Blade element theory
Blade element theory (BET) is an aerodynamic analysis method used to predict the performance of rotating blades, such as those in propellers, rotors, and wind turbines, by dividing the blade into small, independent elements along its span and calculating the local lift and drag forces on each element based on two-dimensional airfoil data.[1][2] Each blade element is treated as a miniature airfoil subjected to the local flow velocity, which combines the rotational speed at that radius and the axial inflow, allowing the angle of attack to be determined and aerodynamic coefficients applied to compute incremental thrust and torque contributions that are then integrated across the entire blade.[3][2] Originating from William Froude's actuator disk and blade element approach in 1878 and further developed by Stefan Drzewiecki in the late 19th and early 20th centuries, BET was advanced through contributions from researchers at the National Advisory Committee for Aeronautics (NACA) and the Royal Aircraft Establishment (RAE) in the 1920s to 1950s, with significant foundational work by Hermann Glauert in the 1920s and 1930s providing equations for induced velocities and efficiency.[2][4] The theory assumes that elements act independently without significant spanwise flow interactions, that the flow is steady and axisymmetric, and that induced velocities are small relative to the freestream, enabling a straightforward numerical integration for overall performance predictions.[1][2] In practice, BET is often combined with momentum theory to form blade element momentum theory (BEMT), which iteratively couples local blade forces with global flow induction effects for more accurate modeling of axial and rotational velocities in the rotor plane.[1] This hybrid approach is widely applied in the design and analysis of aircraft propellers, helicopter rotors, and horizontal-axis wind turbines, where it facilitates optimization of blade twist, chord distribution, and pitch for maximum efficiency, typically achieving propeller efficiencies between 50% and 87% depending on operating conditions.[3][2] Key limitations include its neglect of three-dimensional effects like tip losses and wake rotation at high induction factors, necessitating empirical corrections such as Prandtl's tip-loss model or Glauert's empirical adjustment for axial induction factors exceeding 0.4.[1] Despite these approximations, BET remains a computationally efficient and foundational tool in aerospace engineering, underpinning modern simulation codes like NREL's AeroDyn for wind turbine aerodynamics.[1]Introduction
Definition and Principles
Blade element theory (BET) is a one-dimensional aerodynamic model employed to evaluate the performance of rotating blades, such as those in propellers or rotors, by segmenting the blade into discrete radial elements, or strips, aligned perpendicular to the spanwise direction. Each element functions independently as a two-dimensional airfoil cross-section, enabling the isolated computation of local lift and drag forces without accounting for interactions between strips. This strip method simplifies the analysis of blade aerodynamics by focusing on sectional contributions rather than the entire blade as a unified structure.[5][6] Central to BET are the prerequisite concepts of airfoil aerodynamics, where lift and drag forces on a blade section are quantified using dimensionless coefficients that vary with the angle of attack (the angle between the oncoming flow and the airfoil chord line), the Reynolds number (which governs the ratio of inertial to viscous forces in the flow), and the Mach number (which reflects the influence of compressibility at higher speeds). These coefficients are derived from experimental wind tunnel data or theoretical models tailored to the airfoil's geometry, providing the empirical basis for force predictions in each blade element.[7] The theory's core assumption is that aerodynamic forces on any given element arise solely from local flow conditions at that radial position, including the relative velocity (the vector sum of the blade's rotational velocity and any axial inflow), the local chord length, twist distribution, and collective pitch angle. This locality implies negligible spanwise flow or three-dimensional relief effects, allowing forces to be resolved into normal and tangential components relative to the rotor plane.[5][6] BET originated as an adaptation of fixed-wing airfoil strip theory—initially developed for non-rotating lifting surfaces—to the challenges of rotating systems, with foundational contributions from William Froude in 1878 and extensions by Stefan Drzewiecki in 1892 for propeller analysis. As a complementary approach to momentum theory, which addresses global mass and energy balances in the flow field, BET emphasizes detailed sectional loading to inform blade design and optimization.[5][6][8]Historical Development
The development of blade element theory traces its roots to early efforts in propeller analysis during the 19th century, where foundational momentum-based approaches laid the groundwork but revealed key shortcomings. William John Macquorn Rankine's 1865 formulation of momentum theory modeled propellers as actuator disks that accelerate fluid axially, providing an idealized estimate of thrust and power without considering blade geometry or detailed loading distribution. This theory, however, suffered from significant limitations in accurately predicting blade loading variations along the span, as it treated the propeller as a uniform disk and neglected rotational swirl or three-dimensional effects. The explicit inception of blade element theory emerged in the early 20th century through the work of Stefan Drzewiecki, a Polish engineer who between 1892 and 1920 developed the first systematic blade element approach primarily for marine propellers. Drzewiecki's method divided propeller blades into discrete radial sections, or elements, each analyzed as an independent airfoil contributing to overall thrust and torque via summation of local forces.[4] His 1920 publication, "Théorie générale de l'hélice," formalized this primitive version, emphasizing empirical airfoil data to account for sectional lift and drag, marking a shift from global momentum models to localized aerodynamic predictions.[9] A pivotal refinement occurred in 1926 with Hermann Glauert's application of blade element principles to aircraft propellers, integrating momentum theory to form the blade element momentum (BEM) framework. In his seminal book, The Elements of Aerofoil and Airscrew Theory, Glauert introduced "strip theory," where each blade element experiences a local inflow influenced by induced velocities from the overall momentum balance, incorporating circulation concepts from thin airfoil theory to better capture twist and solidity effects.[10] This combination enabled more precise performance calculations for airscrews, addressing Drzewiecki's oversight of wake-induced velocities and establishing BEM as a practical design tool for aviation.[10] Following World War II, blade element theory saw expanded adoption in rotary-wing aerodynamics, particularly through Theodore Theodorsen's 1940s analyses at the National Advisory Committee for Aeronautics (NACA). Theodorsen's work adapted blade element methods to helicopter rotors, accounting for varying blade angles and forward flight influences on elemental forces.[11] This integration facilitated early performance predictions for emerging helicopter designs, bridging fixed-wing propeller theory with the dynamic challenges of rotorcraft.[11] Key milestones in the 1950s involved extensions to compressible flow regimes, driven by high-speed propulsion demands. NACA investigations, including reports from 1950, incorporated blade element drag-lift ratios and Prandtl-Glauert corrections into the framework to model transonic effects on propeller blades, improving accuracy for supersonic tip speeds without overhauling the core methodology.[12] Subsequent decades witnessed no fundamental paradigm shifts, though computational implementations proliferated in the 1990s, such as Svend Øye's dynamic inflow model for BEM, which added time-dependent wake adjustments via first-order filters to enhance transient simulations in wind turbine and rotor analyses.[13] These numerical advancements solidified blade element theory's role as a standard, efficient tool, later complemented by modifications like Prandtl's tip loss corrections for finite blade effects.[13]Core Concepts
Blade Element Division
In blade element theory, the blade is discretized into a finite number of radial elements, typically between 10 and 20, extending from the hub to the tip to facilitate localized analysis of aerodynamic behavior.[14][15] Each element spans a small radial width dr at a local radius r from the axis of rotation, allowing the blade's geometry to be defined independently for each segment.[14] The local geometric parameters include the chord length c(r), which varies along the radius to optimize lift distribution, and the twist angle \beta(r), which adjusts the blade's pitch to maintain consistent angles of attack across the span.[16][14] Additionally, the local solidity \sigma, defined as \sigma = \frac{B c(r)}{2 \pi r} where B is the number of blades, quantifies the relative blade area at each radius and influences the overall loading.[14][4] The local flow geometry for each element is determined by the interaction between the axial inflow velocity V_{\text{axial}} and the tangential velocity due to rotation. The relative velocity V_{\text{rel}} experienced by the element is given by V_{\text{rel}} = \sqrt{( \Omega r )^2 + V_{\text{axial}}^2}, where \Omega is the angular velocity of the rotor.[14] This velocity vector forms the inflow angle \phi with respect to the plane of rotation, calculated as \phi = \atan\left( \frac{V_{\text{axial}}}{\Omega r} \right). The angle of attack \alpha for the element is then the difference between the inflow angle and the local twist: \alpha = \phi - \beta, under the assumption of small stall angles where two-dimensional airfoil characteristics remain valid.[14] A key assumption in this division is the independence of each element, treating it as an isolated two-dimensional airfoil section with no spanwise flow or three-dimensional effects, such as tip losses or sweep influences, initially considered.[14][16] This simplification enables the use of empirical airfoil data, such as lift and drag coefficients versus angle of attack from wind tunnel tests, to characterize each element's response.[14]Local Aerodynamic Forces
In blade element theory, the local aerodynamic forces acting on each infinitesimal blade element are determined by treating the element as a two-dimensional airfoil section subjected to the local relative flow velocity. The blade is conceptually divided into annular elements at radial position r, each with local chord length c(r), allowing the forces to be computed independently based on the local angle of attack \alpha. The lift force on such an element, perpendicular to the relative velocity \mathbf{V}_{\text{rel}}, is given by dL = \frac{1}{2} \rho V_{\text{rel}}^2 c(r) C_l(\alpha) \, dr, where \rho is the fluid density, C_l(\alpha) is the lift coefficient obtained from two-dimensional airfoil polars for the section's geometry and surface roughness, and dr is the radial width of the element.[17] Similarly, the drag force, parallel to \mathbf{V}_{\text{rel}}, is dD = \frac{1}{2} \rho V_{\text{rel}}^2 c(r) C_d(\alpha) \, dr, with C_d(\alpha) as the corresponding drag coefficient from the same airfoil data. These coefficients are functions of the local angle of attack, which is the angle between the chord line and \mathbf{V}_{\text{rel}}.[17] The elemental lift and drag forces are then resolved into components aligned with the axial (thrust) and tangential (torque-producing) directions relative to the rotor axis. The axial force contribution, or elemental thrust, is dT = dL \cos \phi - dD \sin \phi, where \phi is the local inflow angle between \mathbf{V}_{\text{rel}} and the plane of rotation. The tangential force contribution is dQ = dL \sin \phi + dD \cos \phi, and the associated elemental torque is dM = r \, dQ. This resolution accounts for the orientation of the forces relative to the rotor's motion, enabling the computation of overall propeller or rotor performance when integrated.[17] In some formulations, the lift and drag are combined into a normal force perpendicular to the chord line, dN = dL \cos \alpha + dD \sin \alpha, which simplifies analysis of the pressure distribution along the element but is less commonly used for direct thrust and torque calculations.[18] The accuracy of C_l and C_d depends on the local Reynolds number, Re = \rho V_{\text{rel}} c(r) / \mu, where \mu is the dynamic viscosity, as airfoil polars vary with Re due to viscous effects on boundary layer development. Blade element theory assumes attached flow over the airfoil, meaning no flow separation occurs, which justifies the use of steady, two-dimensional airfoil data without three-dimensional corrections for tip effects or spanwise flow. This assumption holds best at moderate angles of attack and higher Re, typical of full-scale propellers or rotors.[1]Mathematical Formulation
Elemental Force Equations
In blade element theory, the contributions to overall thrust and torque from an individual blade element are derived by resolving the local aerodynamic lift and drag forces into components aligned with the rotor's axial and tangential directions. This resolution relies on the local inflow angle \phi, which is the angle between the blade element's relative velocity vector and the plane of rotation. The lift force dL is perpendicular to the relative velocity, contributing positively to both thrust (via its axial projection) and torque (via its tangential projection), while the drag force dD opposes the relative velocity, subtracting from thrust and adding to torque in the propeller convention.[19] The differential thrust dT produced by all B blades at a radial position r over an elemental span dr is thus dT = B \left[ dL \cos\phi - dD \sin\phi \right], where the cosine and sine terms account for the geometric projection of the forces onto the axial direction. The differential torque dQ, which contributes to the rotational load, is obtained similarly by projecting the forces onto the tangential direction: dQ = B r \left[ dL \sin\phi + dD \cos\phi \right]. The total torque Q is found by integrating dQ from the blade root to tip, and the required power P to drive the rotor at rotational speed n (in revolutions per second) is P = 2\pi n Q. These expressions assume no radial flow and treat each element independently, with the inflow angle \phi determined from the vector sum of axial inflow and tangential blade speed.[20][19] To facilitate comparison across scales and operating conditions, these forces are often nondimensionalized using reference quantities such as air density \rho, rotational speed \Omega = 2\pi n, and rotor radius R. The local thrust coefficient for an element is expressed in differential form as dC_T = \frac{B c r^2}{\pi R^4} \left( C_l(\alpha) \cos\phi - C_d(\alpha) \sin\phi \right) dr, where c(r) is the local chord length, and C_l(\alpha) and C_d(\alpha) are the sectional lift and drag coefficients at the local angle of attack \alpha. This form arises by substituting the expressions for dL and dD (with relative speed W \approx \Omega r in high-speed approximations) into the dimensional equations and normalizing by the reference thrust \frac{1}{2} \rho (\Omega R)^2 \pi R^2. A corresponding local torque coefficient dC_Q follows analogously, enabling the computation of overall performance coefficients C_T = \int dC_T and C_Q = \int dC_Q. Approximations neglecting drag simplify to dC_T \approx \frac{B c r^2}{\pi R^4} C_l(\alpha) \cos\phi \, dr, emphasizing the dominant role of lift in efficient designs.[21][22]Integration and Performance Metrics
In blade element theory, the total aerodynamic forces and moments on a propeller or rotor blade are determined by integrating the contributions from each infinitesimal blade element along the radial span. The total thrust T is computed as the integral of the differential thrust dT from the hub radius r_{\text{hub}} to the tip radius r_{\text{tip}}: T = \int_{r_{\text{hub}}}^{r_{\text{tip}}} dT, where dT arises from the local lift and drag forces resolved in the axial direction. Similarly, the total torque Q is the integral of the differential torque dQ, which accounts for the moment arm r and the tangential force components: Q = \int_{r_{\text{hub}}}^{r_{\text{tip}}} dQ. The required power P is then obtained from the torque and rotational speed n (in revolutions per second) as P = 2\pi n Q, or equivalently by integrating the elemental tangential power contributions: P = \int_{r_{\text{hub}}}^{r_{\text{tip}}} \Omega r \, dF_{\text{tangential}}, with \Omega = 2\pi n as the angular velocity and dF_{\text{tangential}} the tangential force on each element. These integrations aggregate the local aerodynamic effects to yield the overall blade performance, assuming independence of elements along the span.[23][2] Key performance metrics are derived from these integrated quantities to characterize efficiency and non-dimensional behavior. Propeller efficiency \eta is defined as the ratio of useful axial power output to input shaft power: \eta = \frac{T V_{\text{axial}}}{P}, where V_{\text{axial}} is the freestream velocity; this metric, typically peaking around 0.8 for optimized designs, quantifies how effectively the integrated thrust propels the vehicle relative to the power expended. The advance ratio J = \frac{V_{\text{axial}}}{n D}, with D as the propeller diameter, provides a non-dimensional measure of operating condition, influencing the flow angle at each radius and thus the integrated performance. Additionally, the power coefficient C_P = \frac{P}{\rho n^3 D^5}, where \rho is fluid density, normalizes power for scaling and comparison across designs.[23][2][21] For practical computation, especially with variable pitch blades where the twist angle \beta(r) varies radially, the integrals are approximated numerically by discretizing the blade into finite elements (e.g., 10–20 sections). The trapezoidal rule is commonly applied to sum the elemental contributions, ensuring accurate handling of pitch variations by evaluating local angles of attack at each discrete radius. This approach maintains the theory's simplicity while enabling performance predictions for complex geometries.[23][2][21]Applications
Propellers and Rotors
Blade element theory (BET) is extensively applied in the design of fixed-pitch and variable-pitch propellers, where it facilitates the optimization of blade twist angle β(r) and chord distribution c(r) along the radius r to achieve uniform loading across the blade span. This approach minimizes variations in sectional loading, enhancing overall efficiency by distributing thrust evenly and reducing tip losses. For instance, in optimizing for uniform inflow, designers adjust β(r) to maintain optimal local angles of attack and vary c(r) to balance lift generation, often targeting an ideal twist that aligns with the varying tangential velocity Ωr. Off-design performance is assessed by varying the advance ratio J = V/(nD), where V is freestream velocity, n is rotational speed, and D is diameter; BET predictions show efficiency drops at low J due to increased blade angles of attack and stall risks, guiding adjustments for multi-speed operations. In helicopter rotor applications, BET incorporates flap and lag degrees of freedom to model the dynamic response of articulated blades under forward flight conditions. Flapping motion, which allows blades to tilt out-of-plane to counter asymmetric lift, is analyzed by integrating local aerodynamic forces with hinge offsets, while lag motion accounts for in-plane bending that influences stability and power requirements. These extensions enable prediction of rotor trim and control margins by coupling blade element forces with flapping equations. Autorotation, critical for safe descents, is analyzed under zero torque conditions, where BET balances drag-induced deceleration with inflow to sustain rotation, estimating descent rates for various collective pitches.[26] For marine propellers, BET addresses cavitation effects, particularly the reduction in lift coefficient Cl at higher radial positions r due to locally lowered pressures from high rotational speeds. Cavitation inception alters Cl by introducing vapor cavities that disrupt lift, with models combining BET and panel methods to predict performance drops. This informs designs with increased blade area ratios to mitigate pressure minima.[27] Basic integration of elemental forces yields total thrust T and torque Q, essential for performance metrics in these applications.Wind Turbines and Hydrofoils
Blade element theory (BET), when adapted for wind turbines, focuses on power extraction rather than propulsion, analyzing horizontal-axis wind turbines (HAWTs) by dividing blades into elements to predict aerodynamic loading and efficiency. The tip-speed ratio, defined as \lambda = \Omega [R](/page/R) / V_{\text{[wind](/page/Wind)}}, where \Omega is the rotational speed, R is the rotor radius, and V_{\text{wind}} is the freestream wind speed, governs the relative flow angles and optimal blade twist for maximum power coefficient C_P. BET combined with momentum theory enables prediction of blade loading that approaches the theoretical Betz limit of C_P \approx 0.59, representing the maximum extractable power from an ideal actuator disk, though practical HAWT designs using BET achieve lower values due to tip losses and non-ideal flow.[28][28][28] In hydrofoils and tidal devices, BET accounts for the higher density of water (\rho_{\text{water}} \approx 1000 \, \text{kg/m}^3) compared to air, which amplifies hydrodynamic forces on blade elements and requires adjusted airfoil data for lift and drag coefficients at relevant Reynolds numbers. Tidal stream turbines apply BET by modeling blades as summed 2D hydrofoil elements, optimizing chord and pitch distributions to maximize torque under bidirectional flows. Bio-inspired designs, such as those mimicking fish fins, approximate propulsion or maneuvering surfaces as series of elemental hydrofoils, using BET to estimate thrust from flapping or undulating motions while incorporating added mass effects in dense fluids.[29][29][30] Yaw misalignment and wake interactions in wind turbines are addressed in BET through simple corrections that adjust the effective relative velocity V_{\text{rel}} for each blade element, incorporating skewed inflow angles to account for reduced axial induction in yawed conditions. For wake effects, downstream turbines experience velocity deficits, which BET models by modifying V_{\text{rel}} based on wake skew \chi \approx (0.6a + 1)\gamma, where a is the axial induction factor and \gamma is the yaw angle, thereby estimating power losses without full CFD.[28][31] Since the 2000s, BET has remained relevant for early-stage sizing of offshore wind turbines, providing low-fidelity aerodynamic predictions to evaluate rotor scaling, power curves, and levelized cost of energy in farm layouts. As of 2025, ongoing advancements include improved BEMT formulations for urban air mobility applications, such as electric vertical takeoff and landing (eVTOL) vehicles, enhancing robustness for low-advance-ratio operations.[32][33]Limitations and Extensions
Fundamental Limitations
Blade element theory (BET), in its basic form, relies on the strip theory assumption, which treats each blade section as an independent two-dimensional airfoil operating in isolation. This approach neglects three-dimensional flow effects, including spanwise flows, interference between blade elements, and the formation of tip vortices that alter local velocities near the blade tips. As a result, BET overpredicts aerodynamic loading and thrust in these regions, with inaccuracies reaching 5-10% in overall performance estimates and up to 15-20% in local load predictions under uniform inflow conditions.[34][35] A core oversight in BET is its treatment of flow as effectively inviscid beyond the use of precomputed two-dimensional airfoil polars for lift and drag coefficients. While these polars empirically account for viscous effects like boundary layer development and separation in isolated airfoils, the theory ignores spanwise viscous interactions, three-dimensional boundary layer growth along the blade, and associated flow separation in rotating environments. This limitation becomes pronounced at high angles of attack or off-design conditions, where real boundary layer dynamics lead to unmodeled stall delays or early separation, reducing predictive accuracy for drag and moment coefficients.[36] BET further assumes uniform inflow across the rotor disk, implying constant axial and tangential velocities at each blade element without azimuthal or radial variations. This simplification fails in yawed inflow scenarios, where the angle between the rotor axis and freestream creates non-uniform loading and skewed wakes, or during stalled conditions with significant flow unsteadiness. Lacking any inherent wake modeling, BET cannot capture the downstream momentum deficits or vortex interactions that influence upstream elements, leading to errors in induced velocity estimates and overall rotor performance under non-axial flows.[37] Historically, pre-1930s formulations of BET, such as those developed by Drzewiecki around 1900, exhibited significant inaccuracies by disregarding momentum theory principles. These early versions ignored induced velocity losses and slipstream momentum deficits, resulting in overestimations of propeller efficiency by 5-10% compared to experimental measurements—for instance, theoretical efficiencies of 83% versus observed values of 77% in model tests. Such errors stemmed from assuming freestream conditions at the blades without accounting for the axial and rotational momentum changes across the rotor, which only later integrations with momentum theory in the 1920s began to address.[38][21]Key Modifications and Improvements
One significant enhancement to the basic blade element theory (BET) is Prandtl's tip loss correction, introduced in the 1920s to account for three-dimensional flow effects at blade tips arising from finite blade numbers and vortex shedding. This correction modifies the effective solidity by applying a tip loss factor F = \frac{2}{\pi} \cos^{-1} \left( e^{-f} \right), where f = \frac{B}{2} \frac{1 - r/R}{\sin \phi}, with B denoting the number of blades, r the local radius, R the blade radius, and \phi the local inflow angle. By reducing the predicted loading near the tip, this factor addresses overestimation in pure BET, improving tip load predictions by approximately 10-15% in typical rotor applications and enhancing overall thrust and power estimates.[39][40][41] The Blade Element Momentum Theory (BEMT), developed as a hybrid of BET and momentum theory in the 1930s, further refines performance predictions by incorporating induction effects from the rotor's global flow field. In BEMT, the axial induction factor a is iteratively solved by equating the blade element differential thrust to the momentum theory expression dT = 4 \pi \rho V^2 a (1 - a) r \, dr, where \rho is fluid density, V is the far-field velocity, and the blade element forces are computed using the local relative velocity V_{\text{rel}}. This approach balances local blade forces with annular momentum changes, enabling more precise computations of thrust, torque, and power across the span, particularly for lightly loaded rotors where pure BET underestimates wake influences.[1][42] Extensions for compressible flows and unsteady conditions emerged in the 1970s through vortex wake models that integrate trailed vorticity from blade elements into the theory, capturing dynamic wake evolution and transient aerodynamic responses in applications like maneuvering rotors. These models simulate the convective and diffusive behavior of helical vortex sheets, improving load predictions during acceleration or yaw by up to 20% compared to steady-state assumptions. In contemporary practice, hybrid CFD-BET methods couple low-order BET with high-fidelity computational fluid dynamics simulations to resolve complex transient loads, such as those from gusts or turbine interactions, while retaining computational efficiency for design iterations.[43][44] Recent advancements since 2010 have leveraged machine learning to enhance BET's handling of lift coefficient C_l(\alpha) under variable conditions, such as turbulent or yawed inflows, by developing data-driven surrogate models trained on CFD or experimental airfoil data. These techniques predict nonlinear C_l(\alpha) variations, including stall delays and post-stall behavior, enabling BET applications in stochastic environments with reduced reliance on static lookup tables and improved accuracy for fatigue load assessments in wind energy systems. As of 2024, physics-informed neural networks (PINNs) have been integrated as surrogate models for BEMT, bypassing iterative solutions for faster predictions in transient flows while maintaining high fidelity.[45][46][47]Examples
Analytical Example
To illustrate the application of blade element theory (BET) in a straightforward manner, consider a simple two-bladed propeller with radius R = 1 m, rotational speed n = 1000 RPM (corresponding to angular velocity \omega = 2\pi n / 60 \approx 104.72 rad/s), operating in an axial freestream velocity V_\mathrm{axial} = 10 m/s. The blades have uniform chord length c = 0.1 m and constant pitch angle \beta = 20^\circ. Air density is taken as \rho = 1.225 kg/m³. Airfoil performance data are obtained from experimental polars for the symmetric NACA 0012 section at Reynolds number \mathrm{Re} \approx 6 \times 10^6 and low Mach number (M = 0.15).[48] The propeller disk is divided into 10 annular elements of equal radial width \Delta r = 0.1 m, with each element centered at radial position r_k = 0.05 + (k-1) \times 0.1 m for k = [1](/page/1) to $10 (i.e., r = 0.05, 0.15, \dots, 0.95 m). For each element, the local tangential velocity is U_t = \omega r, and the inflow angle is \phi(r) = \tan^{-1}(V_\mathrm{axial} / U_t). The effective angle of attack is then \alpha(r) = \beta - \phi(r). The relative velocity magnitude is W(r) = \sqrt{V_\mathrm{axial}^2 + U_t^2}. Using B = 2 blades, the elemental thrust and torque contributions are calculated as \Delta T = B c \Delta r \cdot \frac{1}{2} \rho W^2 (C_L \cos \phi - C_D \sin \phi), \Delta Q = B c \Delta r \cdot \frac{1}{2} \rho W^2 r (C_L \sin \phi + C_D \cos \phi), where C_L and C_D are the lift and drag coefficients interpolated from the NACA 0012 polars at \alpha(r). Since the airfoil is symmetric, C_L(-\alpha) = -C_L(\alpha) and C_D(-\alpha) = C_D(\alpha). For |\alpha| > 16^\circ (beyond the tabulated stall point), values are held constant at the $16^\circ data for simplicity in this hand calculation.[48][49] The following table presents representative calculations for three elements (inner, mid-span, and near-tip) to demonstrate the process; full summation over all 10 elements yields the totals. Angles are in degrees, and coefficients are interpolated linearly from the experimental table (e.g., for \alpha = 9.2^\circ, C_L \approx 1.005 and C_D \approx 0.0095 via linear fit between $8^\circ (C_L = 0.880, C_D = 0.0080) and $10^\circ (C_L = 1.088, C_D = 0.0105) points at Re=6×10^6). For low |\alpha|, approximate C_L \approx 0.11 \alpha (in radians) and C_D \approx 0.0065.[48]| r (m) | \phi (°) | \alpha (°) | C_L | C_D | \Delta T (N) | \Delta Q (N·m) |
|---|---|---|---|---|---|---|
| 0.25 | 20.9 | -0.9 | -0.100 | 0.0065 | -0.89 | -0.07 |
| 0.50 | 10.8 | 9.2 | 1.005 | 0.0095 | 34.3 | 3.4 |
| 0.95 | 5.8 | 14.2 | 1.465 | 0.0165 | 178.4 | 19.0 |