Fact-checked by Grok 2 weeks ago

Engine efficiency

Engine efficiency is a measure of how effectively an engine converts the in into useful work, defined as the ratio of the work output to the total input from the , often expressed as a . In thermodynamic terms, it represents the fraction of supplied that is transformed into work, with the remainder lost as or other inefficiencies. The thermal efficiency of heat engines, which form the basis for most internal combustion and external combustion engines, is fundamentally constrained by the second law of thermodynamics, with the maximum possible efficiency given by the Carnot formula: \eta = 1 - \frac{T_c}{T_h}, where T_c and T_h are the absolute temperatures of the cold and hot reservoirs, respectively. Real engines, operating on cycles like the Otto or Diesel, achieve lower efficiencies due to irreversibilities such as friction, incomplete combustion, and heat losses, typically ranging from 20% to 40% for automotive engines. Overall engine efficiency is a product of several components, including combustion efficiency (the completeness of fuel burning), thermodynamic efficiency (conversion of heat to work), gas exchange efficiency (work associated with intake and exhaust), and mechanical efficiency (losses to friction and auxiliaries). Key factors influencing engine efficiency include the , fuel properties, operating temperature, and cycle design; for instance, higher compression ratios in the improve efficiency according to \eta = 1 - \frac{1}{r^{\gamma-1}}, where r is the compression ratio and \gamma is the specific heat ratio. Practical limits for advanced internal combustion engines are estimated at around 60% brake , constrained by material durability and irreversible losses, though engines have reached up to 53% as of 2024, for example in heavy-duty engines developed by using advanced technologies including elements of (HCCI). Improving efficiency is crucial for reducing fuel consumption, lowering , and enhancing , driving innovations such as recovery and advanced turbocharging.

Fundamentals

Definition and Importance

Engine efficiency refers to the of useful work output to the total input from , typically expressed as a percentage. This is commonly quantified as , where the work output is the mechanical power delivered (brake thermal efficiency, or BTE), and the input is the content of the , often measured using its lower heating value (LHV). For instance, in internal combustion engines, BTE accounts for losses due to , pumping, and incomplete , distinguishing it from indicated thermal efficiency, which measures only the work done within the cylinders during the and strokes. In practical terms, engine efficiency is governed by thermodynamic principles but limited by irreversible processes such as to and exhaust, mechanical friction, and inefficiencies. As of 2024, modern engines achieve peak BTEs of around 35-45%, while engines can reach 40-50% typically, with records up to 53% under optimal conditions, far below the theoretical Carnot limit but representing significant advancements from early designs that operated at under 10%. These metrics highlight how captures the conversion of fuel's —the maximum useful work potential—into , with the remainder dissipated as . The importance of engine efficiency lies in its direct impact on fuel economy, environmental sustainability, and economic viability in transportation, which accounts for approximately 60% of global oil consumption. Higher efficiency reduces fuel consumption and greenhouse gas emissions; for example, a 1% improvement in light-duty vehicle engine efficiency can cut CO2 emissions by millions of tons annually while lowering operational costs for consumers and fleets. In heavy-duty applications, efficiency gains of 30% or more through advanced designs could enhance commercial vehicle fuel economy by up to 50%, supporting energy security and compliance with stringent regulations like those from the U.S. Environmental Protection Agency. Moreover, as engines evolve to integrate electrification and alternative fuels, prioritizing efficiency remains key to extending the viability of internal combustion technologies amid the transition to lower-carbon systems.

Thermal Efficiency Metrics

Thermal efficiency in engines quantifies the fraction of fuel's converted into useful work, typically expressed as a . It serves as a primary for evaluating performance and utilization, encompassing various subtypes that account for different stages of energy conversion losses. Indicated thermal efficiency measures the work produced within the relative to the input from , focusing on the without accounting for losses. It is calculated as the ratio of indicated (the work done on the during the power stroke, derived from pressure-volume diagrams) to the total supplied by the , often using the formula: \eta_{i} = \frac{W_{i}}{Q_{in}} = \frac{\text{Indicated Work Output}}{\text{Fuel Energy Input}} where W_i is the indicated work and Q_{in} is the heat input, typically m_f \times \text{LHV}, with m_f as fuel mass flow rate and LHV as the lower heating value. This efficiency is higher than overall engine efficiency because it excludes friction and pumping losses, and in modern diesel engines, it can approach 45-50% under optimal conditions due to effective combustion and cycle design. Brake thermal efficiency, the most commonly reported metric for practical engine performance, represents the ratio of brake power (measurable output at the crankshaft) to the fuel's energy input, incorporating all major losses including mechanical friction. It is defined by: \eta_{b} = \frac{W_{b}}{Q_{in}} = \frac{\text{Brake Power}}{\text{Fuel Energy Input}} where W_b is the brake work output. Brake thermal efficiency is related to indicated thermal efficiency through mechanical efficiency (\eta_m = W_b / W_i), such that \eta_b = \eta_i \times \eta_m, with \eta_m typically ranging from 80-90% in well-designed engines. As of 2024, for spark-ignition engines, brake thermal efficiency generally falls between 30-40%, while compression-ignition engines achieve 35-50%, reflecting differences in combustion processes and compression ratios. The overall brake thermal efficiency can be decomposed into component efficiencies to isolate specific loss mechanisms: \eta_b = \eta_c \times \eta_{th} \times \eta_{ge} \times \eta_m, where \eta_c is combustion efficiency (fraction of fuel energy released , often >95% in stoichiometric mixtures), \eta_{th} is thermodynamic efficiency (governed by the cycle's compression and expansion, limited by the Carnot principle but practically 50-60% in ideal Otto or Diesel cycles), \eta_{ge} is gas exchange efficiency (accounting for pumping work during intake and exhaust, typically 95-98%), and \eta_m is mechanical efficiency as noted above. This breakdown highlights how inefficiencies arise cumulatively, with combustion and thermodynamic factors dominating in high-efficiency designs. Volumetric efficiency, while not a direct metric, influences by determining the air- charge inducted per cycle, affecting completeness and . It is defined as the ratio of inducted to the displaced , often enhanced to 100-150% via turbocharging, thereby improving overall performance by optimizing utilization.

Thermodynamic Principles

Carnot Limit

The represents the ideal reversible thermodynamic cycle for a operating between two thermal reservoirs at temperatures T_h (hot) and T_c (cold), establishing the fundamental upper limit on dictated by the second law of thermodynamics. Proposed by Sadi Carnot in , this cycle consists of two isothermal processes and two adiabatic processes, with no net change, making it the most efficient possible configuration for converting heat into work. Any real operating between the same temperatures cannot exceed this , as irreversibilities in practical processes lead to generation and reduced performance. The Carnot efficiency \eta_C is given by the formula: \eta_C = 1 - \frac{T_c}{T_h} where temperatures are in absolute units (). This expression shows that efficiency increases with the temperature difference between the reservoirs, approaching 100% as T_c approaches , though practical constraints limit achievable differences. For example, in a typical context with a peak of about 2000 and an exhaust of around 1000 , the Carnot limit would be approximately 50%, serving as a theoretical . In the context of engine efficiency, particularly for internal combustion engines, the Carnot limit provides a conceptual upper bound but does not directly constrain performance, as these engines operate on open cycles involving chemical and mass exchange rather than closed, steady-state between fixed reservoirs. Unlike ideal Carnot engines, internal combustion processes feature transient , incomplete heat addition at constant maximum temperature, and no mandatory heat rejection for thermodynamic reasons—cooling serves primarily to protect materials. Consequently, while engines achieve thermal efficiencies of 25-40% against a notional Carnot limit of 35-40% based on average operating temperatures, their limits stem more from irreversibilities like , incomplete , and losses than from the Carnot bound itself. As of 2025, advanced designs have reached up to 45-48% in prototypes. Seminal analyses, such as those in Heywood's Internal Combustion Engine Fundamentals, emphasize that optimizing cycle parameters can approach but never surpass this thermodynamic ceiling in idealized models.

Practical Thermodynamic Cycles

Practical thermodynamic cycles model the operation of real engines by simplifying complex processes into idealized steps, typically assuming air as an with constant specific heats. These cycles deviate from the reversible by incorporating irreversibilities such as constant-volume or constant-pressure heat transfer, which limit achievable efficiencies. Key examples include the for spark-ignition engines, the for compression-ignition engines, and the for gas turbines, each optimized for specific engine types while balancing efficiency, power output, and practical constraints like material limits and dynamics. The consists of four processes: isentropic compression from intake to top dead center, constant-volume heat addition during spark-ignition , isentropic expansion to bottom dead center, and constant-volume heat rejection during exhaust. This cycle's depends primarily on the r = V_1 / V_2, where V_1 and V_2 are the volumes at bottom and top dead center, respectively. The ideal air-standard is given by \eta_{\text{Otto}} = 1 - \frac{1}{r^{\gamma - 1}} where \gamma is the specific heat ratio (approximately 1.4 for air). For a typical automotive compression ratio of r = 8, the ideal efficiency reaches about 56.5%, though real engines achieve 25-40% due to losses, , and incomplete . Compared to the Carnot efficiency between the same limits (e.g., 85% for low 300 K and high 2000 K), the Otto cycle is less efficient because addition and rejection occur at varying temperatures rather than isothermally. The modifies the for compression-ignition engines by replacing constant-volume addition with constant-pressure combustion, allowing higher compression ratios (typically 14-20) without knocking. Its processes are: isentropic compression, constant-pressure addition as fuel injects and burns, isentropic expansion, and constant-volume rejection. The incorporates the cutoff r_c = V_3 / V_2, the volume during addition: \eta_{\text{Diesel}} = 1 - \frac{1}{r^{\gamma - 1}} \cdot \frac{r_c^\gamma - 1}{\gamma (r_c - 1)}. For r = 18 and r_c = 2, the ideal efficiency is approximately 63%, with practical values of 40-50% in heavy-duty engines as of 2025, and some exceeding 50% with advanced designs, outperforming cycles at equivalent compression ratios due to leaner mixtures and reduced heat rejection at lower temperatures. This makes Diesel cycles preferable for applications prioritizing fuel economy over , though they remain below Carnot limits owing to non-isothermal processes. The underpins continuous-flow engines, featuring isentropic compression in a , constant-pressure heat addition in a , isentropic expansion through a , and constant-pressure heat rejection in the exhaust. Unlike reciprocating cycles, it operates in a steady-state open system, with depending on the ratio r_p = P_2 / P_1: \eta_{\text{Brayton}} = 1 - \frac{1}{r_p^{(\gamma - 1)/\gamma}}. For ratios of 30-40 common in modern , ideal approach 50-60%, but actual values are 35-40% due to component inefficiencies like and polytropic around 85-90%. The 's increases with higher r_p and inlet temperatures (up to 1900 K with advanced cooling as of 2025), yet it falls short of Carnot because heat transfer occurs at constant rather than isothermally, and work consumes a significant portion (back-work ratio of 40-60%) of output. This excels in high-power, low-weight applications like but requires to approach theoretical limits.

Influencing Factors

Compression Ratio

The (CR) in a reciprocating is defined as the ratio of the volume when the is at bottom dead center (maximum volume) to the volume when the is at top dead center (minimum volume). This geometric parameter directly influences the engine's by determining the pressure and temperature rise during the compression stroke, which in turn affects the conversion of heat energy from into mechanical work. Higher CR values compress the air-fuel mixture more effectively, leading to greater of combustion gases and improved work extraction, but practical limits arise from material strength, , and combustion characteristics. In spark-ignition () engines, modeled by the ideal , is fundamentally tied to through : \eta = 1 - \frac{1}{r^{\gamma - 1}} where r is the and \gamma is the ratio of specific heats (approximately 1.4 for air-standard conditions). This formula demonstrates that increases monotonically with , as higher compression reduces the heat rejected during the exhaust relative to the added during , approaching theoretical limits but never exceeding the Carnot efficiency. For instance, at r = 10 and \gamma = 1.4, the ideal Otto reaches about 60%, though real engines achieve 30-40% due to irreversibilities. However, engines are constrained to values of 8-12 to avoid knocking (autoignition), which can damage components; higher-octane fuels or advanced designs like mitigate this but do not eliminate the limit. In compression-ignition (CI) diesel engines, governed by the , the formula incorporates both and the cutoff ratio \rho (the ratio of volumes at the end and start of addition): \eta = 1 - \frac{1}{r^{\gamma - 1}} \cdot \frac{\rho^\gamma - 1}{\gamma (\rho - 1)} where higher amplifies by elevating pre-combustion temperatures for reliable ignition without . engines typically operate at of 14-22, yielding up to 65-70% and practical values of 40-53%, surpassing SI engines due to the absence of knock constraints. Nonetheless, excessive increases mechanical stresses, emissions from higher combustion temperatures, and pumping losses, necessitating trade-offs in design. For gas turbine engines under the , an analogous pressure ratio (related to volumetric ) follows a similar trend: \eta = 1 - \frac{1}{r_p^{(\gamma - 1)/\gamma}}, where modern turbines achieve ratios of 15-40 for around 35-43%. Overall, optimizing remains a cornerstone of engine efficiency improvements, with historical advancements like turbocharging effectively boosting effective compression while respecting physical limits. Seminal analyses, such as those in the derivations, underscore that CR enhancements can yield 1-2% efficiency gains per unit increase, though and real-world losses cap benefits.

and Losses

Mechanical and friction losses in engines refer to the energy dissipated due to mechanical interactions and within the engine, reducing the conversion of from into useful mechanical work. These losses primarily encompass rubbing from contacting surfaces, pumping losses associated with , and accessory drive requirements. Collectively, they contribute to the difference between indicated (IMEP), which represents the work done on the during the process, and brake (BMEP), the net output at the . The (FMEP) quantifies these losses as FMEP = IMEP - BMEP, typically accounting for 10-15% of the input in internal combustion engines. Rubbing friction losses arise from relative motion between components such as rings and cylinder walls, and main bearings, bearings, and the valve train. The assembly, including rings and , often dominates, contributing up to 40-50% of total losses due to high contact pressures and sliding velocities, with the piston-cylinder interface alone responsible for approximately 20% of overall . Bearings and valve train add further losses through hydrodynamic regimes, where viscous shear in the film dissipates energy. , defined as η_m = BMEP / IMEP_g (gross indicated), is thus lowered by these rubbing components, with FMEP often modeled empirically as tfmep = a + bN + cN², where N is engine speed, reflecting the quadratic increase in at higher speeds. Pumping losses occur during the intake and exhaust strokes, where the engine must work against throttling and flow restrictions to manage the working fluid, particularly pronounced in spark-ignition engines under part-load conditions. Accessory losses include power consumed by auxiliaries like oil and water pumps, alternators, and fans, which can represent 10-20% of total FMEP depending on engine load and speed. In a typical 2.0 L inline-four engine, these components—pumping, rubbing, and accessories—distribute such that rubbing friction forms the largest share at high speeds, while pumping dominates at low loads, collectively reducing overall thermal efficiency by 5-10 percentage points. Efforts to mitigate these losses focus on low-friction materials, optimized lubrication, and variable valve timing to enhance mechanical efficiency without compromising durability.

Combustion Process and Oxygen Utilization

The combustion process in internal combustion engines involves the of hydrocarbons with oxygen from intake air, converting into to drive motion or rotation. Under ideal conditions, this reaction produces (CO₂) and (H₂O), with all and oxygen fully consumed; however, real-world processes often feature incomplete reactions due to limitations in mixing, , and . Oxygen utilization refers to the extent to which available oxygen participates in the oxidation of molecules, directly influencing completeness and . Poor utilization leads to unburned and partial oxidation products like (CO) and hydrocarbons (HC), which reduce the energy extracted from the . The air-fuel ratio (A/F), defined as the mass of air to the mass of fuel in the , governs oxygen availability and is a primary determinant of . The stoichiometric A/F for is approximately 14.7:1, balancing fuel and oxygen for complete without excess reactants. This ratio is quantified using the air excess ratio λ = (actual A/F) / (stoichiometric A/F), or equivalently, the equivalence ratio φ = 1/λ = (actual fuel-air ratio) / (stoichiometric fuel-air ratio). At λ = 1 (φ = 1), oxygen utilization is theoretically optimal, achieving of 95–98% in spark-ignition engines by minimizing unburned species. analysis can measure these ratios and efficiencies, confirming that deviations from degrade performance. In rich mixtures (λ < 1, φ > 1), oxygen is insufficient relative to , resulting in oxygen-deficient zones where is incomplete; this causes a drop in efficiency as some escapes oxidation, forming CO and HC while elevating exhaust temperatures and emissions. For instance, during rich operation in spark-ignition engines, efficiency declines because the lack of oxygen prevents full burnout, with efficiency falling below 90% at φ > 1.2. Conversely, lean mixtures (λ > 1, φ < 1) provide excess oxygen, promoting complete and higher thermal efficiency in diesel engines, which inherently operate lean with λ ≈ 2 for optimal efficiency; here, excess air ensures all burns, though it may dilute the charge and limit power density. Peak efficiency in diesel engines occurs around λ = 2, while spark-ignition engines favor λ ≈ 1.1 for balanced power and efficiency. An excessively low A/F (rich condition) further exacerbates issues like high smoke and reduced efficiency due to incomplete oxygen- mixing. Factors like inadequate mixing or low temperatures can impair oxygen utilization even at near-stoichiometric ratios, as oxygen molecules may not reach all fuel particles in the brief combustion duration (typically 1–2 milliseconds). In diesel engines, stratified charge designs improve local oxygen availability near injectors, enhancing utilization and efficiency. Advanced control systems, such as lambda sensors, maintain optimal A/F to maximize oxygen use, with studies showing that precise stoichiometry can boost indicated thermal efficiency by 2–5% compared to uncontrolled rich operation. Oxygen-enriched intake air (e.g., >21% O₂) further improves utilization by reducing dilution, lowering exhaust losses and increasing efficiency by up to 10% in some configurations, though this is not standard in conventional engines.

Heat Transfer and Exhaust Losses

In internal combustion engines, heat transfer losses primarily arise from the conduction, , and of from the hot gases to the engine's walls, , head, and surrounding components such as the coolant jacket and sump. These losses represent a significant portion of the energy input, typically accounting for 20-30% in conventional and engines, depending on operating conditions like load and speed. For instance, in a heavy-duty operating on the SET cycle, coolant heat losses constitute about 10.6% of energy, with much of this stemming from (EGR) cooling. The process is driven by the steep temperature gradients between the peak temperatures (often exceeding 2000 ) and the cooler metal surfaces (maintained around 400-500 to prevent material degradation), leading to peak heat fluxes of 1-3 MW/m² in spark-ignition engines and up to 10 MW/m² in compression-ignition engines. This parasitic heat rejection reduces the availability of energy for useful work, directly lowering by diverting that could otherwise contribute to expansion in the power stroke. Exhaust losses occur when high-temperature combustion products are expelled from the without fully converting their into mechanical work, carrying away substantial in the form of , , and . In typical internal combustion engines, these losses range from 20-35% of the , with the exact value influenced by factors such as exhaust (often 600-900°C), gas composition, and backpressure from aftertreatment systems. For example, in a reference heavy-duty , exhaust accounts for 35.5% of input over the SET cycle, representing the largest single loss mechanism after brake power output (39.1%). The content of the exhaust—quantifying the theoretically recoverable work—is lower than the total due to generation during expansion and mixing, but it remains a prime target for recovery via bottoming cycles like Organic Rankine systems, which can reclaim up to 20% of this . Heat transfer and exhaust losses are interconnected, as reducing wall heat rejection often increases exhaust energy by concentrating more thermal content in the gases, potentially raising exhaust exergy by up to 18% under optimized conditions. In low-temperature combustion strategies, for instance, port insulation and advanced materials with low thermal conductivity can cut heat losses by 30%, redistributing about 10% of that energy to brake work while enhancing exhaust recoverability. However, excessive mitigation of heat transfer—such as through higher coolant temperatures—must balance against risks like knock in spark-ignition engines or lubricant breakdown, underscoring the thermodynamic trade-offs in engine design. Overall, these losses limit peak brake thermal efficiencies to below 40% in most practical engines, highlighting the need for integrated approaches like waste heat recovery to approach theoretical limits.

Internal Combustion Engines

Spark-Ignition Engines

Spark-ignition (SI) engines, commonly used in gasoline-powered vehicles, operate on the Otto thermodynamic cycle, in which a initiates of a premixed air-fuel charge at constant volume. The cycle consists of isentropic compression, constant-volume heat addition, isentropic expansion, and constant-volume heat rejection, modeled ideally with air as a perfect gas having constant specific heats. The of an ideal is expressed as \eta = 1 - \frac{1}{r^{\gamma - 1}} where r is the (the of maximum to minimum volume) and \gamma is the specific , approximately 1.4 for typical air-fuel mixtures. In practice, SI engines employ s of 8 to 12 to balance efficiency gains with the risk of autoignition (knock), yielding ideal efficiencies of 56% to 62%. Actual brake thermal efficiencies in SI engines are typically 25–35%, with peak values around 30–36% at mid-to-full load conditions and lower values (20–30%) at part-load, leading to an overall average of approximately 25% in typical light-duty driving cycles. This discrepancy arises from several irreversibilities and losses absent in the ideal model: throttling losses during intake, which reduce net work by creating a partial ; heat transfer to cylinder walls and , accounting for up to 20-30% of ; mechanical in pistons, bearings, and valves; incomplete leading to unburned hydrocarbons; and exhaust gas losses carrying away residual . A key limitation in SI engines is engine knock, caused by premature autoignition of the end-gas mixture under high compression, which caps the and thus efficiency potential compared to compression-ignition engines. Modern SI engines mitigate this through higher-octane fuels, advanced , and technologies like (EGR) to cool the charge, enabling compression ratios up to 13-14 in some designs and pushing peak efficiencies toward 40%. Variable valve timing and direct injection further reduce pumping losses and improve charge homogeneity, contributing 4-8% efficiency gains in production engines.

Compression-Ignition Engines

Compression-ignition (CI) engines, also known as engines, achieve higher than spark-ignition engines primarily due to their ability to operate at elevated ratios, typically 14:1 to 25:1, which elevate the temperature of the sufficiently for auto-ignition of injected without a spark. This design enables operation with excess air, minimizing throttling losses and allowing more complete under high expansion ratios. In light-duty vehicles, CI engines demonstrate the highest thermodynamic cycle efficiency among types, often reducing consumption by 25-33% relative to comparable spark-ignition counterparts when paired with advanced transmissions. The ideal thermodynamic cycle for CI engines is the , consisting of isentropic compression of air, constant-pressure heat addition via , isentropic expansion, and constant-volume heat rejection. The of this cycle is expressed as \eta = 1 - \frac{1}{r^{\gamma-1}} \cdot \frac{\rho^\gamma - 1}{\gamma (\rho - 1)}, where r is the , \rho is the cutoff ratio (ratio of volumes at the end and start of heat addition), and \gamma is the specific heat ratio (approximately 1.4 for air). Higher compression ratios directly increase efficiency by improving the and reducing heat rejection relative to work output, though practical limits arise from material strength and combustion noise. In real-world applications, brake thermal efficiencies of modern CI engines range from 40% to 50%, with advanced heavy-duty designs approaching 50% through optimized and reduced losses. As of 2025, some advanced heavy-duty CI engines have exceeded 50%, with records up to 53% brake thermal efficiency. The overall efficiency is the product of efficiency (typically 95-98%, reflecting minimal unburned hydrocarbons), thermodynamic efficiency (governed by the cycle), gas exchange efficiency (enhanced by unthrottled ), and (impacted by and pumping, often 85-90%). mixtures promote efficient oxygen utilization, while the absence of throttling eliminates significant pumping work, contributing to superior part-load compared to throttled spark-ignition engines. Efficiency improvements in CI engines stem from technologies like high-pressure common-rail , which enables precise control over injection timing and quantity for better air-fuel mixing, and turbocharging with intercooling, which boosts and allows downsizing without power loss. (EGR) systems, while primarily for emissions control, can be tuned to maintain by optimizing temperatures. These advancements have enabled indicated thermal efficiencies over 45% in production engines, though trade-offs with emissions aftertreatment (e.g., diesel particulate filters) introduce minor parasitic losses of 2-5%.

Gas Turbine Engines

Gas turbine engines operate on the , a thermodynamic cycle characterized by isentropic compression, constant-pressure heat addition, isentropic expansion, and constant-pressure heat rejection. This cycle forms the basis for both stationary power generation and aeronautical systems, where air is compressed, mixed with fuel and , and then expanded through a to produce work. Unlike reciprocating internal combustion engines, gas turbines feature continuous , enabling high power-to-weight ratios but introducing unique challenges related to airflow and heat management. The ideal of the depends primarily on the ratio r_p across the and the specific heat ratio \gamma of the , typically air. It is given by the : \eta = 1 - \frac{1}{r_p^{(\gamma-1)/\gamma}} For air-standard conditions with \gamma = 1.4, increases with higher ratios, approaching the Carnot limit asymptotically but limited in practice by material constraints and component losses. In real engines, deviations from ideality arise from non-isentropic processes in the and , drops in the , and incomplete , reducing overall . Key factors influencing efficiency include the inlet (TIT), which drives higher cycle efficiency by increasing the mean of heat addition, and the overall pressure ratio, which enhances work recovery. Modern advancements, such as advanced cooling and matrix composites, allow TITs exceeding 1,500°C, boosting simple-cycle efficiencies to over 40% on a lower heating value (LHV) basis for large industrial units (>100 MW). and polytropic efficiencies, typically 85-92%, further determine net output, with intercooling or reheating cycles proposed to mitigate losses but rarely implemented due to . In power generation applications, simple-cycle gas turbines achieve thermal efficiencies of 20-40%, constrained by exhaust heat losses that can exceed 60% of input energy. Combined-cycle configurations integrate a (HRSG) to capture exhaust heat for operation, yielding overall efficiencies up to 60% LHV in modern plants through optimized conditions and duct firing. For aeronautical gas turbines, reaches 50-55% in high-bypass turbofans at cruise, but overall efficiency emphasizes propulsive aspects, with specific fuel consumption (SFC) below 0.5 lb/(lbf·h) in advanced models. Efficiency improvements stem from materials enabling higher TITs and aerodynamic designs reducing losses, as seen in the evolution from early axial compressors to multi-stage, variable-geometry systems. Seminal contributions, such as Frank Whittle's 1930s turbojet patents, laid the groundwork for practical implementations, while ongoing research focuses on compatibility and carbon capture integration to sustain high efficiencies in decarbonized contexts.

External Combustion Engines

Steam Power Cycles

Steam power cycles, particularly the , form the foundational thermodynamic framework for external combustion engines utilizing as the . Developed by Scottish engineer William John Macquorn Rankine in his 1859 textbook A Manual of the Steam Engine and Other Prime Movers, the cycle models the conversion of heat energy into mechanical work through a series of processes involving phase changes in water. Unlike internal cycles, steam power cycles operate externally, where heat is supplied to a to generate high-pressure , which then expands in a or to produce work, before condensing and being pumped back to the . This closed-loop system emphasizes through controlled heat addition and rejection, making it suitable for large-scale power generation in thermal plants. The basic Rankine cycle consists of four idealized processes: (1) isentropic compression in a pump, (2) isobaric heat addition in a boiler, (3) isentropic expansion in a turbine, and (4) isobaric heat rejection in a condenser. The thermal efficiency of the cycle is given by \eta = \frac{W_{\text{net}}}{Q_{\text{in}}} = 1 - \frac{Q_{\text{out}}}{Q_{\text{in}}}, where W_{\text{net}} is the net work output, Q_{\text{in}} is the input during and , and Q_{\text{out}} is the rejected in the . In practice, this is limited by the Carnot efficiency upper bound but reduced by irreversibilities such as and losses, typically ranging from 20% to 40% for conventional steam plants. Higher pressures and lower temperatures enhance by increasing the average temperature of addition and decreasing rejection temperatures, respectively; for instance, operating at supercritical pressures above 22.1 MPa (the critical point of ) can eliminate change and achieve efficiencies up to 47%. Key modifications to the basic Rankine cycle improve efficiency while addressing practical issues like turbine blade erosion from wet steam. Superheating the steam beyond its saturation temperature at boiler pressure raises the average heat addition temperature, boosting efficiency by 5-10% and ensuring exit steam quality exceeds 90% to minimize moisture-related losses. Reheating involves expanding steam in a high-pressure turbine, reheating it isobarically, and then expanding further in a low-pressure turbine, which not only maintains steam dryness but also increases net work output, yielding efficiency gains of approximately 4-5%. Regeneration, through feedwater heaters that preheat boiler feedwater using extracted turbine steam, reduces the temperature difference during heat addition, thereby lowering irreversibilities and improving efficiency by 5-15% depending on the number of heating stages. These enhancements, rooted in Rankine's foundational analysis of heat-to-work conversion, have enabled modern ultrasupercritical steam cycles to approach 50% efficiency, with the highest reported net efficiency of 49.37% as of 2023 in plants like China's Pingshan Phase II.

Stirling Cycle Engines

The Stirling cycle engine, invented by Scottish clergyman Robert Stirling in 1816, operates as an external combustion on a closed regenerative . Patented as an improvement over contemporary steam engines to reduce explosion risks from high-pressure boilers, it initially found application in for mines, leveraging heated air as the working fluid to drive pistons without direct flame exposure. Despite early promise, it struggled to compete with steam engines due to material limitations at the time, though its design emphasized safety and potential for higher through heat regeneration. The engine's core principle involves cyclic and of a sealed working gas—typically air, , or —between hot and cold reservoirs, with a regenerator matrix storing and transferring to minimize losses. Key components include a power for mechanical output, a displacer to shuttle gas between temperature zones, a heater, a cooler, and the regenerator, which acts as a thermal absorbing during and releasing it during . Configurations vary: alpha-type with separate hot and cold pistons, beta-type with and displacer, and gamma-type with cylinders, each influencing mechanical simplicity and . The consists of two isothermal processes ( at low T_c and at high T_h) and two constant-volume regeneration processes, approximating the ideal reversible . Thermodynamically, the ideal Stirling cycle achieves efficiency \eta = 1 - \frac{T_c}{T_h}, matching the Carnot limit, as the regenerator enables near-perfect heat recovery. In practice, deviations arise from non-isothermal processes, finite rates, and losses such as in the regenerator (up to 10-15% efficiency penalty), shuttle heat conduction via displacer motion, and appendage dead volume reducing net work output. The analysis, a sinusoidal for motion, quantifies cycle performance: mean p_m = \frac{m R T_m}{V_{c0} + V_{e0} + V_r} (where m is gas mass, R the , T_m the mean , and V_{c0}, V_{e0}, V_r the swept volumes of / spaces and regenerator), yielding work per W = \frac{p_m V_{c0} b \sin \alpha}{2} (V_{e0} - V_{c0}) with phase angle \alpha. Regenerator effectiveness, ideally 1.0, drops to 0.8-0.95 in real designs, limiting to 30-50% of Carnot. For example, with T_h = 900 and T_c = 300 , theoretical \eta \approx 66.7\%, but imperfect regeneration reduces it to 30-40%. Practical efficiencies demonstrate the cycle's strengths in low-emission, multi-fuel applications. The GPU-3 engine achieved up to 27.2% brake thermal efficiency at 3000 RPM using at 2.76 MPa (400 psia), while modern kinematic engines have achieved up to 38% in prototypes under optimal conditions, dropping at lower speeds due to reduced . Modern free-piston variants, lacking linkages, minimize losses; simulations of a miniaturized double-acting with at 2 W input and 200 K differential yield up to 14% mechanical efficiency, optimized at 7.5 atm and low damping (0.0205 Ns/m). In solar systems, engines convert moderate-temperature heat (500-750 K) to electricity at 20-30% efficiency, outperforming in certain scenarios by recovering . These figures highlight the engine's scalability, with larger units approaching 40% in automotive prototypes, though challenges like high material costs and slow limit widespread adoption. As of 2025, research continues to optimize engines for and renewable applications, with models achieving higher efficiencies through advanced materials and simulation techniques.

Measurement and Improvements

Efficiency Standards and Testing

Efficiency standards for engines establish benchmarks for performance, fuel consumption, and emissions, ensuring comparability across manufacturers and compliance with environmental regulations. These standards typically focus on metrics such as brake thermal efficiency (BTE), which measures the ratio of useful work output to the energy content of the input, calculated as BTE = (brake power × 3600) / ( flow rate × lower heating value of ). Indicated assesses the engine's internal processes before mechanical losses, while specific fuel consumption (SFC) quantifies use per unit of power output, often in grams per (g/kWh). Testing under these standards uses controlled conditions to isolate variables like ambient temperature, pressure, and humidity, typically on dynamometers that simulate load and speed. International Organization for Standardization (ISO) standards provide foundational methods for engine performance evaluation. outlines reference conditions (e.g., 25°C air , 99 kPa , 30% relative ) and procedures for measuring , consumption, and oil consumption in reciprocating internal combustion engines using liquid or gaseous fuels. It includes steady-state and transient test cycles, allowing calculation of SFC and thermal efficiencies through direct measurement of , speed, and flow. For road vehicles, specifies net determination via testing, correcting for atmospheric conditions to ensure repeatability. These methods prioritize accuracy within ±2% for measurements, enabling global harmonization. In the United States, the Environmental Protection Agency (EPA) regulates engine testing under 40 CFR Part 1065, which details procedures for exhaust emissions and fuel mapping across engine operating points. This includes non-road transient cycles (NRTC) and steady-state tests for heavy-duty engines, where fuel efficiency is derived from integrated fuel consumption over the cycle, often yielding BTE values up to 45% for modern diesel engines under optimal loads. For light-duty vehicles, EPA's fuel economy testing in 40 CFR Part 600 employs the Federal Test Procedure (FTP-75), a chassis dynamometer cycle simulating urban and highway driving to compute miles per gallon (MPG), with adjustments for cold-start and evaporative losses. The Society of Automotive Engineers (SAE) complements these with J1349, a test code for net power and in spark-ignition and compression-ignition engines, conducted at 25°C and 99 kPa with corrections for accessories like alternators. This standard ensures power ratings reflect real-world service, indirectly supporting efficiency assessments by providing baseline curves for SFC calculations. Globally, the Worldwide Harmonized Light Vehicles Test Procedure (WLTP), adopted under UN ECE Regulation 83, replaces older cycles like NEDC with a more dynamic profile (up to 131 km/h speeds, 23-minute duration) to better estimate real-world fuel consumption and CO2 emissions, often resulting in 20-30% lower reported MPG compared to prior methods due to its realism. WLTP testing on dynamometers includes gear-specific acceleration and variable payloads, enhancing accuracy for and conventional engines.
StandardScopeKey MetricsTest Conditions
ISO 15550:2016Reciprocating IC engines (transport, non-road)Power, SFC, oil consumption25°C, 99 kPa, 30% RH; steady-state/transient cycles
SAE J1349SI and CI engines net power/torqueNet power, torque curves25°C, 99 kPa; dynamometer with accessory correction
EPA 40 CFR 1065Non-road/heavy-duty enginesEmissions, fuel mappingNRTC/steady-state; integrated over map
WLTP (ECE R83)Light-duty vehiclesFuel economy, CO2Dynamic cycle: urban/highway phases, up to 131 km/h
These standards evolve with technology, incorporating real-driving emissions (RDE) tests via portable analyzers to validate lab results, ensuring efficiencies align with regulatory targets like the EU's 95 g/km CO2 fleet average.

Modern Enhancements and Hybrids

Modern enhancements to efficiency have focused on technologies that optimize air-fuel mixtures, reduce parasitic losses, and enable downsizing without sacrificing power. Turbocharging, combined with (GDI), allows smaller engines to produce equivalent output to larger naturally aspirated ones, achieving fuel economy improvements of approximately 10% in spark-ignition engines through higher boost pressures and better combustion control. Variable valve timing (VVT) further refines this by adjusting intake and exhaust valve operations to improve and enable cycles like the Atkinson or , which prioritize expansion over compression for thermal efficiencies up to 5% higher than conventional cycles in boosted applications. Variable compression ratio (VCR) mechanisms, such as those adjusting piston stroke or volume, allow engines to adapt to varying loads, balancing high for efficiency at low speeds with lower ratios to prevent knocking under , yielding overall brake gains of 3-5% across drive cycles. Opposed-piston designs eliminate cylinder heads to minimize losses, potentially increasing peak efficiencies to over 50% when paired with advanced fuels like , which supports higher rates without formation. Hybrid powertrains represent a paradigm shift by integrating internal combustion engines with electric motors and batteries, recapturing energy via regenerative braking—where braking action charges the battery instead of dissipating heat—and allowing the engine to operate solely in its high-efficiency range. This synergy reduces fuel consumption by 25-35% in urban and combined cycles compared to conventional vehicles, as the electric motor provides torque assist and enables automatic engine shutoff during idling. In diesel hybrid generators, such systems achieve over 40% efficiency gains at low loads (e.g., 200 kW) and up to 60% at very low loads (e.g., 100 kW) by using variable-speed engines and energy storage to avoid inefficient part-throttle operation. Plug-in hybrids extend these benefits with grid-rechargeable batteries, further displacing fossil fuel use in short trips while maintaining long-range capability.

References

  1. [1]
    Engine Efficiency - an overview | ScienceDirect Topics
    Engine efficiency is defined as the amount of energy produced by the engine per unit of fuel energy consumed. AI generated definition based on: Reference Module ...
  2. [2]
    6.2: Engines and Thermal Efficiency - Physics LibreTexts
    Nov 8, 2022 · Engines convert heat transfer between two thermal reservoirs at different temperatures into work. For reasons we will learn later, ...Simple Engine · Real-World Engines · Thermal Efficiency · Otto Cycle
  3. [3]
    Engine Efficiency - DieselNet
    The overall brake thermal efficiency of the engine is a product of the combustion, thermodynamic, gas exchange, and mechanical efficiency.
  4. [4]
    [PDF] Defining engine efficiency limits - Department of Energy
    Thermodynamic analysis provides insight on potential for efficiency gains. • 1st and 2nd Laws of Thermodynamics can be used to provide detailed analysis of how ...<|control11|><|separator|>
  5. [5]
    Advanced Combustion Engines - Stanford
    Dec 9, 2011 · With typical combustion engines operating between 1750°K and 298°K, this equation states that the maximum efficiency for such an engine is 83%.
  6. [6]
    [PDF] Internal Combustion Engines - Department of Energy
    Engine efficiency improvements alone can potentially increase passenger vehicle fuel economy by 35% to 50%, and commercial vehicle fuel economy by 30%, with ...
  7. [7]
    Lab scientists optimizing high performance fuels for advanced ...
    Mar 20, 2018 · "Increasing the efficiency of internal combustion engines is one of the most cost-effective approaches to improving the energy efficiency of ...
  8. [8]
    Indicated Thermal Efficiency - an overview | ScienceDirect Topics
    Thermal efficiency can be divided into two kinds: indicated thermal efficiency and brake thermal efficiency. The brake thermal efficiency (BTE) indicates how ...
  9. [9]
    Engine Thermal Efficiency - an overview | ScienceDirect Topics
    Engine thermal efficiency depends on various factors: such as engine design, combustion chamber design, engine operating conditions (load and speed) and the ...
  10. [10]
    Explained: The Carnot Limit | MIT News
    May 19, 2010 · The Carnot Limit sets an absolute limit on the efficiency with which heat energy can be turned into useful work.Missing: formula | Show results with:formula
  11. [11]
    PHYS 200 - The Second Law of Thermodynamics and Carnot's Engine
    The Carnot heat engine is discussed in detail to show how there is an upper limit to the efficiency of heat engines and how the concept of entropy arises.
  12. [12]
    15.4 Carnot's Perfect Heat Engine: The Second Law of ...
    Since temperatures are given for the hot and cold reservoirs of this heat engine, E f f C = 1 − T c T h can be used to calculate the Carnot (maximum theoretical) ...
  13. [13]
    The Carnot Efficiency | EGEE 102 - Dutton Institute - Penn State
    The Carnot Efficiency is the theoretical maximum efficiency one can get when the heat engine is operating between two temperatures.
  14. [14]
    Heat Engines: the Carnot Cycle - Galileo
    Efficiency = TH−TCTH. This was an amazing result, because it was exactly correct, despite being based on a complete misunderstanding of the nature ...
  15. [15]
    [PDF] Otto and Diesel Cycles
    Diesel versus Otto Cycle Efficiency ... 3-4: Isentropic exp. 4-1: Constant-pressure heat rejection. Page 25. Gas Power Cycles - Page 25. Brayton Cycle Efficiency ...
  16. [16]
    Thermodynamic Foundations – Introduction to Aerospace Flight ...
    Efficiency comparison of Otto, Diesel, Brayton, and Carnot cycles. For the Otto cycle (spark-ignition engines), the ideal air-standard efficiency ranges ...Missing: formulas | Show results with:formulas
  17. [17]
    3.7 Brayton Cycle - MIT
    The Brayton cycle thermal efficiency contains the ratio of the compressor exit temperature to atmospheric temperature, so that the ratio is not based on the ...Missing: Otto | Show results with:Otto<|control11|><|separator|>
  18. [18]
    [PDF] Compression-Ratio-and-Thermal-Efficiency.pdf
    Compression Ratio – Otto Engine. The compression ratio, CR, is defined as the ratio of the volume at bottom dead center and the. volume at top dead center. It ...
  19. [19]
    Otto Cycle Thermodynamic Analysis
    On this page we discuss the basic thermodynamic equations that allow you to design and predict engine performance. In an internal combustion engine, fuel and ...
  20. [20]
    Thermal Efficiency for Diesel Cycle | Equation | nuclear-power.com
    In general, the thermal efficiency, ηth, of any heat engine is defined as the ratio of the work it does, W, to the heat input at the high temperature, QH.
  21. [21]
    The Effects of Compression Ratio and Expansion Ratio on Engine ...
    Mar 9, 2009 · For example, for a compression ratio of 10, expansion ratios of 10 and 30 provided brake thermal efficiencies of about 34% and 43%, respectively ...
  22. [22]
    [PDF] Engine Friction Reduction Technologies - Department of Energy
    Jun 19, 2014 · Overall Project Take-Home Message on Lubricants: Parasitic friction losses in engines and transmissions consume up to 10-15% of fuel.
  23. [23]
    [PDF] Engine friction - – terminology – Pumping loss - MIT OpenCourseWare
    The engine coolant and oil flow losses are provided for by the oil and water pump. The nature of the loss is a pumping loss though. SI engine friction. ( ...
  24. [24]
    Automotive & Aerospace Lubrication - Purdue Engineering
    Up to 60% of the total fuel energy used in a typical automotive engine is lost to various factors and mechanical friction accounts for roughly 15% of these ...
  25. [25]
    [PDF] AP-42, Vol. I, 3.3: Gasoline And Diesel Industrial Engines
    This situation occurs if there is a lack of available oxygen near the hydrocarbon (fuel) molecule during combustion, if the gas temperature is too low, or if ...
  26. [26]
    [PDF] combustion efficiency - DSpace@MIT
    During rich engine operation, the combustion efficiency drops off. Since the mixture is oxygen deficient under these conditions, some of the fuel will not burn.Missing: utilization | Show results with:utilization
  27. [27]
    Air fuel ratio - x-engineer.org
    The best combustion efficiency is obtained at λ = 2.00 for diesel and λ = 1.12 for spark ignition (gasoline) engines. Go back. Air fuel ratio calculator. ma [g]
  28. [28]
    Air-Fuel Ratio - an overview | ScienceDirect Topics
    An excessively low air–fuel ratio may produce the problems of deteriorated combustion efficiency, high smoke, and high exhaust gas temperature. Air–fuel ratio ...
  29. [29]
    [PDF] Oxygen-Enriched Combustion - Department of Energy
    Oxygen-enriched combustion increases oxygen in combustion air, reducing energy loss and increasing efficiency, and can lower emissions and improve heat ...Missing: utilization | Show results with:utilization
  30. [30]
    None
    ### Summary of Heat Transfer in Engines from MIT Lecture Notes
  31. [31]
    [PDF] Heavy-duty vehicle diesel engine efficiency evaluation and energy ...
    The reference heavy-duty diesel engine converted 39.1% of its fuel energy to brake power over the SET engine cycle, with 35.5% lost as exhaust heat, 10.6% lost ...
  32. [32]
    3.5 The Internal combustion engine (Otto Cycle) - MIT
    3.5.1 Efficiency of an ideal Otto cycle. The starting point is the general expression for the thermal efficiency of a cycle:.
  33. [33]
    Gasoline Engine - General Physics II
    Notice that increasing the compression ratio (V1/V2) increases the efficiency. The compression ratio of a typical, common automobile engine is about 8; for ...
  34. [34]
    Otto Cycle - an overview | ScienceDirect Topics
    For a real engine with similar specifications, the thermal efficiency is less than 30%; hence, the Otto cycle is unsuccessful in predicting the thermal ...
  35. [35]
    Spark Ignition - Engine Combustion Network
    •High engine-out HC, CO, and NOx emissions, but low tailpipe emissions (catalyst) •Peak thermal efficiency typically 30%, full load range possible
  36. [36]
    Optimization of a Spark Ignition Engine Knock and Performance ...
    Sep 2, 2022 · Since the advent of the internal combustion engine, knock has been a vital issue limiting the thermal efficiency of spark ignition engines ...3.1. Combustion Model · 4. Optimization Techniques · Glossary
  37. [37]
    Improving Thermal Efficiency of Internal Combustion Engines - MDPI
    This article reviews the recent progress in the improvement of thermal efficiency of IC engines and provides a comprehensive summary of the latest research on ...
  38. [38]
    4 Spark-Ignition Gasoline Engines | Assessment of Fuel Economy ...
    The EPA and NESCCAF both estimate a 4 to 6 percent reduction in fuel consumption (EPA, 2008; NESCCAF, 2004), while EEA estimates a 6.5 to 8.3 percent reduction ...
  39. [39]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Dec 15, 2021 · (1) Typical values for the compression ratio for a compression ignition IC engine are r = 12 to 24 with engine thermal efficiencies of 7 = 40 to ...
  40. [40]
  41. [41]
    The Diesel Engine - HyperPhysics
    It is convenient to express this efficiency in terms of the compression ratio rC = V1/V2 and the expansion ratio rE = V1/V3. The efficiency can be expressed in ...
  42. [42]
    Internal Combustion Engine Basics | Department of Energy
    The cycle includes four distinct processes: intake, compression, combustion and power stroke, and exhaust. Spark ignition gasoline and compression ignition ...
  43. [43]
    Turbine Engine Thermodynamic Cycle - Brayton Cycle
    The Brayton cycle analysis is used to predict the thermodynamic performance of gas turbine engines. The EngineSim computer program, which is available at this ...Missing: fundamentals | Show results with:fundamentals
  44. [44]
    How Gas Turbine Power Plants Work - Department of Energy
    A simple cycle gas turbine can achieve energy conversion efficiencies ranging between 20 and 35 percent. With the higher temperatures achieved in the Department ...Missing: modern | Show results with:modern
  45. [45]
    [PDF] 4.4-1 Introduction Heat Transfer Analysis Frank J. Cunha, Ph.D., P.E.
    Other losses may include mixing and aerodynamic losses, such as profile drag, skin-friction, gas diffusion, secondary flows, tip clearance, boundary-layer ...
  46. [46]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Dec 15, 2021 · The Brayton Cycle and Improvements. The Brayton cycle is a thermodynamic model for gas turbine engines. In a simple gas turbine engine.
  47. [47]
    [PDF] Combustion Turbines - U.S. Environmental Protection Agency
    Simple- cycle gas turbines used in power plants are available with efficiencies of over 40 percent (LHV).
  48. [48]
    3 Aircraft Gas Turbine Engines - The National Academies Press
    These turbines operate at cruise, with motor thermodynamic efficiencies of up to 55 percent and propulsive efficiencies of well over 70 percent, yielding an ...
  49. [49]
    [PDF] Chapter23 - Rose-Hulman
    Review the operation of reciprocating engines. • Solve problems based on the Otto and Diesel cycles. • Solve problems based on the Brayton cycle and the Brayton.Missing: formulas | Show results with:formulas
  50. [50]
    [PDF] The Life and Legacy of William Rankine - Purdue e-Pubs
    A Rankine. Cycle takes heat at a higher temperature and converts some of it to work, using an expansion engine to drive an alternator to produce electricity.
  51. [51]
    RANKINE CYCLE - Thermopedia
    The Rankine cycle is the fundamental operating cycle of all power plants where an operating fluid is continuously evaporated and condensed.
  52. [52]
    [PDF] Chapter 8 - WPI
    The second law of thermodynamics requires the thermal efficiency to be less than 100%. Most of today's vapor power plants have thermal efficiencies ranging ...
  53. [53]
    [PDF] Stirling Engine Design Manual
    Mar 27, 1978 · This is a Stirling Engine Design Manual, second edition, prepared for NASA, covering what a Stirling engine is and major types.
  54. [54]
    Robert Stirling - Linda Hall Library
    Oct 25, 2019 · Although the Stirling engine is more efficient that a steam engine, and thus more economical, it never successfully competed with steam or ...
  55. [55]
    [PDF] Schmidt analysis for Stirling Engines
    Now, since the Schmidt analysis is based on the Ideal Isothermal model, the thermal efficiency should reduce to the Carnot efficiency. The thermal efficiency ...
  56. [56]
    [PDF] Design Exploration of a Miniaturized Stirling Engine - DSpace@MIT
    Stirling engines operate on the principles of thermodynamics, the engine goes through a ... engine efficiency of 20%, which is one of the highest efficiency.<|control11|><|separator|>
  57. [57]
  58. [58]
    ISO 15550:2016 - Internal combustion engines
    2–5 day deliveryISO 15550:2016 specifies standard reference conditions and methods of declaring the power, fuel consumption, lubricating oil consumption and test methods for ...Missing: summary | Show results with:summary
  59. [59]
    ISO 1585:2020 - Road vehicles — Engine test code — Net power
    2–5 day deliveryISO 1585:2020 specifies a method for testing engines in automotive vehicles, evaluating performance and presenting power and fuel consumption curves for net ...Missing: efficiency | Show results with:efficiency
  60. [60]
    40 CFR Part 1065 -- Engine-Testing Procedures - eCFR
    This part describes the procedures that apply to testing we require for the following engines or for vehicles using the following engines.Subpart J —Field Testing and... · Calibrations and Verifications · Title 40 · 1065.530
  61. [61]
    Engine Testing Regulations | US EPA
    The test procedures referenced on this page are related to engine-based exhaust emission standards. All types of engines are subject to the procedures in 40 CFR ...
  62. [62]
  63. [63]
    The Worldwide Harmonised Light Vehicle Test Procedure (WLTP)
    Mar 30, 2021 · The WLTP is mandatory for all new cars powered by an internal combustion engine. The VCA provides details of the purpose of the WLTP and ...
  64. [64]
    EU's WLTP and RDE Emissions Testing | TÜV SÜD
    Testing used to assess emissions levels and fuel economy for light-duty vehicles and engines under various operating conditions in accordance with European ...<|separator|>
  65. [65]
    Fuel-engine collaboration accelerates America's energy ...
    Jul 8, 2025 · Up to a 10 percent boost in engine efficiency with turbocharged spark ignition engines · An additional 9–14 percent gain in fuel economy using ...
  66. [66]
    [PDF] Downsized, boosted gasoline engines
    Oct 28, 2016 · With GDI, the compression ratio in a turbocharged engine can be higher than with port fuel injection (PFI) and engine efficiency is improved.
  67. [67]
    [PDF] Advanced Combustion and Emission Control Roadmap
    • Diesel engines provide improved engine thermal efficiency from part load to high load. Diesel engine has a cost premium driven by aftertreatment, fuel ...
  68. [68]
    Combined Technologies for Efficiency Improvement on a 1.0 L ...
    Apr 2, 2019 · The average brake thermal efficiency (BTE) benefit percentage is 3.47% with 9.6 compression ratio and 5.33 % with 12 compression ratio. However, ...
  69. [69]
    Future of Combustion Engines Promises Efficiency, Green Fuels - TT
    Aug 2, 2024 · Advanced researchers and engine makers say the internal combustion engine's ability to be more efficient and run on greener fuels means an extended life.Missing: definition | Show results with:definition
  70. [70]
    How Do Hybrid Electric Cars Work? - Alternative Fuels Data Center
    Hybrid electric vehicles are powered by an internal combustion engine and one or more electric motors, which uses energy stored in batteries.How Do Plug-In Hybrid Electric... · Fuel Cell Electric Vehicles · How It Works
  71. [71]
    6 Hybrid Power Trains | Assessment of Fuel Economy Technologies ...
    Hybrid vehicles achieve reduced fuel consumption by incorporating in the drive train, in addition to an internal combustion (IC) engine, both an energy storage ...
  72. [72]
    [PDF] Hybrid Power Generation for Improved Fuel Efficiency and ...
    The hybrid system improves fuel efficiency over 40% at low loads and 60% at very low loads, while maintaining similar efficiency to diesel at higher loads.
  73. [73]
    How Do Plug-In Hybrid Electric Cars Work?
    Plug-in hybrid electric vehicles (PHEVs) use batteries to power an electric motor and another fuel, such as gasoline, to power an internal combustion engine ...