Fact-checked by Grok 2 weeks ago

Thrust


Thrust is the reaction force generated by a system through the of mass, such as exhaust gases or air, in one direction, producing an equal and opposite force that propels a forward, in accordance with Newton's third law of motion.
This force is essential for overcoming aerodynamic in and gravitational forces in rockets, enabling sustained motion through the atmosphere or .
In rocketry, thrust is primarily produced by expelling high-velocity gases from a , with the magnitude approximated by the product of the exhaust and the effective exhaust in conditions.
Jet engines generate thrust by ingesting ambient air, compressing and combusting it with to accelerate the exhaust, incorporating both change of the and differences across the .
Key performance metrics include , which quantifies efficiency as thrust per unit of consumed, and , which indicates the system's ability to accelerate a against .
, achieved by directing the exhaust , enhances maneuverability in and missiles by allowing control over the force's direction.

Fundamentals of Thrust

Definition and Physical Nature

Thrust is a generated by accelerating a of gas or in one direction, producing a that propels a vehicle or object in the opposite direction. This is fundamental to systems in , rockets, and other vehicles, where it counteracts or gravitational forces to enable motion. In essence, thrust represents the net forward-directed component of the reaction to the expulsion or deflection of . The physical nature of thrust stems directly from Newton's third law of motion, which states that for every action, there is an equal and opposite reaction. When a propulsion system expels high-velocity exhaust gases rearward, the momentum change of that mass imparts an equal momentum change forward to the vehicle, manifesting as thrust. Quantitatively, thrust magnitude is the product of exhaust velocity relative to the vehicle and the mass flow rate of the expelled fluid, as expressed in the rocket thrust equation \mathbf{T} = \mathbf{v} \frac{dm}{dt}, where \mathbf{v} is the effective exhaust velocity vector and \frac{dm}{dt} is the rate of mass expulsion. This formulation highlights thrust as a rate of change of momentum, underscoring its dependence on both velocity and mass flux rather than pressure alone. As a quantity, thrust has both and , aligned with the vehicle's primary of motion, and is measured in newtons () in the (), equivalent to kg·m/s². Its instantaneous value can vary with operating conditions, such as settings or , but it fundamentally arises from conserved in isolated systems, independent of the surrounding medium in environments like . In air-breathing engines, thrust incorporates additional terms across the and exhaust, but the core mechanism remains the of .

Underlying Principles from Classical Mechanics

Thrust represents the mechanical generated by systems through the of , grounded in Newton's third law of motion, which asserts that every action has an equal and opposite reaction . In rocket engines, hot gases are expelled rearward at high velocity, imparting to the exhaust while producing an equal forward reaction on the vehicle. This principle applies universally in to systems where arises from transfer, independent of the surrounding medium. The quantitative foundation derives from the conservation of linear for variable-mass systems. Consider a in free space with no external forces; the total remains constant. If the ejects a small \Delta m at \mathbf{v_e} rearward, the change in balances the of the exhaust, yielding the thrust as \mathbf{T} = \mathbf{v_e} \frac{dm}{dt}, where \frac{dm}{dt} > 0 denotes the positive exhaust . This relation emerges from applying the \frac{d\mathbf{p}}{dt} = \mathbf{F_{ext}} + \mathbf{u} \frac{dm}{dt}, with \mathbf{u} = -\mathbf{v_e} for mass ejection and \mathbf{F_{ext}} = 0 in ideal isolation. For in atmosphere, extends to include ingested air , where thrust equals the : T = \dot{m_e} v_e - \dot{m_i} v_i, with \dot{m_e} and v_e as exhaust and , and \dot{m_i}, v_i as values; in steady , this simplifies under classical assumptions of incompressible or behavior. Newton's second law complements this by relating thrust to via \mathbf{F} = m \frac{d\mathbf{v}}{dt}, but for , the full equation incorporates the thrust term explicitly. These derivations hold in non-relativistic regimes, as validated by empirical tests since the early 20th century, confirming momentum-based force generation without reliance on external reaction surfaces.

Measurement and Units

Thrust, as a quantity representing , is quantified using units of in standard measurement systems. In the (SI), the unit is the (N), defined as the force required to accelerate a of one at one meter per second squared (kg·m/s²). In the imperial system, commonly used in contexts such as U.S. , thrust is expressed in pounds- (lbf), where 1 lbf equals approximately 4.448 N and corresponds to the force exerted on one pound- under . Direct measurement of thrust typically occurs during ground testing using specialized force transducers. For rocket engines, load cells—precision scales calibrated against known weights—quantify the force generated by exhaust expulsion, often integrated into thrust stands that isolate the engine from external vibrations. thrust stands, recommended by for their accuracy in low-thrust applications like electric , suspend the engine and measure displacement or oscillation induced by thrust, calibrated via hanging weights or electrostatic forces to achieve uncertainties below 1%. In aircraft jet engines, static thrust is assessed on rigs or scales during sea-level tests, while in-flight approximations rely on indirect metrics like (EPR), the ratio of exhaust to inlet pressure, correlated to thrust via manufacturer calibration curves. Thrust magnitude is often normalized for performance evaluation, such as (dimensionless, in g's or as a multiple of ) or (Isp, in seconds or m/s), but these derive from primary measurements rather than replacing them. Calibration ensures traceability to national standards, with errors minimized through simulations for space propulsion to account for effects absent in flight.

Historical Evolution

Ancient and Pre-Modern Origins

The principle of thrust, arising from the reaction to expelled mass or fluid, was first empirically demonstrated in antiquity through rudimentary devices exploiting steam propulsion. In the 1st century AD, described the , a spherical vessel mounted on an axis and heated to produce steam that escaped through tangential nozzles, generating rotational torque via unequal reaction forces from the jets. This apparatus, detailed in Hero's treatise Pneumatica, illustrated the foundational reactive propulsion mechanism without practical application beyond demonstration, as it prioritized curiosity over utility in the Hellenistic engineering tradition. Centuries later, chemical emerged with the invention of in around 850 AD by Taoist alchemists seeking an elixir of immortality, yielding a that enabled expulsive thrust in weaponry. By the , this led to fire s— tubes filled with attached to arrow shafts—providing auxiliary thrust to extend range and speed during flight. The earliest documented military use occurred in 1232 AD during the Siege of , where forces deployed barrages of these "arrows of flying fire" against Mongol invaders, marking the transition from incendiary to propulsive ry. These devices relied on rapid gas expulsion for linear thrust, though stabilization and control remained primitive, limited by inconsistent burn rates and stick guidance. In medieval , knowledge of and rocket-like weapons disseminated via and trade routes, influencing Islamic and Byzantine military texts by the 13th century, yet innovations stagnated without systematic refinement until the . Empirical observations of in cannons and further hinted at reaction principles, but absent a theoretical framework—such as Isaac Newton's third law articulated in 1687—no unified concept of thrust as conserved transfer developed. These pre-modern efforts thus laid groundwork through trial-and-error, prioritizing ballistic efficacy over efficiency or scalability.

Development in the 20th Century

The 20th century marked a pivotal era for thrust generation, transitioning from theoretical concepts to practical propulsion systems enabling supersonic flight and space access. In rocketry, achieved the first liquid-fueled launch on March 16, 1926, near , using and to produce thrust that propelled the vehicle to an altitude of 41 feet for 2.5 seconds. This demonstrated the viability of continuous combustion for sustained thrust, contrasting with earlier solid-fuel limitations. Independently, aviation propulsion advanced through concepts; patented a engine design in 1930, emphasizing axial compression for efficient air intake and exhaust acceleration to generate thrust via Newton's third law. World War II accelerated implementation, with achieving the first jet-powered aircraft flight on August 27, 1939, using Hans von Ohain's HeS 3B centrifugal-flow in the , producing approximately 1,100 pounds of thrust. The became the first operational jet fighter in 1944, powered by two engines each delivering 1,980 pounds of thrust, enabling speeds over 540 mph and highlighting thrust-to-weight advantages over piston engines. Concurrently, German engineers under developed the from 1936 to 1942, with its first successful test flight on October 3, 1942; the liquid-propellant engine, burning alcohol and , generated 60,000 pounds of thrust to reach suborbital altitudes exceeding 50 miles. These systems underscored thrust's role in overcoming drag and gravity, though early designs suffered from inefficiencies like high fuel consumption. Postwar efforts refined thrust mechanisms for commercial and space applications. ’s Gloster E.28/39 flew with Whittle’s on May 15, 1941, validating scalability, while the U.S. and Soviet programs adapted V-2 technology for ballistic missiles and launch vehicles. engines emerged in the 1950s, incorporating a to bypass additional air for thrust augmentation, improving efficiency; by the 1960s, high-bypass variants like those on the 707 (introduced 1958) achieved specific fuel consumption reductions of up to 30% compared to pure . In space propulsion, the Saturn V's F-1 engines, first static-tested in 1964, delivered 1.5 million pounds of thrust per engine using RP-1 and , powering Apollo missions in 1969 and exemplifying clustered liquid-propellant designs for orbital escape. These advancements prioritized verifiable thrust metrics, such as pounds-force and exhaust , driving empirical iterations despite material and control challenges.

Contemporary Advances Since 2000

In aviation propulsion, engines emerged as a key advancement, enabling higher bypass ratios for improved thrust efficiency and reduced fuel burn. The series, introduced in the mid-2010s, incorporates a planetary gear system allowing the fan to rotate slower than the , achieving up to 20% better specific fuel consumption compared to prior high-bypass turbofans while delivering thrust ratings from 24,000 to 35,000 lbf. Similarly, GE Aviation's GE9X engine, certified in 2019 for the , produces 134,300 lbf of thrust—the highest for any commercial —and incorporates like ceramic matrix composites for higher operating temperatures and efficiency gains of 10% over its predecessor. Rocket propulsion saw transformative developments through reusable and high-performance engines. SpaceX's Merlin 1D engines, deployed on since 2013, generate 845 kN of sea-level thrust each using a with and , facilitating over 300 successful orbital launches and booster recoveries by 2025. The engine family, introduced for in the late 2010s, employs a full-flow with and oxygen, yielding over 2,300 kN of thrust per engine in vacuum-optimized variants, with 3 achieving simplified designs and thrust-to-weight ratios exceeding 200 by 2024 static fires. Electric propulsion systems advanced toward higher power handling for deep-space applications. NASA's Evolutionary (NEXT) project, initiated in the 2000s, produced gridded thrusters with up to 4,190 seconds and thrusts of 0.236 N at 6.9 kW, demonstrated in tests by 2017 for potential use in future missions like crewed Mars transfers. The Variable Specific Impulse Magnetoplasma Rocket (VASIMR), developed by , reached 120 kW power levels in tests by 2022 with the VX-200SS prototype, enabling variable exhaust velocities from 30 to 120 km/s for optimized thrust in variable gravity environments. Hypersonic propulsion progressed with scramjet demonstrations, emphasizing sustained supersonic combustion. The Boeing X-51A Waverider, tested in 2010, operated a hydrocarbon-fueled at for over 200 seconds, generating thrust via air-breathing at altitudes up to 70,000 feet. Additive manufacturing further enabled complex geometries in components, such as regenerative cooling channels in nozzles, reducing part counts by up to 85% and accelerating development cycles, as applied in NASA's metal AM rocket engines since the 2010s.

Mechanisms of Thrust Generation

Air-Breathing Propulsion Systems

Air-breathing propulsion systems produce thrust by drawing in atmospheric air, mixing it with fuel for combustion, and accelerating the exhaust gases rearward to impart forward momentum to the vehicle, leveraging the Brayton thermodynamic cycle in most modern variants. These engines eliminate the need for onboard oxidizers, relying instead on ambient oxygen, which reduces mass and enhances efficiency for sustained atmospheric operations compared to rocket systems that must carry both fuel and oxidizer. Thrust arises primarily from the change in momentum of the airflow: ingested air at freestream velocity v_\infty is accelerated to exhaust velocity v_f, yielding net thrust T = \dot{m} (v_f - v_\infty), where \dot{m} is the mass flow rate, augmented by pressure differences at the nozzle exit. Piston-engine and systems, suited for low-speed flight below 0.6, generate thrust indirectly through propellers that accelerate a large to modest velocities, achieving high via T = \dot{m} v_d where v_d approximates the velocity increment. In , a drives the , with the core engine operating on to produce shaft power, as demonstrated in aircraft like the which has utilized such systems since 1952 for long-range cruise at speeds up to 925 km/h. These configurations excel in , with specific fuel consumption around 0.5 lb/hp-hr, due to the 's ability to match exhaust velocity closely to flight speed, minimizing kinetic energy losses. Turbojet and turbofan engines dominate subsonic to supersonic applications, compressing inlet air via rotating machinery before and expansion through turbines that power the , with final acceleration in a . s, as in the General Electric J79 used in the F-4 Phantom II since 1958, expel core flow at high velocity for thrust levels exceeding 79 kN, effective up to but inefficient at low speeds due to high exhaust velocities mismatched to flight speed. Turbofans mitigate this by ducting a portion of airflow around the core ( up to 12:1 in high-bypass variants like the CFM56), blending low-velocity fan thrust with core jet thrust for overall efficiencies improved by 20-30% over pure s, powering aircraft such as the with takeoff thrusts around 120 kN per engine. Ramjets and scramjets extend air-breathing to high supersonic and hypersonic regimes, forgoing mechanical in favor of aerodynamic ram effects from high flight speeds. Ramjets, operational from Mach 3 to 6, ignite fuel in slowed airflow, as tested in the X-15's XLR99 auxiliary since 1960, producing thrusts up to 267 kN through , , and . Scramjets maintain supersonic for 6+ velocities, avoiding thermal choking; NASA's X-43A achieved 9.68 on November 16, 2004, using a hydrogen-fueled with air captured at over 2 km/s, demonstrating sustained thrust via minimized drag and efficient oxygen utilization at altitudes above 30 km. These systems demand initial acceleration by rockets or other means, with challenges in fuel-air mixing and heat management limiting operational durations to seconds in current prototypes.

Rocket and Expulsive Mass Systems

Rocket propulsion systems generate thrust through the expulsion of high-velocity propellant mass, invoking Newton's third law of motion, where the forward force on the vehicle equals the rate of momentum change of the ejected mass. Unlike air-breathing engines, rockets carry both fuel and oxidizer, enabling operation in vacuum or sparse atmospheres. The fundamental thrust equation in vacuum simplifies to T = \dot{m} v_e, with \dot{m} as the propellant mass flow rate and v_e as the effective exhaust velocity relative to the rocket. In atmospheric conditions, an additional term accounts for exhaust pressure differential: T = \dot{m} v_e + (p_e - p_a) A_e, where p_e is nozzle exit pressure, p_a ambient pressure, and A_e exit area. Efficiency in rocket systems is quantified by , defined as I_{sp} = v_e / g_0, where g_0 is (9.81 m/s²), yielding units of seconds. Higher I_{sp} indicates better utilization for velocity change. Chemical s, dominant in current applications, achieve I_{sp} values from approximately 200–300 seconds at for propellants to 400–450 seconds in for bipropellants like and oxygen. rocket motors combust pre-mixed grains for high thrust but limited controllability and lower I_{sp}, while engines pump separate and oxidizer for throttleability and higher efficiency. The Tsiolkovsky rocket equation governs achievable velocity increment: \Delta v = v_e \ln(m_0 / m_f), derived from conservation of momentum assuming constant exhaust velocity and no external forces. Here, m_0 is initial mass and m_f final mass after propellant expulsion. This exponential mass ratio requirement underscores the challenge of multi-stage designs for orbital insertion, as single-stage vehicles struggle to reach escape velocities exceeding 11 km/s due to structural and propellant mass penalties. Expulsive mass systems extend beyond chemical rockets to any mechanism relying on reaction mass ejection, such as nuclear thermal rockets heating to v_e \approx 8–9 km/s for I_{sp} \sim 900 seconds, though undeveloped for routine use. Pure momentum expulsion without , like gas thrusters, yields low I_{sp} (50–100 seconds) suitable for attitude control. Thrust scales with \dot{m} and v_e, but constraints limit scaling, as kinetic imparted to exhaust is P = \frac{1}{2} \dot{m} v_e^2.

Electric and Non-Thermal Propulsion

Electric propulsion systems produce thrust by accelerating ionized using electrostatic or electromagnetic fields powered by , bypassing the thermal expansion of gases characteristic of conventional engines. This approach yields exhaust velocities of 20 to 50 km/s, corresponding to specific impulses of 2,000 to 9,000 seconds, enabling substantial mass savings for long-duration missions despite low thrust levels typically ranging from micronewtons to newtons. Thrust arises from the momentum change of the accelerated ions or , governed by \mathbf{T} = \dot{m} v_e, where \dot{m} is the and v_e the effective exhaust , with electrical P related via P \approx \frac{1}{2} \dot{m} v_e^2 for efficient conversion. Electrostatic variants, such as gridded thrusters, ionize neutral —commonly —via electron bombardment in a chamber. Positive ions are then electrostatically extracted and accelerated through multi-aperture grids separated by 0.5 to 1 mm, with the screen grid at positive potential and accelerator grid at negative, creating fields up to 3 kV. Ions achieve directed without significant component, exiting as a whose imparts thrust; a separate neutralizer emits electrons to prevent charging. NASA's NEXT thruster exemplifies this, delivering 236 mN thrust at 6.9 kW input with 4190 s , validated through over 48,000 hours of ground testing. Electromagnetic systems, including Hall-effect thrusters, generate thrust through closed-drift acceleration in an annular channel. gas flows past a central while a radial (0.01-0.1 T) traps electrons, forming a Hall current that ionizes the gas via collisions and establishes an axial of 100-300 V/cm. Unmagnetized ions are accelerated by this self-sustaining field to exhaust velocities of 10-20 km/s, with thrust transferred via magnetic interaction with the thruster structure. These devices achieve thrust densities up to 0.1 N/kW, higher than gridded ions, and have powered missions like ESA's , producing 68 mN at 1.5 kW. Non-thermal propulsion extends to propellantless mechanisms like solar sails, which exploit transfer from without expelling mass. For a perfectly reflecting , thrust is T = \frac{2 I A \cos^2 \alpha}{c}, where I is solar intensity (1366 W/m² at 1 ), A area, \alpha incidence angle, and c , yielding ~9 μN/m² near Earth. The 2010 mission deployed a 200 m² , generating ~1.1 mN to enable interplanetary cruise and de-spin maneuvers, demonstrating viability for continuous, low-acceleration trajectories. Emerging concepts, such as electric sails using charged tethers to deflect protons, promise similar non-thermal coupling but remain experimental.

Core Analytical Concepts

Thrust Equations and Derivations

The fundamental derivation of thrust equations stems from the conservation of linear within a surrounding the system, as governed by the integral form of the momentum theorem. In steady-state operation, the net axial force (thrust) balances the difference in momentum flux across the inlet and exit boundaries, augmented by pressure forces at the exit if not matched to ambient conditions. This yields the general thrust equation: F = \dot{m}_e v_e - \dot{m}_0 v_0 + (p_e - p_0) A_e, where \dot{m} denotes , v axial , p , and A_e exit area, with subscripts e for exit conditions and $0 for . The terms represent momentum thrust from accelerated exhaust minus inlet ram , plus a pressure thrust correction. For rocket propulsion in vacuum, where no freestream inlet exists (\dot{m}_0 = 0, v_0 = 0), the equation simplifies, and if exit pressure matches ambient (p_e = p_0), thrust approximates T = \dot{m}_e v_e, with v_e as effective exhaust velocity relative to the vehicle. This form derives from the variable-mass form of Newton's second law: m \frac{dv}{dt} = -v_e \frac{dm}{dt} + F_\text{ext}, where for isolated systems (F_\text{ext} = 0), thrust T = v_e \dot{m} with \dot{m} = -\frac{dm}{dt} > 0 as the positive propellant expulsion rate. The pressure term (p_e - p_0) A_e accounts for incomplete expansion, significant in underexpanded or overexpanded nozzles, as quantified in nozzle performance analyses. In air-breathing engines like turbojets, the full general equation applies, with inlet ram drag \dot{m}_0 v_0 subtracting from gross thrust, yielding net thrust T = \dot{m}_e (v_e - v_0) + (p_e - p_0) A_e under matched inlet mass flow (\dot{m}_e = \dot{m}_0). Derivation assumes one-dimensional flow and neglects viscous drag on engine surfaces, validated through control volume analysis where momentum influx from ingested air reduces effective propulsion. For or systems, thrust derives from actuator disk theory, modeling the device as an infinitesimally thin disk imparting to . Axial balance gives T = \dot{m} (v_f - v_\infty), where v_f is far-wake , v_\infty , and \dot{m} = \rho A v_d with disk-averaged v_d = \frac{1}{2} (v_f + v_\infty) from considerations. For static conditions (v_\infty = 0), this yields T = \frac{1}{2} \rho A v_f^2, linking thrust to induced and . Electric propulsion, such as ion thrusters, follows the same principle, with thrust T = \dot{m} v_e from accelerated ions or , where v_e results from acceleration rather than , enabling high exhaust velocities but low \dot{m}. Derivations across systems unify under , with deviations arising from working fluid ( vs. atmospheric air) and acceleration mechanisms.

Relations to Power and Efficiency

The useful delivered by a propulsion system to propel a is the product of and vehicle , P = T v, representing the rate of work done against or to accelerate the vehicle. This relation holds across propulsion types, including jets, rockets, and propellers, as it derives from the fundamental definition of mechanical as times velocity. The power input to the system, however, is the rate at which is added to the exhaust gases or propelled . For engines, thrust is T = \dot{m} v_e + (p_e - p_a) A_e, where \dot{m} is the , v_e the exhaust , p_e and p_a the exhaust and ambient s, and A_e the exit area; under conditions, the pressure term vanishes, simplifying to T \approx \dot{m} v_e. The corresponding jet power is P_j = \frac{1}{2} \dot{m} v_e^2 \approx \frac{T v_e}{2}, reflecting the imparted to the exhaust relative to the vehicle. Propulsive efficiency \eta_p quantifies how effectively this input power converts to useful , given by \eta_p = \frac{T v}{P_j} = \frac{2 v}{v + v_e} for ideal and , where v_e is the effective exhaust velocity relative to the . This efficiency peaks near 100% when v_e slightly exceeds v, as in low-speed systems where fluid acceleration is minimal, but drops to zero at static conditions (v = 0) since all energy becomes wasted exhaust . High-speed applications favor or with higher v_e, trading for sustained thrust despite lower \eta_p at speeds. For air-breathing engines, intake momentum reduces effective v_e, but the formula holds approximately; overall system efficiency also incorporates , typically 30-50% for turbojets. In static or low-speed scenarios, such as hover or takeoff, power requirements scale nonlinearly with thrust due to ambient fluid constraints in air-breathing systems. For static jets, thrust T = \frac{1}{2} \rho A v_f^2 and power P = \frac{1}{4} \rho A v_f^3, yielding P^2 = \frac{T^3}{2 \rho A}, where \rho is and A the effective area; thus, doubling thrust demands over 2.8 times the power. Rockets circumvent this dependence, enabling higher static thrust-to-power ratios via onboard acceleration, though at the cost of lower in atmosphere. Electric propulsion systems exhibit similar relations but prioritize high v_e and over raw thrust, with efficiency tied to input electrical power conversion.

Specialized Metrics and Configurations

Specific impulse (I_{sp}), a fundamental metric of propulsion efficiency, quantifies the thrust generated per unit of propellant mass flow rate, expressed as I_{sp} = \frac{v_e}{g_0} where v_e is exhaust velocity and g_0 is standard gravitational acceleration (approximately 9.81 m/s²). This yields units of seconds, with chemical rockets typically achieving 200–450 seconds and electric systems exceeding 1,000–10,000 seconds due to higher exhaust velocities at lower thrust levels. Higher I_{sp} enables greater change in velocity (Δv) for a given propellant mass via the Tsiolkovsky rocket equation, prioritizing it in mission design despite trade-offs with thrust magnitude. Thrust-to-weight ratio (TWR), defined as TWR = \frac{T}{mg} where T is thrust, m is or , and g is local , assesses capability and structural feasibility. Values exceeding 1 are required for liftoff in vertical ascent vehicles, with launch stages often targeting 1.2–1.5 to overcome and losses; for example, preliminary designs for small missiles specify TWR of 1.5 based on staged estimates. In electric , low TWR (often <<1) suits orbital maneuvers but demands precise . For air-breathing engines, (TSFC) measures efficiency as fuel per unit thrust, typically in g/(kN·s) or lb/(lbf·h), with turbofans achieving 0.3–0.6 lb/(lbf·h) at due to flow reducing fuel needs relative to thrust. configurations, such as (A_e/A_t, exit to throat area), specialize thrust output: sea-level nozzles limit ratios to 10–20 for pressure recovery, yielding higher ambient thrust, while vacuum-optimized ratios exceed 50–100, boosting I_{sp} by 10–20% in space but risking at altitude. Thrust vectoring configurations, implemented via gimbaled nozzles or fluid injection, enable directional control without auxiliary surfaces, critical for stability in single-engine rockets; for instance, differential throttling in clustered engines provides and . Solid rocket grains adopt specialized geometries—cylindrical, star, or finned—to tailor thrust profiles, regressing burn surfaces for neutral, progressive, or regressive curves matching phases. These metrics and setups interlink, as in low-thrust electric systems where high I_{sp} compensates for TWR deficits in deep-space applications.

Applications Across Domains

Aerospace and Aviation

![F-35 Heritage Flight Team performs in Bell Fort Worth Alliance AirShow.jpg][float-right] In and , thrust provides the forward force necessary to propel through the atmosphere, counteracting and enabling takeoff, cruising, and maneuvering. This force is generated by air-breathing systems that ingest ambient air, compress and combust it with to produce high-velocity exhaust, or by propellers that accelerate a larger volume of air at lower speeds. The fundamental mechanism relies on Newton's third law, where the rearward acceleration of air or exhaust gases imparts an equal and opposite forward reaction on the . Early relied on piston-engine-driven propellers, as demonstrated by the ' 1903 Flyer, which generated thrust via two 12-horsepower propellers rotating at 745 RPM to achieve sustained powered flight. Propellers excel in efficiency at subsonic speeds by moving a large of air through a modest velocity change, making them suitable for and aircraft used in regional transport. Transition to occurred with the He 178's first turbojet-powered flight on August 27, 1939, marking the advent of high-speed by accelerating a smaller of air to much higher velocities. Contemporary predominantly employs high-bypass engines, which route a significant portion of ingested air around (bypass ratios often exceeding 5:1) to augment thrust while minimizing compared to pure turbojets. This design enhances by better matching exhaust velocity to flight speed, reducing specific to levels around 0.5-0.6 lb/(lbf·h) for modern wide-body airliners. In military applications, fighter jets prioritize high thrust-to-weight ratios, often exceeding 1:1, to enable , rapid acceleration, and vertical climbs; for instance, such ratios allow sustained pitch-up maneuvers without loss of speed at lower altitudes. Thrust management is critical for metrics like takeoff and climb , with engines rated by static thrust (e.g., tens of thousands of pounds-force for large airliners) and adjusted via variable geometry or afterburners in high- scenarios. gains in turbofans stem from increased mass flow through the , producing up to 80% of total thrust in high-bypass configurations, which has driven fuel savings in long-haul flights since their adoption in the .

Space Exploration and Orbital Mechanics

In , thrust propels beyond Earth's atmosphere and enables by providing the necessary change in , or delta-v, to counteract gravitational forces and achieve stable orbits. Unlike atmospheric , engines generate thrust in solely through the expulsion of onboard at high exhaust , yielding a thrust force T = \dot{m} v_e + (P_e - P_a) A_e, where \dot{m} is the , v_e is the exhaust , P_e and P_a are and ambient pressures (with P_a = 0 in ), and A_e is the area; this results in higher vacuum-specific thrust compared to sea-level operation due to the absence of back-pressure losses. The Tsiolkovsky rocket equation governs achievable delta-v as \Delta v = v_e \ln(m_0 / m_f), where m_0 is initial mass and m_f is final mass after propellant expulsion, highlighting the exponential mass ratio required for significant velocity changes and necessitating multi-stage designs to discard dead weight for missions like lunar transfers, which demand approximately 11-12 km/s total delta-v from low Earth orbit. Specific impulse, defined as I_{sp} = v_e / g_0 (with g_0 as standard gravity), quantifies propulsion efficiency; chemical rockets typically achieve 300-450 seconds, sufficient for high-thrust impulsive burns, while electric systems exceed 1,000-10,000 seconds for low-thrust, continuous acceleration suited to deep-space trajectories. Orbital mechanics relies on thrust for maneuvers such as Hohmann transfers, where two impulsive burns alter semi-major axis: the first increases velocity by \Delta v_1 = \sqrt{\mu / r_1} \left( \sqrt{2 r_2 / (r_1 + r_2)} - 1 \right) at perigee, and the second circularizes at apogee, minimizing use for efficient raising; total delta-v scales with orbital parameters but is independent of thrust magnitude, though higher thrust reduces burn duration and sensitivity to perturbations. Plane changes and , as in , require delta-v proportional to velocity and inclination difference, often performed via attitude thrusters for precise vector adjustments. For interplanetary missions, initial high-thrust escapes achieve hyperbolic excess velocity v_\infty, followed by low-thrust corrections to exploit gravity assists, as thrust-to-mass ratios below 10^{-5} enable spiral trajectories with superior efficiency over impulsive approximations. Limitations arise from the rocket equation's tyranny, constraining fractions to 1-5% for Earth-to-Mars transits without in-orbit refueling.

Marine and Terrestrial Engineering

In marine engineering, thrust is the forward propulsive force generated by devices such as screw propellers or waterjets to overcome hydrodynamic drag and advance the vessel. Propellers produce thrust by accelerating a mass of water rearward, with the magnitude determined by the propeller's rotational speed, blade geometry, advance speed, and fluid density; typical open-water propeller efficiency ranges from 0.5 to 0.7 for merchant ships. The required thrust exceeds the hull's total resistance R_T due to thrust deduction, where the effective thrust at the hull is T(1-t), and t (thrust deduction factor) accounts for energy losses in the propeller wake and hull interaction, typically 0.15–0.25 for single-screw vessels. Thrust calculations often employ blade element momentum theory, integrating local blade forces, or simplified momentum theory akin to actuator disk models, yielding T = \dot{m} (v_e - v_0), where \dot{m} is mass flow rate, v_e exit velocity, and v_0 inflow velocity. Ship propulsion systems convert shaft power P_S to thrust via propeller efficiency \eta, with thrust estimated as T = \frac{\eta P_S}{V}, where V is ship speed; for a given power, thrust decreases with speed due to increased drag. Waterjet systems, used in high-speed craft, generate thrust by pumping and accelerating water through nozzles, offering advantages in shallow drafts but lower efficiency (around 0.3–0.5) compared to propellers at low speeds. Historical advancements include the adoption of controllable-pitch propellers in the early 20th century, enabling variable thrust without speed changes, as seen in vessels like the RMS Queen Mary (1936), which used four blades producing up to 200,000 hp total thrust-equivalent power. Transverse thrust, a side force from propeller rotation (e.g., right-handed screws pushing stern to port in ahead motion), must be countered by rudders or bow thrusters in maneuvering. In terrestrial engineering, thrust manifests as tractive force at the vehicle-ground interface, generated by wheeled, tracked, or legged systems to propel against rolling resistance, grade, and inertial loads; unlike fluid-based marine thrust, it relies on friction or soil shear rather than expelled mass. For wheeled vehicles, maximum tractive force F_t = \mu N, where \mu is the tire-road friction coefficient (0.7–1.0 for dry asphalt, 0.3–0.6 for wet) and N is normal load, limits acceleration; engine torque \tau transmits as F_t = \frac{\tau}{r} \cdot i_g \cdot \eta_t, with r wheel radius, i_g gear ratio, and \eta_t transmission efficiency. Tracked vehicles, such as tanks, derive thrust from grousers shearing soil, modeled by Bekker's equations where gross traction b = c + \sigma \tan\phi (cohesion c, internal friction \phi, pressure \sigma), yielding peak thrust before slip; for example, the M1 Abrams tank (1980) achieves ~70 tons tractive pull via 28-inch tracks on soils up to 0.8 cohesion. Off-road performance emphasizes soil-thrust mechanics, with net traction NT = GT - MR, where gross thrust GT from shear and motion resistance MR from compaction; interference effects reduce effective thrust by 10–20% in soft soils due to adjacent track-soil interactions. Rail vehicles generate longitudinal thrust via wheel-rail adhesion, limited to \mu \approx 0.2–0.3, with high-power locomotives like the GE ES44AC (2002) delivering 4,300 hp to produce 120,000 lbf startup tractive effort through sandboxed sand for enhanced friction. Emerging electric and hybrid systems improve thrust efficiency by precise torque vectoring, reducing slip in autonomous ground vehicles.

References

  1. [1]
    What is Thrust? - Glenn Research Center - NASA
    Jul 21, 2022 · The thrust equation describes how the acceleration of the gas produces a force. The type of propulsion system used on an aircraft may vary from ...
  2. [2]
    General Thrust Equation
    Thrust equals the exit mass flow rate times the exit velocity minus the incoming mass flow rate times velocity plus the exit area times the static pressure ...
  3. [3]
    Rocket Engines – Introduction to Aerospace Flight Vehicles
    Therefore, while the rocket engine delivers higher thrust by expelling mass at a much higher velocity, it does so with significantly lower propulsive ...
  4. [4]
    Rocket Thrust Equation
    If the free stream pressure is given by p0, the thrust F equation becomes: F = m dot * Ve + (pe - p0) * Ae.
  5. [5]
    Thrust - Glenn Research Center - NASA
    Jan 21, 2023 · Thrust is the force that moves an aircraft through the air, generated by accelerating air in one direction, pushing the aircraft in the ...
  6. [6]
    Newton's Third Law of Motion
    May 7, 2021 · A jet engine also produces thrust through action and reaction. The engine produces hot exhaust gases which flow out the back of the engine ...
  7. [7]
    General Thrust Equation
    The thrust is then equal to the exit mass flow rate times the exit velocity minus the free stream mass flow rate times the free stream velocity. F = (m dot * V) ...Missing: physics | Show results with:physics
  8. [8]
    Chapter 3: Gravity & Mechanics - NASA Science
    Jan 16, 2025 · Like an airplane's jet engine, a rocket creates thrust by expelling mass to take advantage of Sir Isaac Newton's third law (see above). In both ...
  9. [9]
    12.3: Rocket Propulsion - Physics LibreTexts
    Jul 20, 2022 · Let F → e ⁢ ⁢ denote the total external force acting on the rocket. We shall use the momentum principle, to determine a differential equation ...
  10. [10]
    Introduction to Rocket Propulsion | Physics - Lumen Learning
    By Newton's third law, this force is equal in magnitude to the thrust force acting on the rocket, so Fthrust=veΔmΔt F thrust = v e Δ m Δ t , where all ...
  11. [11]
    Newton's Laws of Motion | Glenn Research Center - NASA
    Jun 27, 2024 · Whenever one object exerts a force on another object, the second object exerts an equal and opposite on the first. Sir Isaac Newton worked in ...Missing: classical | Show results with:classical
  12. [12]
    The SI units of thrust is same as that of A. Force/Area B ... - Vedantu
    Since N (Newton) is the S.I. unit of force, thrust has the unit same as force. Therefore, the correct option is C.
  13. [13]
    Force - The Engineering ToolBox
    The unit of force in the Imperial or British system is the pound - lb, lbf . 1 lbf = 4.45 N; A pound is the unbalanced force which will give a 1 slug mass an ...<|separator|>
  14. [14]
    How Do You Measure the Thrust of a Rocket Engine? | NIST
    Feb 24, 2025 · Rocket thrust is measured using load cells attached to a stand. The engine's force compresses the cells, which use strain gauges to measure the ...Missing: characteristics | Show results with:characteristics<|control11|><|separator|>
  15. [15]
    Recommended Practices in Thrust Measurements
    Oct 6, 2013 · This paper summarizes recommended practices for the design, calibration, and operation of pendulum thrust stands.
  16. [16]
    How is engine thrust measured in flight? - Aviation Stack Exchange
    Oct 15, 2014 · Engine thrust is measured in flight by EPR - Engine Pressure Ratio. EPR is the ratio of the turbine exhaust pressure divided by the pressure measured at the ...Why do we refer to "power" for turboprop engines and "thrust" for ...How to measure thrust produced from combustion chamber of Gas ...More results from aviation.stackexchange.com
  17. [17]
    [PDF] Precision flow measurement techniques for low-thrust auxiliary ...
    Impeller wheel meters for steady flow and a photometric device for pulsed flow are used to measure propellant flow, achieving better than 1% accuracy.
  18. [18]
    Recommended Practice for Thrust Measurement in Electric ... - NIH
    This paper summarizes recommended practices for the design, calibration, and operation of pendulum thrust stands, which are widely recognized as the best ...
  19. [19]
    Brief History of Rockets - NASA Glenn Research Center
    About three hundred years after the pigeon, another Greek, Hero of Alexandria, invented a similar rocket-like device called an aeolipile. It, too, used steam as ...
  20. [20]
    The ancient invention of the steam engine by the Hero of Alexandria
    Mar 21, 2014 · Heron was the first inventor of the steam engine, a steam powered device that was called aeolipile or the 'Heron engine'.
  21. [21]
    The History of Early Fireworks and Fire Arrows - ThoughtCo
    Feb 17, 2019 · The first use of true rockets as weapons is reported as occurring in 1232. The Chinese and the Mongols were at war with each other, and the ...
  22. [22]
    Rocket History -
    Rockets were first used as actual weapons in the battle of Kai-fung-fu in 1232 A.D. The Chinese attempted to repel Mongol invaders with barrages of fire arrows ...
  23. [23]
    95 Years Ago: Goddard's First Liquid-Fueled Rocket - NASA
    Mar 17, 2021 · Believing that liquid propellants offered the most promise, he successfully launched the first liquid-fueled rocket on March 16, 1926, on a farm in Auburn, ...
  24. [24]
    Engines - NASA Glenn Research Center
    It was Frank Whittle, a British pilot, who designed and patented the first turbo jet engine in 1930. The Whittle engine first flew successfully in May, 1941. ...Missing: key milestones
  25. [25]
    27 August 1939 | This Day in Aviation
    Aug 27, 2025 · This was the very first flight of an aircraft powered only by a jet engine. (Left to right) Erich Karl Warsitz, Ernst Heinrich Heinkel, and Hans ...
  26. [26]
    Germany conducts first successful V-2 rocket test | October 3, 1942
    On October 3, 1942, German rocket scientist Wernher von Braun's brainchild, the V-2 missile, is fired successfully from Peenemunde, as island off Germany's ...
  27. [27]
    The Jet Age | National Air and Space Museum
    The jet engine revolutionized air travel. Powerful and durable, jets enabled aircraft manufacturers to build bigger, faster, and more productive airliners.
  28. [28]
    Is SpaceX's Raptor engine the king of rocket engines?
    May 25, 2019 · The Merlin produces .84 MNs of thrust, the RS-25 produces 1.86 MNs, the Raptor currently is at 2 MNs, the BE-4 is hoping to hit 2.4 ...
  29. [29]
    [PDF] Recent Progress on the VASIMR® Engine IEPC-2022 - 525
    Jun 23, 2022 · The VX-200SS™ advances the. VASIMR® technology readiness level (TRL) to level 5 by integrating a new 120 kW, vacuum compatible, 2nd stage power ...<|separator|>
  30. [30]
    Metal Additive Manufacturing Developments for Propulsion ...
    NASA has been involved in the development and maturation of metal additive manufacturing (AM) for space applications since the 2000's.
  31. [31]
    [PDF] Airbreathing Engine Performance Parameters
    Thermal Efficiency. – for thrust produced using a nozzle. • this is just the cycle efficiency for a cycle that outputs kinetic energy (nozzle) instead of.
  32. [32]
    Fundamentals of Propulsion Systems – Introduction to Aerospace ...
    The principle of thrust generation for a rocket engine is the reaction force associated with accelerating a mass of gas at high velocity, the gas being a ...
  33. [33]
    [PDF] Propulsion Over a Wide Mach Number Range
    Air breathing propulsion systems for hypersonic flight present the engine designer with circumstances that differ in important fundamental ways from those ...
  34. [34]
    3 Air-Breathing Propulsion | A Review of United States Air Force and ...
    The security environment requires the United States to field propulsion systems that cover the entire Mach number range, up to Mach 16.
  35. [35]
    [PDF] An Overview of the NASA FAP Hypersonics Project Airbreathing ...
    Primarily, the RALV mission focuses on Two-Stage-To-Orbit (TSTO) systems utilizing airbreathing combined-cycle-engine propulsion; whereas, the PAES mission ...
  36. [36]
    [PDF] History of Ramjet and Scramjet Propulsion Development for U.S. ...
    history of high-speed airbreathing propulsion ramjet engines and their respective vehicle and weapon systems developed under the support of the U.S. Navy.
  37. [37]
    X-43A: How NASA Hit Mach 9.68 With an Air-Breathing Engine
    Oct 4, 2025 · Scramjets, or supersonic combustion ramjets, offer a compelling alternative by using atmospheric oxygen for combustion, thereby reducing the ...
  38. [38]
    Rocket Propulsion
    Different propulsion systems develop thrust in different ways, but all thrust is generated through some application of Newton's third law of motion. For every ...
  39. [39]
    Specific Impulse
    Thrust is the force which moves a rocket through the air. Thrust is generated by the rocket engine through the reaction of accelerating a mass of gas.
  40. [40]
    [PDF] Rocket Propulsion Fundamentals
    3rd Law: For every action, there is an equal and opposite reaction. In rocket propulsion, a mass of propellant (m) is accelerated (via the combustion process) ...
  41. [41]
    Ideal Rocket Equation | Glenn Research Center - NASA
    Nov 21, 2023 · Ideal Rocket Equation · change in rocket momentum=M(u+du)−Mu=Mdu · change in exhaust momentum=dm(u−v)−udm=−vdm · change in system momentum=Mdu−vdm.
  42. [42]
    [PDF] Fundamentals of Electric Propulsion: Ion and Hall Thrusters
    ... Principles and Scaling.................................329. 7.2.1 Crossed ... Newton's second law, the Hall current force on the magnets is equal and.Missing: mechanics | Show results with:mechanics
  43. [43]
    What is Electric propulsion? - European Space Agency
    The propellant is ejected up to twenty times faster than from a classical chemical thruster and therefore the overall system is many times more mass efficient.
  44. [44]
    Electric Propulsion Basics — proptools 0.0.0 documentation
    Like all rockets, electric thrusters generate thrust by expelling mass at high velocity. In electric (electrostatic) propulsion, a high-velocity jet is produced ...
  45. [45]
  46. [46]
    [PDF] Hall Thrusters - DESCANSO
    Hall thrusters are relatively simple devices consisting of a cylindrical channel with an interior anode, a magnetic circuit that generates a primarily radial.
  47. [47]
    Hall Thrusters — Busek
    Hall thrusters generate thrust by creating and accelerating ionized gas via an electric field. Confined within a finely shaped magnetic field.Missing: generation | Show results with:generation
  48. [48]
    NASA Next-Generation Solar Sail Boom Technology Ready for ...
    Apr 10, 2024 · Solar sails use the pressure of sunlight for propulsion, angling toward or away from the Sun so that photons bounce off the reflective sail to ...
  49. [49]
    [PDF] THE DEVELOPMENT OF SOLAR SAIL PROPULSION FOR NASA ...
    Solar sails are a near term, low thrust, propellantless propulsion technology suitable for orbital maneuvering, station keeping, and attitude control ...Missing: methods | Show results with:methods
  50. [50]
    Space sails for achieving major space exploration goals
    Oct 1, 2024 · Solar sails are a mature technology for propellant-free propulsion. A lightweight reflective membrane provides a large area for receiving ...
  51. [51]
    Rocket Thrust Equations
    Thrust is produced according to Newton's third law of motion. The amount of thrust produced by the rocket depends on the mass flow rate through the engine ...
  52. [52]
    Propeller Thrust | Glenn Research Center - NASA
    Jul 14, 2025 · We can use Bernoulli's equation to relate the pressure and velocity ahead of and behind the propeller disk, but not through the disk. Ahead of ...Missing: derivation | Show results with:derivation
  53. [53]
    Beginner's Guide to Propulsion
    On airplanes, thrust is usually generated through some application of Newton's third law of action and reaction. A gas, or working fluid, is accelerated by ...
  54. [54]
    7.4.1.2. Propulsive and Overall Efficiency
    The power produced by engine thrust is the product of thrust and flight velocity (V). The ratio of this useful power to the increment in kinetic energy ...
  55. [55]
    11.2 Thermal and Propulsive Efficiency - MIT
    It is often convenient to break the overall efficiency into two parts: thermal efficiency and propulsive efficiency.
  56. [56]
    Propulsive Efficiency - an overview | ScienceDirect Topics
    Propulsive efficiency is defined as the ratio of useful thrust power to the total power imparted to the airstream, indicating how effectively the power ...
  57. [57]
    Specific Thrust
    When dealing with a gas, the general thrust equation is given as: F = mdot e * Ve - mdot 0 * V0 + (pe - p0) * Ae. Thrust F is equal to the exit mass flow ...Missing: derivation | Show results with:derivation
  58. [58]
    [PDF] Thruster Principles - DESCANSO
    Electric thrusters propel the spacecraft using the same basic principle as chemical rockets—accelerating mass and ejecting it from the vehicle. The.
  59. [59]
    Specific Impulse - an overview | ScienceDirect Topics
    The specific impulse is a measure of the thrust that is produced per unit fuel expended per unit time and is a measure of the fuel efficiency of the system.
  60. [60]
    4.0 In-Space Propulsion - NASA
    These propellants have the advantages of being ambient-temperature liquids (thus not requiring extensive thermal management) and easily ignited (catalyst for ...
  61. [61]
    [PDF] KSC -'0"' ENG.ME RING
    This equation will give a time-averaged specific impulse value for any rocket propulsion system, particularly ... The thrust-to-weight ratio expresses the ...
  62. [62]
    [PDF] Dr. Donald Edberg - AIAA
    A thrust-to-weight ratio (TWR) of 1.5 was chosen for the launch vehicle and based on preliminary 1st stage mass estimations of 2200 kg (4850 lbm) and 2nd stage ...
  63. [63]
    Low-Thrust Mission Design and Application
    Jul 28, 2010 · Electrothermal thrusters are the most widely used electric propulsion systems to date, but electrostatic systems are the industry's state-of-the ...
  64. [64]
    [PDF] Aircraft Engines - Federal Aviation Administration
    Many aircraft use a form of the gas turbine engine to produce power for thrust. These engines are normally the turboprop, turboshaft, turbofan, and a few ...
  65. [65]
    The Criticality of Thrust Measurement Testing in Aerospace - Interface
    Oct 26, 2023 · The force emitted by a thrust engine dictates the size and speed at which a payload can be lifted off the ground. Enormous amounts of thrust are ...
  66. [66]
    [PDF] 2023 Small Satellite Propulsion Technologies Compendium v1.41
    Any domestic or foreign propulsion system that fell within the following targeted propulsion technologies was identified and reviewed as a credible option.
  67. [67]
    What is Thrust?
    Thrust is the force which moves an aircraft through the air. Thrust is generated by the engines of the airplane. What is thrust? Thrust is a mechanical ...
  68. [68]
    Propeller Thrust
    Propeller thrust is calculated as exit mass flow rate times exit velocity minus free stream velocity, or F = .5 * r * A * [Ve ^2 - V0 ^2].
  69. [69]
    Turbofan Engine - NASA Glenn Research Center
    A turbofan engine is a modern gas turbine with a core, fan, and rear turbine. It uses air that bypasses the core to generate thrust.<|separator|>
  70. [70]
    Turbofan Thrust
    Thrust equals the sum of the exit mass flow rate times exit velocity minus free stream mass flow rate times velocity for both streams.Missing: derivation | Show results with:derivation
  71. [71]
    Beginner's Guide to Propulsion/RangeGames: Thrust to Weight Ratio
    Compare the thrust to weight ratios in airliners, fighter jets, and executive jets. thrust to weight ratio is generally highest in fighters, next highest in ...
  72. [72]
    Turboprop Thrust
    Propellers develop thrust by moving a large mass of air through a small change in velocity. Propellers are very efficient and can use nearly any kind of engine ...Missing: generation | Show results with:generation
  73. [73]
    Why does a rocket engine provide more thrust in a vacuum than in ...
    Oct 18, 2013 · Rocket thrust is given by the equation. F=˙mvexit+Ae(P1−P2). where ˙m is the mass flow rate, vexit is the average exit flow velocity across ...From the General Thrust Equation towards Tsiolkovsky, how to ...Is specific impulse and thrust maximum in vacuum or for exit ...More results from space.stackexchange.com
  74. [74]
    What is Electric Propulsion?
    Specific Impulse: A measure of the effectiveness of a rocket engine relative to the amount of propellant (fuel) it consumes. In some ways, specific impulse can ...<|separator|>
  75. [75]
    Basics of Space Flight: Orbital Mechanics
    An overview of orbital mechanics including types of orbits, mathematical formulae, and example problems.
  76. [76]
    Chapter 7 – Manuevering – Introduction to Orbital Mechanics
    Mar 12, 2023 · Notice that the Delta V is directly proportional to the velocity of the satellite in the orbit, so it is prudent to perform plane change ...
  77. [77]
    Impulsive Maneuvers - Orbital Mechanics & Astrodynamics
    The total Δ v is the sum of the absolute value of the change in velocity required for each impulse event. This means that speeding up or slowing down or ...
  78. [78]
    Breaking the Rocket Equation: How Refuellable Spacecraft Change ...
    One of the chief tenets of space transportation has been the immutability of the Tsiolkovsky rocket equation, placing great emphasis on the specific impulse ...
  79. [79]
    Propeller Thrust - an overview | ScienceDirect Topics
    The thrust generated by the propeller is the consequence of a complex interaction between the forward motion of the propeller, its rotational speed, and ...
  80. [80]
    [PDF] 13.012 Hydrodynamics for Ocean Engineers Marine Propellers - MIT
    Nov 30, 2004 · The thrust deduction coefficient is defined as: 1. /t t. R T. = − where Rt is the total ship resistance and T is the propeller thrust. Note that ...
  81. [81]
    Thrust generated by a ship - Engineering Stack Exchange
    Apr 24, 2017 · Then you can calculate the thrust FT for given power at the shaft PS as FT=ηFrPSu.
  82. [82]
    [PDF] Basic principles of ship propulsion
    This paper will explain the most elementary terms used regarding ship types, di- mensions and hull forms, and clarify some of the parameters pertaining to hull.
  83. [83]
    What is Transverse Thrust in Ships? - Marine Insight
    Mar 16, 2024 · Thrust can be defined as the propulsive force that drives the vessel through the water against the resistive forces, mostly hydrodynamic.
  84. [84]
    [PDF] engine torque / traction force - ASSOCIATION ADILCA
    The traction force means the force acting on the tyres of the drive wheels by contact with the ground to create or maintain the movement of the car(**). © ...
  85. [85]
    [PDF] Evaluating Vehicle Mobility Using Bekker's Equations - DTIC
    When the vehicle develops a thrust, it is the shearing of soil on soil that develops the shearing tractive force. From Eq. 2 it can be seen that this generation ...
  86. [86]
    Chapter 2. Traction Mechanics. Part IV. Track Systems
    The term (GT -- MR int ) is the thrust or tractive force F produced at the track-soil interface. Thus, the net traction, NT, can be expressed as (fig. 2.90):.
  87. [87]
    Assessment of the side thrust for off-road tracked vehicles based on ...
    9 and Fig. 10. The traction force of a tracked unmanned ground vehicle (UGV) depends on the soil thrust generated by the shearing action on the soil-track ...