintaining the necessary 20°C surface-to-depth gradient for efficient operation. The primary advantages of shelf-based systems include enhanced stability against surface waves and currents due to seabed anchoring, which eliminates the need for complex dynamic positioning or mooring systems used in floating designs, thereby lowering long-term maintenance costs and improving reliability. Proximity to coastal infrastructure also facilitates routine inspections, repairs, and workforce access via boat or short-distance subsea operations, contrasting with the logistical challenges of offshore floating plants. These systems align with U.S. Department of Energy priorities post-1981, which shifted focus toward shelf-mounted and land-based to capitalize on nearshore resource potential and mitigate deep-water deployment risks. However, deployment is constrained by site-specific geology, as suitable continental shelves must feature persistent thermal gradients without excessive sedimentation or seismic activity that could damage anchors or pipes. Construction often involves seabed preparation, such as trenching for pipes and foundations, which can resuspend sediments, release nutrients or pollutants, and disrupt benthic habitats, potentially leading to localized ecological shifts in species composition and oxygen levels. High upfront capital costs for underwater structural engineering and pipe installation—estimated at 40–60% of total project expenses—further limit commercialization, with no large-scale operational shelf-based plants documented as of 2021, though conceptual assessments identify viable sites along shelves in regions like the U.S. Exclusive Economic Zone and parts of Brazil. Emerging demonstrations, such as the EU-funded prototype off Gran Canaria installed in 2025, explore shelf-adjacent resilience to storms but remain experimental at sub-megawatt scales.

Floating Platforms

Floating platforms represent a key deployment strategy for ocean thermal energy conversion (OTEC) systems, enabling operations in deep offshore waters where the vertical temperature gradient, or , provides the most consistent and substantial differentials—often exceeding 20°C between surface waters around 25–30°C and deep waters below 5°C at depths of 800–1000 meters. Unlike land-based or shelf-mounted installations, floating platforms decouple OTEC from geographical constraints such as shallow coastal bathymetry or limited land availability, facilitating scalability for multi-megawatt plants in tropical and subtropical regions with suitable ocean conditions. This configuration supports continuous baseload power generation, as the ocean's thermal mass buffers against diurnal and seasonal variations, yielding capacity factors potentially above 90%. Designs for floating OTEC platforms emphasize structural integrity against hydrodynamic loads, including wave excitation, current-induced motions, and mooring tensions, often employing semi-submersible or spar-type hulls optimized for minimal heave and pitch in extreme tropical weather. Storm resistance is critical, with platforms engineered to withstand cyclones and hurricanes through features like deep draft mooring systems and lightweight composite materials to reduce vulnerability to fatigue and biofouling on cold-water pipes extending hundreds of meters downward. Advantages include negligible land footprint and reduced visual intrusion compared to terrestrial renewables, alongside co-production potential for desalination or aquaculture via byproduct cold seawater. However, challenges persist, such as elevated capital costs for offshore fabrication and installation—estimated 20–50% higher than shelf-based systems due to subsea grid cabling and dynamic positioning—and the need for coupled aero-hydro-servo-elastic modeling to predict platform responses under combined environmental forcing. Recent developments underscore advancing feasibility, with the EU-funded PLOTEC project (2022–2025) designing and testing a scaled, storm-resilient floating platform at the PLOCAN test site off Gran Canaria, Spain, incorporating optimized hull forms and power cycles to validate performance in real ocean conditions over a planned year-long trial starting in 2025. Complementing this, Global OTEC's 1.5 MW "Dominique" platform, awarded Approval in Principle by in 2023, features a tension-leg or catenary mooring system and is slated for deployment off São Tomé and Príncipe in 2025, following successful 2024 sea trials of a prototype in Atlantic waters that demonstrated hydrodynamic stability. These initiatives, supported by partnerships like Global OTEC's memorandum with Enogia for turbine subsystems, address prior barriers by prioritizing modular construction and hybrid closed-cycle operations, though commercialization hinges on demonstrating levelized costs below $0.15/kWh through economies of scale beyond 10 MW.

Current Status and Projects

Operational Facilities

As of 2025, operational ocean thermal energy conversion (OTEC) facilities remain limited to small-scale demonstration plants, with no commercial-scale installations producing significant grid power. These systems primarily serve research and validation purposes, demonstrating technical feasibility in controlled environments. In Hawaii, the Natural Energy Laboratory of Hawaii Authority (NELHA) operates a 100 kWe closed-cycle OTEC plant managed by Makai Ocean Engineering. Commissioned around 2011 and upgraded over time, this facility uses ammonia as the working fluid in a closed loop, drawing warm surface seawater and cold deep water (approximately 900 meters depth) via pipelines to drive a turbine generator. It has supplied electricity to the local grid intermittently, achieving net power outputs of up to 100 kW while also providing cold seawater for aquaculture and desalination applications at NELHA. The plant's operations have logged thousands of hours, highlighting reliability in a tropical setting with a consistent thermal gradient exceeding 20°C. In Japan, Saga University maintains a 50 kW closed-cycle OTEC demonstration unit employing a double configuration for enhanced efficiency. Operational since the early 1980s with periodic upgrades, this land-based system utilizes nearby coastal waters for heat exchange, focusing on research into turbine performance, heat exchanger fouling, and system optimization. It generates power for on-site use and experimentation, contributing data to ongoing refinements in OTEC technology amid Japan's interest in baseload for island grids. Additionally, a 100 kW plant on Kumejima Island, operated under the Okinawa Prefecture, has been functional since 2013, employing a similar closed-cycle approach to test scalability and integration with local energy needs. These facilities underscore OTEC's potential for continuous baseload power from ocean thermal gradients but reveal challenges such as biofouling, high capital costs, and low conversion efficiencies (typically 3-5%), which limit scaling without further technological advancements. No offshore or hybrid operational plants exist as of October 2025, though pilot projects in locations like India and historical open-cycle tests in Hawaii (e.g., a 210 kW gross system from 1993-1998) inform current designs.

Proposed and Developmental Projects

In Hawaii, OTEC International LLC is advancing plans for a 1 MW closed-cycle OTEC facility at the Natural Energy Laboratory of Hawaii Authority's (NELHA) HOST Park, with lease negotiations nearing completion as of 2025 to enable construction and testing of scalable power generation integrated with the site's microgrid. This project builds on prior demonstrations at NELHA, aiming to validate higher-capacity operations using local seawater resources for baseload electricity. In Japan, Mitsui O.S.K. Lines (MOL) is pursuing commercialization of a 1 MW system on Kumejima Island by around 2026, expanding from an existing 100 kW demonstration plant operational since 2013, with efforts focused on manufacturing larger heat exchangers and cold water pipes to achieve economic viability through co-production of power, desalination, and aquaculture. The initiative, part of the "Kumejima Model," integrates deep seawater utilization for multiple outputs to offset costs. The European Union's Horizon Europe-funded PLOTEC project deployed a floating, storm-resistant OTEC prototype off Gran Canaria in the Canary Islands in October 2025, initiating sea trials to test structural integrity and energy conversion efficiency under Atlantic conditions simulating tropical extremes. Led by a pan-European consortium including , the platform design emphasizes modular scalability for small island deployment, with data collection ongoing to inform full-scale hybrid systems. Global OTEC, in partnership with ENOGIA and organizations like SIDS DOCK, is developing a 1.5 MW floating platform off São Tomé and Príncipe, with potential expansion to 10 MW and operations targeted for 2025 to supply baseload power to the island nation. This first-of-a-kind initiative addresses energy security in equatorial regions by leveraging consistent ocean thermal gradients. In Mauritius, MOL and collaborators, selected under Japan's Ministry of Economy, Trade and Industry (METI) program in January 2024, are conducting site suitability studies and infrastructure assessments for commercial OTEC deployment in the south-southwest coastal area, drawing on Kumejima's 100 kW experience to accelerate practical implementation without specified capacity yet. Global OTEC joined the DeepStar Phase 21 consortium in September 2025 to evaluate modular OTEC systems for replacing diesel and gas power in deepwater offshore oil and gas platforms, focusing on feasibility studies for baseload renewable integration in tropical operations. These efforts highlight ongoing engineering refinements to overcome deployment hurdles like biofouling and pipe stability.

Economic Viability

Cost Components and Analysis

The primary cost components of (OTEC) systems encompass capital expenditures (CAPEX) for construction and installation, and operational expenditures (OPEX) for ongoing maintenance and fuel-free operations. CAPEX dominates, typically accounting for over 90% of lifetime costs due to the engineering demands of large-scale heat exchangers, deep-sea cold water pipes, and robust platforms or floating structures required to exploit . Heat exchangers, which must handle massive seawater volumes for low-temperature differences (around 20°C), represent 30-40% of CAPEX, driven by corrosion-resistant materials like titanium to withstand biofouling and saltwater exposure. The cold water pipe, often 1,000 meters long and 2-10 meters in diameter, contributes 10-20% of CAPEX, with costs escalating for floating systems due to dynamic mooring and pipe stability challenges. Platform or hull construction adds 20-30%, while turbines, pumps, and power electronics comprise 10-20%. OPEX, estimated at 1.4-2.7% of total investment annually, includes cleaning for biofouling on heat exchangers (which can reduce efficiency by 20-50% if unchecked), corrosion monitoring, and labor in remote tropical sites. Unlike fossil fuels, OTEC incurs no fuel costs, but high OPEX stems from specialized maintenance for submerged components and low plant availability (targeting 80-90%) due to mechanical wear. Recent estimates for a 10 MW closed-cycle floating plant peg OPEX at $10-15 million per year, influenced by site-specific factors like wave exposure and access to ports. Levelized cost of energy (LCOE) analyses reveal OTEC's economic hurdles, with recent projections ranging from $0.145 to $0.63/kWh for utility-scale plants, far exceeding solar PV ($0.03-0.05/kWh) or onshore wind ($0.03-0.06/kWh). For a 10 MW closed-cycle plantship using off-the-shelf components, LCOE falls to $0.37-0.46/kWh under concessionary financing, but scales poorly below 100 MW due to fixed infrastructure costs. Sensitivity analyses indicate 20% CAPEX reductions via modular designs or learning curves could lower LCOE by 15-25%, yet persistent challenges like pipe deployment (costing $5-10 million/km) and efficiency caps (3-7% Carnot-limited) hinder competitiveness. A 400 MW plant might achieve $0.116/kWh by 2045 with aggressive cost cuts, but empirical pilots confirm upfront investments of $200-400 million/MW, underscoring scalability barriers over baseload promise.

Efficiency Metrics and Comparisons

The thermal efficiency of ocean thermal energy conversion (OTEC) systems is constrained by the modest temperature differential between warm surface seawater (typically 25–30°C) and cold deep seawater (4–8°C), yielding a ΔT of 20–25°C in optimal tropical sites. The theoretical maximum efficiency follows the Carnot limit, η_C = 1 - (T_cold / T_hot), where temperatures are in Kelvin; for these conditions, η_C approximates 6–8%. Practical efficiencies are substantially lower, ranging from 1–3% in early experimental plants to 3–5% in advanced closed-cycle configurations, primarily due to irreversibilities in heat exchangers, turbine inefficiencies, and high parasitic loads from seawater pumps that can consume 20–30% of gross output. Net power efficiency, accounting for pumping and auxiliary losses, further reduces output; for instance, optimization studies indicate that evaporator-condenser heat transfer coefficient ratios critically influence this metric, with balanced designs maximizing net work at around 2–4% of thermal input. Cycle variants, such as closed Rankine (using ammonia or low-boiling fluids) versus open cycles, show marginal differences, with closed systems achieving slightly higher efficiencies (up to 4%) through better containment of working fluids but at the cost of added complexity. In comparisons to other renewables, OTEC's thermal-to-electric efficiency lags behind solar photovoltaics (15–22% module efficiency) and wind turbines (35–45% aerodynamic efficiency, with 30–50% capacity factors), reflecting its reliance on low-grade heat rather than direct mechanical or photonic conversion. However, OTEC surpasses intermittent sources in dispatchability, enabling baseload operation with capacity factors over 90%, unlike solar or wind, and competes with geothermal plants (10–20% efficiency) in continuous output potential, though at higher upfront scaling demands. Relative to fossil fuel combined-cycle plants (50–60% efficiency), OTEC's low conversion rate elevates levelized costs, but its fuel-free nature mitigates long-term variability in fuel prices.
TechnologyTypical Efficiency (%)Key Attribute
OTEC3–5Baseload, continuous
Solar PV15–22Intermittent, scalable
Wind35–45 (aerodynamic)Variable capacity factor
Geothermal10–20Baseload, site-limited
Combined Cycle Gas50–60Fuel-dependent

Commercialization Barriers

High capital costs represent the foremost economic barrier to OTEC commercialization, driven by the need for massive heat exchangers, cold water pipes, and robust offshore or platform structures to handle corrosive seawater environments. For a 10 MW closed-cycle plant, estimates range from $21,606 to $27,012 per kW, with installation alone comprising up to 39% of total expenses due to the scale of components and logistical demands in deep-water sites. Scaling to 50 MW can reduce costs to $11,223–$16,578 per kW through economies of scale, yet upfront investments remain prohibitive without subsidies or concessional financing, as operational lives extend to 40 years while power purchase agreements typically span only 15–20 years. Low thermodynamic efficiency exacerbates these cost issues, with OTEC cycles yielding 3–7% net efficiency owing to modest ocean temperature gradients of 20–25°C, far below those of conventional power plants. This inefficiency demands enormous seawater throughput—millions of kilograms per second—for modest power outputs, inflating pumping energy losses and operational expenditures that can consume 30–40% of gross power. Resulting levelized costs of electricity (LCOE) for a 10 MW plant hover at $0.37–$0.46/kWh under favorable financing, dropping to $0.19–$0.28/kWh for larger 50 MW units at optimal sites with 24.5°C differentials, but still exceeding unsubsidized solar or wind LCOE by factors of 2–5. Technological uncertainties compound economic risks, as no utility-scale OTEC plants operate commercially, limiting empirical data on durability, maintenance, and performance degradation from factors like biofouling and pipe fatigue. Cold water pipe deployment poses particular engineering hurdles, with historical prototypes failing under dynamic ocean loads, necessitating costly reinforcements and unproven materials for depths exceeding 1,000 meters. The nascent supply chain for specialized components further elevates risks, while site-specific requirements—proximity to deep thermoclines and shore-based grids—restrict viable locations to tropical regions, complicating permitting and transmission infrastructure. Historical periods of low fossil fuel prices, such as 1984–2009, stalled momentum by undermining incentives for investment in this baseload alternative. Financing barriers persist due to perceived high risks, with developers reliant on government-backed loans, carbon credits ($0.0048–$0.025/kWh), or co-product revenues from desalination to achieve viability; for instance, open-cycle systems require desalinated water sales at $1.50/m³ to offset LCOE exceeding $0.62/kWh. Knowledge gaps in holistic economic modeling, including location variability, learning curves, and integrated systems, hinder investor confidence, as most analyses focus on isolated plants rather than fleets or hybrids that could amortize costs. Despite potential for cost reductions via technological maturation and multi-use platforms, the absence of first-of-a-kind demonstrations perpetuates a cycle of underfunding and delayed commercialization.

Environmental Considerations

Local Marine Ecosystem Impacts

Ocean thermal energy conversion (OTEC) facilities pump substantial volumes of seawater—typically on the order of 10 to 100 cubic meters per second for megawatt-scale plants—through intake pipes, raising concerns about entrainment and impingement of marine organisms, particularly plankton, fish eggs, and larvae. Early life stages are most vulnerable, as intake velocities can draw in and damage or kill entrained biota before they pass through screens or heat exchangers, though the overall population-level effects are projected to be negligible for small-scale operations due to the ocean's high reproductive rates and vast plankton densities. Modeling for a hypothetical 10 MW OTEC plant off Oahu, Hawaii, suggests ichthyoplankton mortality from warm surface water entrainment would remain acceptable, comparable to impacts from existing coastal power plant intakes. Discharge of cold deep seawater, often nutrient-enriched compared to surface waters, forms plumes that locally alter temperature profiles and introduce macronutrients like nitrates and phosphates into the euphotic zone, potentially stimulating phytoplankton growth and altering primary productivity. In a Kona, Hawaii, OTEC test, deep water addition to surface waters showed a six-day lag before phytoplankton response, indicating delayed but possible blooms if mixing persists in sunlit depths; however, rapid dilution in open waters typically limits sustained eutrophication. These nutrient inputs could enhance local food webs by boosting zooplankton and fish biomass, though excessive upwelling might disrupt thermoclines and sensitive pelagic communities adapted to stratified conditions. Cold plumes may also create thermal refugia or barriers, influencing migration patterns of reef-associated species. Biofouling on OTEC pipes, heat exchangers, and platforms—driven by microbial films, algae, and invertebrates—poses operational challenges but can foster artificial habitats akin to reefs, attracting sessile organisms and serving as fish aggregation devices (FADs). In Hawaii's Natural Energy Laboratory of Hawaii Authority (NELHA) facilities, including the operational 105 kW closed-cycle plant by Makai Ocean Engineering since 2016, submerged structures have correlated with elevated fish abundances nearby, suggesting neutral to positive localized effects without reported significant disruptions to surrounding ecosystems. Nonetheless, biocide use to mitigate fouling could introduce trace contaminants, though pilot-scale applications have shown minimal residue dispersion. Empirical data from Hawaii pilots indicate no major adverse impacts on local biodiversity or fisheries to date, with monitoring emphasizing the need for continued observation of plume dispersion and entrainment rates as scales increase. Large-scale deployments could amplify cumulative effects, such as regional nutrient shifts or habitat fragmentation, warranting site-specific assessments informed by hydrodynamic models rather than generalized predictions. Overall, OTEC's localized disturbances appear overshadowed by potential productivity gains in nutrient-limited tropics, provided intake designs incorporate velocity caps and fine-mesh screens to minimize entrainment.

Broader Climatic Effects

Large-scale deployment of (OTEC) systems could induce regional surface cooling in tropical waters by upwelling cold deep seawater, potentially altering local heat fluxes to the atmosphere and influencing evaporation rates and cloud formation patterns. Modeling studies indicate that such cooling, combined with enhanced vertical mixing, might augment the oceanic , thereby redistributing heat and salinity on basin scales with implications for global ocean dynamics. However, these circulation changes could also risk disrupting established patterns if OTEC operations exceed certain thresholds, though quantitative thresholds remain uncertain due to model sensitivities. In simulations of widespread OTEC implementation at multi-gigawatt capacities, tropical surface waters exhibit cooling while higher latitudes experience relative warming, resulting in a net oceanic heat intake that partially offsets atmospheric warming from displaced fossil fuel emissions. Relative to high-emissions scenarios without OTEC, deployments are projected to reduce global surface warming by enhancing ocean heat uptake and lowering greenhouse gas emissions, with sustained cooling effects persisting post-deployment from prior mixing and reduced atmospheric CO2. These outcomes hinge on OTEC scaling to terawatt levels, far beyond current pilot capacities, and assume efficient integration without unintended feedbacks like altered phytoplankton productivity affecting carbon sequestration. Empirical data on climatic effects are limited, as operational OTEC facilities remain small-scale (e.g., under 100 kW), precluding direct observation of broader impacts; thus, assessments rely on coupled ocean-atmosphere general circulation models, which vary in resolution and parameterization of mixing processes. Potential indirect effects, such as modifications to tropical cyclone genesis via stabilized surface stratification, have been hypothesized but lack robust quantification across studies. Overall, while OTEC's emissions displacement offers clear climatic benefits, physical ocean alterations pose risks that necessitate site-specific modeling for equatorial deployments.

Empirical Studies and Data

Empirical studies on OTEC environmental impacts derive primarily from monitoring programs at small-scale pilot facilities, as large commercial deployments remain absent. The at the Keahole Point OTEC plant in Hawaii, operational since the early 1990s, has collected quarterly data on temperature, salinity, nutrients, oxygen, pH, and other water chemistry parameters across surface, mid-water, and bottom transects. These observations indicate rapid dilution of mixed effluent plumes, with nitrate concentrations decreasing from intake levels of approximately 16 µmol/kg to 2-4 µmol/kg within 30-40 meters of discharge, depending on release depth (50-100 m). Entrainment surveys at the facility documented low incidences of organisms, such as Gonostomatidae fish (10 individuals) and Acanthephyra shrimp (43 individuals) over two months of sampling, suggesting minimal impingement for deep-water species due to intake screening and low surface biota densities offshore. Benthic community assessments from hybrid OTEC (H-OTEC) pilots provide direct ecological data. A 2023-2024 seasonal study near a Malaysian H-OTEC discharge outlet measured macro- and meiobenthic abundances and structures at 15 sites, revealing localized reductions: macrobenthos density at 113 ind./m² near discharge versus 952 ind./m² farther away during inter-monsoon periods, and meiobenthos at 314 ind./10 cm² near versus 1,207 ind./10 cm² distant. Temperature drops of 7-10°C occurred within 5 m of the outlet, but anomalies were below 0.3°C beyond, correlating with abundance declines and shifts in community composition (e.g., dominance of gastropods at 77% in meiofauna), though diversity indices (Shannon, Pielou, Margalef) showed no significant differences (P > 0.05). These effects were confined to immediate vicinity, with seasonal variations but no evidence of broader disruption. Nutrient upwelling effects have been tested empirically in , where mixing 50% deep cold water with surface water induced a 6-day lag in phytoplankton growth, attributed to initial low light adaptation despite elevated nutrients like and . Fisheries-focused monitoring from 1980s U.S. pilots, including , found no substantial alterations to pelagic or fish populations, with potential benefits from localized productivity enhancements outweighed by entrainment risks estimated below 1% for larvae in modeled scenarios validated against data. Broader climatic data remains absent empirically, as pilot scales (e.g., 100 kW in ) produce negligible heat extraction; 1980s assessments classified thermal discharges as minor, with no observed long-term or circulation changes beyond plume zones. Overall, available data from these facilities indicate localized physical and biological perturbations—primarily from cold-water plumes and injection—but no systemic degradation, supporting acceptability for scaled operations pending validation through expanded monitoring. Gaps persist in deep-slope and cumulative multi-plant effects, with protocols recommending enhanced tows and acoustic surveys for future sites.

Technical Challenges

Material and Biofouling Issues

Ocean thermal energy conversion (OTEC) systems expose structural components, particularly heat exchangers, to aggressive seawater environments characterized by high salinity, dissolved oxygen variations, and temperature gradients between warm surface water (typically 25–30°C) and cold deep water (4–10°C). These conditions promote corrosion mechanisms such as pitting, crevice corrosion, and galvanic degradation, necessitating materials with high resistance to localized attack and biofouling-induced under-deposit corrosion. Candidate alloys include titanium, aluminum bronzes, copper-nickel, and stainless steels, but titanium alloys exhibit the lowest corrosion rates in both warm and cold seawater, with general corrosion below 0.001 mm/year and minimal pitting under natural conditions. Aluminum alloys, while lighter and cheaper, suffer higher dissolution rates (up to 0.1 mm/year in warm water) and require protective cladding or coatings for seawater-side exposure. Heat exchanger design prioritizes for seawater-facing surfaces to mitigate corrosion-fatigue from cyclic stresses and pressure differentials, as demonstrated in pilot plants where plate-and-frame or shell-and-tube configurations achieved over 20 years of projected service life without significant degradation. Stainless steels (e.g., 316L) and Cu-Ni alloys show acceptable performance in cold water but accelerated in warm, oxygen-rich conditions unless crevices are minimized through welding techniques or . Material costs remain a barrier, with comprising up to 50% of expenses in closed-cycle OTEC designs, prompting into hybrid aluminum- finned structures for enhanced while preserving corrosion resistance. Biofouling compounds material challenges by forming biofilms, , , and tube worms on surfaces, reducing coefficients by 20–50% within months and increasing hydrodynamic drag, which elevates pumping power demands by up to 25%. Microfouling (slime layers from and diatoms) initiates rapid deposition in warm inlets, while macrofouling dominates in tropical sites, with growth rates exceeding 1 mm/month for sessile organisms. In OTEC cold water pipes and evaporators, differential between warm and cold sides exacerbates thermal inefficiencies and localized under fouling deposits, as observed in Hawaii-based tests where untreated surfaces required every 3–6 months. Mitigation strategies include intermittent chlorination (0.1–1 ppm free for 1–2 hours daily), which suppresses larval settlement by 80–90% but risks forming disinfection byproducts and harming if overdosed. brushing or ball cleaning restores surfaces but increases operational downtime, while non-chemical methods like low-voltage electrolytic generation of or ultrasonic treatments show promise in reducing macrofouling by 70% without residues, as tested in prototypes. Antifouling coatings (e.g., silicone-based or copper-releasing) provide short-term protection (6–12 months) but degrade under shear stresses, and their efficacy diminishes in deep-water flows. Integrated designs, such as enhanced surface geometries in exchangers, aim to minimize fouling adhesion while maintaining integrity, though long-term field data from operational plants like Makai Ocean Engineering's 2016 facility indicate ongoing needs for adaptive cleaning protocols.

Energy Losses and Scaling Problems

The thermodynamic efficiency of ocean thermal energy conversion (OTEC) systems is inherently constrained by the Carnot limit, which for typical differentials of 20–25°C ( at approximately 25–30°C and deep water at 4–5°C) yields a maximum of 6–8%. Actual thermal efficiencies range from 3–5%, substantially below this limit due to irreversibilities in processes, including finite differences across heat exchangers and non-ideal working fluid behavior. These losses arise from the second-law inefficiencies in the variants used in closed-cycle OTEC, where entropy generation during and reduces the work extractable from the input. Parasitic energy losses further diminish net power output, with seawater pumping representing the dominant component because of the massive flow rates required—often exceeding 1 m³/s per MW of gross power to maintain adequate heat transfer. Pump efficiencies, typically 80% hydraulic and 95% electrical, combined with frictional head losses in large-diameter pipes and heat exchangers, can consume 20–50% of gross turbine output in pilot-scale plants, though this fraction decreases with scale. Ammonia circulation pumps impose minimal additional losses, but seawater-side pressure drops in heat exchangers exacerbate pumping demands, as higher friction factors, while enhancing heat transfer coefficients, increase required power. Heat exchanger designs optimized for OTEC, such as plate-fin or shell-and-tube configurations, must balance these trade-offs, yet empirical tests show that biofouling and scaling further degrade performance over time, amplifying losses. Scaling OTEC plants to commercial capacities (e.g., 100 MW or greater) intensifies these issues, as the required cold water pipe lengths (1,000 m or more) and diameters (several meters) lead to disproportionate hydrodynamic drag and structural stresses, elevating pumping losses nonlinearly with . While larger plants benefit from economies in relative parasitic loads—potentially reducing them to 10–20% of gross output—the absolute volumes of water handled (thousands of m³/s globally for widespread deployment) demand robust solutions, including for floating platforms, which introduce draws from generators and systems. Low , stemming from the modest , necessitates expansive infrastructure, where incremental scaling reveals : for instance, simulations indicate that beyond 50 MW, additional capacity gains are offset by heightened transmission and intake losses unless mitigate pipe inefficiencies. These factors contribute to net efficiencies often below 2% in projected large-scale systems, underscoring the causal link between thermodynamic constraints and practical deployment hurdles.

Operational Reliability Factors

Operational reliability in ocean thermal energy conversion (OTEC) systems is critical for their viability as baseload power sources, given the continuous nature of the ocean thermal gradient, but it is influenced by marine environmental stresses and component vulnerabilities. Prototypes, such as those at Hawaii's Natural Energy Laboratory (NELHA), have demonstrated high seawater system uptime exceeding 99.9% in recent years, supporting potential for reliable operation through redundancy and monitoring. However, factors like , , and on the cold water pipe (CWP) necessitate robust mitigation strategies to achieve commercial-scale availability. Temperature variations in surface and deep seawater directly impact reliability by altering power output and accelerating component failure rates. Seawater temperatures ranging from 20–30°C at the surface and approximately 4°C in deep layers modulate the thermal efficiency and stress on pumps, turbines, and generators, with higher temperatures increasing failure probabilities via models like the Arrhenius law for electrical components. Monte Carlo simulations integrated into power system reliability assessments, using hourly data from Hawaiian sites, show that OTEC incorporation can enhance overall grid indices like loss of load expectation (LOLE), but reliability declines under rising demand or ΔT fluctuations below 20°C, emphasizing the need for site-specific thermal profiling. Biofouling poses a primary operational challenge by accumulating on heat exchanger surfaces and pipes, reducing heat transfer coefficients by up to 50% within months and elevating pumping power demands due to increased friction. In OTEC heat exchangers, which operate across small ΔT gradients of 3–4°C, this fouling—driven by microbial and macro-organism growth in warm seawater—requires preventive measures like continuous low-level chlorination (0.1–0.5 ppm) or mechanical cleaning, with modular plate designs allowing partial replacement to minimize downtime. Empirical tests at NELHA confirm effective control via coatings and biocides, but unchecked growth can necessitate full system shutdowns for descaling, impacting availability. Corrosion from saline, oxygenated accelerates degradation in metallic components, particularly in evaporators and condensers, where pitting and can compromise structural integrity over 20–30 years. Materials such as or aluminum with exhibit low corrosion rates (<0.01 mm/year), as validated in long-term NELHA exposures, but systems demand vigilant via fiber-optic sensors to prevent leaks that could halt operations. Pumps and turbines, often or composite, achieve high reliability through , with historical data indicating inspection intervals of 5 years initially, transitioning to annual checks for sustained uptime. The CWP, typically 800–1,200 m long and 10–30 m in diameter for multi-megawatt plants, represents a high-risk element due to vortex-induced vibrations, currents, and storm loads, potentially leading to or . Detachable interfaces and fiberglass-reinforced designs mitigate deployment and maintenance issues, drawing from precedents, but resonance modes must be damped to avoid ; simulations indicate survival in extreme events via flexible . Operational pilots, including Makai Ocean Engineering's 105 kW plant grid-connected since 2015, underscore that while simplicity yields high equipment uptime, CWP integrity demands pre-operational dynamic modeling for reliability exceeding 95%.

Co-Products and Integrated Uses

Desalination and Water Production

Open-cycle ocean thermal energy conversion (OTEC) systems inherently produce desalinated water as a byproduct of electricity generation. Warm surface seawater, typically at 25–30°C, is evacuated to low pressure and flash-evaporated into steam, which expands through a low-pressure turbine to generate power. This steam, derived from seawater, is subsequently condensed using cold deep ocean water (around 4–5°C), resulting in fresh water distillate with salinity reduced to potable levels, often below 50 ppm total dissolved solids. Water production rates vary with plant capacity and configuration, but integrated systems demonstrate substantial output. For example, an integrated OTEC and can yield approximately 2.28 million liters of potable water per day per megawatt of gross generated, leveraging the thermal gradient for both and without additional energy-intensive processes like . Hybrid OTEC cycles, combining open- and closed-cycle elements, further optimize co-production; a 1 MW hybrid converts vapor to alongside , with cold discharge facilitating condensation efficiency. Demonstration projects illustrate practical implementation. In 2005, established a low-temperature thermal desalination pilot on Island using OTEC principles, producing fresh water from deep-sea cold sources to address island . Similarly, Japan's Saga University operated a flash-type facility on Kumejima Island starting in 2015, integrating OTEC-derived cold water for evaporation-condensation cycles. These efforts highlight OTEC's potential for dual-output in regions with persistent thermal gradients, such as , where simultaneous power and water generation enhances resource efficiency over standalone methods. Closed-cycle OTEC plants, using a secondary like , can incorporate via auxiliary processes such as multi-stage powered by generated electricity or , though yields are generally lower than in open cycles due to indirect integration. Overall, OTEC capitalizes on abundant thermal resources for base-load production, with empirical data from pilots confirming viability in tropical locales despite scaling challenges.

Aquaculture and Biomass Applications

OTEC systems upwell cold (DOW), typically from depths of 600–1000 meters where temperatures range from 4–6 °C, providing a sterile, nutrient-rich medium low in pathogens and that supports in tropical environments. This water enables temperature regulation for intolerant of warm surface conditions, promotes rapid growth via elevated and levels, and minimizes outbreaks due to its purity. At the Natural Energy Laboratory of Hawaii Authority (NELHA) in Kailua-Kona, Hawaii, DOW piped from approximately 900 meters depth has facilitated commercial aquaculture since the 1990s, including production of amberjack (Seriola dumerili, known locally as kanpachi), abalone (Haliotis spp.), bivalves, and marine ornamental fish. Tenants such as Blue Ocean Mariculture and Ocean Era utilize this water for open-ocean cage farming, yielding premium fish products with reduced antibiotic needs owing to the water's quality. NELHA's infrastructure, which supports OTEC pilots, hosts over 40 aquaculture operations producing microalgae, seahorses, and prawns alongside energy research. In , the Kumejima OTEC demonstration facility, operational since 2013, integrates DOW utilization for under the "Kumejima Model," which combines energy production with . Post-OTEC cold water supports farming of kuruma prawns ( japonicus) and cold-soil , leveraging the water's stability to cultivate temperate species in subtropical conditions. This approach has driven local , with DOW enabling year-round prawn production and nutrient-enhanced feeds. The same nutrient upwelling in OTEC processes enhances biomass production through macroalgal cultivation, as DOW's high dissolved inorganic nutrients—often 10–20 times surface levels—fertilize growth for food, biofuels, or . In (IMTA) systems, OTEC effluent boosts macroalgae yields by supplying bacteria-free nutrients, with pilot models demonstrating increased productivity of species like or spp. for biomass harvesting. Research on deep seawater applications shows sustainable seaweed production rates improved by OTEC-derived DSW, with studies reporting enhanced quality and quantity for species such as Eucheuma denticulatum, suitable for extraction or . These applications mitigate limitations in surface waters, potentially scaling to offset fossil fuel-derived while co-generating OTEC power.

Other By-Products like

Ocean thermal energy conversion (OTEC) systems generate cold deep as a primary , which can be applied to and beyond processes. This nutrient-rich water, drawn from depths exceeding 600 meters at temperatures of 4–10°C, provides a natural coolant for systems in tropical coastal areas, potentially offsetting up to 90% of conventional energy use in facilities like hotels or data centers. In Hawaii's OTEC pilot operations during the , such cold water supported experimental loads equivalent to several megawatts thermal, demonstrating scalability for energy-efficient cooling in island economies. Integration of OTEC electricity with water enables as a co-product, leveraging the baseload power for in equatorial regions where solar insolation drives persistent thermal gradients. A 2025 techno-economic assessment of a 10 MW hybrid OTEC plant coupled to electrolyzers projects levelized costs of $3.50–5.00 per kg, competitive with offshore wind-based systems when factoring in OTEC's 24/7 availability and co-located feedstock. This approach avoids issues of variable renewables, with efficiency reaching 65–70% in modeled OTEC hybrids, producing up to 1,500 tons of H₂ annually per 10 MW unit under optimal conditions of 20–25°C. Earlier concepts from the explored OTEC-derived ammonia synthesis or decomposition for , but modern designs prioritize direct due to material advancements and reduced parasitic loads. In tropical deployments, such as proposed seafloor systems at 3,400 m depth, OTEC could power distributed electrolyzers yielding for or export, with cold by-product water aiding to boost . These applications enhance OTEC's economic viability, though deployment remains limited by upfront exceeding $5,000 per kW installed.