Fact-checked by Grok 2 weeks ago

Propeller theory

Propeller theory encompasses the aerodynamic and hydrodynamic principles that govern the design, analysis, and performance of propellers, which convert rotational from an into linear for in , ships, and other vehicles. These devices typically consist of two or more twisted blades radiating from a central , functioning as rotating airfoils or hydrofoils that generate a differential—lower pressure on the forward-facing side and higher on the aft side—to accelerate fluid (air or water) rearward, thereby producing forward according to Newton's third law. The theory addresses efficiency, which can reach up to 80% in optimized designs, influenced by factors such as blade geometry, rotational speed, and fluid velocity. Fundamental to propeller theory are two primary analytical approaches: momentum theory and . Momentum theory models the as an idealized actuator disk that imparts energy to the fluid stream, deriving as T = \dot{m} (V_e - V_0), where \dot{m} is the , V_e is the exit , and V_0 is the free-stream ; this simplifies the to a pressure jump across the disk, with at the disk given by V_p = \frac{1}{2} (V_e + V_0). , in contrast, divides the into discrete radial segments, each treated as an independent , to calculate local and forces based on , which varies from to due to increasing tangential ; this enables detailed of twist and for optimal across operating conditions. Combined, these theories inform performance metrics like the J = \frac{V_\infty}{n d} (where V_\infty is forward speed, n is rotational speed, and d is diameter) and \eta = \frac{T V_\infty}{P}, where P is input . In aeronautical applications, propeller theory emphasizes minimizing effects by limiting numbers below 0.85 to avoid losses and noise generation, with level as M_{\text{tip}}^{5.5}. types include fixed-pitch propellers, efficient at a single flight condition, and variable-pitch or constant-speed designs that adjust for broad speed ranges, often using composites for lightweight durability. For propellers, the theory incorporates hydrodynamic effects like —where local pressure drops below , quantified by the cavitation number \sigma_v = \frac{2(P - P_{\text{vap}})}{\rho V_a^2}—and uses coefficients such as K_T = \frac{T}{\rho N^2 D^4} and K_Q = \frac{Q}{\rho N^2 D^5} to optimize open-water \eta_0 = \frac{T V_a}{2\pi N Q}. These principles, refined since the early , continue to underpin advancements in for diverse contexts.

Historical Development

Early Innovations

The concept of the propeller traces its origins to ancient devices, particularly the , invented around the as a mechanism for raising water through a rotating helical surface immersed in fluid. This empirical invention, used primarily for in and , demonstrated the principle of generating axial via rotational motion in a medium, serving as a foundational precursor to later systems. In the late 18th century, practical applications for emerged with Joseph Bramah's patent in 1785 for a screw propeller, which proposed a cylindrical rod with a continuous helical thread to drive ships forward by rotating beneath the hull. Although Bramah's design was not immediately implemented on a large scale, it represented an early conceptual shift toward screw-based alternatives to oars and sails. By the 1830s, inventors like advanced these ideas through rigorous testing; Ericsson patented a screw propeller in 1836 and successfully propelled the Francis B. Ogden to speeds of 10 miles per hour in 1837 trials on the Thames River. The breakthrough for widespread adoption came with the SS Archimedes, launched in 1838 and commissioned in 1839 as the first full-sized steamship powered exclusively by a screw propeller, achieving speeds of up to 10 knots and outperforming paddle-wheel vessels in comparative sea trials against the Vulcan in 1839. These tests, including later evaluations on HMS Rattler in the 1840s, confirmed the screw's efficiency advantages, such as reduced vulnerability to damage and better performance under sail, prompting navies to integrate propellers into warships like the USS Princeton in 1843. Ericsson's designs for the Princeton, featuring a submerged propeller, further validated these benefits through operational success in U.S. Navy service. Early experiments extended propeller principles to aviation in the mid-19th century, with Gustave de Ponton d'Amécourt constructing a steam-powered model helicopter in 1863 equipped with to generate lift, though it failed to achieve sustained flight despite successful spring-powered variants. Building on such efforts, Alphonse Pénaud pioneered powered in the 1870s, including a 1870 rubber-band-driven with superimposed that demonstrated stable hovering. His most notable innovation, the Planophore of 1871, featured a tail-mounted two-bladed powered by a twisted , enabling a stable 11-second flight covering 40 meters and highlighting inherent aerodynamic through dihedral wings and a stabilizing tail. These model-based trials marked the initial empirical application of propellers to aerial propulsion.

Theoretical Foundations

The theoretical foundations of propeller theory emerged in the late through pioneering experimental and analytical work that shifted design from empirical trial-and-error to systematic scientific principles. William Froude's 1878 investigations into marine efficiency laid crucial groundwork by introducing the blade element approach, which analyzed performance by considering forces on individual sections of the blades independently of the . His studies also pioneered the concept of wake fraction, quantifying the reduction in water velocity behind a ship's due to viscous effects, and explored - interactions that affect overall propulsive efficiency. These contributions enabled more accurate predictions of thrust in real operating conditions, bridging with . In , early aerodynamic studies in the 1890s by advanced propeller theory through rigorous experimentation on and forces. At the , Langley conducted whirling arm tests on model and , measuring aerodynamic coefficients to inform designs for powered flight vehicles like his series. Building on this, the in 1903 developed highly efficient for their Flyer aircraft, applying data from their self-constructed for testing to optimize blade shapes and achieve of approximately 1:1 for propeller sections, which maximized while minimizing induced . Their iterative testing emphasized the propeller as a rotating system, yielding efficiencies around 66-70% in flight. A seminal theoretical framework was provided by Stefan Drzewiecki in his early 1900s texts on marine propellers, where he formalized specifically for screw propellers. Drzewiecki derived key relationships between blade angle, rotational speed, and generated by integrating local aerodynamic forces along the blade span, assuming each element acts independently like a two-dimensional in helical flow. This approach allowed for the calculation of thrust distribution and without relying on global balances, facilitating practical design optimizations for varying and . Key milestones in simplifying propeller analysis include W. J. M. Rankine's 1865 development of actuator disk theory, which modeled the as an idealized disk imparting to the fluid to produce . Originally applied to marine propellers, this theory was revisited in the for contexts, establishing limits on ideal efficiency based on axial flow acceleration. These tools synthesized early theoretical insights, enabling engineers to select propeller configurations that balanced , power, and speed.

Basic Principles

Mechanism of Operation

A functions as a rotating that accelerates surrounding —typically air for aeronautical applications or water for marine ones—rearward, thereby generating forward thrust through Newton's third law of motion, where the reaction force propels the vehicle ahead. This process converts the rotational energy from the driving engine into linear imparted to the , with the blades acting as twisted airfoils that efficiently handle the . The blades trace a helical path through the fluid as the propeller rotates, creating a screw-like advancement that mimics the action of an Archimedes screw, where each revolution advances the vehicle a distance related to the blade pitch. A key dimensionless parameter governing this operation is the advance ratio, defined as J = \frac{V}{nD}, where V is the forward speed, n is the rotational speed in revolutions per unit time, and D is the propeller diameter; this ratio characterizes the relative influence of translational versus rotational motion on performance. Propellers are classified into fixed-pitch and variable-pitch types, with the former having a constant blade optimized for a specific operating condition, such as , where diminishes at off-design speeds due to changes in the effective of incidence. In contrast, variable-pitch propellers allow in-flight adjustment of the blade , enabling higher across a range of speeds by maintaining an optimal for takeoff (fine ) or efficient (coarse ), often via hydraulic or mechanical governors. The accelerated fluid forms a —a conical wake of higher-velocity flow trailing from the disk—which enhances overall by entraining additional ambient fluid and acting as a momentum exchanger between the and the surrounding medium. This visualization underscores the 's role in continuously imparting rearward to sustain forward progress.

Aerodynamic Forces on Blades

Propeller blades operate as rotating airfoils, interacting with the surrounding to produce aerodynamic forces that drive generation. The angle of attack (\alpha) on a section varies along the radial direction due to the rotational motion, where the tangential component increases linearly with , resulting in higher relative speeds at the compared to the . This variation necessitates careful geometry to maintain efficient production across the . The fundamental aerodynamic forces acting on these blade sections are and , analogous to those on fixed wings but influenced by the helical inflow path. Lift L on a differential blade element is expressed as L = \frac{1}{2} \rho V^2 C_l S, where \rho is the air , V is the local combining axial advance and rotational components, C_l is the section dependent on \alpha, and S is the elemental planform area. D follows similarly as D = \frac{1}{2} \rho V^2 C_d S, with C_d the , which increases nonlinearly at higher angles of attack and contributes to both power absorption and efficiency losses. These coefficients are derived from two-dimensional data, adjusted for three-dimensional effects in propeller operation. Circulation theory provides the underlying mechanism for lift generation on propeller blades, treating each section as a lifting line with bound . The Kutta-Joukowski applies directly, stating that the lift per unit span is \rho V \Gamma, where \Gamma is the circulation around the , induced by the effective angle of attack and trailing vortex sheet. This bound creates the pressure differential essential for , while the theorem's assumptions hold for attached flow conditions typical in efficient designs. Rotation imposes additional inertial forces beyond , notably , which acts radially outward proportional to the square of the rotational speed and radius, generating tensile stresses that dominate blade loading and limit material choices. Coriolis forces emerge from the cross-coupling of radial or axial blade velocities with the rotational velocity vector, producing transverse accelerations that alter local stress distributions and contribute to vibratory responses under off-design conditions. These effects are particularly pronounced in high-speed propellers, influencing both structural integrity and aeroelastic stability. Blade design incorporates and to optimize \alpha distribution and mitigate risks at the high-velocity . progressively reduces the pitch angle from to , aligning the line with the local inflow to sustain a near-constant \alpha for maximum C_l/C_d. , the curvature of the mean line, enhances lift at low \alpha by shifting the zero-lift angle, allowing finer control over sectional performance without excessive penalties. These features ensure uniform loading and across the radius.

Theoretical Modeling

Momentum Theory Basics

Momentum theory, also known as disk theory, models a as an idealized, infinitesimally thin disk that imparts axial momentum to the surrounding fluid without considering the detailed geometry of the blades. This approach, originally developed for marine s, treats the as a permeable disk that creates a discontinuity, accelerating the flow uniformly across its area A. The theory assumes inviscid, with no rotation imparted to the wake, and uniform velocity profiles upstream and downstream of the disk. It was first introduced by William John Macquorn Rankine in 1865 and refined by Robert Edmund Froude in 1889, providing a foundational framework for understanding under ideal conditions. In the actuator disk model, the propeller accelerates the from the V to a far-wake of V + 2w, where w is the induced at the disk plane. The at the disk is the , V + w, ensuring a smooth momentum balance. The through the disk is \dot{m} = \rho A (V + w), with \rho denoting . Applying the axial theorem to a enclosing the disk yields the T as the rate of change in axial : T = \dot{m} \left[ (V + 2w) - V \right] = 2 \dot{m} w = 2 \rho A (V + w) w. This equation relates thrust directly to the induced velocity and flow parameters, highlighting how the propeller extracts momentum from the fluid to generate propulsion. The power required to drive the propeller is the kinetic energy imparted to the fluid per unit time, given by P = T (V + w), as the work is performed at the disk velocity. Substituting the thrust expression, the power becomes P = 2 \rho A (V + w)^2 w. This ideal power input assumes no losses, providing a baseline for evaluating real propeller efficiency. The ideal propulsive efficiency \eta is the ratio of useful thrust power T V to input power P, yielding \eta = \frac{V}{V + w}. Expressing this in terms of thrust and flow conditions involves solving the quadratic relation from the thrust equation, resulting in: \eta = \frac{2}{1 + \sqrt{1 + \frac{T}{\frac{1}{2} \rho A V^2}}}. Here, \frac{1}{2} \rho A V^2 represents the times disk area. As velocity V increases relative to thrust loading, w becomes small, and \eta approaches 1, indicating near-perfect efficiency for high-speed, lightly loaded propellers. Conversely, at static conditions (V = 0), efficiency drops to zero. While powerful for initial design insights, momentum theory has limitations: it assumes uniform inflow and no swirl in the wake, neglecting components and losses, and is most accurate for lightly loaded propellers where induced velocities are small compared to V. These simplifications make it unsuitable for highly loaded or off-design conditions without corrections.

Blade Element Theory

Blade Element Theory (BET) is a foundational method in propeller aerodynamics that models the performance of a by treating each blade as a collection of independent two-dimensional sections. Developed primarily by the Stefan Drzewiecki between 1892 and 1920, the theory provides a for calculating and by analyzing local aerodynamic forces on discrete blade segments, enabling the incorporation of detailed characteristics without relying on global flow assumptions. This approach contrasts with simpler momentum-based models by resolving variations along the blade span, making it suitable for design optimization. In BET, the propeller blade is divided into numerous thin radial elements of infinitesimal width dr at a distance r from the axis of rotation. For each element, the local flow velocity V_\text{local} is the vector sum of the axial advance velocity V (forward speed of the vehicle) and the tangential velocity due to rotation \omega r (where \omega is the angular velocity), yielding a resultant magnitude of V_\text{local} = \sqrt{V^2 + (\omega r)^2}. The inflow angle \phi at this radius is then \phi = \tan^{-1} \left( \frac{V}{\omega r} \right), representing the angle between the local velocity vector and the plane of rotation. These parameters allow the element to be analyzed as a stationary airfoil in a relative wind, using standard two-dimensional lift and drag coefficients C_l and C_d, which depend on the local angle of attack \alpha and Reynolds number Re. The aerodynamic forces on each blade element are computed using airfoil data. The differential thrust dT contributed by all B blades at radius r is given by dT = \frac{1}{2} \rho V_\text{local}^2 (C_l \cos \phi - C_d \sin \phi) B c \, dr, where \rho is air density and c is the local chord length. Similarly, the differential torque dQ is dQ = \frac{1}{2} \rho V_\text{local}^2 (C_l \sin \phi + C_d \cos \phi) B c r \, dr. Here, the lift component C_l contributes positively to both thrust (via its axial projection) and torque (via its tangential projection), while drag C_d subtracts from thrust but adds to torque due to its alignment with the flow. To achieve optimal performance, the blade twist is designed such that the local pitch angle \beta(r) = \phi + \alpha, where \alpha is selected for maximum lift-to-drag ratio at the local Re. The total thrust T and torque Q are obtained by integrating these differentials from the hub radius to the tip: T = \int dT, \quad Q = \int dQ. This integration accounts for radial variations in geometry and flow, providing a complete performance prediction. One key advantage of BET is its reliance on empirical two-dimensional airfoil tables for C_l and C_d as functions of \alpha and Re, allowing designers to incorporate real viscous effects and sectional shapes without full three-dimensional computations. It naturally handles blade twist, variable chord distribution, and non-uniform loading along the span, facilitating iterative design for specific operating conditions like advance ratio and rotational speed. While early formulations neglected three-dimensional effects like induced velocities, BET's modular structure permits extensions, such as coupling with momentum theory for improved accuracy in modern applications.

Performance Equations

Thrust Generation

The thrust generated by a propeller is fundamentally described by the equation T = \rho n^2 D^4 K_T(J), where \rho is the fluid density, n is the rotational speed in revolutions per second, D is the , J = V / (n D) is the with V as the forward velocity, and K_T(J) is the dimensionless thrust coefficient determined from experimental charts or computed via (BEMT). This formulation combines momentum theory, which treats the propeller as an disk imparting axial momentum to the flow, with , which resolves forces on discrete blade sections to yield the coefficient K_T. BEMT equates the elemental from blade , dT = \frac{1}{2} B \rho U^2 c (C_L \cos \phi - C_D \sin \phi) dr, to the momentum change, dT = 4 \pi \rho r V^2 (1 + a) a F dr, where B is the number of blades, U is the local , c is the chord length, C_L and C_D are and coefficients, \phi is the inflow angle, a is the axial induction factor solved iteratively, and F is the Prandtl tip-loss correction factor. For a blade, the thrust contribution arises from integrating the axial component of aerodynamic forces along the radial span, with local \sigma(r) = B c(r) / (2 \pi r) quantifying the blade area fraction relative to the annular disk area, influencing loading distribution. In a simplified assuming small of attack \alpha and neglecting , the single-blade thrust approximates T_{\text{single}} \approx \int_{r_{\text{hub}}}^{R} \frac{1}{2} \rho (\omega r)^2 C_{l_\alpha} \alpha c \, dr, where \omega = 2 \pi n is the , C_{l_\alpha} is the sectional lift curve slope, and the integration accounts for varying \alpha = \beta(r) - \phi(r) with blade twist \beta(r) and local inflow \phi(r) = \tan^{-1}(V / (\omega r)). This elemental lift-based approach, derived from , scales the force with local (\omega r)^2 and emphasizes outer radial contributions due to higher tangential speeds. Scaling to multi-blade propellers involves multiplying the -blade thrust by B, yielding total T = B \cdot T_{\text{[single](/page/Single)}}, but interference effects from adjacent blades modify the induced flow field, particularly through the axial induction factor a in BEMT, which decreases with increasing B as the propeller approximates an ideal disk. Higher blade counts enhance capacity by distributing load and reducing tip losses via the Prandtl correction factor F \approx \frac{2}{\pi} \cos^{-1} \left( e^{-B (1 - r/R)/(2 \sin \phi)} \right), though excessive can increase profile drag. For instance, typical aircraft propellers use 2–4 blades to balance these effects, achieving values of 0.05–0.15. Variations in thrust equations distinguish operational contexts: in marine applications, open-water thrust T = \rho n^2 D^4 K_T(J) is evaluated in unbounded flow to characterize efficiency across advance ratios, often using standardized curves from model tests. For , static thrust at zero advance (J = 0) simplifies to T_{\text{static}} \approx \rho n^2 D^4 C_{T0}, where C_{T0} is the zero-speed , reflecting induced velocities that double the slipstream momentum change from rest. This static case, crucial for takeoff performance, yields higher coefficients (e.g., C_{T0} \approx 0.1–0.2) compared to advancing flight due to the absence of forward inflow.

Torque and Power

In propeller theory, torque represents the rotational force necessary to drive the blades through the , while quantifies the energy input required for . The Q is expressed dimensionlessly through the torque coefficient K_Q, which depends on the J = \frac{V}{nD}, where V is the advance speed, n is the rotational speed in revolutions per second, and D is the . The standard for is Q = \rho n^2 D^5 K_Q(J), with \rho denoting ; this formulation arises from and empirical data for both marine and aeronautical propellers. P is derived directly from as P = 2\pi n Q, linking transfer to the propeller's mechanical input. The relationship between , , and T is mediated by \eta, which measures the conversion of shaft to useful thrust . Efficiency is given by \eta = \frac{J}{2\pi} \frac{K_T}{K_Q}, where K_T = \frac{T}{\rho n^2 D^4} is the ; this highlights how influences overall performance by balancing rotational losses against forward . In ideal conditions from theory, the minimum for a in hover (zero advance) is P_{\text{ideal}} = T \sqrt{\frac{T}{2 \rho A}}, where A = \frac{\pi D^2}{4} is the disk area; this induced arises from the imparted to the and sets a theoretical lower bound for real systems. At the blade level, arises from the of azimuthal components along the . The differential dQ from an elemental section at r is dQ = r (dL \sin \phi + dD \cos \phi), where dL and dD are the and increments, and \phi is the ; total is obtained by summing these over the . This formulation underscores the role of dD, which contributes to power absorption even in high-efficiency designs, as it generates a tangential opposing . In practical applications, torque characteristics differ between variable-speed and constant-speed propellers. Variable-speed systems allow rotational speed to vary with load, resulting in that scales nonlinearly with output to match varying flight or sea conditions. In contrast, constant-speed propellers maintain fixed n by adjusting , which modulates absorption to optimize across speed ranges. For marine propellers, imposes critical limits, as excessive loading leads to vapor bubble formation and breakdown; design guidelines specify maximum K_Q values to avoid these effects, often capping at levels that prevent blade face erosion.

Practical Considerations

Efficiency and Losses

The efficiency of a propeller, denoted as η, is defined as the ratio of useful to input : η = (T V) / P, where T is the , V is the freestream , and P is the required to drive the . In momentum theory, assuming and uniform loading across the disk, the maximum is given by η_max = 1 / (1 + w/V), where w represents the induced at the disk. Real-world propeller performance deviates from this ideal due to various losses that reduce . Profile arises from viscous effects on the surfaces, contributing to efficiency losses through increased coefficients that elevate requirements without proportional gains. Tip vortices, formed at tips due to pressure differences, induce three-dimensional flow effects that effectively reduce the loaded disk area and generate additional , typically accounting for about 5% efficiency reduction in multi- designs. Rotational slip, characterized by non-axial wake rotation where the imparts tangential to the , leads to dissipation in swirl components, adding roughly 7% loss as this rotational does not contribute to forward . To mitigate these effects in theoretical models, corrections such as Prandtl's tip loss factor are applied, which adjusts the effective blade loading near the : F = \frac{2}{\pi} \cos^{-1} \left( e^{-f} \right), where f = \frac{B}{2} \frac{1 - r/R}{\sin \phi}, with B as the number of blades, r/R the normalized radial position, and \phi the local flow angle. Scale effects further influence losses via the (), which impacts the (C_L / C_D) of blade sections; lower , common in small-scale propellers, increases relative viscous and can reduce overall by 7-15% compared to high- conditions. Contemporary (CFD) analyses of optimized , incorporating viscous and three-dimensional effects, validate peak ranging from 85% to 95% under conditions, surpassing earlier empirical models by accurately capturing complex flow interactions.

Design Influences

The of a is fundamentally shaped by geometric that balance , structural integrity, and operational requirements. The D is a primary , where larger values enhance by reducing induced power losses according to momentum theory principles, allowing the propeller to accelerate a greater mass of fluid at lower velocities. However, increases in D are constrained by structural limits, such as tip speed restrictions to avoid effects in or ground clearance issues, as well as material strength to withstand centrifugal forces. The pitch-to-diameter ratio P/D influences the advance ratio J = V / (nD), where higher ratios optimize for cruise conditions by enabling finer load distribution at higher forward speeds, while lower ratios are selected for takeoff and climb to maximize static and . The number of blades B affects dynamic behavior; increasing B distributes more evenly, reducing and through smoother torque pulsations, though it elevates overall demands on the due to higher rotational inertia. Blade loading is quantified by the activity factor (AF), defined for a single blade as \text{AF} = 2600 \sigma c_{l_{0.7R}}, where \sigma = B c / (\pi D) is the solidity, and c_{l_{0.7R}} is the lift coefficient at the 0.7 radius station (typically around 0.5); the total AF is then proportional to B. A higher AF indicates greater blade area relative to the disk, enabling higher thrust loading but potentially narrowing stall margins by increasing local angles of attack. Relatedly, propeller solidity \sigma = B c / (\pi D) modulates stall susceptibility; higher solidity provides broader stall margins by distributing lift but can reduce efficiency at off-design conditions due to increased profile drag. Application-specific adaptations further refine design choices. In marine propellers, skewed blades—where blade angle varies progressively along the span—mitigate cavitation by smoothing pressure gradients and delaying bubble inception, particularly in nonuniform wakes, unlike aircraft propellers where compressibility rather than cavitation dominates. Variable-pitch mechanisms in aircraft allow real-time adjustment of blade angle to match engine RPM to optimal values across flight phases, maintaining constant speed and efficiency without fixed-pitch compromises. Advanced configurations like address residual swirl losses from single rotors by having rear blades counter-rotate, recovering and boosting overall efficiency by 5-10% while neutralizing effects on the . Modern material selections, such as carbon-fiber-reinforced composites, enable weight reductions of 10-30% compared to aluminum or metal alloys, improving useful load and reducing vibrational stresses without sacrificing strength. As of 2025, emerging designs like toroidal propellers, featuring looped blade shapes, promise further efficiency gains of up to 30% alongside reduced noise and vibration, particularly for and electric applications.

References

  1. [1]
    Aircraft Propellers – Introduction to Aerospace Flight Vehicles
    Propellers convert rotational motion into thrust by creating aerodynamic lift forces on their blades, which act as rotating wings.
  2. [2]
    Propeller Analysis
    May 13, 2021 · From airfoil theory, we know that the pressure over the top of a lifting wing is lower than the pressure below the wing. A spinning propeller ...
  3. [3]
    Propeller Thrust | Glenn Research Center - NASA
    Jul 14, 2025 · The details of how a propeller generates thrust is very complex, but we can still learn a few of the fundamentals using the simplified momentum ...
  4. [4]
    None
    ### Summary of Marine Propeller Theory, Key Parameters, Design Aspects, and Fundamental Equations
  5. [5]
    Propellers: A complete history - CJR Propulsion
    Oct 11, 2023 · The first iterations of what can be recognised as an early 'propeller' occur around two millennia ago with Archimedes' Screw (est. 200 BC) which ...
  6. [6]
    The Early History Of The Screw Propeller - U.S. Naval Institute
    The first successful screw propellers for ships were put into use just a little less than one hundred years ago.
  7. [7]
    Hand-cranked model of stern of vessel fitted with Joseph Bramah's ...
    Hand-cranked model of stern of vessel fitted with Joseph Bramah's patented screw-propeller (no. 1478) from 1785 ; Made: 1785-1861 in Millwall ; Related people.
  8. [8]
    The Screw Propeller - who invented it? - Irvine Burns Club
    The design was patented in the U.S. in 1838-39. Both Smith & Ericsson introduced the screw propeller on war vessels in 1843, in their respective countries.
  9. [9]
    Early Helicopter Technology - Centennial of Flight
    In 1863, the Vicomte Gustave Ponton d'Amecourt built a model helicopter with counter-rotating propellers and a steam engine. He patented it in France and ...
  10. [10]
    Alphonse Pénaud - Linda Hall Library
    May 31, 2019 · Alphonse Pénaud, a French aviation pioneer, was born May 31, 1850. In the early 1870s, Pénaud began building model aircraft powered by twisted rubber cords.Missing: Gustave Ponton Amécourt
  11. [11]
    Pénaud's 'Planophore' took off 150 years ago - FAI
    May 18, 2021 · It was the French inventor Alphonse Pénaud who first used twisted rubber to power a flying model in 1870 and adopt its successful propulsion effect.Missing: Gustave Ponton Amécourt
  12. [12]
    [PDF] Wake Scale Effects and Propeller Hull Interaction: State of the Art.
    This survey reviews the progress of research on the dual problems of wake scale effects and hull propeller interactions. The survey covers theoretical,.
  13. [13]
    Langley Propeller, Fixed-Pitch, Two-Blade, Wood and Fabric
    During the 1890s, Langley mounted a substantial aerodynamic research program at the Smithsonian. This research included tests using a large whirling arm ...Missing: Pierpont | Show results with:Pierpont
  14. [14]
    [PDF] An Historical and Applied Aerodynamic Study of the Wright Brothers ...
    May 18, 2005 · Similarly, the Wrights estimated the lift-to-drag ratio of the 1903 Flyer to be 8 based mostly on their wind tunnel data. Full scale tests ...
  15. [15]
    The Wright Brothers and Airplane Propeller Design
    Jan 2, 2018 · Their design worked, and the finished project is estimated to produce a maximum efficiency of about 66 to 70 percent. Today's propellers are ...Missing: lift drag ratio source<|separator|>
  16. [16]
    [PDF] A PROPELLER DESIGN AND ANALYSIS CAPABILITY ... - CORE
    In 1900 and 1901 Stefan .K. Drzewiecki, a Polish mathematician ... utilised the blade element theory of Drzewiecki, they realised that it was not able to.
  17. [17]
    The Wright Brothers propeller theory and design - ResearchGate
    Although nearly 80% efficiency can be obtained from the design by the Wright brothers [1] , the requirements relating to environmental issues, carrying capacity ...
  18. [18]
    Actuator disk theory and blade element momentum theory for the ...
    Oct 1, 2023 · The Actuator Disk Theory and the Blade Element Momentum Theory (BEMT) are widely applied in the field of tidal and wind turbine design.
  19. [19]
    [PDF] Analysis of the Blade Element Momentum Theory - HAL
    Jul 1, 2022 · The Momentum Theory, also known as Disk Actuator Theory or Axial Momen- tum Theory, was introduced by William J. M. Rankine in 1865 [39] and is, ...
  20. [20]
    Actuator disk theory and blade element momentum ... - ResearchGate
    Sep 21, 2025 · (Rankine, 1865), still originally for propellers but later on for turbines. By combining these two fundamental mathematical models, velocity,.
  21. [21]
    Propeller Propulsion
    The details are complex because the propeller acts like a rotating wing creating a lift force by moving through the air. For a propeller-powered aircraft, the ...Missing: mechanism | Show results with:mechanism
  22. [22]
    [PDF] Chapter 7 - Propellers - Federal Aviation Administration
    General. The propeller, the unit that must absorb the power output of the engine, has passed through many stages of development.
  23. [23]
    11.7 Performance of Propellers - MIT
    Figure 11.28: Typical propeller efficiency curves as a function of advance ratio ( $ J=u_0/nD$ ) and blade angle (McCormick, 1979). Image fig7PropEta_web ...
  24. [24]
    Propeller Thrust
    Because the blades rotate, the tip moves faster than the hub. So to make the propeller efficient, the blades are usually twisted. The angle of attack of the ...
  25. [25]
    [PDF] 4. FORCES ON THE PROPELLER BLADES - VTechWorks
    To calculate W we use the vortex theory of a propeller (McCormick, 1967). From the Kutta-Joukowski theorem each element of the propeller blade must have a.
  26. [26]
    [PDF] TECHNICAL NOTE 2851
    Because high rotational speeds are necessary to maintain low advance ratios at high forward speeds, high levels of stress will be experienced in supersonic ...Missing: definition | Show results with:definition
  27. [27]
    [PDF] Structural Optimization Including Centrifugal and Coriolis Effects
    centrifugal force applied to the blade acting as a stiffening effect. This ... Coriolis Forces in Turbine Blade Vibrations", Journal of Engineering for Gas.
  28. [28]
    The Effect of Blade-Section Camber on the Stall-Flutter ...
    Oct 30, 2025 · Report presenting an investigation to determine the effect of blade-section camber on the stall-flutter characteristics of three propellers ...Missing: scholarly | Show results with:scholarly<|control11|><|separator|>
  29. [29]
    W. J. M. Rankine, “On the Mechanical Principles of the Action of ...
    W. J. M. Rankine, “On the Mechanical Principles of the Action of Propellers,” Transactions of the Institution of Naval Architects, 1865, pp. 13-39.
  30. [30]
    Modeling the influence of streamwise flow field acceleration ... - WES
    Jul 28, 2025 · Froude, R.: On the part played in propulsion by difference of fluid pressure, Transaction of the Institute of Naval Architects, 30, 390–405, ...
  31. [31]
    [PDF] Propeller Theories - 1) Momentum Theory 2) Blade Element Theory
    Assumptions for conceptual modeling of a propeller (Fig.) 1) The propeller is assumed to be replaced by an 'actuator disk', a flow energizer.
  32. [32]
    VII. Production of Thrust with a Propeller
    To understand more about the performance of propellers, and to relate this performance to simple design parameters, we will apply actuator disk theory. We model ...Missing: derivation | Show results with:derivation
  33. [33]
    Blade Element Theory - an overview | ScienceDirect Topics
    Although Froude's work of 1878 failed in some respects to predict propeller performance accurately, it was in reality a great advance, since it contained ...Missing: paper | Show results with:paper
  34. [34]
    A short History of the Propeller - MH-AeroTools
    1878, Froudes blade element sketch, William Froude develops the blade element theory. ; 1900-1905, The Wright brothers design and test propellers systematically ...
  35. [35]
    Propeller design I : practical application of the blade element theory
    This report is the first of a series of four on propeller design and contains a description of the blade elements or modified Drzewiecke theory.Missing: explanation | Show results with:explanation
  36. [36]
    [PDF] PROPELLER BLADE ELEMENT MOMENTUM THEORY WITH ...
    Blade element momentum theory is used as a low- order aerodynamic model of the propeller and is coupled with a vortex wake representation of the slip-stream to ...Missing: Stefan 1900
  37. [37]
    [PDF] Basic principles of ship propulsion
    Torque coefficient. The propeller torque Q = P / (2πn) is expressed dimensionless with the help of the torque coefficient KQ, see Fig. 2.08: ρ × n2 × d5. KQ ...
  38. [38]
    [PDF] Reynolds Number Effects on the Performance of Small-Scale ...
    Jun 16, 2014 · Propeller power is calculated from the measured propeller torque by ... dQ = r [dL sin (φ + αi) + dD cos (φ + αi)]. (12). In terms of the lift ...
  39. [39]
    [PDF] Design Considerations for Propellers in a Cavitating Environment
    maximum propeller thrust and torque loading limits are defined for four different blade section shapes, including recommended design limits. Introduction. A ...
  40. [40]
    [PDF] Performance 10. Thrust Models
    We can conclude this section by defining the ideal efficiency of a propeller. The efficiency of the propeller is the useful power out divided by the input power ...
  41. [41]
    [PDF] RESEARCH · MEMORANDUM
    sections in this group, suffers the greatest profile-drag loss (about. 4 ... efficiency of this propeller is about 2 percent less than the efficiency of ...
  42. [42]
    [PDF] NASA TM X-73612
    Apr 1, 1977 · for an initial design, The tip loss for this design is nearly 5 percent ... Figure 5 showed that there was a 7 percent swirl loss with the current ...
  43. [43]
    (PDF) Tip-Losses with Focus on Prandlt's Tip Loss Factor
    Jun 5, 2020 · The largest part of the chapter consists in the derivation of Prandtl's tip-loss factor. The developments are done using general expressions for the velocity ...
  44. [44]
    [PDF] Propeller Performance Data at Low Reynolds Numbers
    Jan 7, 2011 · The rig was designed with the torque transducer placed between the motor housing and support arm of the thrust mechanism, as is shown in Fig. 4.<|control11|><|separator|>
  45. [45]
  46. [46]
    Propeller Diameter - an overview | ScienceDirect Topics
    As the diameter increases, assuming the rotational speed is permitted to fall to its optimum value, the propeller efficiency will increase and hence for a ...Missing: structural | Show results with:structural
  47. [47]
    Aircraft Propeller Theory | AeroToolbox
    The moment of inertia of the blade increases exponentially with the blade diameter, which results in a much larger resulting propeller torque. The engine must ...
  48. [48]
    [PDF] NATIONAL ADVISORY COMMITTEE FOR AERONAUTICS
    Adjustment for propeller blade drag.- Figure 4 is a plot of esti- mated ... order to attain a maximum value of section lift-drag ratio. The induced.<|separator|>
  49. [49]
    Effect of Solidity and Aspect Ratio on the Aerodynamic Performance ...
    Apr 10, 2023 · Overall, the optimal AR for efficiency is equal to 0.4, but the best trade-off between the stall margin, fan pressure, and efficiency is ...
  50. [50]
    Influence of skew angle on the cavitation dynamics and induced low ...
    Jun 1, 2023 · A smaller skew angle gives slightly higher blade efficiency than a highly skewed propeller (Malmir, 2019). Mossad and Yehia (2011) showed that ...Missing: aircraft | Show results with:aircraft
  51. [51]
    aircraft propeller, constant speed propeller, ground adjustable ...
    Feb 28, 2025 · A constant-speed propeller is designed to automatically adjust its blade pitch to maintain a constant rotational speed in every phase of flight.
  52. [52]
    [PDF] Analysis and Experiments for Contra-Rotating Propeller - CORE
    The results show the superiority of the contra-rotating propeller over the traditional single propeller, with increased propeller efficiency of about 9% at the ...
  53. [53]
    What Are the Benefits of Composite Propellers?
    Apr 19, 2018 · In general, composite propellers offer a significant weight reduction over aluminum props. At Hartzell, our structural composite propeller ...