Fact-checked by Grok 2 weeks ago

Area rule

The area rule, also known as the transonic area rule or Whitcomb area rule, is an aerodynamic principle that minimizes on operating at and supersonic speeds by ensuring the cross-sectional area of the varies as smoothly and gradually as possible along its longitudinal axis, often resulting in a shape resembling a cigar or Coke bottle. This approach integrates components like wings and tail into the fuselage to avoid abrupt changes in area distribution, which would otherwise generate shock waves and increase drag near Mach 1. Developed in the early , the rule revolutionized high-speed by enabling significant drag reductions—up to 60% in tests—without increasing engine power. The concept was pioneered by American aerodynamicist Richard T. Whitcomb while working at the (NACA) , where he conducted extensive experiments using a slotted-throat tunnel modification to simulate conditions. Inspired by earlier ideas from German engineer and validated through tests on body-of-revolution models in 1951–1952, Whitcomb's breakthrough addressed the critical drag-rise problem that had stalled U.S. supersonic fighter programs. In 1952, he disclosed the principle to industry partners, leading to rapid redesigns such as the , which achieved supersonic capability after applying the rule to narrow its fuselage midsection. For this innovation, Whitcomb received the prestigious in 1954. The area rule's influence extends to nearly every and built since the , including fighters like the F8U Crusader and Grumman F11F Tiger, as well as bombers such as the . Its principles have also informed modern designs, including business jets and missiles, by providing a foundational method for optimizing equivalent body shapes in . While primarily associated with external , extensions of the rule have been explored for supersonic area distribution to further refine high-Mach performance.

Principles

Core concept

Wave drag represents a significant aerodynamic penalty encountered during high-speed flight, particularly in the transonic and supersonic regimes, where it arises from the formation of shock waves that disrupt the airflow around the aircraft. These shock waves occur when local airflow accelerates to or beyond the speed of sound, compressing the air abruptly and generating pressure discontinuities that contribute to a nonlinear increase in drag. Unlike parasitic or induced drag, wave drag is inviscid in origin but can exacerbate boundary layer separation, further amplifying total drag. The area rule is a fundamental design principle that minimizes by ensuring a smooth and gradual variation in the equivalent cross-sectional area along the aircraft's longitudinal axis, encompassing the , , and . This equivalent area accounts for the combined volume displacement of all components as if they were integrated into a single , aiming to approximate a distribution that yields near-zero rise. By reshaping the —often through waisting or narrowing at mid-—to compensate for the abrupt area increase caused by attachment, the overall profile mimics an ideal, low- shape such as a streamlined . A key metric influenced by area distribution is the , defined as the freestream Mach number at which begins to rise sharply, typically marked by a 20% increase in or a specific in the . Optimal application of the area rule delays this divergence, allowing higher speeds before significant drag penalties occur. The principle applies across and supersonic flight, though its implementation varies with flow regime.

Transonic regime

In the regime, where flight speeds approach 1, airflow over an exhibits mixed and supersonic characteristics, leading to the formation of local supersonic regions on the vehicle's surfaces due to acceleration over curved geometries such as wings and fuselages. These regions are typically terminated by shock waves, which interact with the to cause shock-induced separation, resulting in a rapid increase in drag known as drag divergence. This drag rise is primarily from effects, exacerbated by abrupt changes in the 's geometry that promote stronger shocks and flow unsteadiness. The area rule addresses this by ensuring a , variation in the longitudinal distribution of the equivalent cross-sectional area, treating the entire (fuselage, wings, and ) as a single body of revolution to minimize strength and delay the onset of divergence. Specifically, the cross-sectional area should increase from the nose to a maximum at the location of largest volume (often near the ) and then decrease symmetrically toward the , avoiding sudden expansions or contractions that would otherwise generate excessive near 1. A practical implementation of this principle is the "coke-bottle" or wasp-waisted shape, where the is indented laterally at the junction to compensate for the 's added cross-sectional area, creating an overall smoother area profile. Qualitatively, the area distribution plot for such a design resembles a smooth bell curve: starting near zero at the , rising monotonically through the forward fuselage and integration zone to a peak, and then falling monotonically to near zero at the tail, as opposed to the irregular peaks seen in non-area-ruled configurations. Wind tunnel tests of area-ruled wing-body combinations demonstrated significant aerodynamic benefits, with reductions in the zero-lift drag-rise coefficient of up to 57% at 1.0 and 31% at higher speeds, enabling to achieve higher performance without excessive penalties.

Supersonic regime

In the supersonic regime, where numbers exceed 1, airflow behaves according to linear theory, with disturbances propagating along characteristic cones rather than diffusing isotropically as in flow. This linearity simplifies analysis, revealing that —stemming from pressure discontinuities via shock waves and fans—dominates the total profile, often comprising over 70% of zero-lift for non-optimized configurations. Unlike skin friction or induced , is highly sensitive to the vehicle's longitudinal , making area distribution a critical design parameter. The supersonic area rule, an extension of linear aerodynamic principles, stipulates that is minimized when the cross-sectional area varies along the axis such that the second of area with respect to length is as small as possible, thereby attenuating among Mach waves emanating from disparate body components like wings, , and nacelles. In practice, this involves shaping the equivalent body of revolution—representing the vehicle's total area distribution—to follow smooth, often linear progressions, ensuring that area increments from added elements (e.g., pods) are counterbalanced by indentations to maintain overall uniformity. Non-uniform area distributions exacerbate drag by generating localized waves and Prandtl-Meyer expansion fans that intersect and amplify pressure gradients downstream, leading to higher and energy loss. For instance, a sudden area bulge can initiate a strong that reflects off adjacent surfaces, creating a cascade of disturbances; conversely, gradual area changes distribute these effects, weakening individual waves. Applying the supersonic area rule in cruise conditions yields substantial aerodynamic benefits, including lift-to-drag ratios improved by 15-25% through targeted reductions, as demonstrated in wind-tunnel models at 1.41 where optimized configurations exhibited 16-25% lower compared to unruled designs. These gains enhance by lowering requirements for sustained supersonic flight, enabling longer ranges without proportional increases in size or weight.

Theoretical foundation

Mathematical derivation

The mathematical derivation of the area rule originates from linearized theory for slender bodies in supersonic flow, where the goal is to minimize by optimizing the distribution of cross-sectional area A(x) along the body axis x. The governing equation is the linearized supersonic potential equation for the perturbation potential φ (with total potential Φ = U_∞ x + φ and |∇φ| ≪ U_∞): \beta^2 \frac{\partial^2 \phi}{\partial x^2} - \frac{\partial^2 \phi}{\partial y^2} - \frac{\partial^2 \phi}{\partial z^2} = 0, where β = √(M_∞² - 1) and M_∞ > 1 is the freestream . This hyperbolic admits solutions along characteristics at the Mach angle μ = arcsin(1/M_∞). The Prandtl-Glauert transformation scales the transverse coordinates as y' = y/β and z' = z/β, mapping the problem to an equivalent incompressible Laplace ∇'² φ = 0 in the transformed space, facilitating solution via standard methods like source distributions. For a slender axisymmetric body (|dA/dx| ≪ 1), the body is modeled as a line distribution of sources along the x-axis with strength per unit length σ(x) = U_∞ dA/dx, satisfying the kinematic boundary condition that the radial velocity matches the body slope: ∂φ/∂r ≈ U_∞ dr/dx at the body surface r = r(x), approximated on-axis for slenderness (r ≪ l, where l is body length). The velocity is u = ∂φ/∂x, v = ∂φ/∂y, w = ∂φ/∂z, and the linearized is C_p = -2 (∂φ/∂x)/U_∞. The far-field pressure integration or momentum flux through a control surface (e.g., Trefftz plane at x → ∞) yields the wave drag. At zero lift, the volume-induced wave drag D_v for the body is given by the double integral form from slender body theory: D_v = -\frac{\rho_\infty U_\infty^2}{4\pi} \int_0^l \int_0^l \frac{dA}{dx_1} \frac{dA}{dx_2} \ln |x_1 - x_2| \, dx_1 \, dx_2, where ρ_∞ is freestream density (non-dimensionalization implied for consistency); this arises from evaluating the axial momentum deficit due to the source-induced pressure perturbations, with boundary conditions ensuring no flow disturbance ahead of the (causality via Mach cones) and decay at infinity. For lifting cases, an analogous drag due to lift D_L term appears as D_L ∝ ∫ (dC_L/dx)^2 dx (with C_L(x) the local distribution), but the area rule treats the total effective cross-sectional area (combining volume and lift contributions) to minimize the combined . The area rule criterion for minimum drag follows from minimizing the quadratic form of the double integral subject to fixed volume V = ∫_0^l A(x) dx and length l. A simple approximation using a single integral ∫ (dA/dx)^2 dx (valid in certain limits) and calculus of variations with Lagrange multiplier λ yields the Euler-Lagrange equation d²A/dx² = λ (constant), implying quadratic area variation for the parabolic body, though the full supersonic theory with wave propagation effects yields the optimal Sears–Haack distribution. In the transonic regime (M_∞ ≈ 1), the derivation employs the transonic small-disturbance (TSD) approximation to the full potential equation, retaining nonlinearity essential for mixed sub/supersonic flow: \left( 1 - M_\infty^2 - \frac{\gamma + 1}{U_\infty} M_\infty^2 \frac{\partial \phi}{\partial x} \right) \frac{\partial^2 \phi}{\partial x^2} + \frac{\partial^2 \phi}{\partial y^2} + \frac{\partial^2 \phi}{\partial z^2} = 0, with boundary conditions similar to the supersonic case but now elliptic-parabolic, capturing shock formation via entropy jumps. For slender bodies, axial dominance reduces this to a quasi-one-dimensional form where streamtube area changes A_eff(x) (effective including shocks) determine drag rise via choking (minimum area where local M=1), analogous to isentropic duct flow; configurations with identical A(x) produce equivalent streamtube distortions and momentum losses, yielding the transonic area rule. As the transonic similarity parameter τ = (1 - M_∞²) (l / \bar{c})^{2/3} → ∞ (approaching supersonic), the TSD nonlinearity weakens, and the equation asymptotes to the linearized supersonic form, recovering the drag integral. For a simple conical geometry (A(x) = π (r_b x / l)^2, 0 ≤ x ≤ l, with base area A_b = π r_b²), dA/dx = 2 A_b x / l², and the double integral evaluates to a value ~18–25% higher than the optimal Sears–Haack for typical fineness ratios. For a cylindrical body with hemispherical nose (step change in A at x = r_n, A(x) = π r_n² for x > r_n), the abrupt dA/dx produces localized shocks, inflating ; smoothing via linear A ramp reduces it proportionally to the gradient squaring.

Sears–Haack body

The is a theoretical body of revolution that minimizes in supersonic flow for a specified l and enclosed V. Its geometry features a smooth, symmetric pointed at both ends, with the cross-sectional area distribution given by A(x) = \frac{16 V}{3 \pi l} \left[ 1 - \left( 2 \frac{x}{l} - 1 \right)^2 \right]^{3/2}, where x ranges from 0 to l. This distribution ensures a bulbous midsection, with maximum area at x = l/2, tapering to zero at the ends. The shape derives from supersonic slender-body theory, which models the flow as a distribution of line sources along the axis. The perturbation velocity potential is expanded in a Fourier sine series using the angular coordinate \theta, where the axial position transforms as x = (l/2)(1 - \cos \theta). The volume V fixes the first Fourier coefficient, while the wave drag coefficient is proportional to the sum of the squares of all coefficients. Minimization subject to the volume constraint requires setting all higher-order coefficients to zero, yielding the area distribution A(\theta) = (4 V / \pi l) (\sin \theta - (1/3) \sin^3 \theta), equivalent to the Cartesian form above. This eliminates contributions from nonlinear terms in the drag integral, resulting in the absolute minimum wave drag of D/q = \frac{128 V^2}{\pi l^4}—independent of Mach number in the slender approximation. Compared to other shapes with the same volume and length, the Sears–Haack body exhibits lower wave drag because non-optimal distributions activate higher Fourier modes, increasing the drag integral. For a parabolic body, where A(x) \propto x (l - x), the wave drag is approximately 25% higher, with D/q = \frac{5 \pi^2}{16} (V^2 / l^4). A conical body, with linear radius growth and A(x) \propto x^2, incurs even greater drag, typically 18–25% more than the Sears–Haack for fineness ratios around 10, as its abrupt area changes produce stronger shocks. These differences arise directly from the second derivative of the area distribution in the approximate drag formula C_{D_w} = \frac{1}{2 \beta^2} \int_0^l (dA/dx)^2 dx / S_{\max}, where \beta = \sqrt{M^2 - 1}; the Sears–Haack smooths this derivative optimally. This ideal shape assumes inviscid, irrotational flow and a slender (low volume-to-length ratio, small disturbances relative to ), neglecting viscous effects like layers and skin friction, as well as finite thickness that introduces nonlinear shock interactions. In practice, these assumptions lead to deviations, with real bodies showing higher drag due to (adding 10–20% in typical tests) and non-slender effects amplifying at lower fineness ratios.

Historical development

Pre-World War II research

Early research on high-speed in during laid foundational concepts for managing through careful shaping of aircraft and body geometries, with Ludwig Prandtl's contributions providing key theoretical precursors. Prandtl, director of the Kaiser Wilhelm Institute for Fluid Mechanics, advanced understanding of pressure in streamlined bodies, emphasizing how adverse pressure gradients could lead to and increased , particularly relevant for slender configurations at elevated speeds. His work on boundary layers and early supersonic flow theories influenced subsequent studies linking body area variations to wave and pressure components. A pivotal advancement came from , who studied under Prandtl and worked at the Deutsche Versuchsanstalt für Luftfahrt (DVL) in . In , Busemann presented a seminal paper at the Fifth Volta Congress in , , introducing swept-wing theory to mitigate drag rise in the transonic and supersonic regimes by reducing the component of airflow normal to the wing . This concept, while focused on wings, foreshadowed broader ideas in area distribution for overall vehicle drag reduction, as it highlighted the importance of geometric alignment with local flow velocities to minimize formation and pressure disturbances. Experimental validation emerged through wind tunnel tests at German facilities, including those at the DVL and Luftfahrtforschungsanstalt (LFA), which demonstrated drag penalties from abrupt changes in cross-sectional area during high-subsonic flows. In the early , engineers Heinrich Hertel and Otto Frenzl at aircraft works in conducted wind tunnel experiments on various body shapes, observing that sudden area expansions or contractions caused significant increases in due to localized shock waves and flow disruptions. These tests, using models with varying contours, provided quantitative data showing drag rises in conditions from non-smooth area profiles, informing early efforts to optimize body streamlining. Parallel developments in rocketry further shaped ideas on area distribution, as engineers tested designs for minimal drag at high speeds. The Aggregat-4 (, later V-2) program, initiated in the mid-1930s under , involved extensive evaluations at facilities like the DVL to refine the rocket's cylindrical body with tapered nose and fins, ensuring streamlined shaping to reduce drag and pressure losses. These experiments, conducted from 1936 onward, highlighted aerodynamic optimization for high-speed vehicles.

Post-war advancements

Following , significant advancements in the area rule occurred through empirical research at the (NACA), particularly under Richard T. Whitcomb at . In 1952, Whitcomb conducted tests on wing-body models in the 8-Foot High-Speed Tunnel, which featured a slotted-throat modification for improved flow simulation. These experiments demonstrated that abrupt variations in the aircraft's cross-sectional area distribution caused sharp drag rises near Mach 1 due to formation, while gradual, smooth distributions substantially mitigated this effect, achieving approximately 60% reduction in drag-rise increments. This breakthrough, termed the Whitcomb area rule, built upon earlier theoretical ideas but was validated through these practical tests, marking a pivotal refinement. Throughout the , NACA pursued further experiments to quantify and optimize the area rule's benefits, transitioning from basic models to more complex configurations. Tests on various wing-fuselage combinations, including slender bodies of revolution like the RM-10 model evaluated across multiple facilities, confirmed consistent reductions of 35-60% in the regime, depending on the geometry adjustments. These efforts, supported by declassified wartime reports on swept-wing and area distribution concepts, accelerated U.S. adoption and integration into high-speed research. The decade's work emphasized the rule's applicability to real-world designs, shifting focus from theoretical speculation to measurable performance gains. A key milestone came with the 1956 publication of NACA Report 1273 by Whitcomb, titled "A Study of the Zero-Lift Drag-Rise Characteristics of Wing-Body Combinations Near the Speed of Sound." This report formalized the area rule by presenting detailed experimental data from transonic tests, deriving empirical correlations between area distribution and zero-lift drag, and establishing guidelines for minimizing wave drag without excessive structural penalties. The findings solidified the rule's foundational role in post-war aeronautics. The Whitcomb area rule's influence extended internationally in the late 1950s, with researchers in the United Kingdom and Soviet Union conducting follow-up studies inspired by NACA publications and shared declassified German insights. In the UK, aerodynamicists at the Royal Aircraft Establishment explored analogous area-matching techniques for supersonic configurations, while Soviet programs incorporated similar principles in high-speed vehicle development to address transonic bottlenecks.

Practical applications

Early implementations

The represented one of the earliest practical applications of the area rule in aircraft , undergoing a major redesign in 1953 to incorporate fuselage waisting that smoothed the distribution of cross-sectional area along the aircraft's length. This modification addressed the original prototype's inability to achieve supersonic speeds in level flight, as predicted by tests conducted at NASA's from May to October 1953, transforming a subsonic interceptor into a capable supersonic platform. The resulting YF-102A prototype, built in 118 days, first flew on December 20, 1954, and successfully exceeded 1 in both level flight and climb the following day, attaining a top speed of 1.24 and improving overall performance by approximately 25% compared to the initial . Integrating the area rule into these existing designs posed notable challenges, particularly the need to retrofit "wasp-waisted" fuselages onto airframes already in production or prototyping, which required halting F-102 assembly lines and incurring substantial costs and delays. Engineers faced difficulties balancing the area rule's requirements with structural integrity, engine placement, and weapon systems, often necessitating iterative modifications during development. Additionally, while predictions at facilities like Langley's 8-Foot High-Speed Tunnel accurately foresaw the F-102's initial supersonic shortfalls—confirmed by 1954 flight tests—minor discrepancies arose in real-world conditions due to factors like effects and model scaling limitations, prompting further refinements. These early implementations ultimately enabled the deployment of the first U.S. supersonic interceptors, with USAF trials in 1954-1955 validating the area rule's effectiveness through performance data showing sustained 1+ speeds and reduced drag penalties. The F-102A entered operational service in 1956 as a key interceptor against potential threats, paving the way for broader adoption in subsequent designs and demonstrating the rule's critical role in overcoming barriers.

Notable aircraft designs

The exemplified extreme application of the area rule, with its razor-thin and small, clipped wings designed to minimize cross-sectional area variations for and supersonic flight, achieving a top speed of Mach 2.0. This design, influenced by NACA research, allowed the aircraft to break world speed records shortly after its 1956 debut, with the area-ruled reducing by integrating the wing and body seamlessly. The , the first operational supersonic bomber introduced in 1956, incorporated a full area distribution strategy across its slender "wasp-waist" to enable sustained flight over long ranges. This approach, tested extensively in NACA wind tunnels, optimized the equivalent cross-sectional area for supersonic cruise, contributing to the aircraft's ability to set 19 world speed and altitude records between 1960 and 1970. The design's success in drag reduction allowed for a maximum speed of 1,325 mph at high altitudes, though it required an external pod for fuel and weapons to maintain the smooth area profile. Similarly, the benefited from area ruling in its redesign from the F-102, featuring a "coke-bottle" that smoothed the cross-sectional area distribution and reduced , enabling the interceptor to reach 2.3. This modification, applied in the late 1950s, improved the aircraft's lift-to-drag (L/D) ratio during supersonic dash, supporting its role as a high-speed all-weather defender with a service ceiling exceeding 57,000 feet. In civilian applications, Boeing's 2707 concepts from the integrated the supersonic area rule to enhance aerodynamic efficiency, employing a variable-geometry and "coke-bottled" to achieve an estimated supersonic L/D ratio of approximately 7.6 at 2.7 cruise. These designs aimed for ranges over 4,000 miles but were ultimately canceled in 1971 due to economic factors, though the area rule contributed to projected drag reductions of up to 20% in compared to non-ruled configurations. The Anglo-French , entering service in 1976, applied area ruling principles in its and wing-body blending to minimize supersonic during cruise at 2.04, resulting in notable drag reductions that supported efficient high-altitude operations with an L/D ratio of approximately 7.5. Early design methods combined the area rule with slender-body theory, refining the rear lines to further cut drag by optimizing cross-sectional equivalence, which helped achieve transatlantic crossings in under four hours.

Extensions and limitations

Modern adaptations

The area rule has been extended to supersonic regimes through the supersonic area rule, developed by Robert T. Jones, which adjusts cross-sectional areas based on cone angles to minimize at higher speeds. Further extensions apply similar principles to hypersonic flows for bluff bodies, ensuring equivalent drag through matching cross-sectional area changes in non-axisymmetric shapes. This hypersonic formulation, explored since the , uses theoretical models to refine body shapes under high- conditions. In applications, the area rule has been applied to enhance performance, particularly in afterbody configurations where drag penalties from abrupt area changes are pronounced. Wind-tunnel validations confirm that area-ruled shapes reduce drag by smoothing equivalent-body-of-revolution profiles, allowing efficient acceleration through without excessive thrust requirements. This approach has informed designs for cruise and tactical s, prioritizing low-observable and high-speed in operational envelopes spanning to supersonic speeds. For high-speed unmanned aerial vehicles (UAVs) operating near speeds, area ruling can contribute to by mitigating rises, though most UAVs remain and do not require it. In space vehicle extensions, reentry bodies leverage –Haack approximations for minimization, as seen in configurations balancing lift and for controlled atmospheric interface. Recent advancements as of 2025 incorporate area rule principles into low-boom supersonic designs, such as NASA's X-59 QueSST aircraft, which uses fuselage shaping to reduce while minimizing perception to 75 perceived level decibels. The X-59, which completed its first flight in October 2025, demonstrates integrated aerodynamic optimization for quiet supersonic overland flight. Euler-based CFD analyses revisit Whitcomb's rule for modern aircraft, revealing nuanced shock interactions that guide conceptual designs with partial reductions near 1, while highlighting limitations above 1.05.

Constraints and criticisms

While the area rule effectively reduces in and supersonic regimes, its implementation introduces significant aerodynamic trade-offs, particularly in balancing drag minimization with , surface integration, and internal requirements. The characteristic "coke-bottle" waist, required to achieve smooth cross-sectional area distribution, often narrows the fuselage at the junction, which can shift the center of gravity unfavorably and compromise without compensatory tail or designs. Additionally, this waist restricts space for control surfaces like vertical stabilizers and reduces internal , limiting accommodations, fuel capacity, or payload in designs such as early delta- fighters, where engineers had to prioritize drag reduction over . Viscous and nonlinear flow effects further limit the rule's accuracy, especially at high angles of attack or in configurations with thick s, where inviscid linear theory assumptions break down. At elevated angles of attack, separation and viscous interactions dominate, causing the area rule-based predictions to deviate from idealized slender-body assumptions. (CFD) analyses confirm these discrepancies, showing that equivalent-body methods neglect viscous contributions to total , leading to non-conservative estimates in off-design conditions like maneuvering flight. Critics argue that the area rule overemphasizes reduction at the expense of other components, such as base drag from afterbodies or between lifting surfaces and fuselages, which can negate projected benefits in integrated vehicle designs. Its utility is also limited in regimes, where effects are minimal, and in highly maneuverable requiring large angles of attack, as the rule assumes steady, attached flows and slender geometries that do not align with agile requirements. Post-2000 CFD validations have highlighted the area rule's diminished role in optimized hypersonic designs, where real-gas effects, strong detached shocks, and dominate over linear wave-drag considerations. Studies using Euler and Navier-Stokes solvers demonstrate that while the rule provides qualitative guidance for initial shaping, quantitative predictions falter beyond , with modern optimization tools favoring multidisciplinary approaches over standalone area distribution. In supersonic transport (SST) contexts, environmental constraints exacerbate these limitations, as area-ruled fuselages optimized for low drag often elevate cruise altitudes into the , amplifying and climate forcing from and emissions—potentially 10-20 times greater per seat-kilometer than equivalents. Regulatory limits further restrict viable shapes, rendering pure area rule adherence impractical without hybrid low-boom adaptations.

References

  1. [1]
    The Whitcomb Area Rule: NACA Aerodynamics Research ... - NASA
    The area rule revolutionized how engineers looked at high-speed drag and impacted the design of virtually every transonic and supersonic aircraft ever built.
  2. [2]
    Dr. Richard T. Whitcomb | National Air and Space Museum
    In 1952 Whitcomb solved the problem with a design to reduce drag and increase speed without the addition of power. His idea was labeled the 'area rule'. His ...
  3. [3]
    [PDF] 5. Drag: An Introduction
    Jan 22, 1997 · The principle that aerodynamicists use to achieve low values of wave drag is known as the area rule. Proposed by Richard Whitcomb* at the NACA's ...
  4. [4]
    Bluff Body Flows – Introduction to Aerospace Flight Vehicles
    The area rule states that the cross-sectional area should change smoothly and continuously, rather than abruptly, to minimize drag on an airplane; that is, the ...
  5. [5]
    [PDF] Wall-Resolved Large Eddy Simulations of Transonic Shock-Induced ...
    Jul 10, 2017 · The transonic flow condition gives rise to the formation of a shock in the aft region of the bump. The interaction of this shock with the ...
  6. [6]
    A Study of the Zero-Lift Drag-Rise Characteristics of Wing-Body ...
    A Study of the Zero-Lift Drag-Rise Characteristics of Wing-Body Combinations Near the Speed of Sound Comparisons have been made of the shock phenomena and ...
  7. [7]
    [PDF] RESEARCH MEMORANDUM
    The partial transonic-area-rule modifications resulted in a reduction of minimum-drag-coefficient rise of 57 percent at a Mach number of 1.0 and 31 percent ...<|control11|><|separator|>
  8. [8]
    [PDF] SUPERSONIC-AREA-RULE DESIGN AND ROCKET ... - DTIC
    The difference is brought out to show that average-area comparisons cannot be used for predicting the wave drag at supersonic speeds. ... Whitcomb, Richard ...
  9. [9]
    [PDF] of tfe supersonic area rule to tbe
    This paper presents certain considerations and techniqpes pertinent to the application of the supersonic area rule to the desigh of aiPplane fuselages.Missing: original | Show results with:original
  10. [10]
    [PDF] Supersonic Potential Flow
    The solutions for supersonic potential flow around slender bodies have already been covered in section. (Bok) but some features are valuable to emphasize in ...
  11. [11]
    [PDF] 10. Supersonic Aerodynamics - Virginia Tech
    Jul 31, 2016 · This equation is used to analyze different cross-sectional area distributions and also to find shapes with minimum drag. Since the equation is ...
  12. [12]
    None
    ### Summary of Mathematical Derivation of Supersonic Area Rule
  13. [13]
    [PDF] 3. Drag: An Introduction - Virginia Tech
    Jan 31, 2019 · The integral over II as x → ∞, corresponds to the so-called Trefftz Plane. The integral over III is the wave drag integral, which is zero for ...Missing: (dA/ | Show results with:(dA/
  14. [14]
    [PDF] chapter 6 supersonic aerodynamics - DTIC
    The general solution of a compressible flow problem consists of finding three unknown velocity components and three density and pressure changes with respect to ...
  15. [15]
    [PDF] Revisiting the Transonic Area Rule for Conceptual Aerodynamic ...
    Whitcomb, in his famous NACA Report [4] applied the. Transonic Area Rule to the four basic geometries shown in FIGURE 23. In order from (a) – (d) in the figure, ...<|separator|>
  16. [16]
    [PDF] Bodies having Minimum Pressure Drag in Supersonic Flow
    Aug 19, 2004 · We know that the minimum drag shape is a flat plate in two-dimensional flow ... profiles, discovered independently by Sears(1947) and Haack(1947).
  17. [17]
    [PDF] A Theoretical Trade-Off Between Wave Drag and Sonic Boom ...
    Dec 15, 2022 · In 1956, Whitcomb published a paper on an idea that became known as the “supersonic area rule”. He discovered wave drag can be reduced by ...
  18. [18]
    [PDF] Studies of optimum body shapes at hypersonic speeds
    The present study was directed toward the questions that arise in the application of optimum bodies to the design of hypersonic cruise aircraft.Missing: original | Show results with:original
  19. [19]
    [PDF] Ludwig Prandtl's Boundary Layer
    Indeed, for slender aerodynamic shapes, most of the drag is due to skin friction. Another marked result according to Prandtl is flow separation:3. In given ...
  20. [20]
    [PDF] Reexamining Busemann's Wing Sweep Theory
    IN 1935, Adolf Busemann presented a paper at the Fifth Volta Congress that revolutionized the theory of swept wings at high speeds.[1] 1935 was an ...
  21. [21]
    [PDF] 2-1 historical review of missile aerodynamic developments
    In the 16th century, a German, Conrad Haas, made sketches of multistage rockets in which the first stage consumed itself so that separa- tion of the stages was ...
  22. [22]
    Supersonic Revolution - HistoryNet
    Oct 10, 2017 · Jones, a gifted aerodynamicist at the Langley laboratory, had independently postulated the concept of the sharply angled delta and swept wing ...
  23. [23]
    Richard Whitcomb's Triple Play | Air & Space Forces Magazine
    Feb 1, 2010 · Whitcomb received the 1954 Collier Trophy “for discovery and experimental verification of the Area Rule, a contribution to base knowledge ...
  24. [24]
    [PDF] The Cutting Edge: A Half Century of U.S. Fighter Aircraft R&D - DTIC
    24The area-rule concept predicts drag at transonic and supersonic speeds based on the relationship of fuselage cross section and surface area. NACA ...
  25. [25]
    26 February 1955 | This Day in Aviation
    Feb 26, 2025 · The fuselage incorporated the “area rule,” a narrowing in the fuselage width at the wings to increase transonic performance, similar to the ...
  26. [26]
    [PDF] Orders of Magnitude - NASA
    Essentially, the area rule postulated that the cross~section of an airplane ... design of the Lockheed F~104 Starfighter. The NACA kept involved throughout.
  27. [27]
    NASA Armstrong's Milestones
    ... area rule principle" developed at the NACA by Richard Whitcomb. Credits ... Lockheed F-104 Starfighter. Credits: NASA. Jan. 31, 1958: Explorer 1, the ...
  28. [28]
    Convair B-58A Hustler - Air Force Museum
    Hustlers flew in the Strategic Air Command between 1960 and 1970. Setting 19 world speed and altitude records, B-58s also won five different aviation trophies.
  29. [29]
    10-by 10 Wind Tunnel Performed Critical Analysis for the First ...
    Sep 24, 2020 · In September 1956, the NACA's Lewis Flight Propulsion Laboratory began a critical test for the world's first supersonic bomber, the Convair B-58 Hustler.
  30. [30]
    History | F-106 Delta Dart
    ... area ruling" (Area Rule) in the early 1950's. Area Rule reduces drag at transonic speeds and is reflected in the "coke bottle" or "wasp waist" shaped ...Missing: trials outcomes
  31. [31]
    Convair F-106: The Ultimate Interceptor - HistoryNet
    May 11, 2018 · During the Cold War years, Convair's delta-wing F-106A was the fastest and most lethal all-weather interceptor in the U.S. Air Force inventory.Missing: 2.3 | Show results with:2.3<|control11|><|separator|>
  32. [32]
    [PDF] Reassessing the B2707-100 Supersonic Transport Aircraft
    the early 1960's. [10] It is based on the concepts of Whitcomb's supersonic area rule and estimates the drag associated with formation of shock waves around ...
  33. [33]
    Concorde and the Future of Supersonic Transport - AIAA ARC
    Longitudinal trim was carefully considered to minimize cruise trim drag. This involved suitable design of wing camber and twist and the addition of the fuel ...<|separator|>
  34. [34]
    [PDF] CONSTRAINED AEROTHERMODYNAMIC DESIGN OF ...
    This hypersonic design rule was validated in 2-D using a standard hypersonic. CFD test geometry for an elliptically shaped body. As mentioned, the CFD code.
  35. [35]
    Initial Sizing and Reentry-Trajectory Design Methodologies for Dual ...
    The empirical term EWD is the ratio of the actual wave drag to the minimum possible wave drag (Sears–Haack) and relates to the wave drag efficiency of the ...
  36. [36]
    [PDF] Hypersonic Area Rule, - DTIC
    Nov 29, 1977 · formulates the hypersonic area rule. According to this rule, during flow about thin nonaxisymmetrical bluff bodies with equal values of.
  37. [37]
    Limitations of the Sonic Area Rule - Sci-Hub
    It is very important to know what limitations should be placed on the applicability of the sonic area rule, otherwise ineffective designs and incorrect drag.
  38. [38]
    [PDF] Global Environmental Impact of Supersonic Cruise Aircraft in the ...
    We analyze the climate and ozone impacts of commercial supersonic aircraft using state-of-the-science modeling capabilities ranging from plume to global scale.Missing: rule | Show results with:rule
  39. [39]
    [PDF] Environmental limits on supersonic aircraft in 2035
    Jan 1, 2022 · This study investigates the market potential and atmospheric impacts of two potential supersonic transport (SST) aircraft: a small jet (“Small ...