ART

A swept wing is a wing that angles either backward or occasionally forward from its root rather than in a straight sideways direction.

Swept wings have been flown since the pioneer days of aviation. Wing sweep at high speeds was first investigated in Germany as early as 1935 by Albert Betz and Adolph Busemann, finding application just before the end of the Second World War. It has the effect of delaying the shock waves and accompanying aerodynamic drag rise caused by fluid compressibility near the speed of sound, improving performance. Swept wings are therefore almost always used on jet aircraft designed to fly at these speeds. Swept wings are also sometimes used for other reasons, such as low drag, low observability, structural convenience or pilot visibility.

The term "swept wing" is normally used to mean "swept back", but variants include forward sweep, variable sweep wings and oblique wings in which one side sweeps forward and the other back. The delta wing is also aerodynamically a form of swept wing.

Design characteristics

For a wing of given span, sweeping it increases the length of the spars running along it from root to tip. This tends to increase weight and reduce stiffness. If the fore-aft chord of the wing also remains the same, the distance between leading and trailing edges reduces, reducing its ability to resist twisting (torsion) forces. A swept wing of given span and chord must therefore be strengthened and will be heavier than the equivalent unswept wing.

A swept wing typically angles backward from its root rather than forwards. Because wings are made as light as possible, they tend to flex under load. This aeroelasticity under aerodynamic load causes the tips to bend upwards in normal flight. Backwards sweep causes the tips to reduce their angle of attack as they bend, reducing their lift and limiting the effect. Forward sweep causes the tips to increase their angle of attack as they bend. This increases their lift causing further bending and hence yet more lift in a cycle which can cause a runaway structural failure. For this reason forward sweep is rare and the wing must be unusually rigid.

The characteristic "sweep angle" is normally measured by drawing a line from root to tip, typically 25% of the way back from the leading edge, and comparing that with the perpendicular to the longitudinal axis of the aircraft. Typical sweep angles vary from 0 for a straight-wing aircraft, to 45 degrees or more for fighters and other high-speed designs.
Aerodynamics
Subsonic and transonic flight
Yakovlev Yak-25 swept wing
In the transonic phase, the swept wing also sweeps the shock which is at the top rear of the wing. Only the velocity component perpendicular to the shock is affected.

As an aircraft enters the transonic speeds just below the speed of sound, the pressure waves associated with subsonic flight converge and begin to impinge on the aircraft. As the pressure waves converge the air in front of the aircraft begins to compress. This creates a force known as wave drag. This wave drag increases steeply until the whole aircraft is supersonic and then reduces.

However, shock waves can form on some parts of an aircraft moving at less than the speed of sound. Low-pressure regions around an aircraft cause the flow to accelerate, and at transonic speeds this local acceleration can exceed Mach 1. Localized supersonic flow must return to the freestream conditions around the rest of the aircraft, and as the flow enters an adverse pressure gradient in the aft section of the wing, a discontinuity emerges in the form of a shock wave as the air is forced to rapidly slow and return to ambient pressure.

With objects where there is a sudden reduction in profile/thickness and the local air expands rapidly to fill the space taken by the solid object or where a rapid angular change is imparted to the airflow causing a momentary increase of volume/decrease in density, an oblique shock wave is generated. This is why shock waves are often associated with the part of a fighter aircraft cockpit canopy with the highest local curvature, appearing immediately behind this point.

At the point where the density drops, the local speed of sound correspondingly drops and a shock wave can form. This is why in conventional wings, shock waves form first after the maximum Thickness/Chord and why all airliners designed for cruising in the transonic range (above M0.8) have supercritical wings that are flatter on top resulting in minimized angular change of flow to upper surface air. The angular change to the air that is normally part of lift generation is decreased and this lift reduction is compensated for by deeper curved lower surfaces accompanied by a reflex curve at the trailing edge. This results in a much weaker standing shock wave towards the rear of the upper wing surface and a corresponding increase in critical mach number.

Shock waves require energy to form. This energy is taken out of the aircraft, which has to supply extra thrust to make up for this energy loss. Thus the shocks are seen as a form of drag. Since the shocks form when the local air velocity reaches supersonic speeds, there is a certain "critical mach" speed where sonic flow first appears on the wing. There is a following point called the drag divergence mach number where the effect of the drag from the shocks becomes noticeable. This is normally when the shocks start generating over the wing, which on most aircraft is the largest continually curved surface, and therefore the largest contributor to this effect.

Sweeping the wing has the effect of reducing the curvature of the body as seen from the airflow, by the cosine of the angle of sweep. For instance, a wing with a 45 degree sweep will see a reduction in effective curvature to about 70% of its straight-wing value. This has the effect of increasing the critical Mach by 30%. When applied to large areas of the aircraft, like the wings and empennage, this allows the aircraft to reach speeds closer to Mach 1.

One of the simplest and best explanations of how the swept wing works was offered by Robert T. Jones: "Suppose a cylindrical wing (constant chord, incidence, etc.) is placed in an airstream at an angle of yaw – i.e., it is swept back. Now, even if the local speed of the air on the upper surface of the wing becomes supersonic, a shock wave cannot form there because it would have to be a sweptback shock – swept at the same angle as the wing – i.e., it would be an oblique shock. Such an oblique shock cannot form until the velocity component normal to it becomes supersonic."[1]

One limiting factor in swept wing design is the so-called "middle effect". If a swept wing is continuous – an oblique swept wing, the pressure iso-bars will be swept at a continuous angle from tip to tip. However, if the left and right halves are swept back equally, as is common practice, the pressure iso-bars on the left wing in theory will meet the pressure iso-bars of the right wing on the centerline at a large angle. As the iso-bars cannot meet in such a fashion, they will tend to curve on each side as they near the centerline, so that the iso-bars cross the centerline at right angles to the centerline. This causes an "unsweeping" of the iso-bars in the wing root region. To combat this unsweeping, German aerodynamicist Dietrich Küchemann proposed and had tested a local indentation of the fuselage above and below the wing root. This proved to not be very effective.[2] During the development of the Douglas DC-8 airliner, uncambered airfoils were used in the wing root area to combat the unsweeping.[3][4] Similarly, a decambered wing root glove was added to the Boeing 707 wing to create the Boeing 720.[5]
Supersonic flight
At supersonic speeds an oblique shock exists in front of the leading edge of the wing. The velocity component perpendicular to the shock is different upstream and downstream of the shock. The velocity component parallel to the shock is the same on both sides of the shock.
The delta wing of the Convair F-106 Delta Dart is a form of swept wing.

Airflow at supersonic speeds generates lift through the formation of shock waves, as opposed to the patterns of airflow over and under the wing. These shock waves, as in the transonic case, generate large amounts of drag. One of these shock waves is created by the leading edge of the wing, but contributes little to the lift. In order to minimize the strength of this shock it needs to remain "attached" to the front of the wing, which demands a very sharp leading edge. To better shape the shocks that will contribute to lift, the rest of an ideal supersonic airfoil is roughly diamond-shaped in cross-section. For low-speed lift these same airfoils are very inefficient, leading to poor handling and very high landing speeds.[6]

One way to avoid the need for a dedicated supersonic wing is to use a highly swept subsonic design. Airflow behind the shock waves of a moving body are reduced to subsonic speeds. This effect is used within the intakes of engines meant to operate in the supersonic, as jet engines are generally incapable of ingesting supersonic air directly. This can also be used to reduce the speed of the air as seen by the wing, using the shocks generated by the nose of the aircraft. As long as the wing lies behind the cone-shaped shock wave, it will "see" subsonic airflow and work as normal. The angle needed to lie behind the cone increases with increasing speed, at Mach 1.3 the angle is about 45 degrees, at Mach 2.0 it is 60 degrees.[7] For instance, at Mach 1.3 the angle of the Mach cone formed off the body of the aircraft will be at about sinμ = 1/M (μ is the sweep angle of the Mach cone)[8]

Generally it is not possible to arrange the wing so it will lie entirely outside the supersonic airflow and still have good subsonic performance. Some aircraft, like the English Electric Lightning are tuned almost entirely for high-speed flight and feature highly swept wings that make little to no compromise for the low-speed problems that such a profile creates.[9][10] In other cases the use of variable geometry wings, as on the Grumman F-14 Tomcat and Panavia Tornado, allows an aircraft to move the wing to keep it at the most efficient angle regardless of speed, although the drawbacks incurred of increased complexity and weight have led to this being a rare feature.[11][12]

Most high-speed aircraft have a wing that spends at least some of its time in the supersonic airflow. But since the shock cone moves towards the fuselage with increased speed (that is, the cone becomes narrower), the portion of the wing in the supersonic flow also changes with speed. Since these wings are swept, as the shock cone moves inward, the lift vector moves forward as the outer, rearward portions of the wing are generating less lift. This results in powerful pitching moments and their associated required trim changes.

Disadvantages
Spanwise flow of the boundary layer

When a swept wing travels at high speed, the airflow has little time to react and simply flows over the wing almost straight from front to back. At lower speeds the air does have time to react, and is pushed spanwise by the angled leading edge, towards the wing tip. At the wing root, by the fuselage, this has little noticeable effect, but as one moves towards the wingtip the airflow is pushed spanwise not only by the leading edge, but the spanwise moving air beside it. At the tip the airflow is moving along the wing instead of over it, a problem known as spanwise flow.

The lift from a wing is generated by the airflow over it from front to rear. With increasing span-wise flow the boundary layers on the surface of the wing have longer to travel, and so are thicker and more susceptible to transition to turbulence or flow separation, also the effective aspect ratio of the wing is less and so air "leaks" around the wing tips reducing their effectiveness. The spanwise flow on swept wings produces airflow that moves the stagnation point on the leading edge of any individual wing segment further beneath the leading edge, increasing effective angle of attack of wing segments relative to its neighbouring forward segment. The result is that wing segments farther towards the rear operate at increasingly higher angles of attack promoting early stall of those segments. This promotes tip stall on back swept wings, as the tips are most rearward, while delaying tip stall for forward swept wings, where the tips are forward. With both forward and back swept wings, the rear of the wing will stall first. This creates a nose-up pressure on the aircraft. If this is not corrected by the pilot it causes the plane to pitch up, leading to more of the wing stalling, leading to more pitch up, and so on. This problem came to be known as the Sabre dance in reference to the number of North American F-100 Super Sabres that crashed on landing as a result.[13][14]

The solution to this problem took on many forms. One was the addition of a fin known as a wing fence on the upper surface of the wing to redirect the flow to the rear ; the MiG-15 was one example of an aircraft fitted with wing fences.[15] Another closely related design was addition of a dogtooth notch to the leading edge, as present on the Avro Arrow interceptor.[16] Other designs took a more radical approach, including the Republic XF-91 Thunderceptor's wing that grew wider towards the tip to provide more lift at the tip. The Handley Page Victor was equipped with a crescent wing, featuring substantial sweep-back near the wing root where the wing was thickest, and progressively reducing sweep along the span as the wing thickness reduced towards the tip.[17][18]

Modern solutions to the problem no longer require "custom" designs such as these. The addition of leading-edge slats and large compound flaps to the wings has largely resolved the issue.[19][20][21] On fighter designs, the addition of leading-edge extensions, which are typically included to achieve a high level of maneuverability, also serve to add lift during landing and reduce the problem.[22][23]

The swept wing also has several more problems. One is that for any given length of wing, the actual span from tip-to-tip is shorter than the same wing that is not swept. Low speed drag is strongly correlated with the aspect ratio, the span compared to chord, so a swept wing always has more drag at lower speeds. Another concern is the torque applied by the wing to the fuselage, as much of the wing's lift lies behind the point where the wing root connects to the plane. Finally, while it is fairly easy to run the main spars of the wing right through the fuselage in a straight wing design to use a single continuous piece of metal, this is not possible on the swept wing because the spars will meet at an angle.
Sweep theory

Sweep theory is an aeronautical engineering description of the behavior of airflow over a wing when the wing's leading edge encounters the airflow at an oblique angle. The development of sweep theory resulted in the swept wing design used by most modern jet aircraft, as this design performs more effectively at transonic and supersonic speeds. In its advanced form, sweep theory led to the experimental oblique wing concept.

Adolf Busemann introduced the concept of the swept wing and presented this in 1935 at the 5. Volta-Congress in Rome. Sweep theory in general was a subject of development and investigation throughout the 1930s and 1940s, but the breakthrough mathematical definition of sweep theory is generally credited to NACA's Robert T. Jones in 1945. Sweep theory builds on other wing lift theories. Lifting line theory describes lift generated by a straight wing (a wing in which the leading edge is perpendicular to the airflow). Weissinger theory describes the distribution of lift for a swept wing, but does not have the capability to include chordwise pressure distribution. There are other methods that do describe chordwise distributions, but they have other limitations. Jones' sweep theory provides a simple, comprehensive analysis of swept wing performance.

To visualize the basic concept of simple sweep theory, consider a straight, non-swept wing of infinite length, which meets the airflow at a perpendicular angle. The resulting air pressure distribution is equivalent to the length of the wing's chord (the distance from the leading edge to the trailing edge). If we were to begin to slide the wing sideways (spanwise), the sideways motion of the wing relative to the air would be added to the previously perpendicular airflow, resulting in an airflow over the wing at an angle to the leading edge. This angle results in airflow traveling a greater distance from leading edge to trailing edge, and thus the air pressure is distributed over a greater distance (and consequently lessened at any particular point on the surface).

This scenario is identical to the airflow experienced by a swept wing as it travels through the air. The airflow over a swept wing encounters the wing at an angle. That angle can be broken down into two vectors, one perpendicular to the wing, and one parallel to the wing. The flow parallel to the wing has no effect on it, and since the perpendicular vector is shorter (meaning slower) than the actual airflow, it consequently exerts less pressure on the wing. In other words, the wing experiences airflow that is slower - and at lower pressures - than the actual speed of the aircraft.

One of the factors that must be taken into account when designing a high-speed wing is compressibility, which is the effect that acts upon a wing as it approaches and passes through the speed of sound. The significant negative effects of compressibility made it a prime issue with aeronautical engineers. Sweep theory helps mitigate the effects of compressibility in transonic and supersonic aircraft because of the reduced pressures. This allows the mach number of an aircraft to be higher than that actually experienced by the wing.

There is also a negative aspect to sweep theory. The lift produced by a wing is directly related to the speed of the air over the wing. Since the airflow speed experienced by a swept wing is lower than what the actual aircraft speed is, this becomes a problem during slow-flight phases, such as takeoff and landing. There have been various ways of addressing the problem, including the variable-incidence wing design on the Vought F-8 Crusader,[24] and swing wings on aircraft such as the F-14, F-111, and the Panavia Tornado.[11][12]
Variant designs

The term "swept wing" is normally used to mean "swept back", but other swept variants include forward sweep, variable sweep wings and oblique wings in which one side sweeps forward and the other back. The delta wing also incorporates the same advantages as part of its layout.
Forward sweep
Main article: Forward-swept wing
LET L-13 two-seat glider showing forward swept wing
Grumman X-29 experimental aircraft, an extreme example of a forward swept wing

Sweeping a wing forward has approximately the same effect as rearward in terms of drag reduction, but has other advantages in terms of low-speed handling where tip stall problems simply go away. In this case the low-speed air flows towards the fuselage, which acts as a very large wing fence. Additionally, wings are generally larger at the root anyway, which allows them to have better low-speed lift.

However, this arrangement also has serious stability problems. The rearmost section of the wing will stall first causing a pitch-up moment pushing the aircraft further into stall similar to a swept back wing design. Thus swept-forward wings are unstable in a fashion similar to the low-speed problems of a conventional swept wing. However unlike swept back wings, the tips on a forward swept design will stall last, maintaining roll control.

Forward-swept wings can also experience dangerous flexing effects compared to aft-swept wings that can negate the tip stall advantage if the wing is not sufficiently stiff. In aft-swept designs, when the airplane maneuvers at high load factor the wing loading and geometry twists the wing in such a way as to create washout (tip twists leading edge down). This reduces the angle of attack at the tip, thus reducing the bending moment on the wing, as well as somewhat reducing the chance of tip stall.[25] However, the same effect on forward-swept wings produces a wash-in effect that increases the angle of attack promoting tip stall.

Small amounts of sweep do not cause serious problems, and had been used on a variety of aircraft to move the spar into a convenient location, as on the Junkers Ju 287 or HFB 320 Hansa Jet.[26][27] However, larger sweep suitable for high-speed aircraft, like fighters, was generally impossible until the introduction of fly by wire systems that could react quickly enough to damp out these instabilities. The Grumman X-29 was an experimental technology demonstration project designed to test the forward swept wing for enhanced maneuverability during the 1980s.[28][29] The Sukhoi Su-47 Berkut is another notable demonstrator aircraft implementing this technology to achieve high levels of agility.[30] To date, no highly swept-forward design has entered production.
History
Early history

The first successful aeroplanes adhered to the basic design of rectangular wings at right angles to the body of the machine, but there were experimentalists who explored other geometries to achieve better aerodynamic results. The swept wing geometry appeared before World War I, and was conceived as a means of permitting the design of safe and stable aeroplanes. The best of these designs imposed "self-damping" inherent stability upon a tailless swept wing. These inspired several flying wing gliders and some powered aircraft during the interwar years.[31]
A Burgess-Dunne tailless biplane: the angle of sweep is exaggerated by the sideways view, with washout also present at the wingtips.

The first to achieve stability was British designer J. W. Dunne who was obsessed with achieving inherent stability in flight. He successfully employed swept wings in his tailless aircraft (which, crucially, used washout) as a means of creating positive longitudinal static stability.[32] For a low-speed aircraft, swept wings may be used to resolve problems with the center of gravity, to move the wing spar into a more convenient location, or to improve the sideways view from the pilot's position.[31] By 1905, Dunne had already built a model glider with swept wings and by 1913 he had constructed successful powered variants that were able to cross the English Channel. The Dunne D.5 was exceptionally aerodynamically stable for the time,[33] and the D.8 was sold to the Royal Flying Corps; it was also manufactured under licence by Starling Burgess to the United States Navy amongst other customers.[34]

Dunne's work ceased with the onset of war in 1914, but afterwards the idea was taken up by G. T. R. Hill in England who designed a series of gliders and aircraft to Dunne's guidelines, notably the Westland-Hill Pterodactyl series.[35] However, Dunne's theories met with little acceptance amongst the leading aircraft designers and aviation companies at the time.[36]
German developments
Adolf Busemann proposed the use of swept-wings to reduce drag at high speed, at the Volta Conference in 1935.

The idea of using swept wings to reduce high-speed drag was developed in Germany in the 1930s. At a Volta Conference meeting in 1935 in Italy, Dr. Adolf Busemann suggested the use of swept wings for supersonic flight. He noted that the airspeed over the wing was dominated by the normal component of the airflow, not the freestream velocity, so by setting the wing at an angle the forward velocity at which the shock waves would form would be higher (the same had been noted by Max Munk in 1924, although not in the context of high-speed flight).[37] Albert Betz immediately suggested the same effect would be equally useful in the transonic.[38] After the presentation the host of the meeting, Arturo Crocco, jokingly sketched "Busemann's airplane of the future" on the back of a menu while they all dined. Crocco's sketch showed a classic 1950's fighter design, with swept wings and tail surfaces, although he also sketched a swept propeller powering it.[37]

At the time, however, there was no way to power an aircraft to these sorts of speeds, and even the fastest aircraft of the era were only approaching 400 km/h (249 mph).The presentation was largely of academic interest, and soon forgotten. Even notable attendees including Theodore von Kármán and Eastman Jacobs did not recall the presentation 10 years later when it was re-introduced to them.[39]

Hubert Ludwieg of the High-Speed Aerodynamics Branch at the AVA Göttingen in 1939 conducted the first wind tunnel tests to investigate Busemann's theory.[2] Two wings, one with no sweep, and one with 45 degrees of sweep were tested at Mach numbers of 0.7 and 0.9 in the 11 x 13 cm wind tunnel. The results of these tests confirmed the drag reduction offered by swept wings at transonic speeds.[2] The results of the tests were communicated to Albert Betz who then passed them on to Willy Messerschmitt in December 1939. The tests were expanded in 1940 to include wings with 15, 30 and -45 degrees of sweep and Mach numbers as high as 1.21.[2]

With the introduction of jets in the later half of the Second World War, the swept wing became increasingly applicable to optimally satisfying aerodynamic needs. The German jet-powered Messerschmitt Me 262 and rocket-powered Messerschmitt Me 163 suffered from compressibility effects that made both aircraft very difficult to control at high speeds. In addition, the speeds put them into the wave drag regime, and anything that could reduce this drag would increase the performance of their aircraft, notably the notoriously short flight times measured in minutes. This resulted in a crash program to introduce new swept wing designs, both for fighters as well as bombers. The Blohm & Voss P 215 was designed to take full advantage of the swept wing's aerodynamic properties; however, an order for three prototypes was received only weeks before the war ended and no examples were ever built.[40] The Focke-Wulf Ta 183 was another swept wing fighter design, but was also not produced before the war's end.[41] In the post-war era, Kurt Tank developed the Ta 183 into the IAe Pulqui II, but this proved unsuccessful.[42]

A prototype test aircraft, the Messerschmitt Me P.1101, was built to research the tradeoffs of the design and develop general rules about what angle of sweep to use.[43] When it was 80% complete, the P.1101 was captured by US forces and returned to the United States, where two additional copies with US-built engines carried on the research as the Bell X-5.[44] Germany's wartime experience with the swept wings and its high value for supersonic flight stood in strong contract to the prevailing views of Allied experts of the era, who commonly espoused their belief in the impossibility of manned vehicles travelling at such speeds.[45]
Postwar advancements
Artist's impression of the Miles M.52

During the immediate post-war era, several nations were conducting research into high speed aircraft. In the United Kingdom, work commenced during 1943 on the Miles M-52, a high-speed experimental aircraft equipped with a straight wing that was developed in conjunction with Frank Whittle's Power Jets company, the Royal Aircraft Establishment (RAE) in Farnborough, and the National Physical Laboratory.[46] Despite being envisioned to be capable of achieving 1,000 miles per hour (1,600 km/h) in level flight, thus enabling the aircraft to potentially be the first to exceed the speed of sound in the world,[46] in February 1946, the programme was abrupted discontinued for unclear reasons.[47] It has since been widely recognised that the cancellation of the M.52 was a major setback in British progress in the field of supersonic design.[31]

Another, more successful, programme was the US's Bell X-1, which also was equipped with a straight wing. According to Miles Chief Aerodynamicist Dennis Bancroft, the Bell Aircraft company was given access to the drawings and research on the M.52.[48] On 14 October 1947, the Bell X-1 performed the first manned supersonic flight, piloted by Captain Charles "Chuck" Yeager, having been drop launched from the bomb bay of a Boeing B-29 Superfortress and attained a record-breaking speed of Mach 1.06 (700 miles per hour (1,100 km/h; 610 kn)).[31] The news of a successful straight-wing supersonic aircraft surprised many aeronautical experts on both sides of the Atlantic, as it was increasingly believed that a swept-wing design not only highly beneficial but also necessary to break the sound barrier.[45]
The de Havilland DH 108, a prototype swept-wing aircraft

During the final years of the Second World War, aircraft designer Sir Geoffrey de Havilland commenced development on the de Havilland Comet, which would become the world's first jet airliner. An early design consideration was whether to apply the new swept-wing configuration.[49] Thus, an experimental aircraft to explore the technology, the de Havilland DH 108, was developed by the firm in 1944, headed by project engineer John Carver Meadows Frost with a team of 8–10 draughtsmen and engineers. The DH 108 primarily consisted of the pairing of the front fuselage of the de Havilland Vampire to a swept wing and compact stubby vertical tail; it was the first British swept wing jet, unofficially known as the "Swallow".[50] It first flew on 15 May 1946, a mere eight months after the project's go-ahead. Company test pilot and son of the builder, Geoffrey de Havilland Jr., flew the first of three aircraft and found it extremely fast – fast enough to try for a world speed record. On 12 April 1948, a D.H.108 did set a world's speed record at 973.65 km/h (605 mph), it subsequently became the first jet aircraft to exceed the speed of sound.[51]

Around this same timeframe, the Air Ministry introduced a program of experimental aircraft to examine the effects of swept wings, as well as the delta wing configuration.[52] Furthermore, the Royal Air Force (RAF) identified a pair of proposed fighter aircraft equipped with swept wings from Hawker Aircraft and Supermarine, the Hawker Hunter and Supermarine Swift respectively, and successfully pressed for orders to be placed 'off the drawing board' in 1950.[53] On 7 September 1953, the sole Hunter Mk 3 (the modified first prototype, WB 188) flown by Neville Duke broke the world air speed record for jet-powered aircraft, attaining a speed of 727.63 mph (1,171.01 km/h) over Littlehampton, West Sussex.[54] This world record stood for less than three weeks before being broken on 25 September 1953 by the Hunter's early rival, the Supermarine Swift, being flown by Michael Lithgow.[55]

In February 1945, NACA engineer Robert T. Jones started looking at highly swept delta wings and V shapes, and discovered the same effects as Busemann. He finished a detailed report on the concept in April, but found his work was heavily criticised by other members of NACA Langley, notably Theodore Theodorsen, who referred to it as "hocus-pocus" and demanded some "real mathematics".[37] However, Jones had already secured some time for free-flight models under the direction of Robert Gilruth, whose reports were presented at the end of May and showed a fourfold decrease in drag at high speeds. All of this was compiled into a report published on June 21, 1945, which was sent out to the industry three weeks later.[56] Ironically, by this point Busemann's work had already been passed around.
The first American swept-wing aircraft, the Boeing B-47 Stratojet

On May 1945, the American Operation Paperclip reached Braunschweig, where US personnel discovered a number of swept wing models and a mass of technical data from the wind tunnels. One member of the US team was George S. Schairer, who was at that time working at the Boeing company. He immediately forwarded a letter to Ben Cohn at Boeing, communicating the value of the swept wing concept.[57][58] He also told Cohn to distribute the letter to other companies as well, although only Boeing and North American made immediate use of it.

Boeing was in the midst of designing the B-47 Stratojet, and the initial Model 424 was a straight-wing design similar to the B-45, B-46 and B-48 it competed with. Analysis by Boeing engineer Vic Ganzer suggested an optimum sweepback angle of about 35 degrees.[59] By September 1945, the Braunschweig data had been worked into the design, which re-emerged as the Model 448, a larger six-engine design with more robust wings swept at 35 degrees.[37] Another re-work moved the engines into strut-mounted pods under the wings due to concerns of the uncontained failure of an internal engine could potentially destroy the aircraft via either fire or vibration.[60] The resulting B-47 was hailed as the fastest of its class in the world during the late 1940s,[61] and trounced the straight-winged competition. Boeing's jet-transport formula of swept wings and pylon-mounted engines has since been universally adopted.

In fighters, North American Aviation was in the midst of working on a straight-wing jet-powered naval fighter, then known as the FJ-1; it was later submitted to the United States Air Force as the XP-86.[62] Larry Green, who could read German, studied the Busemann reports and convinced management to allow a redesign starting in August 1945.[37][63][64] The performance of the F-86A allowed it set the first of several official world speed records, attaining 671 miles per hour (1,080 km/h) on 15 September 1948, flown by Major Richard L. Johnson.[65] With the appearance of the MiG-15, the F-86 was rushed into combat, while straight-wing jets like the Lockheed P-80 Shooting Star and Republic F-84 Thunderjet were quickly relegated to ground attack missions. Some, such as the F-84 and Grumman F-9 Cougar, were later redesigned with swept wings from straight-winged aircraft.[66][67] Later planes, such as the North American F-100 Super Sabre, would be designed with swept wings from the start, though additional innovations such as the afterburner, area-rule and new control surfaces would be necessary to master supersonic flight.[68][13]
A MiG-15 of the Soviet Air Forces

The Soviet Union was also intrigued about the idea of swept wings on aircraft, when their "captured aviation technology" counterparts to the western Allies spread out across the defeated Third Reich. Artem Mikoyan was asked by the Soviet government's TsAGI aviation research department to develop a test-bed aircraft to research the swept wing idea — the result was the late 1945-flown, unusual MiG-8 Utka pusher canard layout aircraft, with its rearwards-located wings being swept back for this type of research.[69] The swept wing was applied to the MiG-15, an early jet-powered fighter, its maximum speed of 1,075 km/h (668 mph) outclassed the straight-winged American jets and piston-engined fighters initially deployed during the Korean War.[70] The MiG-15 is believed to have been one of the most produced jet aircraft; in excess of 13,000 would ultimately be manufactured.[71]

The MiG-15, which could not safely exceed Mach 0.92, served as the basis for the MiG-17, which was designed to be controllable at higher Mach numbers.[72] Its wing featured a "sickle sweep" compound shape, somewhat similar to the F-100 Super Sabre, with a 45° angle near the fuselage and a 42° angle for the outboard part of the wings.[73] A further derivative of the design, designated MiG-19, featured a relatively thin wing suited to supersonic flight that was designed at TsAGI, the Soviet Central Aerohydrodynamic Institute; swept back at an angle of 55 degrees, this wing featured a single wing fence on each side.[74] A specialist high-altitude variant, the Mig-19SV, featured, amongst other changes, flap adjusted to generate greater lift at higher altitudes, helping to increase the aircraft's ceiling from 17,500 m (57,400 ft) to 18,500 m (60,700 ft).[75][76]

Germany's swept wing research also made its way to the Swedish aircraft manufacturer SAAB, allegedly via a group of ex-Messerschmitt engineers that had fled to Switzerland during late 1945.[77][78] At the time, SAAB was eager to make aeronautic advances, particularly in the new field of jet propulsion.[79] The company incorporated both the jet engine and the swept wing to produce the Saab 29 Tunnan fighter; on 1 September 1948, the first prototype conducted its maiden flight, flown by the English test pilot S/L Robert A. 'Bob' Moore, DFC and bar,[80] Although not well known outside Sweden, the Tunnan was the first Western European fighter to be introduced with such a wing configuration.[81][82] In parallel, SAAB also developed another swept wing aircraft, the Saab 32 Lansen, primarily to serve as Sweden's standard attack aircraft.[83] Its wing, which had a 10 per cent laminar profile and a 35° sweep, featured triangular fences near the wing roots in order to improve airflow when the aircraft was being flown at a high angle of attack.[83][84] On 25 October 1953, a SAAB 32 Lansen attained a Mach number of at least 1.12 while in a shallow dive, exceeding the sound barrier.[84]
The Avro Vulcan, flying at Farnborough, 1958.

The dramatic successes of aircraft such as Hawker Hunter, the B-47, and F-86 embodied the widespread acceptance of the swept wing research acquired from Germany. Eventually, almost all advanced design efforts would incorporate a swept wing configuration. The classic Boeing B-52, designed in the 1950s, continues in service as a high-subsonic long-range heavy bomber despite the development of the triple-sonic North American B-70 Valkyrie, supersonic swing-wing Rockwell B-1 Lancer, and flying wing designs.[85][86] While the Soviets never matched the performance of the Boeing B-52 Stratofortress with a jet aircraft, the intercontinental range Tupolev Tu-95 turboprop bomber with its near-jet class top speed of 920 km/h, combining swept wings with propeller propulsion, also remains in service today, being the fastest propeller-powered production aircraft.[87] In Britain, a range of swept-wing bombers were designed, these being the Vickers Valiant (1951),[88] the Avro Vulcan (1952),[89] and the Handley Page Victor (1952).[90]

By the early 1950s, nearly every new fighter was either rebuilt or designed from scratch with a swept wing. By the 1960s, most civilian jets also adopted swept wings. The Douglas A-4 Skyhawk and Douglas F4D Skyray were examples of delta wings that also have swept leading edges with or without a tail. Most early transonic and supersonic designs such as the MiG-19 and F-100 used long, highly swept wings. Swept wings would reach Mach 2 in the arrow-winged BAC Lightning, and stubby winged Republic F-105 Thunderchief, which was found to be wanting in turning ability in Vietnam combat. By the late 1960s, the F-4 Phantom and Mikoyan-Gurevich MiG-21 that both used variants on tailed delta wings came to dominate front line air forces. Variable geometry wings were employed on the American F-111, Grumman F-14 Tomcat and Soviet Mikoyan MiG-27, although the idea would be abandoned for the American SST design. After the 1970s, most newer generation fighters optimized for maneuvering air combat since the USAF F-15 and Soviet Mikoyan MiG-29 have employed relatively short-span fixed wings with relatively large wing area.
See also

Delta wing
Theodore von Kármán, first to recognize the importance of the swept wing[91]
Trapezoidal wing
Wing configuration

References
Citations

Sears, William Rees, Stories form a 20th-Century Life, Parabolic Press, Inc., Stanford California, 1994.
Meier, Hans-Ulrich, editor German Development of the Swept Wing 1935–1945, AIAA Library of Flight, 2010. Originally published in German as Die deutsche Luftahrt Die Pfeilflügelentwicklung in Deutschland bis 1945, Bernard & Graefe Verlag, 2006.
Shevell, Richard, "Aerodynamic Design Features", DC-8 design summary, February 22, 1957.
Dunn, Orville R., "Flight Characteristics of the DC-8", SAE paper 237A, presented at the SAE National Aeronautic Meeting, Los Angeles California, October 1960.
Cook, William H. The Road to the 707: The Inside Story of Designing the 707. Bellevue, Washington: TYC Publishing, 1991. ISBN 0-9629605-0-0.
"Supersonic Wing Designs." selkirk.bc.ca. Retrieved: 1 August 2011.
"Supersonic Wing design: The Mach cone becomes increasingly swept back with increasing Mach numbers." Archived 30 September 2007 at the Wayback Machine Centennial of Flight Commission, 2003. Retrieved: 1 August 2011.
Haack, Wolfgang. "Heinzerling, Supersonic Area Rule" (in German), p. 39. Archived 27 March 2009 at the Wayback Machine bwl.tu-darmstadt.de.
Davies 2014, p. 103.
Jones, Lloyd S. U.S. Fighters, Aero, 1975. pp. 272-274.
Woolridge, Capt. E.T., ed. Into the Jet Age: Conflict and Change in Naval Aviation 1945–1975, an Oral History. Annapolis, Maryland: Naval Institute Press, 1995. ISBN 1-55750-932-8.
Spick, Green and Swanborough 2001, p. 33.
"Deadly Sabre Dance". historynet.com. Retrieved 11 November 2020.
Ives, Burl. "Burl Ives Song Book." Ballantine Books, Inc., New York, November 1953, page 240.
Gunston 1995, p. 188.
Whitcomb 2002, pp. 89–91.
Brookes 2011, pp. 6–7.
Lee, G.H. "Aerodynamics of the Crescent Wing." Flight, 14 May 1954, pp. 611–612.
High-Lift Aerodynamics, by A. M. O. Smith, McDonnell Douglas Corporation, Long Beach, June 1975 Archived 7 July 2011 at the Wayback Machine
Handley Page, F. (22 December 1921), "Developments In Aircraft Design By The Use Of Slotted Wings", Flight, XIII (678), p. 844, archived from the original on 3 November 2012 – via Flightglobal Archive
Perkins, Courtland; Hage, Robert (1949). Airplane performance, stability and control, Chapter 2, John Wiley and Sons. ISBN 0-471-68046-X.
Lee, Gwo-Bin. "Leading-edge Vortices Control on a Delta Wing by Micromachined Sensors and Actuators" (PDF). American Institute of Aeronautics and Astronautics. Retrieved 18 October 2018.
Effects of Wing-Leading-Edge Modifications on a Full-Scale, Low-Wing General Aviation Airplane. Nasa TP, 2011.
Bjorkman, Eileen. Gunfighters. Air & Space, November 2015. p. 62.
"Forward swept wings." Homebuiltairplanes. Retrieved: August 1, 2011.
Bedell, Peter A. "Quick Look: Hansa Jet: The ‘German LearJet’ was forward thinking, yet doomed." aopa.org, 1 February 2017.
Sweetman, Bill. "Junkers Ju287 Technology Surprise, 1945-Style." Aviation Week, 1 September 1914.
Green 1970, pp. 493–496.
Gehrs-Pahl, Andreas, ed. (1995). "The X-Planes: From X-1 to X-34". AIS.org. Retrieved 1 September 2009.
Jackson 2000, pp. 457–458.
Hallion, Richard, P. "The NACA, NASA, and the Supersonic-Hypersonic Frontie r" (PDF). NASA. NASA Technical Reports Server. Retrieved 7 September 2011.
Poulsen, C. M. "Tailless Trials." Flight, 27 May 1943, pp. 556–558. Retrieved: 1 August 2014.
Poulsen, C. M. (27 May 1943). "Tailless Trials". Flight: 556–58. Retrieved 27 February 2008.
Lewis 1962, pp. 228–229
Sturtivant 1990, p. 45.
"Issue 9 - North American F-86 Sabre: Swept wing technology". Aviation Classics. Archived from the original on 3 December 2013.
Anderson, John D. Jr. A History of Aerodynamics. New York: McGraw Hill, 1997, p. 424.
"Comment by Hans von Ohain during public talks with Frank Whittle, p. 28." Archived 9 December 2007 at the Wayback Machine ascho.wpafb.af.mil. Retrieved: 1 August 2011.
Anderson 1997, pp. 423–424.
Hermann Pohlmann; Chronik Eines Flugzeugwerkes 1932–1945, 2nd Impression, Motorbuch, 1982, pp. 190-193.
Myhra 1999, p. 4.
Waligorski, Martin. "Pulqui: Argentina's Jet Adventure." Camouflage & Markings: IPMS Stockholm, 22 September 2006. Retrieved: 27 April 2010.
Christopher 2013, pp. 157–160.
Winchester 2005, p. 37.
Ley, Willy (November 1948). "The 'Brickwall' in the Sky". Astounding Science Fiction. pp. 78–99.
Wood 1975, p. 29.
Wood 1975, pp. 34–35.
Wood 1975, p. 36.
Davies and Birtles 1999, p. 10.
Winchester 2005, p. 78.
"Eric 'Winkle' Brown obituary". The Guardian. 22 February 2016. Retrieved 13 August 2016.
Buttler 2007, p. 52.
Wood 1975, pp. 43–46.
"R.Ae.C. Award Winners." Flight International, 5 February 1954. Retrieved: 3 November 2009.
"Speed Record Again Broken?" Saskatoon Star-Phoenix, 25 September 1953.
"Wing Planforms for High-Speed Flight." NACA TN-1033. Retrieved: July 24, 2011.
Von Karman, Aerodynamics: Selected Topics in the Light of their Historical Developments, 1954.
Gunston and Gilchrist 1993, pp. 39–40.
Cook 1991, p. 152.
Gunston and Gilchrist 1993, p. 40.
Fraser November 1949, p. 139.
Werrell 2005, p. 5.
Lednicer, David. "The Incomplete Guide to Airfoil Usage." Archived 20 April 2010 at the Wayback Machine ae.illinois.edu, 15 October 2010. Retrieved: 19 July 2011.
Radinger and Schick 1996, p. 32.
Wagner 1963, .
Knaack 1978, p. 42.
Kinzey 1983, p. 4.
"Archived copy" (PDF). Archived from the original (PDF) on 17 March 2013. Retrieved 4 November 2017.
Gunston 1995, p. 184.
Seidov and Britton 2014, p. 554.
"Mikoyan-Gurevich MiG-15 (Ji-2) Fagot B.", Smithsonian National Air and Space Museum, archived from the original on 20 December 2015
Sweetman 1984, p. 11.
Crosby 2002, p. 212.
Gordon 1997, p. 124.
Belyakov and Marmain 1994, pp. 225–227.
Gunston 1995, pp. 197–198.
Erichs et al. 1988, p. 37.
Dorr 2013, p. 237.
Widfeldt 1966, p. 3.
Flight 1950, p. 558.
Boyne 2002, p. 547.
"1940s." Saab, Retrieved: 27 March 2016.
Saab 30 December 1960, p. 1017.
Gunston and Gilchrist 1993, p. 135.
"B-52 Stratofortress – U.S. Air Force – Fact Sheet Display". af.mil.
Trevithick, Joseph (19 February 2015). "I'll Be Damned, These Boneyard B-52s Can Still Fly". Medium.
Perry, Dominic (19 December 2014). "Russian air force takes first modernised Tupolev bombers". Flightglobal. London. Archived from the original on 27 September 2015. Retrieved 20 November 2015.
Andrews and Morgan 1988, p. 439.
Blackman 2007, pp. 38, 40.
Barnes 1976, p. 503.

Aerodynamics Selected topics in the light of their historical development. Dover publications, New York, 2004. ISBN 0-486-43485-0

Bibliography

Anderson, John D. Jr. A History of Aerodynamics. New York: McGraw Hill, 1997.
Andrews, C.F. and Eric B. Morgan. Vickers Aircraft since 1908. London: Putnam, 1988. ISBN 978-0851778150.
Barnes, C.H. Handley Page Aircraft since 1907. London: Putnam, 1976. ISBN 0-370-00030-7.
Belyakov, R. A. and Marmain, J. MiG: Fifty Years of Secret Aircraft Design. Shrewsbury, UK: Airlife Publishing, 1994. ISBN 1-85310-488-4.
Blackman, Tony. Vulcan Test Pilot: My Experiences in the Cockpit of a Cold War Icon. London: Grub Street, 2007. ISBN 978-1-904943-88-4.
Boyne, Walter J. Air Warfare: An International Encyclopedia, Volume 1. ABC-CLIO, 2002. ISBN 1-5760-7345-9.
Brookes, Andrew. Victor Units of the Cold War. Osprey Publishing, 2011. ISBN 1-84908-339-8.
Buttler, Tony. "Avro Type 698 Vulcan (Database)." Aeroplane, Vol. 35, No. 4, Issue No. 408, April 2007.
Christopher, John (1 June 2013). The Race for Hitler's X-Planes : Britain's 1945 Mission to Capture Secret Luftwaffe Technology. History Press. pp. 157–160. ISBN 978-0752464572.
Cook, William H. The Road to the 707: The Inside Story of Designing the 707. Bellevue, Washington: TYC Publishing, 1991. ISBN 0-962960500.
Crosby, Francis. Fighter Aircraft. London: Lorenz Books, 2002. ISBN 0-7548-0990-0.
Davies, Glyn (2014). From Lysander to Lightning Teddy Petter, aircraft designer. The History Press. ISBN 9780752492117.
Davies, R.E.G. and Philip J. Birtles. Comet: The World's First Jet Airliner. McLean, Virginia: Paladwr Press, 1999. ISBN 1-888962-14-3.
Dorr, Robert F. Fighting Hitler's Jets: The Extraordinary Story of the American Airmen Who Beat the Luftwaffe and Defeated Nazi Germany. MBI Publishing Co, 2013. ISBN 1-6105-8847-9.
Erichs, Rolph et al. The Saab-Scania Story. Stockholm: Streiffert & Co., 1988. ISBN 91-7886-014-8.
Fraser, Jim. "I Fly The World's Fastest Bomber." Popular Science, November 1949. Vol. 155, No. 5. pp. 139–142. ISSN 0161-7370.
Gordon, Yefim. "Mikoyan MiG-19 Variants". Wings of Fame, Volume 9, 1997. pp. 116–149. ISSN 1361-2034. ISBN 1-86184-001-2.
Green, William (1970). Warplanes of the Third Reich. New York: Doubleday. ISBN 978-0-385-05782-0.
Gunston, Bill. The Osprey Encyclopedia of Russian Aircraft: 1875–1995. London: Osprey Aerospace, 1996. ISBN 1-85532-405-9.
Gunston, Bill and Peter Gilchrist. Jet Bombers: From the Messerschmitt Me 262 to the Stealth B-2. Osprey, 1993. ISBN 1-85532-258-7.
Seidov, Igor and Stuart Britton. Red Devils over the Yalu: A Chronicle of Soviet Aerial Operations in the Korean War 1950–53. Helion and Company, 2014. ISBN 978-1909384415.
Jackson, Paul, ed. (2000). Jane's all the World's Aircraft 2000–01 (91st ed.). Coulsdon, Surrey, United Kingdom: Jane's Information Group. ISBN 978-0710620118.
Kinzey, Bert. F9F Cougar in Detail & Scale. Fallbrook, California: Aero Publishers, Inc., 1983. ISBN 9780816850242.
Knaack, Marcelle Size. Encyclopedia of US Air Force Aircraft and Missile Systems: Volume 1 Post-World War II Fighters 1945–1973. Washington, DC: Office of Air Force History, 1978. ISBN 0-912799-59-5.
Lewis, Peter (1962). British Aircraft 1809-1914. London: Putnam Publishing.
Mendenhall, Charles A. Delta Wings: Convair's High-Speed Planes of the Fifties and Sixties. Motorbooks. 1983.
Myhra, David. Focke-Wulf Ta 183 (X Planes of the Third Reich). Atglen, PA: Schiffer Publishing, 1999. ISBN 978-0-7643-0907-6.
Radinger, Willy and Walter Schick. Me 262: Entwicklung und Erprobung des ertsen einsatzfähigen Düsenjäger der Welt, Messerschmitt Stiftung (in German). Berlin: Avantic Verlag GmbH, 1996. ISBN 3-925505-21-0.
"Saab-29: Sweden's new jet fighter." Flight International, 4 May 1950. pp. 556–58.
"Saab: Sweden's Advanced Combat Aircraft." Flight International, 30 December 1960. pp. 1017–20.
Spick, Mike and William Green, Gordon Swanborough. Illustrated Anatomy of the World's Fighters. Zenith Imprint, 2001. ISBN 0-7603-1124-2.
Sturtivant, R. (1990). British Research and Development Aircraft. G.T. Foulis. ISBN 0854296972.
Sweetman, Bill. Modern Fighting Aircraft: Volume 9: MiGs. New York: Arco Publishing, 1984. ISBN 978-0-668-06493-4.
Wagner, Ray. The North American Sabre. London: Macdonald, 1963.
Werrell, Kenneth P (2005). Sabres Over MiG Alley. Annapolis, Maryland: Naval Institute Press. ISBN 1-59114-933-9.
Whitcomb, Randall. Avro Aircraft and Cold War Aviation. St. Catharine's, Ontario: Vanwell, 2002. ISBN 1-55125-082-9.
Winchester, Jim. "Bell X-5." Concept Aircraft: Prototypes, X-Planes and Experimental Aircraft. Kent, UK: Grange Books plc., 2005. ISBN 1-84013-809-2.
Wood, Derek. Project Cancelled. Indianapolis: The Bobbs-Merrill Company Inc., 1975. ISBN 0-672-52166-0.

Further reading
"The High-speed Shape: Pitch-up and palliatives adopted on swept-wing aircraft", Flight International, 2 January 1964

Physics Encyclopedia

World

Index

Hellenica World - Scientific Library

Retrieved from "http://en.wikipedia.org/"
All text is available under the terms of the GNU Free Documentation License