Fact-checked by Grok 2 weeks ago

Internal combustion engine

The internal combustion engine (ICE) is a heat engine in which the combustion of fuel with an oxidizer, typically air, occurs inside a combustion chamber forming an integral part of the working fluid flow circuit, converting chemical energy into mechanical work via expansion of high-temperature, high-pressure gases acting on engine components such as pistons or turbine blades. In reciprocating piston variants, the most common type, this process drives a crankshaft through cyclic motion involving intake, compression, combustion, and exhaust strokes, while rotary forms like Wankel engines use alternative mechanisms for continuous operation. Primarily fueled by hydrocarbons such as gasoline or diesel, ICEs achieve thermal efficiencies ranging from 20% to 40% depending on design, load, and technology, far surpassing early steam engines but limited by thermodynamic constraints like the Carnot cycle. Developed in the late 19th century, the ICE revolutionized transportation and power generation, with Nikolaus Otto's 1876 four-stroke cycle engine marking a pivotal advancement in reliable operation, followed by Rudolf Diesel's 1892 compression-ignition design offering higher efficiency for heavy-duty uses, and Karl Benz's 1885 application in the first practical automobile. These engines power the majority of road vehicles, aircraft, ships, and stationary generators worldwide, enabling global mobility and industrialization on a scale unattainable by prior technologies. Key variants include spark-ignition gasoline engines for light vehicles and compression-ignition diesel engines for trucks and marine propulsion, with ongoing innovations like turbocharging, direct injection, and variable valve timing improving performance and fuel economy. Despite their ubiquity, ICEs produce exhaust emissions including carbon dioxide, nitrogen oxides, particulate matter, and unburnt hydrocarbons, contributing to air pollution and climate forcing, though empirical data show dramatic reductions—over 99% in some pollutants per mile driven in the U.S. since 1970—due to catalytic converters, fuel standards, and engine controls mandated by regulations like the Clean Air Act. Controversies center on their role in anthropogenic CO2 accumulation, prompting shifts toward electrification, yet lifecycle analyses indicate ICEs with biofuels or synthetic fuels can mitigate impacts while retaining advantages in energy density and infrastructure compatibility. Their defining characteristic remains causal efficacy in harnessing combustion for scalable mechanical power, underpinning economic growth despite environmental trade-offs addressed through engineering rather than wholesale replacement.

Principles of Operation

Thermodynamic Fundamentals

The internal combustion engine (ICE) functions as an open thermodynamic cycle heat engine, wherein fuel combustion releases chemical energy as heat within the working fluid—typically an air-fuel mixture—which expands to produce mechanical work, with exhaust gases discarded after each cycle rather than recirculated. This contrasts with closed cycles like the Rankine steam engine, where the fluid is reused; the open nature of ICEs introduces irreversibilities such as incomplete combustion and heat losses, limiting real-world thermal efficiencies to 20-40% despite higher theoretical potentials. Fundamentally, ICE efficiency is bounded by the second law of thermodynamics, with the Carnot efficiency η_C = 1 - T_L / T_H providing an absolute upper limit for any heat engine operating between hot reservoir temperature T_H (combustion gases, often 2000-2500 K) and cold reservoir T_L (ambient or coolant, ~300-600 K), yielding η_C > 80% under ideal conditions; however, practical cycles achieve far less due to non-quasistatic processes and entropy generation. For spark-ignition gasoline engines, the idealized Otto cycle approximates the process: isentropic compression of the air-fuel mixture (process 1-2), constant-volume heat addition via spark-induced combustion (2-3), isentropic expansion (3-4), and constant-volume heat rejection (4-1). The thermal efficiency of the air-standard Otto cycle derives as η_Otto = 1 - (1 / r)^{γ-1}, where r is the compression ratio (typically 8-12 for gasoline engines to avoid auto-ignition) and γ ≈ 1.4 is the specific heat ratio of the working fluid; for r = 10, this yields η_Otto ≈ 60%, though real engines realize ~25-35% owing to pumping losses, friction, and variable γ during combustion. This formula emerges from isentropic relations T_2 / T_1 = r^{γ-1} and T_4 / T_3 = r^{γ-1}, with heat input Q_in = C_v (T_3 - T_2) and rejection |Q_out| = C_v (T_4 - T_1), so η = 1 - |Q_out| / Q_in = 1 - T_1 / T_2. In compression-ignition diesel engines, the Diesel cycle modifies the Otto by replacing constant-volume heat addition with constant-pressure combustion (process 2-3), allowing higher compression ratios (14-25) without knocking, as fuel injects post-compression. Efficiency is η_Diesel = 1 - (1 / r)^{γ-1} \cdot \frac{ρ^γ - 1}{γ (ρ - 1)}, where ρ = V_3 / V_2 is the cutoff ratio (typically 1.5-3, reflecting injection duration); for r = 18 and ρ = 2, η_Diesel ≈ 65% theoretically, enabling real diesel efficiencies of 35-45%, superior to gasoline due to reduced heat rejection at lower expansion ratios but offset by slower constant-pressure burning. These air-standard models assume ideal gases, reversible processes, and constant properties, ignoring real effects like dissociation, variable composition, and heat transfer, which first-principles analysis reveals as primary efficiency reducers via increased entropy production.

Combustion Process

In internal combustion engines, the entails the rapid exothermic oxidation of a by oxygen from compressed within the confined , converting into that expands combustion products to drive on the . This , ideally yielding and , occurs under and conditions, with real-world processes involving finite rates and incomplete mixing. The differs fundamentally between spark-ignition (SI) and compression-ignition (CI) engines, reflecting their respective thermodynamic cycles. In SI engines, operating on the Otto cycle, a homogeneous air-fuel mixture at equivalence ratios near stoichiometric (approximately 14.7:1 by mass for gasoline) is compressed to ratios of 8:1 to 12:1, raising temperatures to 500-700°C, before ignition via an electric spark plug arc reaching 10,000°C. Combustion initiates with kernel formation and propagates as a turbulent premixed flame at 25-50 m/s, causing pressure to peak post-top dead center at 40-60 bar and temperatures of 2000-2500 K, though heat losses and dissociation reduce effective expansion. The near-constant volume heat addition in the ideal model contrasts with real finite-duration burning over 1-2 milliseconds, spanning 40-80° crank angle. In CI engines, such as diesels, air alone is compressed to higher ratios of 14:1 to 25:1, achieving temperatures of 700-900°C and pressures up to 40 bar at the end of compression, enabling auto-ignition upon direct fuel injection. Fuel atomizes, vaporizes during an ignition delay of 0.5-2 ms (5-15° crank angle), followed by premixed combustion—a rapid pressure spike to 100-200 bar—then diffusion-controlled mixing and burning of remaining fuel, with overall combustion extending longer than in SI engines due to heterogeneous charge. Lean overall air-fuel ratios exceeding 20:1 facilitate higher thermal efficiencies but demand precise injection timing to minimize delays and NOx formation. Key physical phenomena include flame quenching near walls, turbulence enhancement of mixing, and potential abnormal combustion like knock in SI (auto-ignition ahead of flame front) or excessive delay in CI leading to rough operation. Fuel chemistry governs ignition: gasoline's higher octane resists premature auto-ignition, while diesel's cetane number (typically 40-55) promotes shorter delays. These processes underpin engine power density and efficiency limits, with real efficiencies of 25-35% for SI and 35-45% for CI arising from incomplete combustion, heat transfer, and pumping losses beyond ideal cycle predictions.

Engine Classifications

Reciprocating Engines

Reciprocating engines, also known as piston engines, constitute the predominant form of internal combustion engines, utilizing one or more pistons that reciprocate within cylinders to convert the pressure from combustion into rotational mechanical work through a crankshaft. In these engines, combustion of fuel-air mixture occurs directly inside the cylinders, driving the pistons linearly before the connecting rods transfer motion to the crankshaft. Key components include the the pistons, for rotary output, linking pistons to the , for and exhaust , and camshafts timing operation. Spark plugs initiate in spark-ignition , while compression-ignition types rely on high ratios to auto-ignite . Reciprocating engines are classified by ignition method and stroke cycle. Spark-ignition engines operate on the Otto cycle, featuring constant-volume heat addition via spark, suitable for gasoline fuels. Compression-ignition engines follow the Diesel cycle with constant-pressure heat addition, achieving higher thermal efficiencies through elevated compression ratios typically exceeding 14:1, and are optimized for diesel fuel. By stroke mechanism, four-stroke engines complete a power cycle over two crankshaft revolutions: intake of air-fuel mixture, compression, combustion-driven power stroke, and exhaust expulsion. This design, patented by Nikolaus Otto in 1876, offers superior fuel efficiency and lower emissions but requires more components like separate valves and a dedicated lubrication system. Two-stroke engines deliver a power stroke every crankshaft revolution, simplifying construction with ports instead of valves, yielding higher power-to-weight ratios ideal for applications like chainsaws and outboard motors. However, two-strokes suffer from incomplete scavenging, leading to higher fuel consumption, oil mixing with fuel for lubrication, and elevated exhaust emissions compared to four-strokes. Four-stroke engines generally exhibit better and at lower RPMs to dedicated and reduced per , though their increases costs. Two-strokes excel in high-RPM but accelerate component , necessitating frequent overhauls. These engines power diverse applications from automobiles and trucks to generators and , with reciprocating designs dominating to scalable efficiency and load response.

Rotary Engines

Rotary engines represent a class of internal combustion engines that achieve cyclic motion through rather than reciprocating pistons, with the serving as the most prominent example. In the Wankel design, a triangular rotor spins within an epitrochoid-shaped , performing , , , and exhaust phases across three faces of the rotor during each eccentric shaft , yielding three power impulses per rotor . This configuration eliminates crankshafts and connecting rods, reducing the number of moving parts to primarily the rotor, eccentric shaft, and seals. Developed by , emerged from in , with practical internal prototypes realized post- through with . NSU produced the first Wankel-powered , the , followed by the 110S as the inaugural commercially successful model after Mazda licensed the technology in . Applications have included automobiles like Mazda's RX-7 series (producing up to 276 horsepower in twin-rotor variants), motorcycles, snowmobiles, and auxiliary power units, though automotive use declined due to efficiency challenges. Key advantages include a compact and high , revs exceeding 9,000 RPM, alongside smoother from continuous . However, drawbacks are significant: and side suffer accelerated from sliding , leading to reliability issues; occurs in a thin, elongated chamber, resulting in incomplete and elevated unburned emissions (up to 10 times higher than engines); and fuel efficiency lags, with specific fuel often 20-30% worse than comparable reciprocating engines due to sealing inefficiencies and total-loss requiring oil injection. These factors contributed to limited adoption beyond niche performance roles. Distinct from Wankel types, historical "rotary" engines in early aviation, such as the 1900s Gnome Omega, featured fixed crankshafts with rotating cylinders and crankcases for cooling, but retained reciprocating pistons and thus differ fundamentally in operation from true pistonless rotaries. Experimental variants like Liquid Piston's high-efficiency rotary X-engine, introduced in the 2010s, aim to address Wankel shortcomings via improved sealing and multifuel capability, targeting diesel and range-extender uses, though commercial scaling remains nascent as of 2025.

Continuous Combustion Engines

Continuous combustion engines differ from intermittent types by sustaining steady-state combustion, where fuel and oxidizer flow continuously into a combustion chamber, generating a persistent stream of high-temperature gases to produce mechanical work. This contrasts with reciprocating engines, which ignite fuel-air mixtures in discrete cycles. The approach enables smoother operation and higher power densities, though it demands precise control of airflow and fuel injection to maintain stability. The archetypal continuous combustion engine is the gas turbine, functioning via the Brayton cycle, which comprises isentropic compression of intake air, isobaric heat addition through continuous combustion, isentropic expansion across turbine blades, and isobaric heat rejection. In operation, an axial or centrifugal compressor raises air pressure to 10-40 times atmospheric levels, elevating its temperature to 400-600 K. Fuel, typically natural gas or kerosene, mixes with this compressed air in an annular or can-type combustor, igniting to yield gases at 1200-2000 K, which then expand through one or more turbine stages. The turbine extracts energy to power the compressor (using 40-60% of output) and delivers net shaft power for propulsion or electricity generation. Variants include turbojets and turbofans for aviation, where exhaust gases provide thrust directly or via a geared fan; industrial turbines for power plants, achieving thermal efficiencies up to 40% in simple cycles and over 60% in combined-cycle setups with steam recovery; and marine drives in ships like frigates. Ramjets and scramjets represent compressor-less continuous designs, relying on vehicle speed for air compression, suitable for supersonic applications above Mach 2 and 5, respectively. Rocket engines also employ continuous combustion in steady-state firing, injecting liquid propellants into a chamber for expansion through nozzles, powering spacecraft with specific impulses of 200-450 seconds. Advantages encompass compact design yielding power-to-weight ratios exceeding 5 kW/kg in aero-derivatives, minimal vibration from steady flow, and rapid startup within minutes for peaking plants. Drawbacks include sensitivity to inlet air quality, with particulates eroding blades, and peak efficiencies only at full load, dropping below 20% at partial output. Material limits, such as turbine inlet temperatures capped at 1700 K by nickel superalloys, constrain performance absent advanced cooling like air film or ceramic coatings.

Historical Evolution

Early Concepts and Precursors

The earliest conceptual precursor to the internal combustion engine (ICE) emerged in the 17th century with experiments harnessing explosive forces within a confined space to generate mechanical work. In 1673, Dutch physicist , collaborating with Godard Reede, designed a prototype consisting of a vertical cylinder partially filled with gunpowder; ignition created an explosion that drove a heavy piston upward against atmospheric pressure, with the piston's motion transferred via a connecting rod to a walking beam for potential pumping or other applications. This device represented an initial attempt at internal combustion by containing the fuel's reaction directly within the working chamber, though it suffered from incomplete combustion, fouling from residue, and lack of a reliable return stroke, rendering it impractical for sustained operation. Subsequent 17th- and 18th-century efforts built on this explosive but yielded due to inefficiencies in fuel delivery, ignition control, and sealing. English inventor Samuel Morland patented a similar gunpowder-based mechanism around 1661, predating Huygens but sharing the challenges of erratic power output and material degradation from byproducts. Various European tinkerers, including attempts documented in the early 1700s, experimented with refined gunpowder charges in cylinders to mimic cannon propulsion on a smaller scale, yet these remained laboratory curiosities, unable to compete with emerging external alternatives like Newcomen's atmospheric of 1712. By the late 18th century, precursors shifted toward gaseous or liquid fuels for cleaner operation. In 1794, English engineer Robert Street patented an ICE design using a volatile liquid fuel vaporized and mixed with air in a cylinder, ignited by a flame to drive a piston; this innovation avoided solid residues but still faced issues with weak power density and valve timing, preventing commercial viability. These early endeavors highlighted fundamental causal challenges—such as achieving controlled, repeatable combustion cycles without excessive heat loss or mechanical wear—that would only be addressed in the 19th century through advances in metallurgy and thermodynamics, paving the way for practical engines.

19th-Century Inventions

The earliest practical internal combustion engine emerged from the work of Belgian inventor Étienne Lenoir, who constructed a single-cylinder, two-stroke device in 1859 that burned a mixture of coal gas and air. This engine, patented in France in 1860 as "an air motor expanded by gas combustion," modified a double-acting steam engine by incorporating slide valves for intake of the air-fuel mixture and exhaust expulsion, with ignition via an electric spark. Operating at low pressure without compression, it delivered about 0.5 horsepower at 100-200 RPM but suffered from low thermal efficiency of roughly 4%, limiting its use to stationary applications like water pumps and printing presses. By 1865, approximately 300 to 500 Lenoir engines had been manufactured and sold, marking the first commercial deployment of such technology. In 1872, American inventor George Brayton patented a distinct constant-pressure cycle engine, utilizing separate single-acting cylinders: one for compressing atmospheric air to about 2-3 atmospheres and another for power production where fuel was continuously injected and burned externally. Known as the "Ready Motor," this two-stroke design employed petroleum or illuminating gas as fuel, ignited by a hot tube or flame, and achieved power outputs up to 3 horsepower in walking-beam configurations for stationary use. The Brayton engine's external combustion chamber allowed steady operation but required bulky air reservoirs, influencing later gas turbine developments despite limited commercial success in the era due to competition from more efficient reciprocating designs. A pivotal advancement occurred in 1876 when German engineer Nikolaus August Otto, building on prior atmospheric engines from 1861-1864, patented the four-stroke cycle internal combustion engine, compressing the air-fuel mixture prior to ignition for greater efficiency. This —intake, compression, power, and exhaust strokes—operated on illuminating gas or vapor, delivering up to 3 horsepower at 150-200 RPM in early models produced by Deutz Gasmotorenfabrik, with thermal efficiencies reaching 12-15%, a marked improvement over Lenoir's design. The engine's success stemmed from its closed-cycle operation and compression ignition timing, enabling scalable production; by 1878, thousands were in use for industrial power, laying the foundation for automotive applications. Toward the decade's end, refinements emphasized higher speeds and portability. In 1885, and developed a compact, high-revolution Otto-derived engine with a vertical single-cylinder design, float-fed carburetion, and hot-tube ignition, capable of 600 RPM and 0.5 horsepower. Patented that year, this "grandfather clock" engine powered the first prototype and subsequent vehicles, prioritizing lightweight and surface carburetion for reliable operation at elevated speeds up to 1,000 RPM in later variants. These innovations bridged stationary engines to mobile propulsion, though widespread vehicle commercialization awaited the 1890s.

20th-Century Commercialization

The of internal combustion engines in the 20th transformed and by and widespread , shifting from experimental prototypes to reliable, scalable power sources. In the automotive sector, Ford's Model T, introduced on , , featured a 20-horsepower, four-cylinder and pioneered affordable through innovative . The of the moving at Ford's in reduced chassis time from more than 12 hours to about minutes, cutting costs and allowing the to be priced as low as $260 by 1925. This efficiency enabled production of over 15 million Model T units by 1927, democratizing personal automobiles and entrenching the four-stroke as the dominant technology for road vehicles. Diesel engines, prized for their higher thermal efficiency and suitability for heavy-duty applications, saw parallel commercial growth. Following Rudolf Diesel's 1892 patent, early marine and stationary installations emerged around 1900, with MAN AG delivering 77 diesel cylinders for commercial use by 1901. High-speed variants suitable for vehicles appeared in the 1920s, powering trucks and buses with fuel economies up to 30% better than gasoline equivalents, which facilitated adoption in commercial trucking fleets. Mercedes-Benz's 1936 launch of the 260 D marked the first series-production diesel passenger car, producing about 85 units initially, though diesel's primary 20th-century foothold remained in industrial, marine, and heavy transport sectors due to higher torque and longevity. In aviation, lightweight gasoline engines drove early powered flight and subsequent commercialization. The Wright brothers' 1903 Flyer employed a custom 12-horsepower, water-cooled inline-four engine weighing 180 pounds, achieving sustained flight and validating internal combustion for aircraft propulsion. World War I accelerated development, with engines like the French Gnome rotary producing up to 100 horsepower by 1910s production runs, equipping thousands of fighters and bombers. Post-war, these advancements supported commercial air travel, as radial and inline configurations powered passenger aircraft from the 1920s onward, with global aviation engine output scaling dramatically by mid-century.

Modern Developments and Adaptations

Recent advancements in internal combustion engine (ICE) technology have focused on enhancing thermal efficiency through innovations such as direct fuel injection, turbocharging, and variable valve timing, which reduce fuel consumption while maintaining performance. These modifications enable engine downsizing, where smaller displacement engines deliver equivalent power via forced induction, achieving up to 10% greater efficiency gains by 2025 compared to prior regulatory baselines. Commercial diesel engines have reached thermal efficiencies exceeding 53%, with research indicating potential for further increases through optimized combustion strategies. Advanced combustion modes, including homogeneous charge compression ignition (HCCI), represent a significant adaptation by promoting lean-burn auto-ignition without spark plugs, yielding improvements in fuel economy of up to 13.4% alongside reduced NOx and particulate emissions. HCCI integrates with turbocharging to expand operational range, though challenges in precise ignition timing control persist, limiting widespread adoption to hybrid-assisted systems. Such technologies underscore causal links between mixture homogeneity and combustion efficiency, prioritizing empirical combustion dynamics over unsubstantiated regulatory assumptions. Hybridization has adapted ICEs for electrified powertrains, incorporating mild-hybrid systems with electric assist for stop-start functionality and regenerative braking, which mitigate urban emissions without full battery reliance. These configurations achieve net-zero carbon potential when paired with renewable fuels, positioning ICEs as scalable for sectors resistant to full electrification, such as heavy-duty transport. Adaptations for alternative fuels include compatibility with e-fuels and , leveraging ICE's fuel-agnostic design to utilize synthetic hydrocarbons produced via carbon capture, thereby reducing lifecycle emissions without engine redesign. combustion in modified ICEs lowers carbonaceous outputs to inherent oxygen , though NOx requires advanced controls. These developments affirm ICE viability amid decarbonization, grounded in verifiable metrics rather than ideologically driven phase-outs.

Configurations and Cycles

Stroke and Valve Mechanisms

The four-stroke cycle, predominant in automotive and aviation reciprocating internal combustion engines, involves four distinct piston strokes over two crankshaft revolutions: intake, compression, power, and exhaust. During the intake stroke, the piston descends from top dead center (TDC) to bottom dead center (BDC), creating a vacuum that draws the air-fuel mixture into the cylinder through the open intake valve while the exhaust valve remains closed. The compression stroke follows, with the piston ascending to TDC, compressing the mixture to increase its temperature and pressure, both valves closed to maintain seal integrity. Ignition then occurs near TDC, initiating the power stroke where expanding combustion gases drive the piston to BDC, delivering torque to the crankshaft. Finally, the exhaust stroke sees the piston rise to TDC, expelling burned gases through the open exhaust valve. Valve mechanisms in four-stroke engines regulate these flows via poppet valves—mushroom-shaped components seated in the or —that open and close precisely to optimize and minimize . The , rotating at half crankshaft speed via timing , , or , features eccentric lobes that actuate valves through followers, , or tappets; springs valves to . valves typically open 10-30 degrees before TDC on the exhaust stroke and close 40-70 degrees after BDC on for overlap, enhancing scavenging and filling at high speeds, while exhaust valves open near BDC on power stroke. In overhead valve () designs, the resides in the , using pushrods and for remote actuation; overhead cam () configurations place it in the head for shorter, stiffer paths enabling higher RPM, with () or () cams handling and exhaust independently for profiles. Two-stroke engines complete the cycle in one crankshaft revolution via two strokes—upward compression/power and downward intake/exhaust—yielding higher power density but lower efficiency due to port timing and charge dilution. Piston motion uncovers intake and exhaust ports in the cylinder wall, obviating poppet valves; scavenging relies on tuned exhaust pulses or crankcase compression to direct fresh charge upward, displacing exhaust. Some designs employ reed or rotary disc valves for intake control, but poppet valves are rare owing to mechanical complexity at high frequencies. Compared to four-strokes, two-strokes fire every revolution, boosting mean effective pressure but increasing emissions from incomplete combustion and oil mixing. Valve timing precision is critical, with deviations causing valve-piston contact; historical fixed profiles suited constant-speed operation, but modern (VVT) systems—emerging commercially in the —adjust , , and via hydraulic or electric actuators for torque across RPM ranges, improving by 5-10% in tests. Early VVT traces to 19th-century steam adaptations, but internal combustion applications prioritized reliability until electronic controls enabled it.

Primary Thermodynamic Cycles

The primary thermodynamic cycles governing reciprocating internal combustion engines are the Otto cycle for spark-ignition gasoline engines and the Diesel cycle for compression-ignition diesel engines. These air-standard models idealize engine operation by assuming a closed system with air as the working fluid, reversible processes, and no friction or heat transfer losses, providing a baseline for efficiency predictions despite real-world deviations from irreversibilities and variable composition during combustion. The Otto cycle consists of four processes: isentropic compression from intake pressure to peak compression, constant-volume heat addition via spark ignition, isentropic expansion driving the piston, and constant-volume heat rejection during exhaust. Nikolaus August Otto patented the first practical four-stroke engine realizing this cycle in 1876, enabling compression ratios typically between 8:1 and 12:1 limited by knock from premixed fuel-air charge. The ideal thermal efficiency is η = 1 - (1/r)^{γ-1}, where r is the compression ratio and γ ≈ 1.4 for air; for r=10, this yields about 60% theoretically, though practical brake thermal efficiencies range 25-35% due to pumping losses and incomplete combustion. The Diesel cycle features isentropic compression, constant-pressure heat addition from fuel injection into hot compressed air causing autoignition, isentropic expansion, and constant-volume heat rejection. Rudolf Diesel patented his engine in 1892, achieving higher compression ratios of 14:1 to 25:1 without knock risk, as fuel is introduced post-compression. Ideal efficiency is η = 1 - (1/r)^{γ-1} \cdot \frac{ρ^γ - 1}{γ(ρ - 1)}, where ρ is the cutoff ratio (volume at end of heat addition over compression volume); this generally exceeds Otto efficiency for equivalent r, with practical diesel engines attaining 40-50% brake thermal efficiency.
CycleIgnition TypeHeat Addition ProcessTypical Compression RatioTheoretical Efficiency BasisPractical Efficiency Range
OttoSparkConstant volume8:1–12:11 - (1/r)^{γ-1}25–35%
DieselCompressionConstant pressure14:1–25:11 - (1/r)^{γ-1} \cdot \frac{ρ^γ - 1}{γ(ρ - 1)}40–50%
A variant, the dual cycle, approximates modern high-speed diesels with both constant-volume and constant-pressure phases, bridging and models for better in finite combustion rates, though and remain foundational for .

Advanced and Experimental Cycles

The modifies the by employing a than , typically achieved through late closing, which reduces pumping losses and enhances at the expense of output. This , originally patented in 1882, allows for better utilization of combustion energy, with modern implementations in engines demonstrating brake thermal efficiencies up to 41% under part-load conditions. The Miller cycle, patented by Ralph Miller in the mid-20th century, similarly delays closing but incorporates supercharging or turbocharging to compensate for reduced volumetric efficiency, enabling higher overall while maintaining efficiency gains; experimental studies on turbocharged gasoline engines have shown reductions in fuel consumption by 5-10% compared to conventional cycles. Homogeneous Charge Compression Ignition (HCCI) represents an experimental low-temperature combustion strategy that combines premixed gasoline-like charge preparation with diesel-like compression ignition, autoigniting the homogeneous mixture without a spark to achieve diesel-level efficiencies (around 45-50% indicated thermal efficiency) and near-zero NOx and soot emissions. Challenges include narrow operating ranges limited by knocking at high loads and misfire at low loads, with research focusing on variable valve timing and exhaust gas recirculation to extend controllability; prototype engines have demonstrated up to 15% better fuel economy than port-fuel-injected Otto engines in steady-state tests. Reactivity Controlled Compression Ignition (RCCI), a dual-fuel variant using high-reactivity diesel for direct injection and low-reactivity gasoline for port injection, stratifies reactivity gradients to precisely control ignition timing and heat release, yielding NOx and soot reductions over 90% relative to conventional diesel while targeting 55-60% thermal efficiency in heavy-duty engines. Drive-cycle simulations indicate 7% fuel economy improvements over diesel with aftertreatment, though real-world implementation requires advanced controls for cycle-to-cycle variability. Novel experimental configurations include six-stroke cycles, which extend the four-stroke process with additional strokes for secondary combustion or water injection to recover exhaust heat, potentially reducing fuel consumption by 30-40% and emissions through cooler operation; recent patents, such as Porsche's 2024 design incorporating planetary gear mechanisms for extra compression and power phases, aim to integrate these into production hybrids. Free-piston engines eliminate the crankshaft, allowing linear piston motion driven by combustion and returned by gas springs or linear alternators, simplifying mechanics and enabling variable compression ratios for efficiencies potentially exceeding 50% in generator applications. Prototypes have achieved stable operation with two-stroke cycles, though challenges persist in precise motion control and durability without traditional bearings. These cycles remain largely developmental due to control complexities and integration hurdles, with empirical data underscoring their potential for efficiency gains amid regulatory pressures on emissions.

Ancillary Systems

Ignition and Starting Mechanisms

Ignition mechanisms in internal combustion engines initiate combustion of the air-fuel mixture through either spark ignition or compression ignition. Spark ignition engines, typically gasoline-powered, premix fuel and air in the intake manifold or cylinder, then generate an electric spark across a spark plug gap to ignite the mixture near the end of the compression stroke. Compression ignition engines, such as diesels, compress only air to high temperatures (around 500-700°C), then inject fuel directly into the hot compressed air, causing auto-ignition without a spark. Early spark ignition systems relied on magneto generators, which produce high-voltage pulses mechanically driven by the engine crankshaft, eliminating the need for batteries. Magnetos, common in pre-1920s engines and aviation, use permanent magnets rotating near coils to induce current, stepped up via interrupters or transformers for spark timing. By the 1910s, battery-coil systems emerged, with Charles Kettering's 1910 Delco ignition using a low-tension magneto for initial spark and battery for consistent operation. Breaker-point distributors, introduced in the early 1900s, mechanically timed sparks via cam-driven contacts opening to collapse magnetic fields in ignition coils, producing 20,000-40,000 volts. Electronic ignition systems, popularized from the 1970s, replaced mechanical points with solid-state switches or Hall effect sensors for precise timing, reducing wear and enabling higher RPM operation. Distributorless systems, introduced in the 1980s, use crankshaft position sensors and engine control units to fire multiple coil-on-plug units directly, improving efficiency and emissions control. Compression ignition avoids dedicated ignition hardware, relying on precise fuel injection timing and glow plugs for cold starts to aid vaporization and initial combustion. Starting mechanisms cranks the engine to achieve initial rotation for self-sustaining combustion. Prior to electric starters, hand-cranking via a protruding crankshaft handle was standard from the late 1800s, risking injury from kickback if timing misfired. Charles Kettering developed the first practical electric starter in 1911, featuring a small DC motor with reduction gears to torque-multiply battery power for cranking, debuting on the 1912 Cadillac. This solenoid-engaged system, refined with Bendix's 1910 pre-engaged pinion drive to mesh gears before full motor torque, became ubiquitous by the 1920s, enabling reliable starts regardless of weather or user strength. Modern starters integrate with start-stop systems for frequent cycling, using higher-efficiency motors and lithium batteries, while diesels often employ glow plugs and higher cranking speeds (200-300 RPM) to build compression heat. Alternative methods, like compressed air starters in heavy-duty or marine applications, inject air to spin pistons, avoiding electrical dependency in hazardous environments.

Forced Induction Technologies

Forced induction technologies enhance the power output of internal combustion engines by compressing the intake air, thereby increasing its density and enabling greater fuel combustion within the cylinders without enlarging the engine displacement. This process counters the limitations of naturally aspirated engines, where air intake is restricted by atmospheric pressure, typically around 1 bar at sea level, resulting in volumetric efficiencies below 100%. By boosting intake pressure to 1.5–3 bar or higher, forced induction achieves power densities up to 50% greater than comparable naturally aspirated designs, while also improving fuel efficiency in downsized engines through better thermodynamic efficiency. Superchargers, the earliest form of forced induction, are mechanically driven compressors powered directly by the engine's crankshaft via belts, gears, or chains, imposing a parasitic load that reduces net efficiency by 10–20% at full boost. The Roots-type supercharger, patented in 1867 by Philander and Francis Roots, uses two lobed rotors to trap and displace air, achieving boost ratios up to 2:1 but with notable inefficiency due to internal leakage and heat generation. Centrifugal superchargers, akin to turbo compressor wheels, spin at 50,000–100,000 rpm to impart kinetic energy to air via impeller blades, offering higher efficiency (70–80%) and adiabatic compression but requiring higher engine speeds for peak boost. Screw-type superchargers, developed in the 1930s by companies like Lysholm, employ intermeshing helical rotors for near-isentropic compression with minimal pulsation, commonly used in high-performance applications for their compact design and boost delivery from low rpm. Early adoption occurred in aviation during World War I, with Mercedes equipping aircraft engines like the D.IIIa in 1917, yielding 20–30% power gains at altitude. Turbochargers, exhaust-gas-driven devices, recover waste energy from the engine's turbine to drive a compressor, eliminating mechanical parasitic losses and enabling efficiencies exceeding 35% in modern units, compared to superchargers' 15–25%. Swiss engineer Alfred Büchi patented the turbocharger concept in 1905 for diesel engines, with the first operational prototype tested in 1915 on a diesel locomotive, though commercial viability emerged in the 1920s for marine applications. The first automotive use appeared in 1938 on Saurer trucks, boosting diesel efficiency by reducing specific fuel consumption by up to 20%. In gasoline engines, widespread adoption began post-World War II, with Oldsmobile's 1962 Jetfire featuring a turbo for 230 kW from a 3.5 L displacement, though early designs suffered from detonation issues requiring water-methanol injection. Turbo lag, the delay in boost buildup due to inertial spool-up (typically 0.5–2 seconds), remains a drawback, mitigated by variable-geometry turbines (VGT) introduced in the 1980s for diesels, which adjust vane angles to optimize exhaust flow across rpm ranges, improving low-end torque by 30–50%. Twin-scroll turbos, dividing exhaust pulses for reduced interference, further enhance transient response in multi-cylinder engines. Hybrid systems like twincharging combine superchargers for instant low-rpm with turbos for high-rpm , as in Volvo's T5 engine, delivering 160 kW from 2.0 L with minimal . Electric-assisted turbos, emerging in the , use to preemptively spool compressors, reducing lag to under 0.1 seconds and 48V mild- for overall gains of 5–10% in transient . Despite benefits, forced increases component stresses, necessitating reinforced pistons, , and cooling, with risks of knock in engines requiring higher-octane fuels or retarded timing, which can cut by 10–15%. Diesel applications dominate due to inherent ratios (16:1–22:1) tolerating without pre-ignition, achieving brake thermal over 40% in turbo-diesel trucks.

Cooling, Lubrication, and Valvetrain

Internal combustion engines generate substantial heat from , necessitating effective cooling to avoid damage, maintain material integrity, and optimize ; typical operating temperatures range from 80–110°C for cylinder walls and up to °C for exhaust valves. Two primary cooling methods exist: , which relies on fins on cylinder heads and barrels with forced from fans or vehicle motion, and cooling, predominant in automotive and high-performance applications for superior uniformity. systems circulate a mixture—typically 50/50 water and —through water jackets in the block and heads via a centrifugal pump driven by the crankshaft, with a radiator dissipating heat via and a thermostat modulating flow to achieve rapid warmup and stable temperatures. , used in early motorcycles, aircraft like the radial engines of World War II, and select vehicles such as the Volkswagen Beetle until 2003, offers simplicity, lighter weight, and no freeze risk but struggles with even cooling in multi-cylinder setups, limiting power density. Lubrication systems minimize friction between moving parts, remove wear debris, and provide secondary cooling by absorbing and dissipating heat from bearings and pistons, with engine oils formulated to withstand shear forces and temperatures up to 150°C in critical areas. Modern engines employ full-force pressure-feed lubrication, where a gear or gerotor pump draws oil from a wet sump (pan below the crankcase) or dry sump (remote reservoir for racing) and delivers it at 2–6 bar to crankshaft main and rod bearings, camshaft lobes, and cylinder walls via drilled passages, with splash lubrication aiding pistons and timing chains. Oil filters remove contaminants, while additives like detergents and anti-wear agents (e.g., zinc dialkyldithiophosphate) enhance longevity; synthetic oils, introduced commercially in the 1970s, offer better viscosity stability across temperatures compared to mineral-based ones. Dry sump systems, standard in high-performance engines since the 1930s, prevent oil starvation under high-g cornering by scavenging oil back to a reservoir. ![Overhead cam engine with forced oil lubrication (Autocar Handbook, 13th ed., 1935)](./assets/Overhead_cam_engine_with_forced_oil_lubrication_Autocar_Handbook%252C_13th_ed%252C_1935 The valvetrain manages the opening and closing of intake and exhaust valves to control air-fuel mixture entry and combustion byproduct expulsion, with components including the camshaft, lifters, pushrods (in overhead valve designs), rocker arms, valves, keepers, and springs that return valves to seats at speeds up to 20,000 cycles per minute in high-revving engines. Overhead valve (OHV) or pushrod systems position the camshaft in the block, using long pushrods and rockers for valve actuation, enabling compact packaging and lower production costs but incurring higher inertial losses and limited maximum RPM around 6,000–7,000. Overhead camshaft (OHC) configurations place the cam directly above valves in the head—single (SOHC) actuating both intake and exhaust via rockers, or dual (DOHC) with separate cams for each—reducing mechanical complexity, enabling shorter valve timing durations, and supporting RPMs exceeding 8,000, as in many post-1980s automotive engines. DOHC designs facilitate variable valve timing (VVT), introduced by Honda in 1989 with VTEC, which hydraulically adjusts cam phasing or lift to optimize low-end torque and high-end power, improving efficiency by 5–10% through better volumetric efficiency and reduced pumping losses.

Fuels and Additives

Hydrocarbon and Conventional Fuels

fuels for internal combustion engines consist primarily of petroleum-derived mixtures of alkanes, cycloalkanes, alkenes, and aromatics, which combust with oxygen to release energy through exothermic producing , , and . These conventional fuels, refined from crude via and cracking, dominate applications due to their high , portability, and compatibility with engine designs optimized for liquid injection and . and account for the majority of usage, with global petroleum for transportation exceeding 60% of total as of 2018. Gasoline, employed in spark-ignition engines, is a volatile blend of hydrocarbons typically spanning C4 to C12 chains, with a boiling range of 32–210 °C to facilitate carburetion or fuel injection. Its key performance metric, the octane rating (anti-knock index or AKI), quantifies resistance to premature auto-ignition under compression; regular grade in the U.S. averages 87 AKI, while premium reaches 91–93 AKI, determined via standardized engine tests comparing to iso-octane and n-heptane blends. Energy content approximates 114,000–125,000 BTU per gallon, though blending with 10% ethanol reduces this by 3–4%. Diesel fuel, suited for compression-ignition engines, features heavier C9–C20 hydrocarbons, yielding superior volumetric energy density at roughly 129,000–138,000 BTU per gallon—113% higher than gasoline equivalents. The cetane number, measuring ignition quality via delay time in a CFR test engine, ranges from 40 to 55 for on-road grades, with higher values from straight-chain paraffins promoting smoother combustion and reduced noise. Refining specifications, such as ASTM D975, limit sulfur to 15 ppm in ultra-low sulfur diesel to minimize emissions while preserving lubricity. Other conventional fuels include (C10–C16 blends) for certain and aviation engines, offering a around 45 and above 38 °C for , and liquefied gases (LPG) like propane-butane mixtures for dual-fuel or dedicated engines, with densities of 91,000 BTU per equivalent. These fuels' stems from their and tunable , though variations in affect ; for instance, aromatic content in inversely correlates with .

Alternative Fuels Including Synthetics and Hydrogen

Alternative fuels for internal combustion engines encompass a range of non-conventional options designed to reduce reliance on petroleum-derived hydrocarbons, including alcohols such as and , (CNG), (LPG), and biofuels like , alongside synthetic hydrocarbons and . These fuels often require minimal or no engine modifications for compatibility, though performance varies; for instance, alcohols offer higher ratings for spark-ignition engines but lower , leading to reduced unless compensated by larger . Gaseous fuels like CNG provide cleaner with lower (PM) and (CO) emissions compared to gasoline, achieving up to 20-30% reductions in some tests, but infrastructure limitations persist. Synthetic fuels, or e-fuels, are chemically identical drop-in replacements for gasoline and diesel, produced via processes like Fischer-Tropsch synthesis from syngas (carbon monoxide and hydrogen) derived from biomass, natural gas, or captured CO2 combined with green hydrogen from electrolysis. Fischer-Tropsch diesel exhibits a cetane number exceeding 70—higher than conventional diesel's 40-55—enabling superior ignition quality, reduced PM emissions by up to 50%, and lower unburned hydrocarbons (HC) without engine alterations; sulfur and aromatic content near zero further minimize SOx and soot formation. However, lifecycle CO2 neutrality depends on renewable feedstocks; fossil-derived syngas undermines this, and overall well-to-wheel efficiency remains low at 10-20% due to electrolysis and synthesis losses, compared to 70-90% for battery electric vehicles. NOx emissions mirror those of fossil fuels, as combustion temperatures and oxygen content are similar, limiting air quality benefits. Production costs exceed 3-4 euros per liter as of 2030 projections, constraining scalability despite compatibility with existing infrastructure. Hydrogen combustion in internal combustion engines involves direct burning of H2 in modified spark-ignition or compression-ignition setups, yielding near-zero CO2 emissions—limited to trace amounts from lubricating oil—but generating elevated NOx due to flame temperatures reaching 2500 K, potentially 2-6 times higher than hydrocarbon fuels without advanced controls like lean-burn or exhaust gas recirculation. Engine modifications include reinforced pistons for preignition resistance, cryogenic storage or high-pressure tanks, and optimized injection to manage backfiring; port fuel injection yields smoother operation and lower NOx than direct injection but at the cost of power density. Brake thermal efficiency can reach 35-40% in prototypes, surpassing gasoline engines' 25-30%, with blending ratios up to 20% H2 in gasoline reducing CO and HC by 20-50% while slightly boosting torque. Drawbacks include reduced volumetric power output—up to 20% lower without turbocharging—and NOx mitigation challenges, as catalytic converters are less effective than in fuel cells, which achieve 50-60% efficiency without combustion byproducts. Demonstration engines, such as Cummins' 6.7L prototypes, validate feasibility for heavy-duty applications but highlight NOx as the primary pollutant, requiring aftertreatment like selective catalytic reduction.

Oxidizers and Combustion Enhancers

In conventional internal combustion engines, atmospheric air serves as the primary oxidizer, supplying oxygen at approximately 21% concentration by volume to facilitate the exothermic oxidation of fuel hydrocarbons during combustion. This oxygen reacts with fuel molecules to produce carbon dioxide, water, and heat, with nitrogen from air acting primarily as a diluent to moderate flame temperatures and prevent excessive thermal stress on engine components. Nitrous oxide (N₂O), when injected into the manifold, functions as a chemical oxidizer and enhancer, decomposing endothermically above °C (572°F) into gas and nascent oxygen, thereby increasing the oxygen available for oxidation beyond that provided by ambient air. This augmentation allows for a richer -air , typically yielding 50-100% instantaneous increases in spark-ignition engines, as demonstrated in applications where systems deliver 100-500 horsepower equivalents for bursts of 10-15 seconds. However, prolonged use risks , component fatigue, and elevated pressures exceeding 200 , necessitating engine reinforcements and precise enrichment to maintain air- ratios near stoichiometric levels (around 14.7:1 for gasoline). Nitrous oxide systems, popularized since the 1940s in aviation and later in drag racing, are not viable for continuous operation due to storage limitations—N₂O is liquefied at -88°C or 50 —and its contribution to nitrous oxide emissions, a greenhouse gas with 298 times the global warming potential of CO₂ over 100 years. Oxygen-enriched combustion, involving intake air augmented to 23-30% oxygen via separation technologies like pressure swing adsorption or membrane permeation, has been experimentally applied to enhance thermal efficiency by 5-10% and reduce unburned hydrocarbons and particulates in both gasoline and diesel engines. Studies on single-cylinder engines show peak torque improvements of up to 15% at oxygen levels of 25%, attributed to faster flame propagation and complete fuel oxidation, though adiabatic flame temperatures rise by 200-300°C, amplifying nitrogen oxide (NOx) formation by factors of 2-4 via the Zeldovich mechanism. Practical implementation remains limited to stationary or industrial ICEs, such as diesel generators, due to the energy-intensive oxygen production (requiring 0.2-0.3 kWh per kg O₂) and risks of pre-ignition or material oxidation; automotive adoption is constrained by system complexity and NOx aftertreatment demands. Chemical combustion enhancers, including alkyl nitrates (e.g., 2-ethylhexyl ) and metal-based catalysts like nanoparticles, promote oxidation by lowering ignition and stabilizing fronts in diesel fuels, enabling 1-3% in emissions and improved cold-start . In methanol-fueled compression-ignition engines, such additives facilitate dual-fuel operation by accelerating autoignition, with cetane numbers boosted from near-zero to 40-50, though they introduce trace byproducts requiring catalytic converters. Niche applications, such as injection at 1-5% by , have demonstrated 20-30% in engines by enhancing OH formation for faster carbon oxidation, but scalability is hindered by costs and byproduct peroxides. Overall, these enhancers prioritize or emissions trade-offs over primary oxidizer , with verified through testing rather than deployment.

Performance Evaluation

Efficiency and Power Metrics

Brake (BTE) measures the of fuel's converted into useful output, calculated as BTE = ( ) / (fuel mass flow rate × fuel lower heating value). This metric accounts for real-world losses including incomplete combustion, to and exhaust, pumping losses during , and . Typical BTE for spark-ignition gasoline engines ranges from 25% to 35% under optimal conditions, limited by lower ratios (around 10:1 to 12:1) to avoid knock. Compression-ignition diesel engines achieve higher BTE of 35% to 45%, benefiting from higher ratios (14:1 to 25:1) that enable leaner burns and reduced rejection during . Advanced diesel engines have pushed BTE boundaries; Weichai Power's 2024 prototype attained 53.09% intrinsic through optimized bowl design, low-friction materials, and precise timing to minimize heat losses and maximize combustion completeness. Similarly, Mercedes-AMG's 1.6-liter turbocharged 1 engine reached over 50% BTE in 2017 via high compression, , and anti-lag turbo systems that recover exhaust energy. engines lag behind but have improved; Nissan's e-POWER series engine hit 50% BTE in 2019 by integrating with electric supercharging to enhance ratios while controlling peak pressures. These peaks contrast with average production values, where diesels offer 20-35% better economy than counterparts due to inherent cycle advantages and higher energy density of . Power metrics quantify output per engine attribute, independent of size. Brake mean effective pressure (BMEP) represents the average pressure required to produce measured , derived as BMEP = (brake × 4π) / for four-stroke engines, serving as a design yardstick. High-performance naturally aspirated engines achieve BMEP around 10-12 , while turbocharged diesels exceed 20 through and intercooling that charge density without excessive . Specific , output per (kW/L), highlights volumetric ; modern turbo gasoline engines reach 100-150 kW/L via direct injection and variable geometry turbos, surpassing diesels' 50-80 kW/L due to diesels' emphasis on over peak .
Engine TypeTypical BTE (%)Peak BMEP (bar)Specific Power (kW/L)
Gasoline (SI)25-3510-1580-150
Diesel (CI)35-4515-2550-80
Advanced/Record>50>20>100
Factors elevating these metrics include higher compression for thermodynamic gains per the Otto or Diesel cycle formulas—η = 1 - (1/r)^{γ-1} for ideal Otto, where r is compression ratio and γ is specific heat ratio—but practical limits arise from material strength, NOx formation, and fuel autoignition. Reduced friction via roller bearings and honed cylinders, along with waste heat recovery, further boosts net output, though gains diminish at part-load where throttling or pumping losses dominate.

Fuel Consumption Measures

Brake specific (BSFC) serves as the primary for evaluating in internal combustion engines, defined as the mass of consumed per unit of output produced over a given time . It is calculated as BSFC = ( rate in g/h) / ( in kW), yielding units of grams per kilowatt-hour (g/kWh), where lower values indicate superior efficiency. This measure isolates engine performance by focusing on shaft output, excluding downstream losses in vehicle applications. Typical BSFC values for spark-ignition engines range from to g/kWh under optimal load and speed conditions, reflecting their lower ratios and reliance on spark timing for . In contrast, -ignition engines achieve 190 to 220 g/kWh, benefiting from higher efficiencies to elevated ratios exceeding 14:1 and leaner air-fuel mixtures. Modern turbocharged diesels can approach 170 g/kWh at , though real-world often exceeds these minima to transient loads and part-throttle inefficiencies. For vehicular applications, fuel consumption is alternatively expressed in distance-specific units such as miles per gallon (MPG) for imperial systems or liters per 100 kilometers (L/100 km) for metric, integrating engine output with transmission, aerodynamics, and rolling resistance. These metrics are derived from standardized dynamometer or chassis tests, such as the U.S. EPA's Federal Test Procedure (FTP-75) for city driving or the Highway Fuel Economy Test (HWFET), which simulate real-world cycles but yield combined ratings like 25-35 MPG for efficient gasoline sedans and 40-50 MPG for diesels. The Worldwide Harmonized Light Vehicles Test Procedure (WLTP), adopted globally since 2017, provides more representative values by incorporating higher speeds and accelerations, often resulting in 10-20% lower MPG estimates than older NEDC protocols. BSFC contours, plotted against engine speed and torque, reveal efficiency islands where minimum consumption occurs, typically near 75-100% load for diesels and mid-range RPM for both types, guiding control strategies like variable valve timing to sustain these regimes. Indicated specific fuel consumption (ISFC), measured at the cylinder before mechanical losses, is higher than BSFC by 10-20% in gasoline engines, highlighting frictional and pumping inefficiencies. These measures enable cross-engine comparisons, with diesel designs consistently outperforming gasoline by 20-30% in BSFC due to thermodynamic advantages, though hybridization and turbocharging narrow gaps in modern spark-ignition variants.

Losses and Optimization Strategies

The primary energy losses in internal combustion engines (ICEs) arise from incomplete , heat transfer to coolant and exhaust, , and gas pumping work, collectively accounting for 60-80% of the fuel's being dissipated without producing useful work. In a typical gasoline engine, carries away 30-40% of the as losses, while 25-35% is rejected via the cooling system due to from gases to walls and components. losses, including ring- interactions and bearing , consume 5-10% of indicated , with higher shares (up to 20-30% of losses) from ancillary components like and fuel pumps. Pumping losses, resulting from throttling and exhaust backpressure in throttled spark-ignition engines, can represent 5-15% of total losses, exacerbating inefficiency at part-load conditions common in vehicular operation.
Loss TypeTypical Percentage of Fuel EnergyPrimary Causes
Exhaust Thermal30-40%High-temperature gas expulsion before full expansion
Cooling/Heat Transfer25-35%Conduction from to walls and fluids
Mechanical Friction5-10%, rings, valves, and bearings
Pumping5-15%Throttling and mismatches
Incomplete Combustion/Chemical2-5%Unburned hydrocarbons and
Optimization strategies target these losses through design refinements and control systems. To mitigate heat transfer losses, strategies include retarding combustion phasing to lower peak gas temperatures and heat transfer coefficients, potentially reducing wall heat losses by 10-20% while maintaining output, as demonstrated in controlled engine tests. Ceramic coatings on pistons and cylinder heads minimize conductive losses by reflecting heat back into the , with studies showing 2-5% gains in engines operating at higher compression ratios. Friction reduction employs low-viscosity synthetic lubricants, (DLC) coatings on rings and pistons, and rollerized cam followers, achieving up to 20% lower frictional (FMEP) in modern engines compared to baseline designs from the 1990s. Pumping losses are addressed via (VVT) and lift systems, such as continuous VVT or Atkinson-cycle operation, which reduce throttling by optimizing intake/exhaust overlap and effective compression, yielding 5-10% fuel economy improvements in production vehicles under part-load. Advanced combustion modes like (HCCI) or stratified further minimize incomplete losses by promoting more complete fuel oxidation at lower temperatures, though challenges in control limit widespread adoption to specific operating regimes. Integrated approaches, such as engine downsizing paired with high-efficiency turbocharging, amplify gains by operating at higher loads where losses scale less adversely, with real-world implementations in passenger cars demonstrating 15-20% reductions in specific fuel consumption since 2010. recovery via exhaust gas heat exchangers or thermoelectric generators recaptures 5-10% of thermal losses for , as validated in heavy-duty prototypes, though system complexity and backpressure penalties constrain commercial viability. These strategies, grounded in empirical testing, underscore that while theoretical Carnot limits cap efficiency below 60% under practical constraints, iterative reductions in losses have incrementally raised peak brake thermal efficiencies to over 50% in laboratory engines by 2022.

Applications and Implementations

Automotive and Light Vehicle Uses

The internal combustion engine (ICE) powers the majority of automobiles and light vehicles, providing on-demand propulsion through controlled combustion of fuels. The first practical application in a self-propelled road vehicle occurred in 1885, when Karl Benz fitted a single-cylinder, four-stroke engine producing 0.75 horsepower to a three-wheeled , achieving speeds up to 10 . This design demonstrated the feasibility of liquid-fueled mobility, offering higher energy density than batteries or steam systems of the era, with refueling via simple fuel transfer rather than lengthy boiling or charging processes. By the early , by Henry Ford's Model T from 1908 onward standardized the four-stroke ICE, enabling widespread adoption for personal transport due to its reliability and scalability. In contemporary passenger cars, sport utility vehicles (SUVs), and light trucks, spark-ignited gasoline engines operating on the Otto cycle remain dominant, typically featuring multi-cylinder configurations such as inline-four or V-six arrangements with displacements from 1.0 to 6.0 liters. These deliver power outputs ranging from 80 horsepower in compact economy models to over 700 horsepower in high-performance variants, with thermal efficiencies improved to 30-35% through technologies like variable valve timing, direct fuel injection, and turbocharging. Compression-ignition diesel engines, prevalent in Europe until the 2010s, provide superior torque and fuel economy—up to 40% thermal efficiency—for diesel light-duty applications, though their new sales share fell below 10% in the EU by 2023 amid stringent nitrogen oxide regulations following emissions test discrepancies revealed in 2015. Globally, ICE-equipped vehicles, encompassing conventional, hybrid, and plug-in hybrid variants, comprised 78% of new light-duty vehicle sales in 2024, far outpacing battery-electric models at 22%, as liquid fuels sustain longer ranges of 300-500 miles per tank versus 200-300 miles for most electrics. Hybrid powertrains integrate downsized s with electric motors to enhance overall , allowing the to operate at optimal loads during acceleration and highway cruising while relying on for low-speed urban driving; the , introduced in 1997, pioneered this approach, achieving combined fuel economies exceeding 50 in later models. Refueling advantages persist, with ICE vehicles replenishing in 3-5 minutes at ubiquitous stations, enabling seamless long-haul travel without the or extended charging times—often 30 minutes or more for 80% capacity—that constrain electric alternatives, particularly in rural or cold climates where performance degrades. U.S. light-duty fleet average fuel economy advanced from 24 in 2000 to 28 by 2024, driven by regulatory standards and engineering refinements reducing parasitic losses and improving combustion control. In motorcycles and smaller light vehicles, compact two-stroke or four-stroke ICEs predominate for their high power-to-weight ratios, with displacements under 1.0 liter suiting urban commuting and off-road use, though two-strokes face phase-outs in some markets due to higher emissions per unit of power. ICE reliability in automotive contexts stems from decades of , yielding exceeding 200,000 miles in modern units, supported by standardized fuels and service networks that outstrip emerging electric in coverage and speed.

Heavy-Duty, Marine, and Aviation Applications

Heavy-duty internal combustion engines, predominantly variants, power trucks, buses, and construction equipment due to their superior output and compared to counterparts. These engines typically range from 400 to 600 horsepower with peak torque between 1,650 and 2,050 pound-feet, enabling effective hauling of heavy loads. For instance, ' 15-liter and 12-liter advanced engines are widely used in demanding applications like long-haul trucking, offering reliability under high loads. In construction, models such as Diesel's 60 Series drive equipment like excavators and generators, with power outputs suited for continuous operation. Marine applications favor large low-speed two-stroke diesel engines for their high power density and efficiency in propelling ships. The Wärtsilä-Sulzer RTA96-C, a 14-cylinder two-stroke engine introduced in 2006, represents the largest reciprocating type, delivering 80.08 megawatts (approximately 107,390 horsepower) at 102 rpm while weighing over 2,300 tons. These engines achieve thermal efficiencies up to 50% when burning heavy fuel oil, powering container ships like the Emma Mærsk. Four-stroke diesels serve smaller vessels, often with turbocharging to boost output without added fuel costs, while two-strokes dominate deep-sea operations for their compact power delivery. Category 1 and 2 marine diesels range from 500 to 8,000 kilowatts (700 to 11,000 horsepower), providing auxiliary propulsion. In , reciprocating piston engines—primarily horizontally opposed configurations—remain the standard for , converting fuel's into mechanical power via piston reciprocation to drive propellers. These engines operate on four-stroke cycles, with cylinders arranged for and reduced in small planes, unlike radial types in older designs. They suit low-speed flight below 20,000 feet, prioritizing reliability over the higher thrust of turbines used in commercial jets.

Stationary and Industrial Power Generation

![Montreal power backup generator installation][float-right]
Reciprocating internal combustion engines serve as a primary technology for stationary power generation in industrial and commercial settings, providing reliable electricity for prime power, peaking, and backup applications. These engines, operating on the Otto or Diesel cycles, convert chemical energy from fuels such as diesel, natural gas, or biogas into mechanical power that drives generators. In combined heat and power (CHP) systems, they achieve overall efficiencies exceeding 70% by recovering waste heat for thermal uses like steam production or space heating.
Historically, stationary internal combustion engines emerged in the late 19th century, with Nikolaus Otto's four-stroke patented in enabling initial industrial applications, though early units produced limited power and competed with steam engines. Rudolf Diesel's compression-ignition engine, introduced in 1897, gained prominence for stationary use by the 1920s due to its higher efficiency and ability to handle heavy loads, supplanting lower-pressure hot-bulb engines in high-power scenarios. By the mid-20th century, diesel and gas s powered factories, mines, and remote installations, with modern units scaling to capacities over 5 MW per engine. In the United States, over 2,000 CHP installations provide nearly 2.3 gigawatts of capacity as of recent assessments. In industrial contexts, these engines support continuous operations in sectors like , and gas, and , often fueled by digester gas or for on-site generation. Diesel generator sets, for instance, range from 6 kW to over 7,000 kVA, offering fuel efficiencies around 40% at 70-80% load, with consumption rates such as approximately 50 gallons per hour for a 500 kW unit at full load. systems, critical for data centers, hospitals, and utilities, leverage the engines' rapid startup—reaching full load in seconds to minutes—and black-start capability without external power. variants provide flexibility for baseload , complementing intermittent renewables by delivering dispatchable power with low emissions when equipped with aftertreatment. Advantages include modular , allowing multiple units for , and proven in harsh environments, with intervals supporting high uptime. In , reciprocating engines outperform turbines in part-load and multi-fuel adaptability, making them suitable for industrial demands. Despite regulatory pressures on emissions, advancements in turbocharging and maintain their viability for reliable, on-demand power where grid stability is paramount.

Environmental and Health Impacts

Emissions Profiles and Technological Reductions

Internal combustion engines (ICEs) primarily emit (CO₂), (NOₓ), (CO), unburned hydrocarbons (HC), and (PM) from incomplete combustion and high-temperature reactions. CO₂ arises from complete fuel oxidation, typically comprising 10-15% of exhaust volume, with modern gasoline passenger vehicles emitting around 120-150 g/km under standardized cycles like WLTP, while diesels emit slightly less (e.g., 109-139 g/km) due to higher . NOₓ forms from and oxygen at elevated temperatures (>1500°C), with diesels producing higher levels (often >50% of total pollutants, up to 8 g/kWh in unregulated heavy-duty cases) than gasoline engines due to operation. CO and HC result from incomplete combustion, more prevalent in gasoline engines (e.g., CO up to 35 g/kg fuel in some diesels but generally higher in rich gasoline mixtures), while PM—solid carbon and sulfates—is dominant in diesels, contributing to respiratory risks. Gasoline engines exhibit higher CO and HC emissions but lower NOₓ and PM compared to diesels, reflecting stoichiometric combustion versus lean diesel operation; for instance, spark-ignited engines account for most HC and CO from mobile sources, while diesels dominate NOₓ and PM contributions. Unregulated emissions have declined dramatically due to regulatory standards: U.S. vehicles achieved 20-60-fold reductions in CO (to ~5 g/mile), NOₓ (~0.2 g/mile), and HC (~0.3 g/mile) over 50 years through technology, with new 2023 models emitting less than half the CO₂ per mile of 1975 counterparts and a 31% tailpipe drop since 2004. These profiles vary by load, fuel quality, and maintenance, with cold starts elevating HC and CO by factors of 10-100. Technological reductions target combustion control and aftertreatment. Exhaust gas recirculation (EGR) lowers NOₓ by 20-50% in diesels via cooled exhaust dilution, reducing peak flame temperatures without lean NOx traps. Three-way catalytic converters, mandated since the 1970s U.S. Clean Air Act, convert >90% of CO, HC, and NOₓ in gasoline engines under stoichiometric conditions using platinum, palladium, and rhodium. Diesel particulate filters (DPF) capture >90-99% PM via wall-flow ceramics with periodic regeneration, while selective catalytic reduction (SCR) injects urea (AdBlue) to achieve >90% NOₓ reduction via ammonia reactions over vanadium or zeolite catalysts. Advanced engine designs further mitigate emissions: direct injection and turbocharging improve efficiency (reducing CO₂ by 10-20%), stratified charge enables with lower HC/CO, and low-sulfur fuels (<15 ppm since 2006 EPA rules) enhance catalyst durability. Combined systems like EGR+SCR yield synergistic NOₓ cuts (e.g., 24-56% in transient cycles), enabling compliance with Euro 6/VI standards that slashed NOₓ by 35-56% from prior eras. These interventions, driven by mandates like the 1970 Clean Air Act's 90% reduction goal by 1975, demonstrate causal efficacy in curbing per-vehicle emissions despite rising vehicle miles traveled.
Emission TypeKey Reduction TechnologyTypical Effectiveness
NOₓEGR, SCR20-90%+
PMDPF90-99%
CO/HCCatalytic converters>90%
CO₂Efficiency improvements (e.g., turbo, injection)10-25% per tech iteration

Noise, Vibration, and Localized Pollution

Internal combustion engines generate significant primarily from , , mechanical impacts of pistons and valves, and airflow over components, with overall levels historically reaching around 100 (A) for early 20th-century industrial engines and modern automotive examples typically measuring 80 to 110 (A) at 1 meter from the engine surface. These levels contribute to (NVH) characteristics that affect occupant comfort and exterior , prompting regulatory limits such as the Union's phased reductions of 3 to 4 in standards between 2016 and 2026, equivalent to halving perceived . Vibration in internal combustion engines arises from reciprocating masses like pistons and connecting rods, cyclic pressure fluctuations, and rotational imbalances, transmitting forces that can degrade component , limit output, and propagate through structures to cause harshness felt by occupants. strategies include dynamic balancing of crankshafts, flexible engine mounts to isolate vibrations at the chassis interface, and viscous dampers on accessories, which reduce by absorbing torsional and linear oscillations without fully eliminating root causes tied to the engine's . Localized pollution from internal combustion engine exhaust encompasses high concentrations of ultrafine particles (UFPs, <0.1 micrometers), nitrogen oxides (NOx), carbon monoxide (CO), and unburned hydrocarbons immediately downwind of the source, with UFPs comprising 80 to 95% of diesel soot mass and capable of deep lung penetration due to their small size and high surface area for adsorbing toxins. These near-field emissions differ spatially from broader pollutants like CO2, forming elevated gradients near roadways—e.g., UFPs and NOx hotspots within tens of meters—that correlate with elevated cardiovascular and respiratory risks in proximate populations, as evidenced by studies linking ambient UFP exposure to systemic inflammation independent of larger particulate matter. Engine design factors, such as fuel injection timing and aftertreatment like particulate filters, influence UFP nucleation in cooling exhaust, but combustion inherently produces these nanoparticles, with diesel engines yielding higher yields than gasoline due to richer mixtures and soot formation pathways.

Full Lifecycle Assessments vs. Alternatives

Manufacturing of battery electric vehicles (BEVs) generates higher upfront greenhouse gas (GHG) emissions than internal combustion engine (ICE) vehicles, primarily due to energy-intensive battery production involving mining and refining of lithium, cobalt, nickel, and other materials; estimates indicate BEV production emissions are about 40% greater than those for comparable ICE vehicles, with batteries contributing 5-14 tons of CO₂-equivalent (tCO₂e) per vehicle depending on battery size. Operationally, BEVs produce zero tailpipe emissions, but their well-to-wheel (WTW) emissions depend heavily on electricity grid carbon intensity; in the average U.S. grid mix as of 2021-2023, BEV use-phase emissions range from 0.14-0.28 kg CO₂e per mile, often lower than ICE vehicles' combined WTW emissions of 0.30-0.40 kg CO₂e per mile for gasoline models. End-of-life recycling can offset some emissions for both, but battery recycling rates remain low (under 5% globally as of 2023), limiting credits to 1-2 tCO₂e savings. Cradle-to-grave LCAs, using models like Argonne National Laboratory's GREET, show BEVs achieving 46-52% lower total GHG emissions than vehicles over a 200,000-mile lifetime in the U.S., assuming a 2023-2024 grid mix and medium-sized sedans; for example, a BEV might emit 39 tCO₂e total versus 56 tCO₂e for an vehicle.
Vehicle TypeTotal Lifecycle GHG (tCO₂e, 240,000 km)Production (tCO₂e)Use Phase (tCO₂e)Key Assumption
BEV391426 (electricity-dependent)U.S./EU grid mix, 2021 data
(HEV)471036Blended gasoline-electric
561045, average efficiency
These advantages for BEVs erode in regions with coal-dominant grids (e.g., parts of or , where emissions parity may require over 100,000 miles or never occur), and sensitivity analyses reveal that optimistic assumptions about future grid decarbonization (projected 50-70% cleaner by 2035) drive many pro-BEV conclusions, while current global averages show smaller gaps of 20-30%. Hybrids often perform intermediately, with 15-20% lower emissions than ICEs without relying on charging infrastructure. Beyond GHGs, BEV LCAs reveal higher non-GHG impacts from supply chains, including depletion (up to 2.2 million liters per ton of ) and localized from in regions like the of , where cobalt extraction has caused and habitat loss; these are often underexplored in GHG-focused studies from advocacy groups. ICE alternatives like biofuels or e-fuels can reduce WTW emissions by 70-90% in compatible engines, potentially matching or exceeding BEVs in flexible, high-utilization scenarios without dependencies. Empirical data underscores that no universal superiority exists; outcomes hinge on local energy sources, vehicle duty cycles, and maintenance realities, with ICEs retaining advantages in total for low-mileage or off-grid applications.

Societal, Economic, and Policy Dimensions

Innovations Driving Economic Growth

The development of practical internal combustion engines (ICEs) in the late , building on Nikolaus Otto's four-stroke cycle patented in , enabled portable, high-power density prime movers that surpassed engines in efficiency and flexibility for mobile applications, catalyzing industrialization by powering mechanized , early , and stationary equipment that boosted crop yields and farm by up to 50% in adopting regions during the early . This shift from animal and power reduced labor intensity in farming and , contributing to a surge in global output; for instance, U.S. agricultural productivity grew at an average annual rate of 1.5% from 1900 to 1940, partly attributable to ICE-driven that displaced draft animals and enabled larger-scale operations. The advent of reliable, affordable gasoline engines culminated in Henry Ford's Model T, introduced in 1908, which leveraged standardized components and assembly-line to the vehicle's from $850 to $260 by 1925, making personal automobiles accessible to the and rural farmers, thereby expanding markets and stimulating ancillary industries like , rubber, and refining. This mass automobility spurred infrastructure investments, including over 3 million miles of U.S. roads built by 1930, and created multiplier effects: the automotive sector generated jobs in supply chains, with alone employing 300,000 workers at peak, while fostering and suburban development that amplified GDP growth by facilitating goods transport and labor mobility. Rudolf Diesel's compression-ignition engine, commercialized in 1897, further propelled by offering 30-50% higher than counterparts, powering ships, locomotives, and factories that reduced transport costs and enabled global expansion; for example, diesel marine engines cut shipping fuel expenses by half compared to by the 1920s, supporting a tripling of merchandise trade volume from 1900 to 1950. In the U.S., the automotive industry's reliance on ICE innovations has historically contributed about 3% to GDP, with direct and indirect economic output exceeding $1.5 trillion annually as of 2024 through , , and servicing that employ millions and drive in related sectors. Subsequent refinements, such as electronic introduced in passenger cars by the and turbocharging widespread by the , enhanced fuel economy by 20-30% and power output, sustaining ICE dominance in light-duty vehicles and averting supply constraints that could have hampered post-war economic booms; these advancements supported the sector's role in generating one-third of U.S. GDP growth during recovery periods like the . Overall, ICE innovations have underpinned a causal chain from localized power generation to networked economies, with empirical data showing correlations between vehicle ownership rates and growth rates exceeding 2% annually in industrializing nations from 1920 onward.

Reliability and Practical Superiorities

Internal combustion engines demonstrate exceptional long-term durability, with well-maintained engines routinely achieving 200,000 to 300,000 miles (320,000 to 480,000 km) of before major overhaul, and some exceeding 1 million km (620,000 miles) in fleet applications such as and commercial . This reliability stems from mature manufacturing processes, widespread availability of replacement parts, and mechanical simplicity that allows for cost-effective repairs by technicians globally, contrasting with the specialized expertise required for systems. In heavy-duty sectors like and industrial generators, ICE units have powered equipment for decades under continuous operation, with documented cases of engines surpassing 1 million hours of through routine servicing. A key practical superiority lies in the superior energy density of hydrocarbon fuels, enabling ICE vehicles to store far more energy per unit mass and volume than lithium-ion batteries; gasoline provides approximately 12.7 kWh/kg, over 50 times the gravimetric density of typical EV batteries at 0.25 kWh/kg. This allows for ranges exceeding 500 miles (800 km) on a single tank without the weight penalties that reduce EV efficiency and payload capacity, particularly advantageous for long-haul trucking and aviation where battery mass would compromise performance. Refueling times further enhance practicality, typically requiring 2 to 5 minutes at standard pumps to achieve full capacity, compared to 20 to 30 minutes for EVs to reach 80% charge via DC fast chargers, enabling seamless integration into high-mobility workflows without extended downtime.
Fuel/Battery TypeGravimetric Energy Density (Wh/kg)Volumetric Energy Density (Wh/L)
Gasoline12,700~9,700
~250~750
ICE systems also exhibit robustness in extreme conditions, maintaining consistent power output in sub-zero temperatures or high altitudes where performance degrades by up to 40% due to chemical limitations, as evidenced by their dominance in logistics and applications. The global fueling infrastructure, comprising over 100 million stations worldwide as of 2023, supports instantaneous scalability for remote or developing regions, bypassing the grid dependencies and installation delays inherent to widespread charging networks. While routine maintenance costs for ICE can accumulate from components like timing belts, overall repair benefit from modular designs where individual parts—such as pistons or valves—cost under $1,000 to replace, versus potential $5,000 to $20,000 for modules outside warranty. This entrenched ecosystem underscores ICE's practical edge in affordability and adaptability for diverse economic contexts.

Regulatory Mandates and Technological Debates

Regulatory mandates targeting internal combustion engines (ICE) primarily focus on reducing tailpipe emissions of criteria pollutants such as , , and , alongside greenhouse gases (GHG) like CO2. In the , Euro 7 standards, agreed upon in 2024 and set to apply from 2026, impose stricter limits on light-duty vehicles, including PM down to 10 nanometers and controls on non-exhaust sources like and wear, though tailpipe limits for legacy pollutants see minimal tightening from Euro 6. In the United States, the Environmental Protection Agency (EPA) under the administration in 2025 initiated rollbacks of prior GHG standards for model years 2027-2032, which had effectively mandated that two-thirds of new light-duty vehicles be electric or by 2032, citing overreach and feasibility concerns. These standards historically drove ICE technological upgrades, such as advanced exhaust aftertreatment systems including for NOx control, achieving over 90% reductions in some pollutants from pre-2000 baselines. Phase-out mandates for new vehicle sales have proliferated globally, with over 70 countries announcing timelines by 2035-2040, often allowing hybrids or compatibility as flexibility measures. The EU's 2035 ban on non-zero-emission combustion engines permits e-fuels for existing fleets post-ban, reflecting debates over outright prohibition versus fuel-neutral approaches. In the , federal actions in 2025, including resolutions, revoked California's waiver for its 2035 sales ban under the Advanced Clean Cars II program, halting state-level enforcement. Such mandates face criticism for overlooking infrastructure constraints, with electricity demand projected to double by 2050 under aggressive , potentially straining fossil-fuel-dependent and raising lifecycle emissions if coal or backups persist. Technological debates center on whether ICE can achieve deep decarbonization through fuels and efficiency gains or if battery electric vehicles (BEVs) represent the sole viable path. Proponents of ICE retention argue that synthetic e-fuels, produced via electrolysis of water and CO2 capture using renewable electricity, enable drop-in compatibility with existing engines and infrastructure, potentially yielding lower lifecycle CO2 than BEVs in scenarios prioritizing vehicle efficiency over upstream energy losses. Studies indicate e-fueled ICE or hybrids could reduce emissions more than BEVs when accounting for battery production's embedded carbon (up to 15-20 tons CO2 per vehicle) and electricity sourcing, though e-fuel scalability remains limited by high costs—currently 3-5 times gasoline—and electrolysis efficiency under 60%. Critics, often from academia and environmental advocacy, contend e-fuels inefficiently divert renewables from direct electrification, inflating overall energy demand by factors of 2-4 compared to BEV charging. Empirical data shows hybrid ICE variants already cut CO and CO2 by 20-50% over pure gasoline ICE without infrastructure overhaul, bridging gaps until e-fuel commercialization advances. These debates underscore causal trade-offs: mandates accelerating BEV adoption risk supply chain vulnerabilities from rare earth mining and battery degradation (20-30% after 10 years), while ICE pathways leverage proven durability—over 200,000 miles with minimal degradation—and fuel flexibility, including biofuels reducing carbon intensity by 70-90% without engine redesign. Mainstream sources favoring often underemphasize these, potentially influenced by funding ties to green tech subsidies, whereas analyses prioritize total system over tailpipe-only metrics. Ongoing innovations, like high-efficiency ICE cycles exceeding 50% in prototypes, suggest regulatory frameworks should incentivize multi-path decarbonization rather than singular technology bets.

References

  1. [1]
    Internal Combustion Engine Basics | Department of Energy
    In an internal combustion engine (ICE), the ignition and combustion of the fuel occurs within the engine itself. The engine then partially converts the energy ...
  2. [2]
    Internal Combustion Engine - NASA Glenn Research Center
    As the name implies, the combustion process of an internal combustion engine takes place in an enclosed cylinder.Missing: definition | Show results with:definition
  3. [3]
    The History of the Internal Combustion Engine - EngiClean.co.uk
    Apr 9, 2023 · The Benz Patent Motorwagen, which was built in 1885 by the German engineer Karl Benz, was the first practical internal combustion engine.
  4. [4]
    What Is an Internal Combustion Engine? - J.D. Power
    Mar 18, 2024 · Two primary types of internal combustion engines exist: a spark ignition gasoline engine and a compression ignition diesel engine. Spark ...
  5. [5]
    Electric Vehicle Myths | US EPA
    FACT: The greenhouse gas emissions associated with an electric vehicle over its lifetime are typically lower than those from an average gasoline-powered vehicle ...
  6. [6]
    The scope for improving the efficiency and environmental impact of ...
    Engine strategies could reduce CO2 emissions by 30%, hybridisation and light-weighting by 50%. •. Implications for transportation policy are discussed. Abstract.
  7. [7]
    Health and Environmental Effects - DieselNet
    Emissions from combustion engines can be a significant contributor to air pollution by a vast array of pollutants—including particulate matter, NOx, HCs, ...
  8. [8]
    [PDF] Otto and Diesel Cycles
    • Open Cycles: the working fluid is exhausted after each cycle (e.g., internal combustion engines). • Closed Cycles: the working fluid is recirculated within ...
  9. [9]
    Engine Efficiency - DieselNet
    For internal combustion engines, the upper limit of thermodynamic efficiency is usually determined with Otto and Diesel cycles calculations. Combustion ...
  10. [10]
    Explained: The Carnot Limit | MIT News
    May 19, 2010 · Today's car engines have efficiencies of 20 percent or less, compared to their Carnot Limit of 37 percent.
  11. [11]
    3.5 The Internal combustion engine (Otto Cycle) - MIT
    The Otto cycle is a set of processes used by spark ignition internal combustion engines (2-stroke or 4-stroke cycles).
  12. [12]
    11.4: The Otto Cycle - Physics LibreTexts
    Sep 9, 2020 · Efficiency = 1 − ( V C V B ) γ − 1 = 1 − T B T C . In principle the efficiency could be very large if the temperature at C, at the end of the ...
  13. [13]
    Thermal Efficiency for Otto 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.
  14. [14]
    Otto Cycle - an overview | ScienceDirect Topics
    From Eq. (13.15), it can be observed that the efficiency of the Otto cycle is mainly the function of CR for the given ratio of Cp ...
  15. [15]
    Engine combustion process explained
    Tutorial on the combustion process of an internal combustion engine, details about mixture formation and pollutant emissions.
  16. [16]
    Combustion in Diesel Engines - DieselNet
    In contrast with spark ignition engines, where combustion starts by an electrically energized spark at one location, combustion in diesel engines starts by ...
  17. [17]
    Engine Fundamentals - DieselNet
    Reciprocating piston engines are perhaps the most common form of internal combustion engine known. They power cars, trucks, trains and most marine vessels. They ...
  18. [18]
    What is Reciprocating Engine?- Types, and Working
    A reciprocating engine is an engine that uses one or more pistons in order to convert pressure into rotational motion.Reciprocating Engine Parts · 4) Crankshaft · Inline Engines
  19. [19]
    [PDF] Reciprocating Internal Combustion Engines - EPA
    Reciprocating engines have higher electrical efficiencies than gas turbines of comparable size, and thus lower fuel-related operating costs. In addition, the ...
  20. [20]
    How an internal combustion engine works in Turbo3 we will tell you
    Jan 13, 2021 · The most basic components of the internal combustion engine are the cylinder, piston, and crankshaft. To these are attached other components ...How An Internal Combustion... · Intake Blow · Compression Stroke<|separator|>
  21. [21]
    Stationary Reciprocating Internal Combustion Engines
    Stationary reciprocating internal combustion engines use either compression ignition (CI) or spark ignition (SI) in order to induce combustion within the ...
  22. [22]
    Otto Cycle and Diesel Cycle (With Diagram) | IC Engine | Thermal ...
    The Otto cycle is employed for petrol and gas engines, Diesel cycle is used for slow speed oil engines and the dual cycle is employed for high speed Diesel ...
  23. [23]
    Diesel cycle, Otto cycle and Dual Cycle - Merchant Navy Decoded
    Jun 22, 2023 · The Diesel cycle uses constant pressure combustion, the Otto cycle uses constant volume combustion, and the dual cycle combines both.
  24. [24]
    Two-Stroke vs. Four-Stroke Differences? - AMSOIL Blog
    Aug 24, 2022 · However, four-stroke engines produce more torque at lower rpm, generally have better durability than high-revving two-stroke engines and also ...Missing: reciprocating | Show results with:reciprocating
  25. [25]
  26. [26]
  27. [27]
    How Rotary Engines Work - Auto | HowStuffWorks
    Jun 21, 2024 · Rotary engines do have some advantages over other engine types. They can be more reliable because their parts and pieces move at a slower speed.<|separator|>
  28. [28]
    Rotary Engine - an overview | ScienceDirect Topics
    The two main disadvantages of the rotary engines are the increased emission of unburned hydrocarbons and higher fuel consumption. Both disadvantages are related ...
  29. [29]
    Wankel Engine: A Triangular Revolution from Past to Future
    Felix Wankel's journey continued in 1938 with research on rotary compressors, evolving into the development of a practical internal combustion engine post-World ...
  30. [30]
    The Re-Emergence of Rotary Engines
    Jul 6, 2022 · Rotary engines may have low-grade combustion, which leads to emission issues, especially high carbon monoxide and hydrocarbon emissions.Missing: history | Show results with:history
  31. [31]
    Advantages and Disadvantages of a Rotary Engine - National Speed
    Jan 29, 2009 · The Rotary Engine is very simple. It's a motor design that utilizes way less moving parts than it's piston counterpart.Missing: combustion types principles history
  32. [32]
    Alternative Combustion Engines - DieselNet
    Dec 15, 2021 · These alternative designs include rotary engines, such as the Wankel engine, two-stroke engines, as well as six-stroke and split-cycle engines.
  33. [33]
    Combustion Engine - an overview | ScienceDirect Topics
    Gas turbine engines exemplify the continuous-combustion engine. The intermittent-combustion engine is characterized by periodic ignition of fuel and air.
  34. [34]
    [PDF] Conventional Type Combustion 3.2.1.1-1 Introduction
    The processes that occur within a gas turbine combustor (e.g., injection of the air and fuel, mixing of the air and fuel, combustion reaction) are “continuous” ...
  35. [35]
    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 ...
  36. [36]
    Brayton Cycle - Chemistry LibreTexts
    May 6, 2025 · The Brayton Cycle is a thermodynamic cycle that describes how gas turbines operate. The idea behind the Brayton Cycle is to extract energy ...Introduction · Ideal Brayton Cycle · Efficiency of the Brayton Cycle · References
  37. [37]
    [PDF] CLASSIFICATION OF INTERNAL COMBUSTION ENGINE
    Internal combustion engines are classified as reciprocating (two-stroke, four-stroke), rotary (Wankel), and continuous combustion (gas turbine, jet, rocket).
  38. [38]
    Leibniz, Papin and the Steam Engine - Schiller Institute
    1673: Huygens successfully demonstrates his gunpowder-fueled engine, suggesting that his invention "permits the discovery of new kinds of vehicles on land ...
  39. [39]
    The First Internal Combustion Engine - went with a bang
    Dec 23, 2012 · The 1680 the Dutch scientist Huygens created the first “internal combustion engine”. Interestingly it was powered by gunpowder.Missing: 1673 | Show results with:1673
  40. [40]
    Untitled
    The first effort to substitute a piston for a cannon-ball, that of Leonardo da Vinci, used gunpowder for fuel, as did Samuel Morland's patent of 1661, Huygens' ...
  41. [41]
    A Brief History of the Internal Combustion Engine
    Jan 15, 2012 · The first internal-combustion engine, according to our modern ideas, was that of Robert Street, patented in England in 1794.Missing: concepts | Show results with:concepts
  42. [42]
    Internal Combustion Engines
    Feb 11, 2025 · The first engine that caused a great deal of public interest and that was manufactured for sale was conceived only in 1858 and patented in 1860, ...Multiple Beginnings · Otto And Langen · The Diesel Engine
  43. [43]
    Inventor of first practical internal combustion engine is largely ...
    Aug 4, 2025 · Lenoir built a two-stroke, one-cylinder internal combustion engine in 1859. He used a converted steam engine with slide valves to take in a ...
  44. [44]
    Birth of an idea: Etienne Lenoir and the internal combustion engine
    Oct 29, 2020 · In 1860, Lenoir received a patent for “an air motor expanded by gas combustion” from Conservatoire National Des Arts Et Métiers, no. N.43624, ...
  45. [45]
    [PDF] Timeline Of The History Of Cars
    - 1860: Belgian inventor Étienne Lenoir created the first internal combustion engine powered by coal gas, laying the groundwork for future gasoline engines. - ...
  46. [46]
    George Brayton - ASME
    Apr 3, 2012 · His engine was explained this way: "Air is compressed in the single-acting pump, which has a volume one-half that of the power cylinder. The ...
  47. [47]
    Patent Page: The Brayton Cycle Engine - Gas Engine Magazine
    Jul 15, 2020 · A detailed drawing of George B. Brayton's engine as shown in his 1872 patent for “improvement in gas engines”.
  48. [48]
    Otto (1832) - Energy Kids - EIA
    Before year's end, Otto had built an internal combustion engine, utilizing a four-stroke piston cycle. Now called the 'Otto cycle' in his honor, the design ...
  49. [49]
    Nicolaus August Otto and the Four-Stroke Engine | Museumsblog
    Jun 10, 2022 · The new four-stroke engine made its first run at the beginning of March 1876, with Deutz AG able to file for a patent the same year.
  50. [50]
    April 3, 1885 - The internal combustion engine is patented
    Gottlieb Daimler patented his grandfather clock engine on this day in 1885. It is recognized as the first practical internal combustion engine.
  51. [51]
    The Daimler riding carriage from 1885. - Mercedes-Benz
    On 3 April 1885, Gottlieb Daimler filed a patent application for the high-speed engine with vertical cylinder. The nickname because of its characteristic ...
  52. [52]
    The Ford Model T | Articles | Ford Motor Company
    Discover how the 1908 Ford Model T, with its affordable price and innovative design, became the first mass-produced car, changing the automotive industry.
  53. [53]
    Henry Ford: Case Study of an Innovator - Blog - Henry Ford Museum
    Jul 21, 2021 · The Model T was lighter than its competitors, allowing its 20-horsepower engine to give it performance equal to that of more expensive cars. The ...
  54. [54]
    Early History of the Diesel Engine - DieselNet
    High-speed diesel engines were introduced in the 1920s for commercial vehicle applications and in the 1930s for passenger cars. Rudolf Diesel's Invention ...Missing: century | Show results with:century
  55. [55]
    Internal Combustion Engine - Otto Cycle | Glenn Research Center
    Jul 28, 2023 · In an internal combustion engine, fuel and air are ignited inside a cylinder. The hot exhaust gas pushes a piston in the cylinder which is ...Missing: definition | Show results with:definition
  56. [56]
    The Resurgence of Internal Combustion Engines in a ... - SelfLube
    Mar 18, 2025 · Modern ICEs now feature direct fuel injection, turbocharging, and variable valve timing, reducing fuel consumption and emissions while improving performance.Missing: hybridization | Show results with:hybridization
  57. [57]
    [PDF] Efficiency technology and cost assessment for U.S. 2025–2030 light ...
    Our analysis indicates 8%–10% greater efficiency improvement is available and cost effective for vehicles by 2025, compared to the latest U.S. regulatory ...Missing: modern | Show results with:modern
  58. [58]
    (PDF) Future of Internal Combustion Engines - ResearchGate
    Oct 10, 2025 · Recent breakthroughs have achieved thermal efficiencies exceeding 53% in commercial diesel engines, with research pathways suggesting potential ...
  59. [59]
    Chapter: Appendix M: HCCI Projects - The National Academies Press
    A 13.4 percent improvement in fuel economy was achieved through improvements in the engine management system (which included GDI, VVT, cooled EGR, stop/start, ...
  60. [60]
    HCCI: New Life for the ICE | MOTOR
    HCCI can generate more horsepower, gain higher fuel efficiency and give off cleaner emissions. But it also presents a difficult challenge of balancing internal ...
  61. [61]
    Advancements in combustion technologies: A review of innovations ...
    HCCI engines represent a promising advancement in combustion technology, offering significant improvements in fuel efficiency and emissions reduction. While ...
  62. [62]
    [PDF] Internal Combustion Engine Technical Sector Team
    New engine efficiency targets will need to be established for renewable fuels, depending on fuel properties and degree of carbon reduction. in-cylinder ...
  63. [63]
    Powering the Future: The role for Internal Combustion Engines in a ...
    Jul 29, 2024 · Continuous improvements in efficiency and lower emissions, as well as significant opportunities for using lower carbon fuels that position it to ...Missing: 2020-2025 | Show results with:2020-2025
  64. [64]
    Future of internal combustion engines using sustainable, scalable ...
    The inherent oxygen improves combustion efficiency and lowers carbonaceous emissions. Also, particulate emissions and NOx emissions are reduced compared to ...
  65. [65]
    Fuel-Agnostic Innovation: ICE's Role in a Clean Future | MOTOR
    Oct 8, 2025 · Cleaner ICEs with hydrogen and alternative fuels offer a scalable path to decarbonization, complementing EVs in hard-to-electrify sectors.
  66. [66]
    The Future of Internal-Combustion Engines
    Hydrogen, propane, hybrid applications and e-fuel alternatives will help ensure an ICE future for motorsports, long-distance drivers, farmers and commercial ...
  67. [67]
    4-Stroke Internal Combustion Engine - NASA Glenn Research Center
    A four-stroke engine has four piston movements: intake, compression, power, and exhaust, before the engine firing sequence repeats.
  68. [68]
    The Four Stroke Engine Cycle | AeroToolbox
    A four-stroke engine takes two crankshaft revolutions to complete one engine cycle with the piston moving through 180° to complete each step in the cycle.Piston Engine Cycles · Cycle Nomenclature · The Four Stroke Cycle · Compression
  69. [69]
    The four stroke cycle | Leading Edge Only
    Dec 2, 2020 · There are four strokes in one four stroke cycle of the engine. They are: the intake stroke, compression stroke, power stroke and exhaust stroke.
  70. [70]
    What is a Camshaft? Diagram, Functions, Parts, Working, Types
    A camshaft controls the opening and closing of the internal combustion engine's valves. Learn its diagram, functions, parts, working, types and applications ...<|separator|>
  71. [71]
    Difference Between 2 Stroke & 4 Stroke Engines | Castrol® USA
    A 2- cycle engine goes through the same process of induction, compression, power and exhaust but does so with the piston undertaking 2 strokes (up and then ...
  72. [72]
    History Of VVT (Variable Valve Timing) Technology
    Mar 30, 2012 · The first variable valve timing systems came into existence in the nineteenth century on steam engines. Stephenson valve gear, as used on ...
  73. [73]
    Tech Through Time: How Variable Valve Timing Works | CarsGuide
    The need for variable valve timing was recognised as early as 1924, when Americans patented a valve with a variable opening duration for an internal combustion ...
  74. [74]
    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 ...
  75. [75]
    How Nikolaus August Otto created the 4-stroke internal combustion ...
    Nov 5, 2020 · In 1876, Otto developed a gaseous fuel, compressed charge 4-stroke cycle that would become known as the Otto-Cycle. This is the principle that ...
  76. [76]
    Diesel Engines
    The diesel engine, invented in the late 19 th century by Dr. Rudolf Diesel, is the most energy efficient powerplant among all type of internal combustion ...Types Of Diesel Engines · Idi Engines · Efficiency & Ghg Emissions
  77. [77]
    Research and application of over-expansion cycle (Atkinson and ...
    This paper provides a critical review of current state-of-the-art research on ACE and MCE. Their similarities and differences are clarified.
  78. [78]
    [PDF] experimental study of a miller cycle based approach for an efficient ...
    Sep 3, 2017 · At the end, this thesis presents an Miller cycle based approach for maximising fuel conversion efficiency of a gasoline engine by combining ...
  79. [79]
    Miller cycle application to improve lean burn gas engine performance
    Aug 15, 2016 · Miller cycle is applied to improve performance and emission characteristics of a gas engine. There are different types of Miller cycle ...
  80. [80]
    Improving the Thermal Efficiency of the Homogeneous Charge ...
    The experimental study of HCCI in a motorcycle engine with n-heptane fuel was conducted with regards to engine performance and combustion characteristics.<|separator|>
  81. [81]
    RCCI Engine Operation Towards 60% Thermal Efficiency 2013-01 ...
    30-day returnsApr 7, 2013 · The present experimental study explored methods to obtain the maximum practical cycle efficiency with Reactivity Controlled Compression ...
  82. [82]
    [PDF] Reactivity controlled compression ignition drive cycle emissions and ...
    In-cylinder blending of gasoline and diesel to achieve reactivity- controlled compression ignition (RCCI) has been shown to reduce NOX and soot emissions ...
  83. [83]
    [PDF] A Review: Six Stroke Internal Combustion Engine - IOSR Journal
    The main advantages of six stroke engine are two power strokes in the six stroke cycle, reduction in fuel consumption by 30-40 percent, dramatic reduction ...<|separator|>
  84. [84]
    Six Education: Inside Porsche's Patent for a Six-Stroke Engine
    Dec 26, 2024 · Six Strokes, Seven Cycles. Porsche's six strokes include the four familiar ones: intake, compression, power, and exhaust. But it inserts an ...
  85. [85]
    Advances in free-piston internal combustion engines
    May 5, 2021 · Free-piston internal combustion engines (FPICEs) have many potential advantages, such as simple structure, high conversion efficiency, ...
  86. [86]
    A review of free-piston linear engines - ScienceDirect.com
    Sep 15, 2016 · Unlike conventional internal combustion engines, a free-piston linear engine has no a crankshaft, and thus the pistons move freely in the ...
  87. [87]
    How Diesel Engines Work: Explaining the Function of Compression ...
    Jan 11, 2018 · Gasoline engine are spark ignition engines and diesel-fueled engines are compression ignition engines. Four Stroke, Internal Combustion Engine ...
  88. [88]
    How The Internal Combustion Ignition System Works - MotorTrend
    Feb 26, 2016 · Heat initiates the internal combustion process. Diesel engines utilize the temperature buildup from extremely high compression (pressure) to ignite the air/ ...
  89. [89]
    Why Magneto Ignition? - Hot Rod
    Magneto ignition is self-contained, generates its own energy, and produces high-intensity sparks at high speeds, unlike battery systems. It also has long ...Missing: internal | Show results with:internal
  90. [90]
    From points to spark: the history of electronic ignitions for classic cars
    Jun 13, 2024 · In 1910, Kettering invented the first practical ignition system, known as the Kettering ignition system or the Delco ignition system. This ...
  91. [91]
    How Ignition Systems Work | Champion Tech Tips
    Conventional breaker-point ignition (in use since the early 1900s) · Electronic ignition (popular since the early 1970s) · Distributorless ignition (introduced in ...
  92. [92]
    Electric starter's inventor Kettering was no crank - Hagerty Media
    Aug 29, 2019 · Unlike other early electric starters, Kettering's starter had reduction gears between itself and the engine's crankshaft, providing mechanical ...
  93. [93]
    Charles F. Kettering, inventor of electric self-starter, is born | HISTORY
    Kettering's key-operated electric self-starting ignition system, introduced on Cadillac vehicles in 1912 and patented three years later, made automobiles ...
  94. [94]
    History of the Starter Motor - crankSHIFT
    Oct 1, 2013 · This is a starter mechanism that was invented by Vincent Bendix in 1910, and first saw use in the 1914 Chevrolet “Baby Grand.” It would go on ...
  95. [95]
    All about starter motor history | SEG Automotive
    From the first sliding-armature starter motor in 1914 to the modern start/stop system, the starter motor has evolved from a simple mechanical starting aid to an ...
  96. [96]
    The Evolution of the Starter Motor | Elreg
    Jan 16, 2020 · Vehicle starters have quite an extensive history. These mechanisms date back to the early 1900s, and since then, have evolved into the vehicle technology we ...
  97. [97]
    How forced induction resulted in greater power density, efficiency ...
    Nov 25, 2020 · The only way to increase its air mass flow is to artificially force a higher density into the air receiver, where the cylinders breath.
  98. [98]
    Forced Induction Technologies in an IC Engine: A Review
    Aug 7, 2025 · Forced induction technology like turbocharging and supercharging can enhance the performance of an internal combustion engine by compressing inlet air charge.
  99. [99]
    The Rise and Fall of the Supercharger - Motor1.com
    Jun 21, 2024 · The machine that became the supercharger predates the practical internal-combustion car by decades. Created by brothers Philander and Francis Roots.
  100. [100]
    The History Of The Blower - MotorTrend
    Sep 20, 2005 · From 1925-38, the Grand Prix formula led to the supercharging of virtually all racing engines; there was a steady increase in power, coupled to ...
  101. [101]
    Turbochargers: A History - MotorTrend
    Jun 11, 2018 · Early ideas came from Daimler and Diesel. Alfred Buchi patented a turbo in 1905, with first successful use in 1925. Early marine and truck ...
  102. [102]
    Turbochargers - USC Viterbi School of Engineering
    A Brief History of the Turbocharger. The first turbocharger was patented in 1905 by Alfred Bucchi of Switzerland, who then built the first prototype in 1915.
  103. [103]
    History of the turbo - MasterTurbo
    The first application of turbos in the commercial vehicle sector came in 1938, when the Swiss truck factory Saurer started using the turbo. Emphasis was placed ...
  104. [104]
    E-boost - The Future of Turbochargers - Claytex
    Mar 24, 2020 · Apart from improving the overall ICE efficiency, it delivers big advantages to the driving experience such as a reduction on the turbo lag and ...<|control11|><|separator|>
  105. [105]
    What are the pros and cons of forced induction in engines? - Quora
    Mar 4, 2020 · Forced induction, (turbocharging) increases the volumetric efficiency of the engine. At higher speeds, a naturally aspirated engine cannot take ...What are the benefits and drawbacks of forced induction ... - QuoraWhat is the purpose of using forced induction in internal combustion ...More results from www.quora.com
  106. [106]
    Turbos & Superchargers - Forced Introduction - MotorTrend
    Jun 1, 2011 · There are advantages and disadvantages to this. Smaller turbos have the ability to reach maximum boost sooner, (which reduces boost lag) ...Centrifugal · Turbos · Intercoolers
  107. [107]
    Engine cooling system: how it works and main components
    The cooling system uses a thermostat to regulate the normal operating temperature of the internal combustion engine. When the engine reaches standard operating ...
  108. [108]
    Cooling System in IC Engine: Working, Types, Advantages, and Uses
    The cooling system comprises passages within the engine block and heads, a water pump for circulating coolant, a thermostat for temperature control, a radiator ...
  109. [109]
    [PDF] COOLING SYSTEM
    There are two types of cooling systems: (i) Air cooling system and (ii) Water-cooling system.
  110. [110]
    How an engine cooling system works | How a Car Works
    This article explains how a car cooling system works. Understand overheating problems, and the role of water, air and fan-based engine cooling systems.<|separator|>
  111. [111]
  112. [112]
    [PDF] Lubrication & Cooling Systems
    Pressure System. The oil pressure branch of the engine lubrication system is pressurized by a gear-type pressure pump located in the oil pump and accessory ...
  113. [113]
    Lubrication Systems Types and Features - SVI International
    Oct 20, 2023 · A pressure lubrication system distributes lubricant to the engine under pressure feed. The oil moves from the sump into the pump to lubricate ...
  114. [114]
    [PDF] LUBRICATION SYSTEM IN IC ENGINES - DAV University
    Lubrication is the action of applying a substance such as oil or grease to an engine or component so as to minimize friction and allow smooth movement.
  115. [115]
    Engine Parts 101: Valve Train Components | Blog | Enginetech
    Overhead valve (OHV) engines have valve train components that consist of valve springs, rocker arms, pushrods, and lifters. Overhead camshaft (OHC) engines have ...Missing: combustion DOHC<|separator|>
  116. [116]
    OHV, SOHC And DOHC Valve Trains: How They Work, And Their ...
    Feb 4, 2019 · Explore the workings and advantages of OHV, SOHC, and DOHC valve trains. Understand their mechanisms and weigh their respective pros and ...
  117. [117]
    DOHC vs SOHC vs OHV: Engine Types & Key Differences - Auto Parts
    Sep 12, 2025 · An OHV (Overhead Valve) engine—often called a pushrod engine—features a camshaft located inside the engine block, beneath the cylinder head.
  118. [118]
  119. [119]
    Fuels of the Diesel-Gasoline Engines and Their Properties
    Feb 26, 2020 · Self-ignition is the basic rule of the combustion process in diesel engines. SIT value is desired to be high in gasoline engines and low in ...
  120. [120]
    Fueling Internal Combustion Engines - Let's Talk Science
    Jan 10, 2023 · Combustion is when fuel reacts with oxygen to produce heat. Chemical reactions that produce heat we call exothermic reactions. Most vehicles use ...
  121. [121]
    Is it really the end of internal combustion engines and petroleum in ...
    Sep 1, 2018 · Currently around 95% of transport energy comes from liquid fuels derived from petroleum and around 60% of all oil produced goes to make ...
  122. [122]
    Gasoline | C18H25NO | CID 6435060 - PubChem - NIH
    Gasoline consists of over 150 identifiable C4-C12 hydrocarbons with boiling range 32-210 °C(1). The individual components of gasoline include paraffins, olefins ...
  123. [123]
    Gasoline explained - octane in depth - U.S. Energy Information ... - EIA
    Octane ratings are measures of fuel stability. These ratings are based on the pressure at which a fuel will spontaneously combust (auto-ignite) in a testing ...
  124. [124]
    Fuel Properties Comparison - Alternative Fuels Data Center
    1 gallon of gasoline has 97%–100% of the energy in 1 GGE. Standard fuel is 90% gasoline, 10% ethanol. ; 1 gallon of diesel has 113% of the energy in 1 GGE due to ...
  125. [125]
    Cetane - an overview | ScienceDirect Topics
    Cetane number varies with hydrocarbon types present in diesel fuels. Normal paraffins have high cetane numbers that increase with carbon number.
  126. [126]
  127. [127]
    Diesel Fuels Derived from Renewable Sources - OSU Extension
    Diesel fuels containing large amounts of shorter-chain and isomerized species have lower cetane numbers while diesel fuels consisting mainly of long-chain ...
  128. [128]
    Detailed Composition-Based Model for Predicting the Cetane ...
    We have developed a simple composition-based model for predicting the cetane number of diesel fuels with general applicability to any diesel fuel regardless ...
  129. [129]
    Alternative Fuels for Internal Combustion Engines - MDPI
    These may be non-carbon fuels such as hydrogen or ammonia, or biofuels such as alcohols, ethers or esters, including synthetic fuels. There are also other types ...
  130. [130]
    [PDF] Alternative Fuels for Internal Combustion Engines
    Methanol. (CH3OH) is an alcohol fuel. Methanol is methane with one hydrogen molecule replaced by a hydroxyl radical. (OH). The alternative fuel currently being ...
  131. [131]
    [PDF] Comparative analysis of alternative fuels for internal combustion ...
    This comparative analysis of biofuels, synthetic fuels, and hydrogen for internal combustion engines reveals that each alternative fuel category offers distinct ...
  132. [132]
    Synthetic Diesel Fuel - DieselNet
    Synthetic diesel fuels can be made from carbon containing feedstocks, such as natural gas or coal, in a process developed by Fischer and Tropsch in the 1920s.Missing: internal | Show results with:internal
  133. [133]
    Fischer-Tropsch Diesel Fuels - Properties and Exhaust Emissions
    30-day returnsMar 2, 2003 · Fischer-Tropsch diesel fuel can be used neat without engine modification or can be blended in any proportion with conventional diesel fuel. A ...
  134. [134]
    Fischer-Tropsch Diesel - an overview | ScienceDirect Topics
    Fischer-Tropsch diesel is defined as a synthetic fuel produced through the Fischer-Tropsch synthesis process, characterized by its higher cetane number and ...
  135. [135]
    The Role of E-fuels in Decarbonising Transport – Analysis - IEA
    Dec 22, 2023 · Fuels obtained from electrolytic hydrogen, or e-fuels, could be a viable pathway and scale up quickly by 2030, underpinned by a massive ...
  136. [136]
    Why e-fuels are not the solution for cars | T&E
    Running cars on e-fuels will not alleviate air pollution. Cars powered by synthetic fuel emit as much poisonous nitrogen oxides (NOx) as fossil fuel engines, ...
  137. [137]
    E-fuels won't save the internal combustion engine
    Jun 23, 2020 · We've found that significant volumes of renewable e-fuels won't be made for less than 3 or 4 Euros per liter in 2030.
  138. [138]
    Hydrogen internal combustion engines and hydrogen fuel cells
    Jan 27, 2022 · Hydrogen engines release near zero, trace amounts of CO2 (from ambient air and lubrication oil), but can produce nitrogen oxides, or NOx. As a ...
  139. [139]
    Does the use of hydrogen produce air pollutants such as nitrogen ...
    Current research has shown that hydrogen combustion via turbines can achieve comparable performance and NOx emissions to those of today's turbines running on ...
  140. [140]
    Analysis of performance, emissions, and lubrication in a spark ... - NIH
    Additionally, hydrogen causes a 2.9% increase in engine BTE. Hydrogen's more efficient combustion process allowed for reducing CO, HC, and smoke opacity ...
  141. [141]
    Assessing fuel economy and NOx emissions of a hydrogen engine ...
    Jul 15, 2023 · NOx emissions are the only gaseous emissions of concern in hydrogen combustion because of the carbon neutrality of the hydrogen fuel [35]. As ...
  142. [142]
    Hydrogen fuel cells and combustion engines - Volvo Group
    Hydrogen fuel cells are more efficient than traditional internal combustion engines because they convert chemical energy directly into electrical energy, ...<|separator|>
  143. [143]
    Combustion
    The original substance is called the fuel, and the source of oxygen is called the oxidizer. The fuel can be a solid, liquid, or gas, although for airplane ...
  144. [144]
    How Nitrous Oxide in Cars Works - Auto | HowStuffWorks
    Feb 21, 2024 · When nitrous oxide is heated to 570 degrees F, its molecules split into nitrogen and oxygen. This means more oxygen is produced for combustion, ...What is Nitrous Oxide? · Components of Nitrous Oxide... · Safety and Risks
  145. [145]
    Nitrous Oxide 101: The Basics Of The Gas Everyone Loves To Hate
    Aug 26, 2021 · By adding oxygen to the combustion process, Nitrous Oxide allows more fuel to be burned in the combustion process, and more power to be made.
  146. [146]
    Nitrous Oxide & the Diesel - Banks Power
    Consequently, nitrous oxide is only injected for short spurts at full throttle, usually lasting no more than 10-15 seconds at a time.<|separator|>
  147. [147]
    The Effect of Oxygen Enriched Air on the Performance and ...
    30-day returnsSep 30, 1993 · A single-cylinder, 4-stroke, spark-ignition engine was used to evaluate the effect of oxygen enriched air on engine performance and exhaust ...
  148. [148]
    Investigating the effects of oxygen enrichment with modified zeolites ...
    Nov 1, 2021 · Oxygen enrichment of intake air in internal combustion engines (ICE) is one of the effective methods used to reduce exhaust emissions and ...
  149. [149]
    Oxygen-Enriched Combustion for Industrial Diesel Engines | ICEF
    Jan 12, 2016 · Oxygen-enriched combustion (OEC) is used in industrial combustion applications to increase the adiabatic flame temperature.Missing: ICE | Show results with:ICE
  150. [150]
    [PDF] STUDY OF USING OXYGEN-ENRICHED COMBUSTION AIR FOR ...
    Oxygen enrichment with its attendant higher combustion temperatures, reduces emissions of particulates and visible smoke but increases NO emissions (by up to ...<|control11|><|separator|>
  151. [151]
    Effects of the Combustion Enhancer Containing Alkyl Nitrate (CEN ...
    30-day returnsAddition of a combustion enhancer containing alkyl nitrate (CEN) allows the use of methanol in a direct injection compression ignition (DICI) engine without any ...
  152. [152]
    Ozone used to improve combustion efficiency in Internal ...
    This finding can help reduce emissions of soot and nitrogen oxide from internal combustion engines. The results are relevant to both biodiesel and conventional ...
  153. [153]
    A review of fuel additives' effects and predictions on internal ...
    Additives such as alcohol, hydrogen, and metal oxides are proven to be successful in improving performance or reducing emissions. Results from selected papers ...
  154. [154]
    Engine Thermal Efficiency - an overview | ScienceDirect Topics
    An initial appraisal of the engine thermal efficiency at full load capacity indicates a heat-to-power ratio of 1.56:1, representing a thermal efficiency of 39.1 ...
  155. [155]
    [PDF] Thermal efficiency - McGraw Commons
    Mar 27, 2013 · However, practical Diesel engines are 30% - 35% more efficient than gasoline engines.[5] This is because, since the fuel is not introduced to ...
  156. [156]
    The Diesel Dilemma: How Efficient Are They Really?
    Aug 1, 2025 · Diesel engine fuel efficiency is significantly superior to gasoline engines, delivering 20-35% better fuel economy through higher compression ...Why Diesel Engine Fuel... · Fuel With More Force · Real-World Diesel Efficiency...
  157. [157]
    China unveils world's 1st diesel engine with 53.09% thermal efficiency
    Apr 22, 2024 · Chinese firm Weichai Power unveils an advanced diesel engine that achieves a remarkable 53.09 percent intrinsic thermal efficiency.
  158. [158]
    Mercedes Created The World's Most Efficient Racing Engine
    Jan 10, 2018 · Mercedes Achieved A 50% Thermal Efficiency Engine Awesome Car Products: https://www.amazon.com/shop/engineeringexplained Subscribe for new ...
  159. [159]
    e-POWER's internal combustion engine achieves 50% thermal ...
    Nissan's e-POWER's internal combustion engine achieves 50% thermal efficiency: Nissan's various technologies, including electrification, autonomous drive, ...
  160. [160]
    Brake Mean Effective Pressure (BMEP) — A Yard Stick for Power
    Jan 6, 2022 · BMEP is a theoretical average cylinder pressure calculated using a constant, torque, and displacement, used to compare engines of different ...Missing: metrics | Show results with:metrics
  161. [161]
    Brake Mean Effective Pressure (BMEP): The Performance Yardstick
    Jun 8, 2019 · BMEP is the average pressure that, if applied uniformly, would produce the measured brake power output. It's a theoretical tool for evaluating  ...Missing: internal | Show results with:internal
  162. [162]
    Brake Specific Fuel Consumption (BSFC) - x-engineer.org
    Aug 19, 2017 · Brake specific fuel consumption (BSFC) is a parameter that reflects the efficiency of a combustion engine which burns fuel and produces rotational power.
  163. [163]
    Brake Specific Fuel Consumption - an overview | ScienceDirect Topics
    Brake specific fuel consumption is one of the important parameters of an engine and is defined as the consumption per unit power in a unit of time. Most of the ...
  164. [164]
    Fact of the week: Internal combustion cars still waste 70 to 88 ...
    Aug 31, 2018 · Another 4 to 6 percent is lost to parasitic engine loads such as water, fuel, and oil pumps. And 5 to 6 percent is lost to friction, lash, and ...
  165. [165]
    Friction and Ancillary Losses - DieselNet
    Ancillary losses can represent about 20-30% of mechanical losses and include the work required to operate critical engine components such as the lubricating oil ...
  166. [166]
    Mechanical efficiency and friction mean effective pressure (FMEP)
    In the table below you can see the distribution (in percentage) of energy losses of an internal combustion engine at part load and at full load. ... Stephen on ...<|separator|>
  167. [167]
    [PDF] Defining engine efficiency limits - Department of Energy
    Pumping losses. 30%. Diesel engines have relatively low pumping losses, but improved volumetric efficiency through optimized ports, manifolds, and ducting and ...
  168. [168]
    A Review of Energy Loss Reduction Technologies for Internal ...
    In this paper, literature surveys are presented that investigate the relative advantages of technologies mainly focused on minimizing energy loss in engine ...
  169. [169]
    (PDF) Improving Thermal Efficiency of Internal Combustion Engines
    Aug 10, 2022 · thermal efficiency of up to 60% was needed with reduced frictional and pumping losses. The findings demonstrate that, with optimization ...<|control11|><|separator|>
  170. [170]
    Benz Patent Motor Car: The first automobile (1885–1886)
    The first stationary gasoline engine developed by Carl Benz was a one-cylinder two-stroke unit which ran for the first time on New Year's Eve 1879.
  171. [171]
    A brief history of the internal combustion engine
    Jun 23, 2023 · This article explores the fascinating journey of the internal combustion engine, covering its key milestones, inventors, technological advancements,
  172. [172]
    How Do Gasoline Cars Work? - Alternative Fuels Data Center
    Internal combustion engine (spark-ignited): In this configuration, fuel is injected into either the intake manifold or the combustion chamber, where it is ...
  173. [173]
    [PDF] VEHICLES IN USE EUROPE 2023 - ACEA
    Diesel-powered light commercial vehicles are still dominant in the EU with 91% of the fleet running on diesel and just 0.6% of vans being battery electric. • ...
  174. [174]
    Tracking global data on electric vehicles - Our World in Data
    Globally, over 1-in-5 (22%) of new cars sold were electric in 2024. This share was 92% in Norway, and in China, it was almost 50%. In the chart below, you can ...
  175. [175]
    What are the benefits of using an internal combustion engine (ICE ...
    Apr 16, 2023 · ... EV battery, an ICE vehicle would lose over half their driving range. Makes one wonder just how much further an EV could travel if the ...
  176. [176]
    2024 EPA Automotive Trends Report Greenhouse Gas Emissions ...
    Preliminary data suggest that the average new vehicle CO2 emission rate and fuel economy will continue to improve in model year 2024, and that the impact of ...
  177. [177]
    Let's Turn Down the Heat on ICE vs. EV - Engine Technology Forum
    Aug 20, 2024 · Gas and diesel internal combustion engine (ICE) car and pickup truck owners are sometimes shocked at the tone and temperament of some EV advocates.<|separator|>
  178. [178]
    Complete Guide Of Truck Engines Makes And Models
    Jan 7, 2020 · There are six variations, ranging in horsepower from 430-605 and peak torque ranging from 1650-2050 pound-feet. The engine brake comes with 600 ...
  179. [179]
    Advanced Diesel Engines | Cummins Inc.
    Apr 18, 2023 · Heavy Duty: Their 15-liter and 12-liter advanced diesel engines are some of the most popular and reliable ever produced. For the most demanding ...
  180. [180]
  181. [181]
    Most Powerful Ship Engine-The Gigantic Wärtsilä-Sulzer
    Aug 29, 2025 · It produces 80,08 MW or 107,390 horsepower, making it the world's largest reciprocating engine. This version entered commercial service in ...Missing: power | Show results with:power
  182. [182]
    Understanding 30 Marine Diesel Engine Types - Nautilus Shipping
    Aug 9, 2022 · Supercharged engine​​ This increases the power output of the engine at no additional operating cost. We use this method in most two and four ...
  183. [183]
    Marine Diesel - ECO, Inc.
    Category 1 and Category 2 marine diesel engines typically range in size from about 500 to 8,000 kW (700 to 11,000 hp). These engines are used to provide ...<|separator|>
  184. [184]
    Piston Engines – Introduction to Aerospace Flight Vehicles
    These reciprocating piston internal combustion engines, which drive propellers, remain the powerhouses of modern general aviation. Their use in relatively low- ...
  185. [185]
    [PDF] Aircraft Engines - Federal Aviation Administration
    Small general aviation aircraft use mostly horizontally opposed reciprocating piston engines. While some aircraft still use radial reciprocating piston engines ...
  186. [186]
    The Beating Heart - by Federal Aviation Administration - Medium
    Mar 3, 2021 · Specifically, it means a reciprocating engine, a term that simply denotes the back and forth motion of the pistons.<|separator|>
  187. [187]
    Reciprocating Internal Combustion Engine - ScienceDirect.com
    These engines can generate power ranging from a few kilowatts to over 5MW and can achieve overall efficiencies exceeding 70% in combined heat and power (CHP) ...
  188. [188]
    Diesel Generator Sets
    4.8 26 · Free deliveryCat diesel generator sets rated 6 to 5,720 ekW (7.5 to 7,150 kVA) for standby or prime power provide efficiency, low fuel consumption and emissions compliance.
  189. [189]
    Complete Guide: How Much Diesel Does a Generator Use?
    A 20 kW diesel generator might use around 2 gallons per hour. A 500 kW industrial diesel generator may require close to 50 gallons per hour at full capacity.
  190. [190]
    Most Efficient Diesel Generator (Reliable Power & Best Engines)
    Oct 9, 2025 · A diesel generator reaches peak efficiency when operated at about 70–80% of its rated load. At this range, fuel is burned cleanly, and the ...
  191. [191]
    Power Generation | Engine Technology Forum
    Electric generators powered by internal combustion engines are an available, affordable, and proven means to maintain electric power.
  192. [192]
    Benefits of Reciprocating Engines in Power Generation
    Sep 1, 2020 · Many experts believe power plants constructed with reciprocating engines are the ideal complement to intermittent renewables.
  193. [193]
    Reciprocating Engines: Powering the Future - Stanley Consultants
    Aug 5, 2025 · RICE power plants can start quickly and reach full capacity within minutes, making them ideal for providing immediate power during outages or ...<|control11|><|separator|>
  194. [194]
    What is a Reciprocating Engine Generator? - Microgrid Knowledge
    Aug 16, 2016 · Reciprocating internal combustion engines provide incremental energy almost instantaneously at a relatively low capital cost.
  195. [195]
    Engines? Turbines? Both? Choosing Power for CHP Projects | Cat
    Gas reciprocating engines and gas turbines both perform well in cogeneration service. The choice of technologies depends on the application characteristics. ...Missing: internal | Show results with:internal
  196. [196]
    [PDF] A global comparison of the life-cycle greenhouse gas emissions of ...
    Jul 1, 2021 · ... g CO2/km for small, 109–122 g CO2/ km for lower medium, and 132–134 g CO2/km for SUV segment gasoline and diesel cars (Díaz et al., 2020) ...
  197. [197]
    [PDF] Fact Sheet: Gasoline vs. Diesel
    May 7, 2019 · Testing both vehicles according to the conditions of the new WLTP resulted in higher CO2 emission levels, at 139 g/km for the TDI and 126 g/km ...
  198. [198]
    THE EXHAUST EMISSIONS FROM DIESEL ENGINE AND THIER ...
    Jan 19, 2025 · NOx has the highest proportion of diesel pollutant emissions with a rate of more than 50 %. After NOx emissions, PM has the second highest ...
  199. [199]
    Real-Time Analysis of NOx Emissions in Heavy-Duty Diesel Vehicles
    May 10, 2025 · 60% of the HDDVs exceeded the PROCONVE P-7 NOx limits, reaching 8.3 g kWh-1. •. 90% of NOx emissions occur at low speeds (up to 40 km/h).Missing: quantitative | Show results with:quantitative
  200. [200]
    Comparison of the emission factors of air pollutants from gasoline ...
    Jul 15, 2022 · EFs for CO, HC, and NO were compared for the four fuels. Diesel vehicles had the highest EF for CO, with an average value of 35.12 ± 21.37 g/kg ...
  201. [201]
    Vehicle emissions of CO, NOx and HCs? - Thunder Said Energy
    Vehicle emissions of CO, NOx and HCs have all fallen by 20-60x over the past 50-years, to 5 grams/mile, 0.2 grams/mile and 0.3 grams per mile, respectively.
  202. [202]
    50 Years of EPA's Automotive Trends Report | US EPA
    Jan 15, 2025 · New vehicles in 2023 emitted less than half the CO2 per mile of 1975 models, with a 31% reduction in tailpipe emissions since 2004. Line chart ...
  203. [203]
    [PDF] A historical review of the U.S. vehicle emission compliance program ...
    The provision protects vehicle owners from the cost of repairs for emissions-related failures that cause the vehicle or engine to exceed emission standards. The ...
  204. [204]
    Effect of EGR on Emissions and Engine Performance - DieselNet
    EGR is an effective strategy to control NOx emissions from diesel engines. NOx emissions may be further reduced by cooling of the EGR stream.
  205. [205]
    Experimental Study on NOx Reduction of Diesel Engine by EGR ...
    Feb 8, 2024 · Exhaust gas recirculation (EGR) and selective catalytic reduction (SCR) are crucial technologies for mitigating nitrogen oxide (NOx) emissions ...
  206. [206]
    Timeline of Major Accomplishments in Transportation, Air Pollution ...
    Nov 22, 2024 · 1970 Congress passes the first major Clean Air Act, requiring a 90 percent reduction in emissions from new automobiles by 1975.
  207. [207]
    EGR, SCR, and DPF Technology to Reduce Diesel Engine Emissions
    Apr 4, 2024 · SCR is highly effective in reducing NOx emissions, with efficiency levels reaching more than 90%. DPF effectively reduces particulate emissions ...
  208. [208]
    Investigation of the impact of the configuration of exhaust after ...
    Jun 1, 2020 · Two technologies are commonly applied to reduce NOx emission, namely the Lean NOx Trap (LNT) and the Selective Catalytic Reduction (SCR) device.
  209. [209]
    Highlights of the Automotive Trends Report | US EPA
    Nov 25, 2024 · On this page: Average new vehicle CO2 emissions and fuel economy continue to improve, aided by growing production of electric vehicles.
  210. [210]
    Effect of Exhaust Gas Recirculation Combined with Selective ... - NIH
    Jun 22, 2022 · This paper studied the effect of EGR combined with SCR on the NO x emission characteristics of a heavy-duty diesel engine based on the engine bench test.
  211. [211]
    Impact of Modern Vehicular Technologies and Emission ... - MDPI
    Jul 18, 2023 · Compared to Euro 6, NOx emissions will be reduced by 35% for cars and vans, and by 56% for buses and lorries. Particle emissions from the ...
  212. [212]
    4 Power Train Technologies for Reducing Load-Specific Fuel ...
    Daimler's Detroit Diesel estimates that, with the installation of SCR in 2010, fuel consumption will be reduced 3 to 4 percent by a combination of higher engine ...<|separator|>
  213. [213]
    Combustion Engine Noise – a Historical Review - ResearchGate
    Measurements of vehicles produced around 1910 emit-ted sound power levels of their industrial engines around 100 dB(A) with an acous-tical efficiency of 1 - 2× ...
  214. [214]
    [PDF] Analyzing the Sources of Noise in Internal Combustion Engines
    Feb 28, 2025 · Engine noise levels usually range from 80 to 110 dBA when measured at a distance of 1 meter from the engine surface [5],The intensity of this ...
  215. [215]
    Highly effective sound insulation for internal combustion engines
    Between 2016 and 2026 the noise limits are to be reduced by a total of three to four decibels (dB), depending on the vehicle type. Three decibels is equivalent ...<|separator|>
  216. [216]
    Engine Vibration: What It Is and How to Control It. - Fluidampr
    Engine vibration limits power and durability. Learn how to control it with proper damping to protect your build and maximize performance.
  217. [217]
    [PDF] review on internal combustion engine vibrations - ijeset
    The vibration reduction will be possible by minimizing unbalanced forces and by providing the anti vibration mounts at the engine-vehicle interface. Many.
  218. [218]
    5 Innovations That Reduce Engine Vibrations - Auto | HowStuffWorks
    Jul 20, 2012 · In a sense, engine mounts are the first line of defense in suppressing engine shakes from propagating throughout the entire vehicle. First, we ...Flexible Engine Mount Pieces · Better Ignition Components
  219. [219]
    Near-highway pollutants in motor vehicle exhaust: A review of ...
    Aug 9, 2007 · The most widely reported pollutants in vehicular exhaust include carbon monoxide, nitrogen and sulfur oxides, unburned hydrocarbons (from fuel ...Missing: internal | Show results with:internal
  220. [220]
    Diesel Engines & Public Health | Union of Concerned Scientists
    Sep 26, 2024 · Ultrafine particulates, which are small enough to penetrate the cells of the lungs, make up 80 to 95 percent of diesel soot pollution.
  221. [221]
    Air Pollution, Ultrafine Particles, and Your Brain: Are Combustion ...
    Feb 22, 2022 · There is a growing body of evidence suggesting UFP may play a bigger role inflicting adverse health impacts than has been recognized.
  222. [222]
    A Work Group Report on Ultrafine Particles (AAAAI) Why Ambient ...
    Studies have shown that ambient ultrafine particles have detrimental effects on both the cardiovascular and respiratory systems.
  223. [223]
    Emissions of Carbonaceous Particulate Matter and Ultrafine ... - MDPI
    As the emission exhaust cools in the exhaust manifold and associated pipework, carbon particles agglomerate, forming high-surface-area material onto which ...
  224. [224]
    Life Cycle Emissions: EVs vs. Combustion Engine Vehicles
    Jun 23, 2023 · The production emissions for BEVs are approximately 40% higher than those of hybrid and ICE vehicles. · Electricity production is by far the most ...
  225. [225]
    [PDF] Life Cycle Analysis Comparison - Transportation Energy Institute
    This is a Life Cycle Analysis Comparison of electric and internal combustion engine vehicles from January 2022.
  226. [226]
    [PDF] R&D GREET 2023 Annual Update: Light Duty Vehicles
    Overall, R&D GREET shows that the 2024 EV produces 52% less GHG emissions than a comparable ICE vehicle. R&D GREET can also track progress and projected ...
  227. [227]
    The Environmental Impact of Battery Production for Electric Vehicles
    Jan 11, 2023 · The environmental impact of battery production comes from the toxic fumes released during the mining process and the water-intensive nature of the activity.
  228. [228]
    Lifecycle carbon footprint comparison between internal combustion ...
    The findings indicated that EVs, in comparison to ICEVs, might offer significant reductions in greenhouse gas emissions if recycled properly. Table 1 summarizes ...
  229. [229]
    Invention of the Internal Combustion Engine
    The internal combustion engines developed by Nikolaus Otto (1832-1891) in the 1860s and Rudolf Diesel (1858-1913) in the 1890s.
  230. [230]
    The Impact of the Combustion Engine on the Industrial Revolution
    May 18, 2023 · Its introduction revolutionized industries, transportation, and agriculture, leading to increased productivity, economic growth, and the ...
  231. [231]
    How the Model T paved the road to profit - The CFO
    May 26, 2024 · Driven forward by revolutionary manufacturing methods, the Model T's sticker price plummeted to $260 by 1925, thrusting personal transportation ...
  232. [232]
    The Ford Model T: America was once an emerging market
    Aug 17, 2016 · The Model T, for instance, spurred significant investment in American infrastructure by providing a tax base to support construction of roads.
  233. [233]
    Wheels of Change: The Automotive Industry's Sweeping Effects on ...
    Model T Fords, and similar cars of the era, were put to good use performing farm chores -- pulling plows and towing wagons. A few years after the Model T Ford ...
  234. [234]
    The Impact of Internal Combustion Engines on Society ... - LinkedIn
    Dec 12, 2023 · The economic impact of internal combustion engines cannot be overstated. The automotive industry, in particular, has become a cornerstone of ...
  235. [235]
    US Economic Contributions | American Automotive Policy Council
    Automakers and their suppliers are America's largest manufacturing sector, responsible for 3% of America's GDP. 2 No other manufacturing sector generates as ...
  236. [236]
    Driving US Economy - Alliance For Automotive Innovation
    2024 Sales vs.​​ Auto manufacturing drives more than $1.5 trillion into the economy each year through the sales and servicing of autos and flows through the ...
  237. [237]
    Internal Combustion Engine Market Size, Share | Growth [2032]
    Innovations such as variable valve timing, turbocharging, and advanced combustion technologies improve ICEs' efficiency, driving the demand for internal ...
  238. [238]
    The Impact of the Auto Industry on the Economy
    This output boosted the macroeconomy substantially, contributing over one-third of all growth in gross domestic product (GDP) during the fourth quarter. In fact ...The Auto Slowdown As An... · 3. Vehicle Sales From... · Tracking Midwest...
  239. [239]
    The Model T | The Journal of Economic History | Cambridge Core
    Jan 21, 2025 · The Model T was revolutionary both because of the mass production methods used to produce it and because of Ford's unusual decision to design a ...
  240. [240]
    How Many Miles Can an Engine Really Last? The Definitive Guide
    Sep 23, 2024 · The average engine lifespan is around 150,000 to 200,000 miles. However, with proper maintenance and care, some engines can last much longer.
  241. [241]
  242. [242]
  243. [243]
    The Future of Internal Combustion Engines | Jaytrac
    Durability – internal combustion engines provide a high degree of durability; they can withstand heavy loads and difficult terrain.
  244. [244]
    Theoretical energy density of different batteries and gasoline
    For instance, gasoline exhibits an energy density of 12,700 Wh/kg, which is approximately 63 times greater than that of a Li-ion battery [22]. While several ...
  245. [245]
    Refueling Times of Combustion Engine vs Electric Vehicles
    May 31, 2024 · While ICEVs have traditionally held the edge in terms of refueling speed, EVs are rapidly closing the gap with advancements in charging ...
  246. [246]
    Waiting for Superbatteries - IEEE Spectrum
    Dec 1, 2022 · The best energy density now commercially available in very large quantities for lithium-ion batteries is at 750 Wh/L, which is widely seen in ...
  247. [247]
    EV vs ICE: Surprising differences in efficiency, cost, and impact
    In fact, for every dollar spent on gasoline, only 20 cents of it is used to move an ICE vehicle along the road. EVs, however, operate at about 87% - 91% ...
  248. [248]
    How does EV repair differ from an ICE vehicle?
    Dec 20, 2023 · Lithium-ion battery replacement costs, when not covered by insurance or warranty, vary from $5,000 to $20,000, based on size, manufacturer, and ...
  249. [249]
    Internal Combustion Engine | Dorleco
    Jul 15, 2025 · Engines with continuous combustion; These engines are constantly fed fuel and oxidizers, which create a steady flame. Among the examples are ...
  250. [250]
    Summary of Euro 7 emission standards for passenger cars & trucks
    Summary of the near-final Euro 7 Regulation for passenger cars and heavy-duty vehicles. Minimal changes to LD limits, particles down to 10nm, limits for tires ...
  251. [251]
    [PDF] The stringency of the proposed Euro 7 regulation for cars and vans
    To assess the stringency of the proposed Euro 7 emission limits for passenger cars and vans, the following analysis compares the emission limits and test ...
  252. [252]
    EPA Announces Plan to Reverse Federal EV Mandates
    The EPA will reconsider its March 2024 vehicle emissions regulation for model-years '27-'32, which effectively requires two out of three new vehicles sold to be ...
  253. [253]
    Trump EPA aims to repeal vehicle emission rules after revoking ...
    Jul 24, 2025 · The U.S. Environmental Protection Agency plans to repeal all greenhouse gas emission standards for light-duty, medium-duty, and heavy-duty ...
  254. [254]
    Final Rule: Multi-Pollutant Emissions Standards for Model Years ...
    These standards will phase in over model years 2027 through 2032. Final Rule: Multi-Pollutant Emissions Standards for Model Years 2027 and Later Light-Duty and ...Missing: combustion | Show results with:combustion
  255. [255]
    Global Car Bans: Ambition Meets Reality - Transport Energy Strategies
    Apr 25, 2025 · Key Takeaways from the 2025 Update · 70+ countries have announced ICEV phaseouts or ZEV sales targets. · Hybrid flexibility and compliance credit ...<|separator|>
  256. [256]
    Will synthetic fuel save the internal combustion engine?
    Feb 29, 2024 · Synthetic fuels are available to buy now. Allowed for some ICE cars post-2035 in EU. Cost and scale still an issue.
  257. [257]
    Keeping Up With the Trump EPA: September 2025 Update
    When President Trump signed into law the CRA resolutions earlier this year, the federal government reversed California's internal combustion engine (ICE) bans.
  258. [258]
    Congress Blocks California's Gasoline Car Ban: A Legal and Policy ...
    Jun 1, 2025 · In December 2024, the Biden administration's EPA granted such waivers, allowing California to proceed with its Advanced Clean Cars II program, ...Missing: timelines | Show results with:timelines
  259. [259]
    What Nobody Is Telling You About E-Fuels And The Threat To EVs
    Feb 11, 2025 · E-fuels or electrofuels, also known as synthetic fuels could end up being the future of the future, and threaten the dominance of Battery Electric Vehicles ( ...
  260. [260]
    Comparison of potentials for Life Cycle CO2 Emission Reduction
    Aug 6, 2025 · Results shows that e-fueled ICE and e-fueled hybrid vehicles have greater CO₂ emissions reduction potential than EVs. ResearchGate Logo.
  261. [261]
    On the pursuit of emissions-free clean mobility – Electric vehicles ...
    Jun 1, 2023 · There is an ongoing debate between proponents of so-called e-fuels (being synthetic fuels made from atmospheric carbon dioxide, water, and ...
  262. [262]
    Automotive industry experts split over role of e-fuels - Fastmarkets
    May 13, 2024 · The viability of e-fuels as an alternative to electric power is debated due to their high cost and unclear role in decarbonization efforts.Missing: feasibility | Show results with:feasibility
  263. [263]
    Synthetic fuel debate rages in Europe - GoAuto
    Jun 8, 2023 · The argument centres around the point that the production of e-fuel will divert sustainably sourced electricity from “more pressing uses”, ...Missing: feasibility | Show results with:feasibility<|separator|>
  264. [264]
    Comparison between EV, HEV, PHEV, and ICE vehicles to achieve ...
    Nov 1, 2023 · Note that the data of Internal Combustion Engine (ICE) vehicle is classified into two types: gasoline and diesel engine. ... use and greenhouse ...
  265. [265]
    Decarbonizing Combustion Vehicles – A Critical Part in Reducing ...
    Reducing fuel carbon intensity, using biofuels, and improving fuel production processes are key to decarbonizing ICEVs. 73% of emissions come from fuel ...Missing: debates | Show results with:debates
  266. [266]
    [PDF] exploring risk perceptions of fossil fuels, e-fuels, and electric drives ...
    Aug 5, 2024 · E-fuels and electric drives are perceived more positively than fossil fuels, with e-fuels having the lowest risk perceptions. Car affinity ...