Fact-checked by Grok 2 weeks ago

Projected area

The projected area of an object is the area of its orthogonal onto a to a specified , equivalent to the or area the object would cast under illumination from that . This is central to fields like and , where it quantifies the effective cross-sectional area influencing interactions with fluids, , or forces. In , projected area most commonly refers to the frontal projected area for calculations on bodies, such as or , defined as the cross-sectional area to the oncoming . For instance, in the force equation D = \frac{1}{2} \rho V_\infty^2 C_D A_{\rm ref}, the reference area A_{\rm ref} is typically this projected frontal area, which determines the volume of fluid displaced and thus the pressure component. For bodies like a of d, it [equals \pi](/page/Equals_Pi) d^2 / 4, while for a of side length l, it is l^2. For lifting surfaces like aircraft , the projected area is the planform area, the shadow of the wing viewed from above (perpendicular to the wing's ), rather than its actual wetted surface area. This area S is calculated as the wing b times the average chord length c_{\rm avg}, or equivalently S = b^2 / AR where AR is the , and it features in the L = \frac{1}{2} \rho V^2 S C_L. Examples include the 247's wing with S = 836 ft², which influences its during cruise at approximately C_L = 0.23. Beyond , projected area applies in and analysis tools, such as NASA's software, where it computes the silhouette area of vehicle components in custom directions for tasks like estimating rotor disk downloads on helicopters or overall profiles. It also correlates with coefficients for irregular particles in high-velocity flows, where larger projected areas indicate lower slenderness and higher . Overall, the choice of direction ensures the area reflects the relevant physical interaction, making it indispensable for performance predictions in engineering design.

Fundamentals

Definition

The projected area of a three-dimensional object or surface refers to the of that object onto a , forming a two-dimensional that corresponds to the shadow it would cast under illumination to the . This effectively captures the object's as seen from the of , independent of its depth or internal features. In contrast to the actual surface area, which measures the total extent of an object's exterior including all curvatures and facets, the projected area disregards these details and considers only the relative to the direction, yielding a simplified for effective exposure. This distinction is crucial in analyses where the projected area provides an of the "frontal" exposure without requiring over the full . The concept emerged in early 20th-century contexts to streamline calculations, such as loads on structures. By the 1930s, similar provisions were adopted internationally, as seen in the 1935 Model Building Bylaw, where pressures on cylindrical s were computed as fractions of the projected area. This approach facilitated practical design without complex surface modeling. Projected area serves as a foundational for evaluating and s on irregular geometries, enabling engineers to apply uniform coefficients to the silhouette rather than performing detailed integrations over curved or oriented surfaces. In applications like , it underpins simplified models for fluid interactions.

Geometric Interpretation

The projected area of a three-dimensional object onto a arises from , a method where all rays are directed perpendicular to the , forming the object's without distortion from effects. In this setup, an imaginary transparent is positioned between the observer and the object, and the object's features are projected onto it via lines parallel to the viewing direction, capturing the spatial extent as if viewed from . This geometric process ensures that distances and shapes in the accurately represent the object's outline in the chosen direction, independent of the object's distance from the . The size of the projected area depends critically on the alignment between the projection direction and the object's orientation. When the projection direction is normal to the object's broadest face, the projected area is maximized, fully exposing the cross-sectional extent. Conversely, as the viewing angle tilts toward edge-on alignment—where the projection direction lies parallel to the face—the projected area decreases, reaching zero for a perfectly flat or infinitely thin profile, as the silhouette collapses to a line. For opaque objects, the projected area is defined by the region bounded by the contour in the , corresponding to the total shadowed region under parallel illumination. In contrast, for transparent objects or surface-based analyses, the projected area involves integrating the contributions from all visible surfaces, computed as the sum of differential areas weighted by the of the cosine of the angle between each surface and the projection direction: A_p = \int_S |\cos \theta| \, dA, where S is , \theta is the angle between the normal and the axis, and dA is the differential surface area; this accounts for the foreshortening of each element without overlap considerations inherent to opacity. Illustrative diagrams commonly depict these principles using simple shapes like a . When the projection direction aligns with the 's , the forms an (a if the is to the ), highlighting the base's . Rotating the projection direction to the yields a rectangular , emphasizing the lateral extent and demonstrating how alters the geometric .

Mathematical Formulation

General Expression

The projected area of an arbitrary surface S onto a plane perpendicular to a fixed unit direction vector \mathbf{d} is given by the surface integral A_{\text{projected}} = \int_{S} \cos \beta \, dS, where \beta is the angle between the outward unit normal vector \mathbf{n} to the surface at each point and the projection direction \mathbf{d}, and dS denotes the differential surface area element. This expression quantifies the effective area as seen from the direction \mathbf{d}, accounting for the orientation-dependent foreshortening of each surface element. The formula derives from the geometric projection of infinitesimal surface elements. Consider a small area element dS with normal \mathbf{n}; its projection onto the plane normal to \mathbf{d} has area dS_{\text{proj}} = dS \cos \beta, since \cos \beta represents the scaling factor along the direction perpendicular to the . With unit vectors, \cos \beta = \mathbf{n} \cdot \mathbf{d}, so the total projected area is the \int_{S} (\mathbf{n} \cdot \mathbf{d}) \, dS. This captures the of the oriented area vector d\mathbf{A} = \mathbf{n} \, dS onto \mathbf{d}. The derivation assumes a fixed projection direction \mathbf{d} and applies to both open and closed surfaces as appropriate; for simple cases without self-occlusion, the is taken over the relevant portion of the surface (typically where \mathbf{n} \cdot \mathbf{d} \geq 0) to avoid cancellation on closed surfaces. The formula assumes diffuse (, treating all rays parallel along \mathbf{d}; for self-shadowing complex surfaces, additional occlusion handling—such as ray tracing or visibility determination—is necessary to exclude hidden elements from the .

Projections for Common Shapes

For common geometric shapes, the projected area onto a plane perpendicular to a given direction can be derived from the general formulation by considering the orientation angle \beta, defined as the angle between the surface normal and the projection direction. For shapes with flat or uniformly oriented surfaces, this simplifies to the original area multiplied by \cos \beta. A flat rectangle of length L and width W, with original area A = L \times W, has a projected area A_{\text{proj}} = L \times W \cos \beta. This holds when the rectangle lies in a plane, and the maximum projected area is A at \beta = 0^\circ (normal incidence), decreasing to 0 at \beta = 90^\circ. Similarly, a circular disc of radius r, with original area A = \pi r^2, projects to an ellipse with area A_{\text{proj}} = \pi r^2 \cos \beta. The projection is maximal at \beta = 0^\circ and vanishes edge-on at \beta = 90^\circ. For a sphere of radius r, with total surface area A = 4\pi r^2, the projected area is the silhouette area, which is always A_{\text{proj}} = \pi r^2, independent of \beta due to rotational symmetry. This corresponds to the area of the circular shadow cast by the sphere in any direction. For a right circular of r and height h, the projected area depends on the angle \phi between the cylinder axis and the projection direction \mathbf{d}: A_{\text{proj}} = \pi r^2 |\cos \phi| + 2 r h |\sin \phi|. End-on (\phi = 0^\circ): \pi r^2; side-on (\phi = 90^\circ): $2 r h. The of a right circular of radius r and slant height l forms a bounded by two straight generators from the to the points on the and the elliptical arc between those points. The exact area depends on the angle and orientation angle \beta (between and \mathbf{d}); end-on (\beta = 0^\circ): \pi r^2; side-on (\beta = 90^\circ): triangular . For general \beta, if the is outside the , the area involves the elliptical plus triangular parts; a precise formula is: if t \cos \beta < r \sin \beta (where t is half- angle related), A_{\text{proj}} = \pi r^2 \sin \beta; otherwise, A_{\text{proj}} = \frac{\sin \beta}{2} [\pi r^2 + 2 r^2 \sec^{-1} (\frac{t \cot \beta}{2 t^2 \cot^2 \beta - 2 r}) + r^2 t^2 \cot^2 \beta - 2 r t \cot \beta] (adjusted for parameters).
ShapeOriginal AreaProjected Area FormulaAngle DependencyMaximum ValueMinimum Value
Flat RectangleL \times WL \times W \cos \betaVaries with \cos \betaL \times W (\beta=0^\circ)0 (\beta=90^\circ)
Circular Disc\pi r^2\pi r^2 \cos \betaVaries with \cos \beta\pi r^2 (\beta=0^\circ)0 (\beta=90^\circ)
Sphere$4\pi r^2\pi r^2Independent\pi r^2\pi r^2
Cylinder$2\pi r h + 2\pi r^2$\pi r^2\cos \phi+ 2 r h\sin \phi
Cone\pi r (r + l)Bounded by generators and base arc; see text for formulaVaries with \beta and apex angle\pi r^2 (end-on)0 (edge-on)
The table above summarizes the angle dependencies, with maximum values at normal incidence where applicable and minima at grazing angles.

Applications

Fluid Dynamics

In fluid dynamics, the projected area plays a crucial role in quantifying drag forces experienced by objects moving through air or water, particularly in aerodynamics and hydrodynamics. It represents the effective cross-sectional area perpendicular to the flow direction, simplifying the analysis of complex three-dimensional shapes by focusing on the silhouette that interacts most directly with the fluid. This concept is integral to the drag equation, which models the resistive force opposing motion: F_d = \frac{1}{2} C_d \rho v^2 A_{\text{proj}} where F_d is the drag force, C_d is the dimensionless drag coefficient dependent on shape and flow conditions, \rho is the fluid density, v is the relative velocity, and A_{\text{proj}} is the projected area. The projected area thus scales the drag magnitude, with larger values increasing resistance for a given velocity and coefficient. For objects falling under gravity, such as skydivers or projectiles, the projected area determines the terminal velocity, the constant speed reached when drag balances gravitational force. At terminal velocity, F_d = mg, leading to: v_t = \sqrt{\frac{2mg}{C_d \rho A_{\text{proj}}}} where m is mass and g is gravitational acceleration. Increasing A_{\text{proj}} reduces v_t, as seen in parachutes, where the canopy's inflated projected area—typically modeled as a circular disk—maximizes drag to safely slow descent rates to around 5-6 m/s. In aircraft, the wing planform area often serves as the reference projected area for overall drag and lift calculations, while the frontal projected area of the fuselage and components estimates parasite drag contributions. Variations in angle of attack—the angle between the oncoming flow and an object's chord line—alter the effective projected area, thereby influencing both lift and drag in vehicle design. For instance, increasing the angle of attack on an airfoil can enhance the projected area normal to the flow, boosting induced drag while initially increasing lift until stall occurs around 15-20 degrees. This effect is critical in optimizing aircraft performance, where minimizing drag at cruise angles (typically 2-4 degrees) reduces fuel consumption. The adoption of projected area concepts in aerodynamics gained prominence during World War II, particularly in modeling drag for projectiles and aircraft; seminal analyses, such as those on the Messerschmitt Bf 109 fighter, used frontal projected areas to predict parasite drag coefficients around 0.18 based on 0.84 m² reference.

Structural Engineering

In structural engineering, the projected area plays a pivotal role in evaluating wind and pressure loads on buildings and other fixed structures, ensuring designs account for effective exposure to environmental forces. The wind pressure p is determined using the formula p = q C_p, where q represents the velocity pressure dependent on wind speed, height, and exposure, and C_p is the external pressure coefficient that varies by surface orientation and shape. The total wind force F on the structure is then computed as F = p A_{\text{proj}}, with A_{\text{proj}} denoting the perpendicular to the wind direction, which captures the structure's silhouette as seen from the wind's approach and avoids overcounting sloped or irregular surfaces. This approach is essential for quasi-static load assessments in building design. Building codes like ASCE 7-22 standardize the use of projected area within wind load calculations, integrating it with gust effect factors G and exposure categories (B for urban, C for suburban, D for open terrain) to adjust the velocity pressure exposure coefficient K_z or K_{zt} for topographic effects. These provisions ensure that projected areas inform the directional procedure for main wind-force resisting systems (MWFRS), applying pressures normal to vertical projections for walls and horizontal projections for roofs, thereby scaling loads realistically for height and site conditions. For instance, in low-rise buildings under 60 feet, simplified methods still rely on projected tributary areas to distribute forces efficiently. For curved surfaces such as arched roofs or domes, engineers apply projected area to mitigate overestimation of loads on inclined planes, converting the curved geometry into an equivalent flat projection orthogonal to the force vector for conservative yet accurate pressure application. This method aligns with ASCE 7 guidelines for non-standard roofs, where the projected horizontal or vertical area simplifies force integration without detailed CFD analysis, particularly for uplift and suction on dome crowns. In contexts involving seismic resilience and axial stress analysis, projected area is employed in hardness testing to characterize material properties critical for structural components. The Vickers hardness test, for example, calculates hardness HV as the applied load F divided by the surface area of the diamond indenter's residual indentation, providing a measure of resistance to plastic deformation under load—key for assessing steel or concrete reinforcements in earthquake-prone designs. This indentation-based metric informs allowable stresses in axial members, ensuring materials withstand combined dynamic and static demands. A historical case study is the Empire State Building's design in the early 1930s, where engineers used projected facade areas to estimate wind forces, supplemented by wind tunnel testing on scale models to measure pressure distributions and validate load paths in the steel frame. This approach, predating modern codes, relied on empirical gust factors and uniform pressures applied to projections, demonstrating the projected area's enduring role in high-rise wind engineering despite lacking contemporary exposure categorizations.

Optics and Radiation

In optics and radiation transfer, the projected area plays a crucial role in determining the effective surface exposed to incoming radiant flux, particularly through Lambert's cosine law, which governs illuminance on a surface. The illuminance E from a point source of intensity I at distance d and incidence angle \theta is given by E = \frac{I \cos \theta}{d^2}, where the cosine term accounts for the reduction in effective area perpendicular to the rays, equivalent to the projected area A_{\text{proj}} = A \cos \theta for a surface of area A. This law ensures that the received intensity scales with the projected area, as oblique angles foreshorten the intercepting surface, a principle foundational to radiative transfer calculations. In solar energy applications, the projected area directly influences photovoltaic panel efficiency by modulating the incident solar irradiance. For a panel of actual area A_h, the effective projected area is A_p = A_h \cos \theta, where \theta is the angle between the panel normal and the sun's rays; this cosine projection maximizes power output when the panel faces the sun directly, minimizing losses from off-normal incidence. The power received P is thus P = F \cdot A_p, with F as the solar flux, highlighting how orientation adjustments can enhance annual energy yield by up to 20-30% in varying latitudes. Photometry employs projected area in lighting design to characterize luminaire performance and luminance, as standardized by the Illuminating Engineering Society of North America (IESNA). In IESNA LM-37, the projected luminous area of a fixture's lens or opening, viewed at angle \theta, is calculated to derive average luminance from goniophotometric candela data, enabling accurate prediction of light distribution and glare in interior spaces. This approach ensures that fixture output ratings reflect the effective emitting area, supporting compliant designs for uniform illumination in applications like offices and roadways. For thermal radiation between surfaces, projected areas approximate view factors in heat exchange models, quantifying the fraction of radiation leaving one surface that intercepts another. The , for instance, computes the view factor as the projected area of the receiving surface onto a hemisphere's base divided by the base area \pi r^2, providing a geometric shortcut for complex enclosures without full integration. Such approximations are essential in engineering radiative heat transfer, as in furnace design or building envelopes, where they simplify the double integral over surface orientations. A practical example is satellite solar array orientation, where maximizing the projected area toward the sun optimizes power generation during orbits. For low-Earth orbit satellites, algorithms derive tracking laws to continuously align arrays such that the instantaneous projected area is maximized, yielding up to 15-20% higher average power compared to fixed orientations, critical for mission longevity. This involves real-time adjustments to counter orbital geometry, ensuring the array's effective area A \cos \theta remains near-optimal relative to solar incidence.

Computation Methods

Analytical Approaches

One analytical approach to computing the projected area of complex polyhedra involves decomposing the object into its constituent faces and summing the contributions from each, leveraging the property of convex bodies. For a convex polyhedron, the projected area A_p(\mathbf{u}) in direction \mathbf{u} (a unit vector) is given by A_p(\mathbf{u}) = \frac{1}{2} \sum_f A_f |\mathbf{n}_f \cdot \mathbf{u}|, where the sum is over all faces f, A_f is the area of face f, and \mathbf{n}_f is its outward unit normal. This formula arises because the projections of front-facing and back-facing faces contribute equally to the silhouette area without overlap in the measure-theoretic sense for convex shapes. For curved surfaces, vector calculus provides a pathway by parameterizing the surface and evaluating the surface integral of the absolute cosine of the angle between the surface normal and the projection direction. The general expression for the projected area of a closed convex surface S is A_p(\mathbf{u}) = \frac{1}{2} \int_S |\mathbf{n}(\mathbf{x}) \cdot \mathbf{u}| \, dS, where \mathbf{n}(\mathbf{x}) is the unit outward normal at point \mathbf{x} \in S. For analytically tractable surfaces like , this integral yields a closed-form solution. Specifically, for a triaxial with semi-axes a \geq b \geq c > 0 aligned with the coordinate axes, the projected area in direction \mathbf{u} = (l, m, n) (with l^2 + m^2 + n^2 = 1) is A_p(\mathbf{u}) = \pi a b c \sqrt{\frac{l^2}{a^2} + \frac{m^2}{b^2} + \frac{n^2}{c^2}}. This result follows from parameterizing the via spherical coordinates adjusted for the axes and integrating the projected differential area elements. For more intricate curved surfaces like , parameterization (e.g., using coordinates with major radius R and minor radius r) allows analytical evaluation of the integral, though it typically involves elliptic integrals that must be computed explicitly for given orientations. Exploiting simplifies computations for rotationally symmetric objects, such as cylinders or cones, by reducing the three-dimensional problem to a two-dimensional cross-section. The projected area can then be found by integrating the projected length of the cross-sectional to the of , multiplied by the appropriate width factor, avoiding full surface parameterization. This approach is particularly efficient when the projection direction aligns with or is to the , yielding expressions in terms of elementary functions like arcsines or logarithms for the boundary contributions. For irregular objects where exact computation is infeasible, approximations based on error analysis often rely on the mean projected area over all directions, which for any convex body equals one-quarter of its total surface area S, i.e., \bar{A}_p = S/4. This Cauchy-Crofton-type result provides a direction-independent estimate with bounded error relative to the maximum projected area (typically within a factor of 2 for compact shapes), making it suitable for preliminary assessments in applications like drag estimation; the relative error decreases for more spherical-like forms. A concrete example is the analytical projection of a cube with side length a. Using the polyhedral decomposition, with face areas A_f = a^2 and normals along the axes \pm \mathbf{e}_i (for i=1,2,3), the projected area simplifies to A_p(\mathbf{u}) = a^2 (|\mathbf{u} \cdot \mathbf{e}_1| + |\mathbf{u} \cdot \mathbf{e}_2| + |\mathbf{u} \cdot \mathbf{e}_3|), since opposite faces pair to contribute a^2 |l_i| each (with l_i the direction cosines). At certain angles, such as along the body diagonal \mathbf{u} = (1,1,1)/\sqrt{3}, this yields a hexagonal silhouette with exact area a^2 \sqrt{3}, confirming the formula through direct summation of the six visible triangular projections from the faces.

Numerical Techniques

Numerical techniques enable the computation of projected areas for complex or irregular objects, where exact analytical solutions are infeasible due to geometry intricacies. These methods typically involve discretizing the object representation, such as meshes or point clouds, and applying algorithmic approximations to estimate the area on a . They balance computational efficiency with accuracy, often requiring trade-offs in or sampling . A prominent approach is ray tracing, which approximates the projected area by launching a large number of parallel rays from the projection direction toward a reference plane encompassing the object's bounding box. The fraction of rays that intersect the object, multiplied by the reference plane's area, yields the estimate; convergence improves with more samples, making it suitable for simulations in fields like . This method handles arbitrary geometries without explicit meshing but introduces variance that diminishes as the square root of the sample count. In (CAD) environments, mesh projection techniques project the vertices of a finite element surface onto the desired plane, then reconstruct and measure the polygon's area. For instance, employs a General Projection operator within its Module to compute projected areas via surface flux integration over discretized directions, effectively capturing the illuminated for imported CAD geometries. Similarly, NASA's software projects mesh-based aerospace models to generate outlines, supporting analyses like drag estimation. For opaque convex objects, convex hull algorithms provide an efficient deterministic solution: project the 3D vertices onto the 2D plane, compute the of these points using algorithms like or , and calculate the enclosed polygon's area via the . This yields exact results for convex shapes, as the projection remains , and is computationally lightweight with O(n log n) complexity. Handling self-occlusion in non-convex objects is critical to prevent overestimation from overlapping projections. resolves visibility by maintaining a for the , updating only closer surfaces during rasterization of the ; the final filled pixels delineate the true for area computation. Alternatively, depth sorting orders polygons by average depth before , excluding rear-facing or elements. These techniques, rooted in hidden surface removal, ensure accurate exclusion of occluded regions. Software implementations like exemplify practical trade-offs: projected area accuracy scales with mesh density, where coarser grids enable but introduce approximation errors, while finer resolutions—often with millions of elements—achieve sub-percent precision at higher computational cost. Such tools validate numerical results against analytical benchmarks for simpler geometries when possible.

References

  1. [1]
    Aerodynamic Drag - The Physics Hypertextbook
    (I would further simplify this by calling it the projected area .) Take the cross section of the object in the direction of its motion. This is the area of the ...
  2. [2]
    Projected Area - OpenVSP Ground School - NASA
    Dec 20, 2022 · The Projected Area analysis produces an area and outlined “silhouette” of a model in preset or custom directions and can also produce the projected area of ...Missing: physics aerodynamics
  3. [3]
    Bluff Body Flows – Introduction to Aerospace Flight Vehicles
    A_{\rm ref} is defined as the reference area, usually the projected frontal area. Using a sphere of diameter {d} as an example, its projected area is {A_ ...
  4. [4]
    Chapter 1. Introduction to Aerodynamics - Pressbooks at Virginia Tech
    Note that the planform area is not the actual surface area of the wing but is “projected area” or the area of the wing's shadow. Also note that some of the ...
  5. [5]
    Projected area and drag coefficient of high velocity irregular ...
    The drag coefficient is positively correlated with the projected area. Higher projected area means lower slenderness and the drag coefficient is found to ...
  6. [6]
    Projected Area and Projected Solid Angle - SPIE Digital Library
    Projected area A p is the rectilinear projection of a surface of any shape onto a plane. In differential form, its expression is dA p =cosθdA.
  7. [7]
    OpenVSP Projected Area - NASA
    Sep 30, 2025 · The Projected Area analysis produces an area and outlined “silhouette” of a model in preset or custom directions and can also produce the ...
  8. [8]
    The Drag Equation
    The drag equation states that drag D is equal to the drag coefficient Cd times the density r times half of the velocity V squared times the reference area A.
  9. [9]
    Projected Outline - an overview | ScienceDirect Topics
    The projected area diameter is defined as the diameter of a spherical particle with the same projected area as the particle viewed in a direction perpendicular ...
  10. [10]
    Searching for Simplicity The Evolution of Wind Provisions in ...
    May 15, 2008 · This article will provide an historical overview of the evolution of wind provisions in standards and codes in the United States.
  11. [11]
    [PDF] WIND LOADING
    This paper is presented to outline the features of the revised Wind. Loading Code NZS4203/303 and to give some background on the history of ... projected area.".
  12. [12]
  13. [13]
    [PDF] CS667 Lecture Notes: Radiometry - Cornell: Computer Science
    If we call the inclined surface area dA0 and continue to call the perpendicular surface area dA, then dA = dA0 cos θ. So the radiance can be written. L(x,ω) ...
  14. [14]
    [PDF] Radiometry - CS@Cornell
    Doing only the area integral gives intensity: We could invent a “projected area” measure in which to hide the cosine factor, but since this computation is not ...
  15. [15]
    [PDF] Fluid-Dynamic Drag - HVL
    The principles of aerodynamic drag, many detailed data, and some special sections in this book should, however, find interested readers in several other fields ...
  16. [16]
    Projected Areas of Cylinder and Cone
    The area is composed of the union of the ellipse projected by the base of the cone and a triangle from the apex of the cone to points tangent to this ellipse.
  17. [17]
    Drag Forces | Physics - Lumen Learning
    For larger objects (such as a baseball) moving at a velocity v in air, the drag force is given by F D = 1 2 C ρ A v 2 , where C is the drag coefficient (typical ...
  18. [18]
    The Drag Coefficient
    The drag coefficient Cd is equal to the drag D divided by the quantity: density r times half the velocity V squared times the reference area A. Cd = D / (A * .5 ...
  19. [19]
    Inclination Effects on Drag
    The angle between the chord line and the flight direction is called the angle of attack. Angle of attack has a large effect on the drag generated by an aircraft ...
  20. [20]
    [PDF] Fluid Dynamic Drag; Drag Analysis of a Fighter Airplane
    Historical survey of parasite drag coefficient (on wetted surface area) of airplanes. Note that the contribution of induced drag (not shown in the graph) ...Missing: adoption projected
  21. [21]
    ASCE 7-10 Wind Load Calculation Example | SkyCiv Engineering
    Apr 4, 2024 · Effective wind area = 26ft*(2ft) or 26ft*(26/3 ft) = 52 ft2 or 225.33 sq.ft. Effective wind area = 225.33 sq.ft. The positive and negative ( ...Missing: early | Show results with:early<|control11|><|separator|>
  22. [22]
  23. [23]
    Vickers Hardness Testing - Buehler - Metallography Equipment ...
    Nov 16, 2021 · It is possible to calculate the Vickers hardness based on the projected area of the impression, which can be measured by image analysis. While ...Missing: structural | Show results with:structural
  24. [24]
    [PDF] Wind pressure on a model of the Empire State Building
    ABSTRACT. Measurements have been made of the distribution of wind pressure over a model of the Empire State Building for the purpose of comparing the ...
  25. [25]
    Lighting Lecture 1 - Cornell University Ergonomics Web
    Cosine cubed law - E = I cos3theta / h^2 distance of source 'd' can be replaced by h/cos theta where h is perpendicular distance of source from plane in which ...
  26. [26]
  27. [27]
    Cosine Losses - SPIE Digital Library
    The effective collection area, or projected area, is calculated as AP=Ah⋅cos(θ) A P = A h · cos ( θ ) where Ah is the actual heliostat area.Missing: panel | Show results with:panel
  28. [28]
    power_flux_area.html - UNLV Physics
    P = F*A*cos(θ) ,. where θ is the inclination angle show in the figure. With with increasing θ, the power decreases monotontically: i.e., θ ↑, P ...
  29. [29]
    None
    ### Summary: Use of Projected Areas in View Factor Approximations for Thermal Radiation
  30. [30]
    Optimal Sun-tracking law for remote sensing satellites operating ...
    To derive an optimal tracking law, the projected solar array area at each instant is maximized, resulting in optimal electrical power generation.
  31. [31]
    [PDF] The Average Projected Area Theorem - arXiv
    Nov 11, 2012 · projected area to surface area might offer insight on ... with S denoting an integral over the solid's surface and AS the solid's surface.
  32. [32]
  33. [33]
    Ray Tracing: The Rest of Your Life
    Whereas, a Monte Carlo program will give a statistical estimate of an answer, and this estimate will get more and more accurate the longer you run it. Which ...
  34. [34]
    Computing Projected Area - COMSOL
    This model presents a way to compute projected area using the Surface-to-Surface Radiation interface of the Heat Transfer Module. By computing illuminated ...
  35. [35]
    [PDF] The Quickhull Algorithm for Convex Hulls
    The convex hull of a set of points is the smallest convex set that contains the points. This article presents a practical convex hull algorithm that ...
  36. [36]
    [PDF] Projections and Z-buffers - UT Computer Science
    The Z-buffer' or depth buffer algorithm [Catmull, 1974] is probably the simplest and most widely used of these techniques. Here is pseudocode for the Z-buffer ...Missing: mesh | Show results with:mesh