Fact-checked by Grok 2 weeks ago

Parasitic drag

Parasitic drag, also known as parasite drag, is the component of total aerodynamic drag on an or other body moving through a that is unrelated to the generation of , arising instead from the viscous interaction between the fluid and the body's surface as well as differences caused by its . It constitutes a major portion of the drag at higher speeds and includes three primary subcomponents: , form drag, and interference drag. Unlike induced drag, which results from the creation of and , parasitic drag affects every part of the vehicle and increases proportionally with the square of the , making it a critical factor in determining maximum achievable speeds and . Skin friction drag originates from the shear stress exerted by the air's on the vehicle's wetted surfaces, where a of slower-moving air forms, with the magnitude depending on , , and whether the is laminar or turbulent. Form drag, sometimes called or profile drag, results from the separation of around non-streamlined shapes, creating low-pressure wakes behind protrusions like antennas, , or engine cowlings, and is minimized through aerodynamic shaping. Interference drag emerges at junctions where components meet, such as the wing-fuselage , where conflicting patterns generate additional and drag greater than the sum of individual components. These elements are quantified in the C_{D_0}, often estimated using methods like the equivalent flat-plate area approach, where C_{D_0} \approx C_F \cdot (S_{wet}/S_{ref}) \cdot FF, with C_F as the , S_{wet} the wetted area, S_{ref} the reference area, and FF a . In aircraft design, reducing parasitic drag is paramount for enhancing performance, as even small reductions—such as a 0.0001 increase in C_D on the —could necessitate removing passengers to maintain range, while redesigns like the F-102's area ruling cut drag by 25 counts, boosting top speed by 10%. Techniques to mitigate it include streamlining fuselages, using flush rivets and smooth finishes for skin friction, fairings at junctions for , and retractable gear or vortex generators to prevent . Parasitic drag buildup analysis, often performed early in the design process using tools like component summation or , directly influences vehicle sizing, requirements, and overall operational costs, underscoring its role in balancing aerodynamic with structural and mission constraints.

Fundamentals

Definition

Parasitic drag, also known as parasite drag, is the component of total aerodynamic that results from the physical presence and shape of an object moving through a , independent of any generation. It arises primarily from differences across the object's surfaces and stresses due to along those surfaces. This acts opposite to the direction of motion and affects all objects in flow, whether they produce or not. The concept of parasitic drag developed in early 20th-century as researchers sought to understand and minimize on and other vehicles. A foundational contribution came from Ludwig Prandtl's theory, introduced in 1904, which described the thin layer of fluid near the object's surface where viscous effects dominate and contribute significantly to . In physical terms, parasitic drag is quantified as a force, typically measured in Newtons in the (SI). Its magnitude is proportional to the of the fluid (one-half the fluid density times the square of the velocity), the object's reference area (such as a projected frontal area or wetted surface area), and a dimensionless parasitic drag coefficient that encapsulates the object's geometry and surface properties. For lifting bodies like wings, total is the sum of parasitic drag and induced drag, the latter arising from the generation of . Representative examples of purely parasitic drag occur on non-lifting bodies, such as a moving through or , where the drag stems entirely from the object's form and surface interaction with the , or a flat plate aligned parallel to the flow at zero , dominated by viscous along its surface.

Relation to Total Drag

In , the total drag force D_{\text{total}} acting on an is the sum of parasitic drag D_{\text{parasitic}} and induced drag D_{\text{induced}}, where induced drag results from the generation of through and . This decomposition is expressed in terms of drag coefficients as C_D = C_{D_0} + C_{D_i}, with C_{D_0} representing the associated with parasitic drag. Parasitic drag, often termed zero-lift drag, is quantified at zero and embodies the inherent resistance due to the vehicle's and surface interactions with the , independent of production. The proportional contributions of parasitic and induced drag to total drag vary significantly with operating conditions. Induced drag predominates at low speeds and high lift coefficients, such as during , where the aircraft requires substantial relative to its speed, leading to stronger vortices and . Conversely, parasitic drag becomes dominant at higher speeds and lower lift coefficients, as seen in flight, because parasitic drag scales with the square of while induced drag decreases inversely with velocity squared. For a typical commercial in cruise, parasitic drag constitutes approximately 60% of total drag, with induced drag accounting for the remaining 40%. This relationship has key practical implications for vehicle design and performance, particularly in and automotive contexts. Reducing parasitic drag enhances by lowering the thrust required to overcome ; for instance, in long-range commercial jets, even small reductions in parasitic drag can yield substantial fuel savings over extended flights due to its large share in cruise conditions. In ground vehicles like automobiles, aerodynamic drag—equivalent to parasitic drag—can represent up to 50% of total road load at highway speeds, making its minimization essential for improving overall and reducing consumption.

Components

Form Drag

Form drag, also known as pressure drag, arises from the unbalanced pressure distribution around an object in fluid flow, primarily due to the effects of viscosity that prevent complete pressure recovery on the rearward-facing surfaces. In viscous flows, the boundary layer thickens and can separate from the surface, leading to the formation of a low-pressure wake behind the body; this separation creates a region of turbulent eddies where the static pressure remains lower than on the forward-facing surfaces, resulting in a net rearward force. Unlike inviscid flow assumptions where pressure forces would cancel out symmetrically, the viscous-induced separation disrupts this balance, making form drag the dominant contributor to total resistance for many shapes. Key factors influencing form drag include the object's , particularly the distinction between blunt and streamlined shapes. Blunt or bodies, such as spheres or cylinders, promote early due to adverse gradients, forming large separation bubbles and expansive low-pressure wakes that amplify the pressure imbalance. In contrast, streamlined shapes like airfoils are designed to delay separation by gradually decelerating the , minimizing wake size and allowing better recovery, thereby reducing form drag significantly. The role of separation bubbles is critical, as they act as localized regions where the detaches and reattaches, but in bodies, persistent separation leads to persistent low-pressure zones that sustain high drag levels. Representative examples illustrate the impact of shape on form . For a bluff body like a at moderate Reynolds numbers, the is approximately 0.47, with nearly all stemming from differences and wake formation. Conversely, a streamlined exhibits a much lower of about 0.01 to 0.05 at low angles of attack, where flow remains attached over most of the surface, limiting the wake and to a minor fraction of total resistance. Historically, early quantification of form drag on various shapes was advanced through experiments conducted by in the 1910s. Using his at the base of the , operational from 1909 to 1912, Eiffel systematically tested models to measure forces, demonstrating how shape influences distribution and separation, which laid foundational for aerodynamic design. These tests, involving thousands of runs, highlighted the dramatic reduction in form achievable with streamlined profiles compared to blunt ones.

Skin Friction Drag

Skin friction drag arises from the viscous shear stresses acting tangentially on the surface of an object moving through a fluid, resulting from the momentum transfer within the velocity gradient of the boundary layer adjacent to the surface. This drag component occurs without flow separation and is directly tied to the fluid's viscosity, where the no-slip condition at the wall causes the fluid velocity to vary from zero at the surface to the free-stream velocity farther away. The boundary layer, a thin region near the surface where viscous effects dominate, is the site of this shear stress, with the wall shear stress \tau_w defined as \tau_w = \mu \left( \frac{\partial u}{\partial y} \right)_{y=0}, where \mu is the dynamic viscosity and u is the velocity component parallel to the surface. The nature of the boundary layer significantly influences skin friction drag, with two primary types: laminar and turbulent. In a laminar boundary layer, flow is smooth and orderly, producing lower shear stresses and thus reduced friction drag, typically occurring at Reynolds numbers \mathrm{Re} < 5 \times 10^5 based on the distance from the leading edge. Conversely, a turbulent boundary layer features chaotic mixing and eddies, leading to higher shear stresses and increased friction drag—often 3 to 5 times that of a laminar layer at comparable conditions—which predominates at \mathrm{Re} > 5 \times 10^5. The transition between these states depends on factors like surface roughness, pressure gradients, and free-stream turbulence, and maintaining laminar flow can substantially lower overall drag. Skin friction drag is particularly dominant on flat plates, where it constitutes the entirety of the drag force, and on streamlined bodies such as modern fuselages, where it constitutes the majority of the component's total due to minimal . On a typical commercial in , skin contributes approximately 50% of the overall , with the being a major wetted surface contributor. Ludwig Prandtl's foundational work in the , particularly his theory introduced in 1904 and applied industrially in the , provided the equations for analyzing these effects, including the skin C_f = \frac{\tau_w}{0.5 \rho V^2}, where \rho is fluid and V is free-stream . This non-dimensional parameter enables prediction of drag from solutions, revolutionizing aerodynamic design by isolating viscous effects near surfaces.

Interference Drag

Interference drag arises from the aerodynamic interactions between different components of the , where streams from one part (e.g., and ) mix and create additional , vortices, or flow disruptions that increase beyond the simple sum of the individual parts. This type of drag is particularly prominent at junctions such as the wing-root- , attachments, or struts, where differing flow directions and speeds conflict, leading to localized separation or eddy formation. In aircraft , interference drag can contribute 5% to 10% of the total in conventional configurations, though poor can increase this significantly, up to 20% in some cases. It is minimized through the use of fairings, fillet shaping, and optimized junctions to smooth airflow transitions and reduce these interactions.

Influencing Factors

and

The geometry and shape of an object profoundly influence parasitic drag by determining the patterns of separation and wake formation, with streamlined forms minimizing differences that contribute to form . Parasitic , primarily driven by shape-induced form , arises from the inability of air to follow abrupt contours, leading to low- wakes that increase overall resistance. Optimizing geometric parameters allows for significant reductions in this component, enabling more efficient designs across scales from to . Key geometric parameters affecting parasitic drag include the fineness ratio, defined as the ratio of an object's length to its maximum , cross-sectional area, and overall streamlining. The fineness ratio plays a critical role in balancing pressure drag; for axisymmetric bodies, an optimal ratio around 6:1 minimizes total drag by promoting gradual flow acceleration and deceleration, as demonstrated in wind tunnel tests on afterbodies, where square-base configurations exhibit drag levels up to 2.5–3.3 times higher than optimal streamlined ones. Reducing cross-sectional area perpendicular to the direction lowers the deficit in the wake, directly decreasing form drag, while streamlining—such as teardrop profiles—accelerates smoothly over the surface to prevent early separation, achieving drag coefficients as low as 0.05, nearly 90% less than a sphere's 0.47. Specific features like leading-edge radius and afterbody taper further refine these effects. A larger leading-edge radius reduces the peak suction pressure and adverse pressure gradients, delaying and thereby lowering ; for instance, increasing the radius on rotor airfoils can postpone onset and cut by enhancing attached flow over a wider range. Similarly, tapering the afterbody, as in boattail designs, contracts the wake by recovering pressure through gradual area reduction, with circular-arc boattails at fineness ratios of 3-5 yielding reductions of 20-50% at speeds compared to blunt bases. Practical examples illustrate these principles in engineering applications. In automobiles, streamlining via aero kits—such as rounded noses and tapered rear ends—can halve the from 0.5 for boxy sedans to 0.25 for optimized models, improving by 5-7% at highway speeds through reduced . For aircraft, fuselage shaping emphasizes elongated, teardrop-like contours to cut parasitic drag; helicopter fuselages with streamlined profiles and positioning the maximum diameter forward helps minimize wake formation and the fuselage's contribution to overall drag. These shape effects exhibit qualitative independence from at high values typical of full-scale operations, where coefficients plateau and trends like separation delay from streamlining persist across scales, though absolute magnitudes scale with regime.

Surface Characteristics

Surface characteristics of an or vehicle play a critical role in influencing the development, primarily affecting , which constitutes a major portion of parasitic . The texture, material, and condition of the surface determine the at the wall, with smoother surfaces promoting and lower , while irregularities disrupt the and elevate . Surface roughness accelerates the transition from laminar to turbulent boundary layers, thereby increasing due to heightened intensity and momentum transfer near the wall. In turbulent flows, rough surfaces amplify the effective , leading to higher overall parasitic drag compared to polished equivalents. For instance, corroded or weathered surfaces exhibit elevated skin friction coefficients, whereas maintained, smooth conditions sustain lower drag levels. A notable example of controlled roughness is the dimpled surface of a , where shallow indentations promote early transition to , which delays and reduces form drag; however, this comes at the cost of increased relative to a smooth . To mitigate roughness-induced drag, laminar flow control techniques such as suction or specialized coatings are employed to delay transition and maintain over larger surface areas, potentially reducing by significant margins. Real-world surface degradation, such as insect residue accumulation on leading edges, exemplifies the drag penalty from unintended roughness, with studies indicating increases in fuel consumption—and thus —up to 30% due to premature and elevated skin . Conversely, engineered anti-drag coatings like riblets, which mimic shark skin with micro-grooves aligned in the flow direction, can reduce the skin coefficient by approximately 8% in turbulent layers on fuselages and wings. The quantifiable impact of roughness is often assessed through the dimensionless parameter k/δ, the ratio of roughness height k to δ; values exceeding 0.1 typically trigger early transition to , substantially raising .

Modeling and Calculation

Drag Coefficient

The parasitic drag coefficient, commonly denoted as C_{D_p} or C_{D_0} (the ), is a dimensionless that quantifies the magnitude of parasitic drag relative to and reference area. It is defined by the relation C_{D_p} = \frac{D_p}{\frac{1}{2} \rho V^2 S}, where D_p represents the parasitic drag force, \rho is the fluid density, V is the freestream velocity, and S is the reference area (typically the wing planform area for aircraft). This formulation normalizes the drag force to enable comparisons across different scales and conditions, capturing the combined influence of form drag and skin friction drag in the absence of lift-induced effects. In the context of profile drag, C_{D_0} specifically refers to the at zero or zero , encapsulating all parasitic contributions from the aircraft's and surface properties. This coefficient is fundamental in aerodynamic modeling, as it forms the baseline for total predictions in the C_D = C_{D_0} + k C_L^2, where deviations from zero introduce induced drag terms. Experimental determination of C_{D_0} relies on measurements, where forces are recorded under controlled conditions to isolate parasitic components, often using force balances and distributions. The parasitic drag coefficient is not constant but varies with operating conditions, particularly and . As increases toward regimes, effects lead to formation and thickening, causing C_{D_p} to rise nonlinearly; for instance, profile drag increases significantly beyond Mach 0.7 due to onset. Similarly, influences C_{D_p} through scale-dependent transition and skin friction: higher s generally reduce the coefficient by promoting turbulent flow with delayed separation, though viscous effects dominate at low speeds. These dependencies are critical for scaling data to full-scale flight. Typical values for C_{D_0} in sleek modern , such as high-subsonic jets with streamlined fuselages and designs, range from 0.015 to 0.025, with 0.02 serving as a representative benchmark for efficient configurations like the P-51 Mustang or contemporary transports. These values are derived from empirical campaigns and flight tests, underscoring the coefficient's role in optimizing and .

Estimation Techniques

Empirical methods for estimating parasitic drag rely on component build-up approaches, where the total is calculated by summing contributions from individual aircraft elements such as the , , nacelles, surfaces, and protuberances, with adjustments for interference effects. This technique draws from extensive experimental data and provides semi-empirical formulas tailored to specific components; for instance, drag is often approximated using relations like C_D = C_f (1 + 1.5(d/l) + 7(d/l)^2), where C_f is the skin , d is the , and l is the , based on wetted area. Similarly, profile drag can be estimated as C_{D_s} = 2 C_f (1 + 1.5(t/c) + 60(t/c)^3), with t/c as the , allowing designers to iteratively refine configurations during early design phases. Hoerner's handbook remains a seminal reference for these methods, emphasizing statistical correlations from and flight data across to hypersonic regimes. Analytical approximations simplify parasitic drag estimation by breaking it into skin friction and form drag components, often using theory for initial predictions. For turbulent over a smooth flat plate, the skin friction coefficient is given by C_f = 0.074 / \mathrm{Re}^{1/5}, where Re is the based on plate length, serving as a for wetted surface contributions in components. This empirical relation, derived from methods and velocity profile assumptions, assumes fully turbulent conditions starting from the and is integrated over surfaces to yield total skin drag. Form drag estimates then add pressure drag penalties for bluff shapes, such as base drag on fuselages approximated as \Delta C_D = 0.029 (d_B / d_f)^2 C_f, with d_B and d_f as base and forebody diameters, respectively. These approximations prioritize rapid assessment in , though they require corrections for real and transitions. Computational fluid dynamics (CFD) tools provide detailed predictions by solving the Reynolds-averaged Navier-Stokes equations to capture viscous effects in the and wake, enabling full-configuration analysis of parasitic drag. These simulations discretize the flow field using finite volume or finite element methods, incorporating turbulence models like k-ε or Spalart-Allmaras to resolve skin friction and separation-induced form drag without relying on empirical breakdowns. For , structured or approaches yield pressure and distributions integrated over surfaces to compute the , with applications in optimizing shapes to minimize parasitic components. Validation of these estimation techniques involves comparing predictions with measurements, where parasitic drag is isolated from total using stabilized flight maneuvers or accelerometer data. Empirical and analytical methods typically show errors up to 10% high or 22% low relative to flight tests in subsonic conditions, while CFD achieves better consistency with refined grids and , though scatter in flight data can reach ±20% due to measurement uncertainties. Overall, for flows, these approaches yield accuracies within 5-10% for well-characterized configurations when cross-validated against experimental results.

References

  1. [1]
    Drag | Glenn Research Center - NASA
    Jan 21, 2023 · Drag is the aerodynamic force opposing an aircraft's motion through air, generated by the interaction of a solid body with a fluid.
  2. [2]
    [PDF] Chapter 5: Aerodynamics of Flight - Federal Aviation Administration
    There are three types of parasite drag: form drag, interference drag, and skin friction. Form Drag. Form drag is the portion of parasite drag generated by the.
  3. [3]
    [PDF] 3. Drag: An Introduction - Virginia Tech
    Jan 31, 2019 · Drag is central to aerodynamic design, and even minor changes can be critical. It can be viewed from different perspectives, including pressure ...
  4. [4]
    [PDF] EAE 130A Project 2 Report: Airplane Drag and Performance Analysis
    One of the important parameters in aircraft design and performance evaluation is drag buildup and analysis. The drag of an aircraft essentially determines ...
  5. [5]
    Parasite Drag | SKYbrary Aviation Safety
    Definition. In aviation, Parasite (Parasitic) Drag (DP) is defined as all drag that is not associated with the production of lift.
  6. [6]
    Boundary Layer | Glenn Research Center - NASA
    Jul 17, 2024 · The theory which describes boundary layer effects was first presented by Ludwig Prandtl in the early 1900's. The general fluids equations had ...
  7. [7]
    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.
  8. [8]
    Drag | SKYbrary Aviation Safety
    Parasitic drag increases with the square of the airspeed, while induced drag, being a function of lift, is greatest when maximum lift is being developed, ...Parasite Drag · Induced Drag · Wave Drag
  9. [9]
    Induced Drag Coefficient | Glenn Research Center - NASA
    Jul 27, 2023 · This additional force is called induced drag because it faces downstream and has been “induced” by the action of the tip vortices.
  10. [10]
    [PDF] Aircraft Drag Reduction: An Overview - Chalmers Publication Library
    parasitic drag and/or power-addressable in many cases via creative overall aircraft configuration design, as discussed in a subsequent section. Bushnell ...
  11. [11]
    [PDF] introduction to drag and automobile drag
    ... aerodynamic drag is a small percentage of total drag at low speeds. • Mileage estimates on the highway differ more significantly, since aerodynamic drag is.
  12. [12]
    Form Drag - an overview | ScienceDirect Topics
    Form drag is caused by differences between the pressure distribution over a body in viscous flow and that in an ideal inviscid flow.
  13. [13]
  14. [14]
    Drag of Blunt Bodies and Streamlined Bodies
    Cylinders and spheres are considered bluff bodies because at large Reynolds numbers the drag is dominated by the pressure losses in the wake. The variation of ...
  15. [15]
    Shape Effects on Drag | Glenn Research Center - NASA
    Sep 9, 2025 · A quick comparison shows that a flat plate gives the highest drag, and a streamlined symmetric airfoil gives the lowest drag, by a factor of almost 30!
  16. [16]
    Bluff Body Flows – Introduction to Aerospace Flight Vehicles
    The drag coefficient for a sphere is also shown for reference. The drag reduces quickly with increasing Reynolds number Re (note the logarithmic scales) ...
  17. [17]
    Drag Coefficient - The Engineering ToolBox
    The drag coefficient quantifies the drag or resistance of an object in a fluid environment. ; Streamlined body, 0.04, π / 4 d2 ; Airplane wing, normal position ...
  18. [18]
    Eiffel Drop Test Machine and Wind Tunnel - ASME
    In 1903, Eiffel built a device to test the drag on various bodies by dropping them along a vertical cable hung from the second level ofthe tower that bears his ...
  19. [19]
    A century of wind tunnels since Eiffel - ScienceDirect.com
    in 1908, before the building of its first wind tunnel, Gustave Eiffel conceived a very ingenious drop test machine to study the drag of solids [2], [3]. He ...
  20. [20]
    Boundary Layer Flows – Introduction to Aerospace Flight Vehicles
    History. Historically, Ludwig Prandtl introduced the concept of the boundary layer. His work on the subject originated from experiments made with his students ...Missing: parasitic | Show results with:parasitic
  21. [21]
    Direct numerical simulation of skin friction drag reduction on ...
    Sep 24, 2024 · The skin friction drag of a turbulent boundary layer is 3–5 times that of a laminar boundary layer.3 Therefore, it is expected that a ...
  22. [22]
    On the drag reduction mechanism of hypersonic turbulent boundary ...
    Feb 21, 2023 · Studies have shown that for a subsonic aircraft in cruise state, the friction drag of turbulent boundary layer accounts for 50% of the total ...
  23. [23]
    [PDF] Ludwig Prandtl's Boundary Layer
    boundary-layer theory occurred in the late 1920s when de- signers began to use the theory's results to predict skin- friction drag on airships and airplanes.
  24. [24]
    [PDF] NACA
    The minimum-drag afterbody is that of a fineness ratio of about 6 and its drag will be approximately 30 to 40 percent of that for a square-base body of fineness ...Missing: parasitic | Show results with:parasitic
  25. [25]
    Effects of Leading Edge Radius on Stall Characteristics of Rotor Airfoil
    Jun 12, 2024 · A larger leading edge radius helps to reduce the suction pressure peak and adverse pressure gradients, thus delaying the leading edge separation and stall of ...
  26. [26]
    [PDF] EFFECT OF FINENESS RATIO ON BOATTAIL DRAG OF CIRCULAR ...
    An investigation has been conducted to determine the effect of fineness ratio on the drag of circular-arc boattails at subsonic and low supersonic speeds.Missing: parasitic | Show results with:parasitic
  27. [27]
    Drag reduction technology and devices for road vehicles
    Jul 15, 2024 · Passive devices modify the shape of a vehicle without the need for an energy input or active control. Examples include wheel covers, vortex ...
  28. [28]
    Experimental Study of Helicopter Fuselage Drag | Journal of Aircraft
    Mar 14, 2016 · The total drag of a helicopter, at level cruise flight, comprises parasite drag, profile drag of rotor blades, and induced drag due to lift ...
  29. [29]
    Effect of Surface Roughness on the Aerodynamic Characteristics of ...
    It is shown that as the surface roughness increases, the minimum drag also increases due to the increase of the skin friction and the lift decreases. Surface ...
  30. [30]
    (PDF) Effect of Surface Roughness on Aerodynamic Performance of ...
    Feb 25, 2025 · Comparisons with previous studies corroborate these findings, proving that surface roughness generates increased turbulence and skin friction, ...
  31. [31]
  32. [32]
    Mechanism of drag reduction by dimples on a sphere - AIP Publishing
    Apr 17, 2006 · The drag reduction by dimples has been explained such that dimples on a sphere induce a turbulent boundary layer on its surface and lower drag ...
  33. [33]
    Drag Reduction by Laminar Flow Control - MDPI
    The greatest potential for aerodynamic drag reduction is seen in laminar flow control by boundary layer suction.
  34. [34]
    NASA Researchers to Flying Insects Bug Off
    Nov 1, 2013 · Studies show that bug splatter and other factors that cause drag can increase airplane fuel consumption as much as 30 percent.
  35. [35]
    Riblets - an overview | ScienceDirect Topics
    From an aerodynamic perspective riblets were identified as a mature technology which could provide modest reductions (7–8% skin friction drag reduction for ...
  36. [36]
    Direct Numerical Simulation of Boundary Layer Transition Induced ...
    A key dimensionless parameter governing transition onset is the relative roughness height k / δ , defined as the ratio of roughness height k to boundary layer ...Missing: criteria | Show results with:criteria
  37. [37]
    [PDF] Fluids – Lecture 10 Notes - MIT
    Drag breakdown. The drag on a subsonic aircraft can be broken down as follows. D = Do + Dp + Di where. Do = “parasite” drag of fuselage + tail + landing gear + ...
  38. [38]
    Aerodynamics of Finite Wings – Introduction to Aerospace Flight ...
    The non-lifting part depends primarily on the airplane's overall shape and is typically referred to as the form drag or parasitic drag coefficient. The lifting ...
  39. [39]
    [PDF] Independent Effects of Reynolds Number and Mach Number on ...
    (b) Effect on usable lift coefficient. Figure 32.—Effect of Reynolds number on drag coefficients for clean and iced wing with WB33 artificial ice shapes.
  40. [40]
    [PDF] The Variation of Profile Drag with Mach Number - AERADE
    He quotes the aircraft lift coefficient and the Reynolds number for each Mach number; these are tabulated in Table la together with his drag coefficients. His ...
  41. [41]
    [PDF] Aircraft Drag Polar Estimation Based on a Stochastic Hierarchical ...
    Dec 7, 2018 · While the zero drag coefficient contains the parasitic drag of the whole aircraft, the wing is mainly responsible for the lift-induced drag.Missing: sleek | Show results with:sleek
  42. [42]
    [PDF] Fluid-Dynamic Drag
    ... DRAG OF AIRPLANE CONFIGURATIONS. (a) Induced Drag of Wings drag due to wing twist, induced interference_15-33. (b) Parasitic Drag of Airplanes drag of “Meteor ...
  43. [43]
    [PDF] Hermann Schlichting (Deceased) Klaus Gersten Ninth Edition
    However a new fifth section on numerical methods in boundary–layer theory has been added. The partition into chapters had to be somewhat modified in order to.
  44. [44]
    [PDF] Technical Status Review Drag Prediction and Analysis from ... - NATO
    Methods based on the Euler equations and Reynolds Averaged Navier-Stokes equations, at least for simple configurations, are approaching this status. The status ...
  45. [45]
    [PDF] Prediction of Aerodynamic Drag. - DTIC
    13. The aerodynamic cleanness factor accounts for form drag and interference effects riot treated by fuselage shape parameters. Examples are form and.