Category Archives: US

Boeing Model 299G

United States of America

Heavy Bomber Design Study

Model 299G as it appeared in a microfilm document from a private collection on the B-17C. Note this is a recreation. (Bill Stanczak)

Introduction

There are very few planes in military and aviation history that have garnered as much attention or popularity as the Boeing B-17 “Flying Fortress”. The aircraft has been called by some “the best bomber of the Second World War”, although there are other contenders for that title. Opinions aside, one cannot dismiss the impact of the B-17 on military history and the evolution of strategic bombing. The development of the B-17 initially started with Boeing Model 299, often incorrectly called the Boeing XB-17 in various sources. Surprisingly, the B-17 was initially not selected for production, as the Model 299 prototype was destroyed in an accident and the US Army Air Corps’ limited budget did not allow for the purchase of the more expensive bomber. As this is such a popular aircraft, one would assume that quite a good portion of what there is to know about the plane and its development has already been researched, and documented. However, as is often the case, there are always discoveries waiting to be found, such as one particular obscure variant of the B-17, the Boeing Model 299G

To say that there is little to no information on this Model 299G would be quite an understatement as, aside from a few drawings, there is nothing that indicates why this aircraft was designed and what its exact purpose was. However, after studying the documents as well as consulting with several well-known aviation experts, it has become somewhat clear that what the Model 299G represents is not just an attempt to create a new and more effective variant of the B-17 based on the already in-production B-17B but, in fact, was a design concept that proved extremely influential in the design of the B-29 “Superfortress”.

Boeing: The American giant and a leader in aircraft design

The Boeing Company exemplifies the idea of rising from small beginnings. It was founded in 1916 on the shores of Lake Washington by a young timber baron by the name of William Boeing, who had an interest in aircraft. The first Boeing aircraft, a seaplane, took off from the shores of Lake Washington in January 1916. However, the company did not really take off until the 1920s and 30s, when Boeing achieved many great and public feats, including designing and building the first all-steel tube fuselage with its then-innovative arc welding process and even becoming one of the first companies to build dedicated mail aircraft. It was also during this time that Boeing would design and build some of its most legendary aircraft, such as the P-26 Peashooter, which, when introduced, was considered to be one of the fastest fighter aircraft in existence.

The company would gain even more fame and recognition with its construction of the Boeing Model 247 in the early 1930s, which allowed Boeing to dominate the early modern airliner market until the introduction of the Douglas DC-2 and the later Douglas DC-3. The Model 247 was considered to be extremely technologically advanced for the time and represented Boeing’s shift to all-metal aircraft construction. Boeing received even greater fame with its development and construction of the Boeing Model 299, which later became the B-17 Flying Fortress, an aircraft that was very well-liked by the top staff of the US Army Air Corps. The B-17 design would also later allow Boeing to create several other highly influential and popular designs based on the Model 299. These designs included civilian aircraft, such as the Boeing Model 307 Starliner and the famed Boeing Model 314 Clipper, which saw great fame while flying for Pan-American Airways in the late 1930s and even saw service as a Presidential transport aircraft for Franklin Delano Roosevelt. The Model 314, while externally quite different, used the same wing structure and design as the Model 299. The Model 299 design also brought forth various military variants which would see various uses, including transport aircraft in the case of the C-108 and VB-17. The Model 299’s development would ultimately culminate with the so-called “ultimate B-17”, the B-17G, which went on to become one of the most popular and well-known variants.

The Birth of Boeing’s big bombers

The development of the Boeing B-17 began in February of 1934, with a US Army Air Corps request for proposals for a new bomber with a range of 5000 miles (8046 kilometers) and a bomb load of at least 2000 pounds (907 kilograms). This request, designated “Project A”, was only a feasibility study for a production aircraft to these requirements. Even with it being a proposal, there was a chance the aircraft would be built, and Boeing put its best designers and engineers on the project and was clearly interested in developing the design. These designers and engineers soon found success, as they were able to successfully design and later build a very good aircraft. Initially, Boeing submitted the XBLR-1 (Experimental Bomber Long Range) for this program, which was later re-designated XB-15 upon its construction. Their competitor, Martin, also submitted a project, the XB-16, but that was later canceled before it actually left the drawing board, nor was a final design for it completed. Because of this, the XB-15 would remain the only bomber built in the XBLR program and was the largest until the Douglas XB-19 was built. The XB-15, while never serving as a bomber as intended, would eventually see service as a transport under the designation XC-105 and would serve until its retirement in 1944. Following its retirement, the aircraft would be partially dismantled and dumped in the so-called “Diabalo Dump”, where it remains to this day.

The Boeing Model 299 and the B-17

 

Boeing Model 299, often incorrectly called XB-17. (National Museum of the USAF)

In May of 1934, the US Army Air Corps announced a second competition, this time for a multi-engine bomber capable of carrying a ton of bombs, having a range of 2000 miles (3219 kilometers), and capable of flying at over 200 miles per hour (173 knots or 321 km/h). Unlike the previous competition, however, this aircraft would be built and brought into limited service, with a potential for full production. For this competition, Boeing decided to design and build what, in essence, was a scaled-down Model 294 (XB-15) under the designation Model B-299. The Model B-299 took many of the base features of the Model 294 and improved on them while scaling down the aircraft. In this regard, it was much like the 294, a twin-wing monoplane with four engines, but it also combined elements of Boeing’s successful Model 247 passenger aircraft. The prototype Model 299 first flew on 28 July 1935 and was very quick to impress the US Army Air Corps as well as the assembled press, with one reporter describing it as a “Flying Fortress”, and the US War Department describing it as an “Aerial Battle Cruiser”. On August 20, the Model 299 was flown to Wright Field, where it would spend the next two months being tested against the Martin 146 and the Douglas DB-1 (B-18 Bolo), where it eventually performed above and beyond the base requirements. The 299 would eventually get the US Army’s stamp of approval as well as an order for 65 YB-17s. However, on 30 October 1935, disaster struck and the Model 299 crashed and burned on takeoff. While the official cause was deemed to be a pilot error, as the pilot had forgotten, due to the lack of a checklist, to unlock the control surfaces (it was this accident that introduced checklists as standard equipment on aircraft), the US Army would cut the order to only 13 planes, designated Y1B-17, and instead ordered the production of 133 Douglas B-18 Bolos. The reason for this decision was twofold. While the destruction of the Model 299 did impact this decision, it was ultimately the US Army’s limited budget and their lack of funding that led them to ultimately choose the B-18 Bolo, as it was the only aircraft they could really afford a large number of. Despite this setback, the US Army was still enthusiastic about the design and allowed Boeing to submit another prototype for evaluation, which they did in the form of a modified Y1B-17 with more powerful engines and a crew of 6 instead of 7.

Boeing Model 299, note the distinctive nose with a small turret. (National Museum of the USAF).

The Boeing Y1B-17 did not differ too much from the original Model 299, however, some improvements were made, including switching the engines to the more powerful Pratt and Whitney R-1820s and changing the design of the landing gear arms. It was this prototype that ultimately won Boeing the contract and would go on into production as the Boeing B-17B.

The Model 299G: A modified B-17 or Something More?

When looking at the B-17’s lineage, one will notice that the very first mass-produced variant of the bomber was the Boeing B-17B or, as the Boeing Company knew it, the Model 299E (later changed to 299M). The B-17B followed a long line of prior limited or prototype variants, including the base Model 299, later Y1B-17, and Y1B-17A. The production run of the B-17B only ran for a total of 39 aircraft before it was switched to the B-17C (Model 299H). According to the documentation and the drawings found, the Model 299G was considered to be a very heavily modified B-17B which was re-engined with the Pratt and Whitney R-2180 Twin Hornets instead of the Pratt and Whitney R-1820-51 Cyclone. Beyond this, unfortunately, the drawings give very little information on this aircraft or really what exactly it was supposed to be. However, according to historians such as Mike Lavelle, this variant may be a link in the greater chain of designs that led to the Model 345, better known as the B-29 Superfortress.

The Design of the Model 299G

Boeing model 322 study, this one lacking the forward gun position but otherwise closely related to the 299G. (Lavell)

The Model 299G is unique compared to other B-17 variants and designs based on the B-17. It shares very little similarity with the Model 299 and Model 299M (B-17B) designs it is based on. Outside of the tail section and some other components, such as the general design of the wings, the rest of the aircraft is almost a completely different design from the B-17B on which it is based. Among the interesting features is the tricycle landing gear arrangement (one wheelset in the front, two on the wings). The aircraft also features a cockpit section very similar to that of the Boeing Model 307, completely eliminating the turtle deck. It shares a similar fuselage to the Stratoliner as well, as it was designed to test the feasibility of pressurization for use in bombers. Another major aspect that stands out about the aircraft is that it appears to have been both wider and longer than the B-17, with a slightly larger wingspan. Also featured were 4 defensive weapon blisters that almost seem like a cross of those on the early B-17s and those featured on the later PB4Y-2 “Privateer”. These were situated on the dorsal and ventral sections of the waist, with the ventral one just behind the wings and the dorsal one farther aft and closer to the tail.

Boeing Model 307 Stratoliner on display at the National Air and Space Museum Udvar Hazy Center. This is the sole preserved complete example. (Smithsonian)

The Model 299G also did not feature an astrodome. Rather, it featured what appears to have been a dedicated observation area above the cockpit. Perhaps the only major similarity it shared with the B-17 was that the 299G was a monoplane and, the wheels still receded into the engine nacelles. The design, as such, does not really seem to have been that of a B-17, bearing the most similarity to Boeing’s Model 307 Stratoliner, which was later adopted into US Army Air Force service as the C-75 Stratoliner. It also more clearly resembles Boeing’s later bomber designs, such as the Model 322, which eventually led to the development of the Model 345, better known as the B-29 Superfortress.

Conclusion

Boeing Model 316, a later design study possibly derived from the Model 299G. These two, and several others would go on to influence the further development of the B-17 and B-29. (Lavell)

While, ultimately, the Model 299G never left the drawing board, it certainly represents an interesting insight into the developmental history of Boeing’s large bomber projects. Based on conversations with several Aviation historians, it has been presented as a possibility that this Model 299G could also have been a very early attempt to design a sort of “Superbomber” that members of the so-called “Bomber Mafia”, including Jimmy Doolittle and General Hap Arnold, had been searching for. This conclusion would indeed make sense, as many of the features of the Model 299G do seem to correspond with later Boeing bomber designs. It has also been suggested that the Model 299G might have been a link in the greater developmental chain of the Boeing Model 345, which eventually saw service as the B-29 Superfortess. Some, however, have also suggested that this aircraft instead represented a link between Boeing Airliner development and their Military Aviation development. However, as of this writing, there is no concrete information or documentation that directly links the Model 299G to the Model 345, though it and other projects were part of the B-29 program’s design studies. Aside from general appearance, there is also really nothing concrete to link the 299G to airliner development either. What is undeniable though is that Model 299G does offer deeper insight into the continued development of the B-17 Flying Fortress and the influence, if indirect, it had on future projects.

Variants

Model 299G – The Boeing Model 299G was designed by Boeing and based on the Boeing B-17B. It never went past the design stage and was not selected for production.

Operators (Projected)

 

  • United States of America
    • US Army Air Corps (Presumed) – The Model 299G was designed by Boeing but never made it past the design stage.

Illustration

 

Credits

  • Article written by J. Manuel
  • Edited by  Henry H. & Stan L.
  • Ported by Henry H.
  • Illustrated by Ed Jackson

Sources

Baugher, J. (1999, July 25). Retrieved from http://www.joebaugher.com/usaf_bombers/b17_1.html

The Boeing Company (2020, December 20). Retrieved from

http://www.boeing.com/history/#/legacy

Harris, S. M., & Angelucci, E. (1983). The Rand McNally Encyclopedia of military aircraft: 1914-1980. New York: Military Press.

Model 299 Crash. (2009, June 25). Retrieved from https://www.nationalmuseum.af.mil/Visit/Museum-Exhibits/Fact-Sheets/Display/Article/610002/model-299-crash/

Model 299 Press Release. (2009, June 25). Retrieved from https://www.nationalmuseum.af.mil/Visit/Museum-Exhibits/Fact-Sheets/Display/Article/610003/AFmuseum/

Lavelle, Mike. War on the Home Front: Building the B-29 Superfortress. Chester River Press, 2011.

Lavelle, Mike, and Matzelle, Liz. “Fwd Boeing 299G.” Received by Jonathan Manuel, 21 Nov. 2021

Simons, Graham M. The Boeing B-29 Superfortress: The Giant Bomber of World War Two and Korea. Pen Et Sword Aviation, 2012.

North American P-51B Mustang

sweden flag USA (1943)
Fighter Aircraft – 3,738 Built

 

A P-51B undergoes testing at a Lockheed reassembly plant in Liverpool, UK. December 1943. [National Archives]
Initially developed to provide an export alternative to the P-40 for France and the UK, North American’s P-51 would prove to be a superb aircraft that would rank among the most decisive weapons of the Second World War. With its streamlined airframe and highly efficient cooling system, the aircraft would reach new heights when equipped with the far more advanced Packard Merlin engine. Though its early years would prove troublesome, it would solve long standing issues regarding the lack of long range bomber escorts, and achieve a level of performance beyond its Axis contemporaries.

Interwar Fighter Developments

The Merlin powered P-51’s share the distinction of being among the most successful fighter aircraft ever developed, but also having one of the convoluted development paths of any mass production fighter. While the aircraft would make its first flights in 1943, it had its roots in the late interwar period where many of the technologies it incorporated were first established.

US interwar fighter development saw rapid technical advancement, but a comparatively small build up of planes. Here an XP-40 undergoes wind tunnel testing, the design would go through a number of changes that would result in the P-40. [This day in Aviation]
The general environment of interwar fighter development for the US Army Air Corps was one of high theoretical advancement, but comparatively slow practical development. While major milestones were made in regards airframe and powerplant design, there was considerably less urgency to develop and mass produce fighters for use by the Air Corps. This was mostly a result of an isolationist foreign policy, which limited availible resources, and to a lesser degree, a desire within the Air Corps to focus on bomber procurement. While the development of new fighters was limited, the Air Corps had great freedom in procuring aircraft for testing purposes. While funding was still limited, they were allowed to procure up to 14 examples of an aircraft through their budget before they would need to petition Congress for additional funding. While a large build up of the Air Corps during this period was a financial and political impossibility, it would prove sufficient for exploring aircraft design and development. This environment would exist into the late 1930’s as the political situations in both Europe and Asia destabilized, and subsequently, the order was given to continue the development of the XP-38, XP-39, and XP-40 into new fighters for use with the Air Corps (Ethell 9).

While these aircraft were being prepared for service, vital new developments were being made in regards to airframe design. At the National Advisory Committee for Aeronautics (NACA) offices at Langley field, efforts had been made to produce airfoils which could achieve laminar flow. In short, this effect is characterized by minimal disruptions to the airflow of the surfaces of the wings and adjoining fuselage. In the context of fighter aircraft, this allowed for a much lower drag coefficient, which would permit better acceleration and would lessen the instability encountered at higher Mach numbers. They would achieve this by June of 1938 when an airfoil displayed laminar flow characteristics in wind tunnel tests (Ethell 10).

Europe Ablaze

The escalation to and the outbreak of hostilities in Europe would completely dispense with the interwar malaise and saw the US begin a massive arms build up. The most notable shift in policy was President Franklin Delano Roosevelt’s call for 50,000 aircraft in January of 1940. The resulting surge of orders would end up leaving most US aircraft manufacturers at capacity, and though they would satisfy domestic demand, the fulfillment of export orders was not a priority. This represented a serious issue facing the Allies in Europe. At the outbreak of the war, the French and British air forces were still largely in the process of expanding and modernizing. While they both possessed examples of modern fighter aircraft, such as the Dewoitine D.520 and Supermarine Spitfire Mk.I’s respectively, they also employed a large number of outdated aircraft in comparison to the better equipped German Luftwaffe. The expedient solution to this problem seemed to be to purchase aircraft abroad, and the US was by far the best source.

To this end British and French interests were served by the British Purchasing Commission. While they had decided on the ideal candidate being the Curtis-Wright P-40, they found the at-capacity firm unwilling to compromise its contracts to the US Army. They were soon negotiating with other firms for P-40’s which would be manufactured under license, and by 1940 had placed as many orders as they could. It was clear to all parties involved that any of the larger firms that were involved in US rearmament would be unable to deliver any sizable number of aircraft to the Allies. In January of 1940, Oliver Echols, in charge of Air Corps procurement, would suggest to the Purchasing Commission to approach a manufacturer that lacked any major contracts involved with US rearmament (Ethell 10).

This suggestion would see the British Purchasing commission returning to older offers from firms that they had turned down the previous year. The most important of these would be North American Aviation. North American had earlier proposed to build P-40’s under license for the Allies, though the offer was given little consideration (Ethell 10). They were likely turned down over their relative inexperience in the field of fighter aircraft, having previously built advanced trainers, like the AT-6 Texan, and the crude NA-50 and NA-68 export fighters. In spite of this, and finding few options among other US aircraft manufacturers, the British Purchasing commission would once again approach North American. This time however, North American was given the option to either produce license-built P-40’s, or instead to design a new aircraft with the aid of research data acquired from Curtiss-Wright on the XP-46 fighter prototype. NAA’s small, but enthusiastic team would choose the latter, and prepared to design a new fighter built around the Alison V-1710 engine.

Enter North American

North American’s greatest claim to fame before the Mustang was the AT-6, arguably the best advanced trainer of its day. [Wikimedia]
By the standard’s of most US industries of the time, North American Aircraft was a fledgling company, though one with great promise. It was originally formed as a holding company in 1929 to purchase stock in other aviation concerns, and was later incorporated under General Motors’ General Aviation branch. As a holding company, North American would gather a considerable amount of resources in these early years, of particular note was the firm’s acquisition of Fokker. In 1934, as a result of new regulations on air mail carriers, General Motors was required to divest itself of North American, which then became an independent firm. Thereafter, North American incorporated its parent company, General Aviation, and continued under the direction of its president James H. ‘Dutch’ Kindelberger (O’Leary 9). He would subsequently take the company west in 1936 where they would open a new facility at Mines Field, California. Prior to the war they would develop the O-47 reconnaissance and observation aircraft, which had begun under General Aviation, and the AT-6 advanced trainer, which was among the most successful designs of its type. They would also produce a set of unsuccessful export fighters which were altogether unimpressive. With this in mind it’s understandable North American was initially passed over, they were in fact, inexperienced in fighter development and their only real foray into that field was a disappointment. However, when the British Purchasing commission returned to the company in 1940, they found the firm more than ready to meet their needs. Their contract was worked out for 400 planes at a price no higher than $40,000 dollars a unit, and spare parts in the amount of 20% of the value of the aircraft.

The first step in developing the new fighter was purchasing the most recent data on fighter design from Curtiss-Wright’s XP-40 and XP-46 prototypes, and acquiring the new breakthrough aerofoil designs recently developed under NACA (Ethell 10, 11). This information was made available to the design team headed by Edgar Schmued, a German born aeronautical engineer who had previously been a GM field service manager for their Brazil branch. The work soon began on a new fighter under the designation NA-50B, later changed to NA-73, under a common and straightforward design strategy. Schmued would work to build a plane that would excel by incorporating all of the most recent developments in fighter design to produce an aircraft that was both cutting edge, yet conventional (Douglas 252). The Curtiss-Wright prototypes were a starting point that was quickly surpassed, with engineer and aerodynamicist Ed Horkey considering the prototypes too dated for use on the new project, and the data was discarded (Forsyth 13). This came as somewhat of a blow considering they were forced to pay about $50,000 for the test data. The same cannot be said for the data acquired from NACA.

Edgar Schmued would join North American through its parent company’s acquisition of Fokker. He would lead the team responsible for designing the Mustang which would be developed continuously through the Second World War. [alchetron]
Horkey would come across NACA’s research through a confidential release for American industrial use, and was convinced that it would make an excellent addition to the new fighter’s design. NAA would send a representative to collect the data from NACA at Langley Field, and they would go on to receive minor technical support. While the design did not possess true laminar flow characteristics, it did drastically reduce drag and improve the performance of the aircraft (Ethell 11). Further streamlining was achieved through the mounting of a low drag, centerline radiator which incorporated the work of British scientist, Dr. F.W. Meredith. This divergent-convergent duct was capable of using the heat ejected by the radiator to actually produce thrust and offset some of the speed loss incurred by drag incurred by the radiator’s air scoop (Douglas 252).

 

Great care was taken to build the prototype in good time. The NA-73X, would make use of a number of components from North American’s AT-6 trainer, including its landing gear, hydraulics, and electrical systems. Remarkably, the construction of the prototype was completed on the 102nd day of the project, but it would have to wait another 20 days for its Allison V-1710-39/F3R engine (Marshall & Ford 94). The supply of Allison engines at the time was constrained, and resulted in the project having to delay its deliveries to the British. Despite this, the fast pace of the program, and the fall of France would see the British order another 320 aircraft before the prototype even flew. With the program approaching testing, the British were awaiting the results and readying their own test pilots to become acquainted with the new plane. The prototype was first flown by American test pilot Vance Breese on the 20th of October, 1940. It would go on to make several more test flights before having to be repaired after an accident with test pilot Paul Balfour. The accident was a result of pilot error, who failed to switch over from an empty fuel tank, and as such the incident did not reflect poorly on the design itself (Marshall & Ford 151). As the sleek new fighter was taking shape, the British Purchasing Commission would notify NAA that the aircraft’s RAF designation was to be the ‘Mustang’ in a communique sent in December 1940 (O’Leary 24).

This prototype NA-73 was delivered to the US Air Corps for testing, though they would not place orders for Mustangs until a later date. [This day in aviation]
Among the last modifications to the NA-73 regarded its armament, fuel capacity, and reinforcement of its wings. Several proposals for its armament were considered, but for the British Mustang they installed a pair of .50 caliber guns in the nose cowling with another two .30 caliber guns in each wing. With these last additions made, the British soon received several of the new aircraft, which now bore the more familiar title of Mustang. The first, AG345, would be put through tests to find any issues from the transition from the NA-73. Several issues arose over the stiffness in the ailerons, power surges in dives, and overheating. These were subsequently addressed, though more drastic changes were needed in the case of the engine, which required installing a new carburetor scoop, and altering the scoop for the radiator (Marshall & Ford 165). The culmination of these new changes would result in the finalized Mustang Mk.I, and a second development prototype, NA-83.

While the aircraft’s development was proceeding at a rapid pace for the British, the USAAC would show very little initial interest in the Mustang. The aircraft the USAAC had dubbed the XP-51 was largely overshadowed by other developments and comparatively little effort was made to conduct exhaustive tests on the XP-51 prototypes at Wright Field to correct their faults. Their interest in the aircraft would be piqued only after the U.S. entrance into the second World War.

Mustang Mk. I

In British service the Mustang would take a different developmental path than what was proceeding in the United States. While the British were receiving their Mustang fighter aircraft, the US had been forced to develop the aircraft into a dive bomber, the A-36, as funds for fighter development had been expended for 1942. In the case of the RAF, the Mustang Mk. I went into service as soon as was practicable and saw their first squadrons, numbers 161 and 613, receive supplies of the new aircraft in April of 1942. They would first be employed as reconnaissance aircraft before later taking on more dangerous work during Operation Jubilee in which they undertook offensive recon sorties over the raid area in Dieppe, France. Beyond this they would be subsequently used to fly nuisance raids and fighter sweeps across the low countries. Its long range, high speed, and effective armament were used to great effect over these areas as they harassed rail and road communications, while also remaining quite capable against enemy fighters wherever they were encountered (Ethell 24, 25). Even by this early mark, the once uncertain contract they signed with North American had already paid off.

It was during this period that the aircraft’s faults and strengths would make themselves evident. The nose mounted guns were troublesome and complicated maintenance; they were often removed from operational planes and were eliminated from the succeeding models of the aircraft. The radiator still presented teething issues, as under certain conditions the oil could freeze over and would fail to circulate, and eventually cause the radiator to boil over. Visibility too would become an issue, as the canopy frame of the cockpit severely restricted the pilot’s view. However despite its faults, the plane was fast, possessing good acceleration and a high top speed that made it capable of outrunning all fighters in the theater at sea level (Ethell 24).

The Mustang Mk.I would prove an exceptional fighter with the RAF, if at first, a little rough around the edges. [wikimedia]
While the radiator issues would be addressed and a new bubble canopy was developed, another, more serious drawback of the design would require far more resources to address. The Allison engines that the early Mustangs were equipped with were considerably lacking when it came to high altitude performance due to their single stage, single speed superchargers. While the aircraft received good marks for its low altitude performance from pilots in the RAF, above the 15,000ft the Allison V-1710 suffered considerable power loss. Though this was by no means surprising, it represented an area where performance could be significantly improved. At higher altitudes the aircraft was outpaced by both contemporary models of the Fw 190 and Bf 109. At low altitudes, it was made somewhat redundant by the RAF’s new Hawker Typhoon, which both flew faster at low altitudes and was better armed. It wasn’t long until the idea arose to fit the Mustang with an engine possessing better high altitude performance, a combination that might well produce an exceptional fighter that was as capable at high altitude as it was down low (Douglas 253).

The first major step toward this came on April 29, 1942, when Wing Commander Ian Campbell-Orde invited one of Rolls Royce’s test pilots, Ronald W. Harker, to test the new aircraft. Harker was impressed by its performance and he believed that if the aircraft was fitted with the new Merlin 61, it would be able to outpace a similarly equipped Spitfire by a considerable margin (Marshall & Ford 215). The Merlin 61 was the obvious choice for many reasons, chief of which was its two stage, two speed supercharger which stood to offer the plane exceptional high altitude performance. To this end, a Mustang Mk.I was provided to Rolls Royce at Hucknall to undergo the necessary modifications. By the beginning of June 1942, the British had correctly projected that the Mustang’s top speed would be increased to 430mph at an altitude of 25,000ft, which was roughly twice as fast as the Allison powered Mustang at that altitude (Douglas 254). When the test aircraft was complete the results were quite impressive, as during a fly off between a Spitfire Mk. IX and a Mustang, both fitted with Merlin 61’s, the Mustang quickly outpaced the Spitfire.

Across the Atlantic, a parallel development began underway after a study of the Mustang’s combat debut with the RAF. The new United States Army Air Force, no longer constrained by funding, rushed to acquire supplies of the Mustang, and sought to re-engine the fighter to improve its high altitude performance. To this end, two P-51’s were set aside for conversion. By the early half of the Summer of 1942, both British and American Mustang experiments were underway. While the Mustang was previously seen as a side project which was never a wholly American or British effort, it was by then extremely clear that the design had tremendous potential and the development of which was of immense importance to the Allies.

Shoeing the Mustang

Orders for various Mustang types for the USAAF would begin in 1942, including this P-51 armed with four Hispano 20mm cannons. These orders were quickly overshadowed by developments to get the Packard engine into the aircraft. [Wikimedia]
Re-engining the Mustang was by no means an easy task, as the Merlin was considerably heavier than the Allison and required a larger cooling system. To achieve this, the radiator was reworked, with the oil cooler moved apart from the radiator matrix to a forward position, and the ducting of the entire scoop assembly being redesigned. Earlier aerodynamic and buffeting issues caused by the radiator intake were also resolved by moving the scoop out of the boundary layer under the fuselage. The resulting set up would also achieve the earlier described Meredith effect, which produced thrust that offset the drag caused by the scoop (Marshal & Ford 97, 219). Additionally, the carburetor’s intake duct was moved beneath the nose which also necessitated lowering the wing to accommodate the lower cowl.

In addition to higher cooling requirements, the new Merlin engine weighed 350lbs more than the Allison and would mount a larger, heavier propeller, which would represent a significant shift in weight. To compensate, 61lbs of ballast was added, the primary fuselage longerons were strengthened, and the wings were strengthened and moved lower and forward. These changes would also help to compensate for the stronger vortex generated by the propeller and the greater forces generated by the improved ailerons (Marshall & Ford 219). The new engine and the subsequent operations would also result in some yaw instability. Adding a fin ahead of the horizontal stabilizer seemed an adequate solution, but it would not be undertaken until far later.

While the testing for most of these modifications was done through a variety of converted air frames, the prototype that brought them all together was the XP-51B, which first flew on October 1, 1942. The importance placed on this aircraft was considerable, as several months prior, a large order for 1200 P-51A’s was placed by the US government on the provision that their production could be switched for P-51B’s, given advanced notice (Marshall & Ford 230).

The first of two XP-51B’s would be ready in October of 1942, however, a long and difficult development process would delay serial production until the summer of the following year.[Thisdayinaviation]
The XP-51B would prove promising but it was troubled by radiator issues which would remain with the aircraft through January of 1943. These were tracked down to a chemical reaction which was found to be degrading the coolant tubes, and was resolved by a lacquer liner. There were also air flow issues within the radiator, which were solved through moving its aftercooler core to improve airflow through the scoop. The prototype’s last major issue was the tendency for its air scoop to produce loud, and worrying, vibrations at high speed. Resolving the problem once again required them to change the geometry of the scoop (Marshall & Ford 258, 311). This was solved by the aforementioned modification that moved it out of the boundary layer below the wing, and further improved as the depth of the gutter was increased and the inlet size was reduced (Matthews 7).


Most of the issues with re-engining the P-51 involved its cooling systems and air scoop, which were revised several times. [NACA]
All of the production models of the Merlin powered P-51’s would fly with engines produced under license by Packard. It was a matter of good fortune that Packard was already engaged in the mass production of their version of the Merlin engine prior to the demand for the engine for the new Mustangs. Packard had built its first V-1650-1’s, a license built Merlin 28, in August of 1941 which were later destined for use in Canadian built Avro Lancasters, DeHavilland Mosquitos, and the updated Curtis Wright P-40F (Marshall & Ford 176). Changing production to suit the needs of the P-51B would however not be easy, and matters were made worse by a general strike at the main plant which, alongside slow development at Wright Field, made for considerable delays. Some of the supply issues would be addressed as the new Mustangs would receive the first priority in terms of supplies, superseding the P-40F and L, and denying its use on the P-38. However, beyond these were the predictable teething troubles, and combined with the less predictable hurdles, they saw widespread deployment of the P-51B delayed considerably. Packard would go on to supply North American with engines, however they would never fully be able to meet the massive demands of both the United States and Great Britain (Marshall & Ford 347).

My Kingdom for a Horse

While development on the Merlin powered P-51’s proceeded, the USAAF had formulated and launched a strategic bombing campaign dedicated to destroying industries vital to the German war effort. The theoretical foundations of this strategy had been set in the interwar era and were initially seen as a means to expand the Army Air Corps into a force with greater autonomy. Many early interwar theorists, such as Maj. Harold George, would describe a vague ‘economic web’ that could be destroyed and force an industrial and morale collapse, but in 1943 these theories were put to the test. The practical details of the campaign were laid out at the Allied conference at Casablanca. There a series of targets was decided upon, but later altered to a plan that favored targeting aircraft and submarine production, in addition to ball bearing plants (Overy 45, 305). However, the main concern for USAAF bombing operations was that thus far, all daylight strategic bombing campaigns had ended in failure after formations of unescorted bombers were shredded by fighters.

The USAAF bombing campaign against Germany began in earnest in early 1943, it was based on a number of untested theories which planners hoped would bring an early end to the war. [National Archives]
Since before the war, it was commonly believed among the Air Corps senior officers that a formation of well armed bombers was capable of defending itself from whatever threats it might face. This assertion would be disproven, as even the small raids against targets in France and the low countries sustained casualties that made consistent raids impossible. In early 1943, the next step of the campaign would be far more ambitious, moving on to targets deeper within Germany itself. The need for a long range escort fighter had already become apparent before this point, and work was underway to produce external fuel tanks for existing fighters, but the offensive would be continued without a fighter aircraft able to accompany raiders for the full duration of their missions.

Throughout the summer and autumn of 1943, the USAAF would launch numerous raids against targets in Western Germany, though the bombers could only be escorted over the low countries by P-47’s and P-38’s. It wasn’t long until these range limitations were understood, and soon after, exploited by the Luftwaffe. Wherever Luftwaffe fighters were untroubled by Allied fighters, they were free to make use of their most effective anti-bomber tactics.

Generalmajor Adolf Galland’s prescribed method of attack for single engine fighters was to make head on, or oblique, attacks from slightly above the bomber formation, carried out by at least a Schwarm, or two pairs of fighters (Marshall & Ford 267). This achieved two things, it increased the closure rate to reduce the likelihood of being hit by defensive gunners, and it was from this position that both the pilot and copilot of the bomber were most vulnerable. In the absence of escort fighters, Luftwaffe pilots would be able to regroup, fly ahead of the formation, climb, and repeat the attack. The lack of escort fighters also meant the Luftwaffe was safe to employ its two engined fighters against bomber formations, which with their heavier armaments, were much better equipped to bring down bombers. Over time their tactics grew even more complex as dedicated aircraft, typically Ju 88’s, were tasked with shadowing bomber formations to pass their altitude, course, and speed to flak and fighter control services.

Many Luftwaffe aircraft would be re-equipped to take on heavy bombers, like this Bf 109G-6 with its underwing 20mm gun pods. [Bundesarchiv]
Prior to the arrival of the P-51’s, the USAAF had two suitable fighters for the purposes of escorting bombers at high altitude, the P-38 and P-47. While they had the high altitude performance, they did not have the range to reach deep into the continent. The issue would be partially resolved through the addition of external fuel tanks, which had been discussed at a conference with the Material Division at Wright Field in March of 1942 (Ethel 51). Work however, was slow and the 108 and 75 gallon drop tanks were not delivered in large numbers until the end of summer, 1943.  These tanks would allow the shorter ranged P-47 to be able to cover bombers over their flight over the low countries, and the P-38, over the Rhineland. It should also be noted that the escort range was considerably lower than the maximum combat range of the aircraft, as the planes flew in a zig-zag pattern overhead so as not to out pace the bombers. Supplies of larger volume fuel tanks which would take the fighters further into German air space would not be available until the spring of the following year. External fuel tank development and procurement had been mismanaged by Army Air Force leadership who were still largely convinced that the bomber’s defensive capabilities were adequate. Had there been a greater supply, and larger volume tanks initially available, the P-47 and P-38 could have escorted bombers over most of Germany. To make matters worse, the P-38, which by then handled the most important leg of the trip, was troubled by a number of technical issues. While the P-38 possessed good high altitude performance, an exceptional climb rate, and a heavy armament, it was handicapped by a cockpit that pilot’s rated the worst of any US fighter in service and had flying characteristics that made it difficult for pilots to aggressively pursue Luftwaffe aircraft (Dean 164). The large, twin engine Lightning also had an unmistakable appearance, such that Luftwaffe pilots would almost always spot and identify the Allied plane before Lightning pilots could do likewise. With this benefit, Luftwaffe pilots were typically the ones who dictated the engagement, and would depart when conditions were unfavorable. On the defense they would have another advantage, both the Fw 190A and the Bf 109G were capable of out maneuvering the P-38 in high speed dives. The P-38 encountered severe compressibility issues at speeds significantly lower than those encountered on the two German fighters (Marshall & Ford 441). Thus, while the P-38 was capable of performing long range escort missions, its pilots would be forced to employ more conservative tactics than those used in the P-47.

By the start Autumn of 1943, USAAF planners were hoping to accelerate their progress on Operation Pointblank. This plan would see bombers raid targets that were vital to the German aviation industry in order to achieve air supremacy over Western Europe before an invasion of the continent. While losses for these raids were still extremely high, it was hoped that dispatching a larger force capable of inflicting serious damage would make it worth it. On August the 17th, the 8th Air Force prepared for its largest raid yet, with 376 B-17’s dispatched to attack the ball bearing works, at Schweinfurt, and a Messerschmitt factory, at Regensburg. Both of these facilities were located deep within Germany and most of the journey would see the B-17’s outside the area where they could be escorted. To compensate for this, the flight over Regensburg would continue over the Alps and into Allied controlled Tunisia. It was hoped that flight over the Alps would prove easy, and in the case of the Schweinfurt force, they believed that the German fighter squadrons would still be on the ground refueling after their first attacks while the bombers made their return. Both would be met with disaster as the Luftwaffe would hit both forces after their escort fighters turned for home, and the Luftwaffe fighters had taken to the air again as the Schweinfurt raiders made the return trip.

The bombers of the USAAF flew in staggered formations in order to maximize the the defensive capabilities of the aircraft. These tactics alone proved totally inadequate to protect bomber formations from fighters and were revised several times to compensate for flak. [National Archives]
Of the 376 bombers to leave England, 60 would be shot down, 176 were damaged, and 30 remained in North Africa where they awaited repairs at the overburdened facilities in Tunisia. Losses in combat and written off airframes amounted to 31% of the dispatched force; in contrast the Germans lost only 28 fighters (Overy 340, 341). Following the disaster, the 8th Air Force would carry out raids only where there was full escort cover and the next deep incursion into German airspace would only be conducted in the spring of the following year. The winter of 1943 would spell uncertainty for the campaign, as it was clear that for all intents and purposes, much of German industry lay beyond striking range. With this limitation threatening to seriously cut back the USAAF’s campaign, they would request that Lockheed, Republic, and North American increase the internal fuel capacity of their fighters, and hoped that a suitable long range escort would materialize.

Leaving the Stable

Col. Charles McCorkle, 15th AF with pilots. The P-51B proved the solution to the problems plaguing the ailing strategic air campaign. [National Archives]
As a result of the pressure to produce new, long range fighters for the escalating campaign in Europe, the first P-51B’s were produced before the prototype had gone through its testing and modification cycles. The first plane, a P-51B-1, was completed March 31, 1943 and would include several features that would later be found unsound on prototype. As a result, these initial planes would have to be altered accordingly and would have many parts that were non-interchangeable with later models (Marshall & Ford 316). In addition to reworking the air scoop and radiator, they would also have their ailerons modified, both to improve their effectiveness and to remove a steel diaphragm which would interfere with the plane’s magnetic compass. Most importantly, the decision was made that the aircraft would incorporate an additional 85 gallon fuselage fuel tank which would provide the aircraft with phenomenal range.

With this new aircraft, the USAAF would finally possess what they had been searching for. With the addition of the new internal fuel tank, the aircraft would be capable of deep incursions into German airspace, and it would deliver on what was promised back in the spring of 1942. They were excellent fighters, especially at high altitude. The early P-51B’s would use the Packard V-1650-3 engine, a license production of the British Merlin 61, which produced 1410 hp at 29,300 ft and 1630 hp at 16,400ft at War Emergency Power (P-51 operation manual 31). This engine would later be replaced with the Packard V-1650-7 in later models of the aircraft, which was geared for better performance at medium altitude. These engines, combined with the low drag fuselage and laminar designed wings would provide the aircraft with a superb climb rate, a high top speed at altitude, and exceptional high speed maneuverability.

While the aircraft had taken a largely completed form with the P-51B-5 and P-51C-1, it would be continuously modified in the field and on the production line, throughout its service with the air force. The most notable of these changes were the additions of a fuselage tank, booster motors for its ammunition belts, a vertical fin extension, and field retrofits for a perspex canopy dubbed the Malcom Hood. However, only the 85 gallon fuel tank would be a universal addition.

The fuselage tank would enable the P-51B’s to reach much of central Europe from England, but it was not present in the first deliveries of the aircraft, as was the case with the 59 P-51B’s active in England at the end of November 1943. The installation kits were first sent out in September of 1943, and the tank was later incorporated into the production run with the first long range P-51B being accepted by the Army in December of the same year (Marshall & Ford 393, 407).

The next addition to the aircraft was intended to solve a major issue with the plane’s machine guns, which were found to be prone to jamming when the pilot pulled turns of over 1g. This issue was a result of the canted position of the guns in the wings which put stress on the ammunition belts. The ideal solution was to reposition the guns, but seeing as that would necessitate a considerable redesign, engineers would instead work in a stop gap measure in the form of boost motors for the ammunition belts. These were issued as kits like the fuel tank, though unlike those for the fuselage fuel tank, they were issued in more limited numbers and the issue persisted well into 1944 (Ethell 64).

The Mustang had long had a tendency to yaw in the opposite direction of a roll, which affected its handling since its earliest models, and this was made significantly worse when fuel was carried in the fuselage tank. Despite the problem being an evident and considerable inconvenience, its solution wouldn’t materialize until much later. Eventually, it was decided to fit the aircraft with a fin extending from its vertical stabilizer, along with adding reverse rudder boost tabs. However, these kits arrived very late, having begun production in April of 1944, and later incorporated into the design of late P-51C’s and the subsequent P-51D (Marshall & Ford 306).

The Malcom Hood provided far better visibility than the earlier ‘birdcage’, and was added to a number of P-51B’s based in Northern Europe. [National Archive]
Many long standing issues revolved around the ‘birdcage’ canopy of P-51 since the aircraft’s inception, and as was the case with the engine, an improvement was found in British service. With the RAF, many Mustangs received a new frameless bubble canopy. This canopy vastly improved visibility, especially to the rear of the aircraft, which was virtually non-existent from within the birdcage, and it could be drawn back on landing and take-off. Dubbed ‘Malcom Hoods’ after their manufacturer, a plexiglas works named Robert Malcom Ltd. they were subsequently sought after by the USAAF for use with their P-51’s in Europe.

Breaking the Stalemate

The new P-51B’s would make their first major debut with the 8th Air Force in early 1944, though the introduction was not as smooth as had been hoped. Squadrons reported a number of issues with the new aircraft, which included high altitude fuel transfer failures with external tanks, glycol reserve tanks that leaked and froze, radiator corrosion and coolant leaks, radios and spark plugs failing, and excessive oil loss (Marshall & Ford 425). However the USAAF hadn’t the time to immediately resolve these teething issues, and with these problems passed along to the manufacturer and Air Force maintenance services, the P-51’s would soon play a key role in the escalating bomber offensive.

Through the winter of 1943, both the day and night bombing campaigns were facing withering losses which spelled serious trouble for maintaining the pace of operations over Europe. With less than one thousand bombers stationed in England, the USAAF would lose 200 in September alone (Douglas 326). In the face of these losses, the Combined Bomber Offensive was failing to carry out the Pointblank directive, which aimed to cripple the Luftwaffe before an invasion of Europe was conducted. During this period the Luftwaffe had actually built up the strength of its fighter force and had reorganized and improved its defenses into a centralized command structure. To make matters worse, the head of RAF’s Bomber Command, Arthur Harris, would ignore orders to attack German industries involved in aircraft production. Instead, he would order Bomber Command to continue to carry out an ineffective area bombing campaign of Germany’s cities believing it would bring an end to the war without the need for an invasion (Overy 343, 344). It was under these bleak circumstances that the US’s Eighth and Fifteenth Airforces were tasked to cripple the Luftwaffe and establish air superiority over much of Europe before the invasion, now only a few months away. However, they would soon see a change in leadership and the delivery of new equipment that would put them on the path to controlling the skies over Europe.

Escort fighters typically flew a few thousand feet above their charges when they weren’t independently seeking the enemy. They weaved back and forth over the bombers in order to not speed past them. Here a flight of four P-51’s flies overhead at roughly 30,000 ft. This tactic declined in use when the relay system came to prominence. [National Archives]
In December of 1943, the USAAF established a joint strategic air command to consolidate their bomber forces over both the European and Mediterranean theaters, and drive them towards a unified objective. With Gen. Spaatz in command of all strategic bomber forces, and Maj. Gen. James Doolittle in command of the Eighth Airforce in England, the USAAF would now have clear strategic direction, and more aggressive leadership. Doolittle would take a pivotal role in revising the existing strategy into one which proved instrumental in undermining, and dismantling the Luftwaffe in the coming weeks. Crucially, he recognized the inadequacy in trying to undermine the Luftwaffe’s fighter strength solely through targeting the production of new aircraft. To hold to this existing, overly conservative strategy was hopeless, and the invasion of France was scheduled for five months after he took office. Targeting the factories alone wasn’t enough, and thus Doolittle would give the order for returning escort fighters to perform fighter sweeps and seek out enemy planes in the air and on the ground (Overy 361). Among the first and most important moves was to create a more effective relay system for the fighters, further increasing the time they could spend over enemy territory.

By the start of 1944, Maj. Gen. Kepner, 8th Air Force, would also play a major role in implementing this new strategy, as he officially untethered the Eighth’s fighters from the bombers and allowed them to seek out the enemy at their discretion. The P-51 would play a pivotal role, as its excellent high altitude performance and range meant it was able to take up the last position of the fighter relay, and was more than a match for whatever it found. Beyond the existing penetration, target, and withdrawal relay positions, the P-51 was also able to take up a fourth mission. These units would perform sweeps 50 to 70 miles ahead of the bomber formation and attack German fighters as they were climbing, assembling, or transiting towards the bomber formation. Their efforts were greatly aided by British signals intelligence services that provided the assembly points for the Luftwaffe’s fighter groups (Marshall & Ford 425, 425; Overy 362).

This change in tactics would have immediate and profound impacts as they began to be widely implemented in February and March of 1944. The first major achievement of the new strategy were the widespread losses inflicted on the twin engined fighter forces, which had earlier proven themselves as potent anti-bomber weapons. Against the new long range fighters, they were almost defenseless, and were withdrawn in March (Overy 366). Similar effects were felt throughout the Luftwaffe’s fighter forces, which thanks to the new P-51’s, were left without any safe haven. Whenever the bombers were over Germany, their escort fighters could make their appearance. While the new strategy often meant that the bomber formations were often less protected, this was counterbalanced in that it placed the German fighters on a defensive footing. The days of Luftwaffe fighters leisurely climbing alongside a formation before diving at it head on were over, now whenever they reached a formation they were forced to conduct hit and run attacks, or face off against the escorts.

Luftwaffe attrition escalated as airfields that were once ignored were now periodically harassed by fighters that attacked transiting and grounded aircraft. Doolittle did all he could to promote these attacks, and would allow for the destruction of aircraft on the ground to count towards a pilot’s ace status (Marshall & Ford 423). These attacks would prove costly to the USAAF, but well worth it as Luftwaffe operational losses for all aircraft increased sharply and it robbed them of the ability to train new pilots in secure airspace. This shift in strategy and subsequent success would prove instrumental to the USAAF in the following months, as their responsibilities were soon to broaden when the Allies landed in France.

When equipped with external fuel tanks, the P-51B could operate over any part of Germany. This proved disastrous for the Luftwaffe as transiting aircraft and those on the ground were now vulnerable, no matter how far they were from Allied air bases. [National Archive]
While the Eight and Fifteenth air forces were still occupied with the task of destroying the Luftwaffe in the air and on the ground, they would soon be given additional missions. The most unexpected of which came in the form of Operation Crossbow, which called upon the Eighth Air Force to disrupt Germany’s use of the new V-1 bomb from coastal bases. Then came the task long awaited, which called upon the Eighth to begin the preparations for Operation Overlord. To meet these new objectives, the Pointblank raids were accelerated, culminating in ‘Big Week’ in February of 1944.

Between the 19th and the 26th, the Eighth and Fifteenth air forces would fly roughly 6,200 sorties against 18 aircraft assembly plants and two ball bearings plants, at a loss of 247 bombers and 28 fighters. Undoubtedly steep, but sustainable in comparison to the Luftwaffe which lost roughly one third of its single engine fighters (Overy 369). The success of the raids themselves was difficult to judge, as fighter production still increased, though at a significantly reduced rate which saw a shortfall of roughly 38.5 percent (Overy 370). During these operations the P-51 would provide the USAAF deep penetration cover and perform strafing attacks against German airfields. However, there weren’t enough long range escorts for full coverage until the summer of 1944. The situation was further complicated when all P-51B’s were grounded between the 10th through the 15th of March in order to address structural issues with the aircraft’s engine mounts, wings, and tail. These were subsequently resolved by replacing the retaining bolts for the engine, reinforcing the tail empennage and ammunition doors, and installing landing gear locks to prevent their uncontrolled release at high speed (Marshall & Ford 442, 446). These issues would however not present a long term obstacle during the early months of 1944 as the tempo of operations and list of targets grew in the following months.

With the major push against the German aviation industry mostly over, the USAAF would soon set its sights on two major targets, rail communications across much of Northwestern Europe, and Germany’s oil industries. The first was an immediate necessity for the success of Operation Overlord, crippling German strategic mobility was essential for an invasion which would require considerable time after the first landings to build up a force on the continent. The formalities were worked out in March when the Transportation Plan was decided upon. It would fortunately have the support of RAF Bomber Command, as Harris’s evident failure to end the war on his terms would see him temporarily divert his force into supporting the preparations for the invasion of France. The subsequent offensive against fuel production would start far less formally. Spaatz was convinced of its necessity, but due to the months it would need to take effect, he was at first unable to convince his superiors to divert resources to it. However, in a matter of weeks, he was able to argue for its necessity under the Pointblank Directive and was then allowed to conduct attacks against Germany’s synthetic fuel industry whenever resources permitted (Overy 371).

Between the now crippling fuel shortage and marauding allied fighters, the Luftwaffe soon found themselves completely overwhelmed by the autumn of 1944. Here a P-51 lines up on an He 177 heavy bomber, as the one beside it continues to burn. [National Archives]
With these new policies in place, the Luftwaffe would be thoroughly disrupted as a result of Spaatz’s strategy, and Doolittle and Kepner’s tactics. The USAAF would end up inflicting punishing losses on the Luftwaffe in the air, disrupting the manufacturing of new aircraft, and eventually causing chronic fuel shortages that severely limited their ability to conduct large scale operations of any kind. In this, the P-51 would prove essential with its exceptional high altitude performance, and its endurance that could take it anywhere over Germany.

In many ways, the bombing of factories alone was a largely ineffective means of inflicting serious damage to the German war economy, as many industries proved to be exceedingly resilient. Fighter production proved a particularly difficult target, as apart from the later targeted aero engine industry, production and final assembly plants could be dispersed and were largely safe from raiders. When fighter production was further streamlined and resources were diverted to support it, Germany would end up vastly expanding fighter production during the period in which those industries were the most frequently raided (Zeitlin 59). This was, however, was achieved only by reducing the rate of modifications and improvements, and transferring resources away from the production of bombers. In comparison, the later targeting of fuel production and rail transportation proved key, as the inability to reliably move material by rail combined with chronic fuel shortages proved a fatal military and economic obstacle. As a result, establishing air supremacy over Western Europe before Operation Overlord was as much an achievement of long range fighter operations as it was of the bombers. The Luftwaffe could sustain itself when aircraft deliveries did not meet expectations, but it quickly found itself struggling when it lost scores of pilots and found itself hard pressed to train new ones once they had lost control of the skies over Germany.

Pre-war military theorists envisioned fleets of bombers destroying vital war industries with the near pin-point accuracy they achieved in controlled tests. The reality of the campaign revealed this as hopelessly optimistic when even the most accurate raids resulted in large amounts of collateral damage. [National Archives]
In the end it must also be said that the civilian costs of the raids were steep, and while the Eight and Fifteenth Airforces were not involved in a campaign directed against the civilian populace, as was the case with Bomber Command and the USAAF elsewhere, the technical limitations of the time meant that bombs frequently fell on civilian areas. Even under ideal circumstances, the dimensions of a bomber formation were larger than their targets and it was physically impossible to strike factories, railyards, and refineries without causing significant damage to the surrounding area. The realities of the campaign would also prove worse than predicted. Targets were frequently obscured by bad weather and smoke generators, and formations typically took heavy anti-aircraft fire on the approach. As a result, bombs were often released by the best estimate from the bomb sight or at the direction of a ground mapping radar system (Overy 347). Even outside of Germany, the civilian costs of these operations were heavy as the Allied air forces carried out the transportation plan. In France alone, between March and June of 1944, French officials placed the figure of civilians killed by Allied bombing at 25,266 (Overy 574).

The 4th Fighter Group ‘Debden Eagles’

When the US entered the Second World War, few American airmen had any combat experience, with the notable exceptions being volunteer airmen in service with foreign armies. The Debden Eagles were one such group, having volunteered to serve with the RAF and entered service in late 1940 and 1941. While they were among the few Americans fighting against Nazi Germany at the time, they had garnered a somewhat unfortunate reputation as glory-seekers and primadonnas thanks to their unique position (Bucholtz 6). Their tendency of excessive overclaiming of victories during this period would prove particularly irritating to their superiors. With the US entry into the war, the Eagle squadrons, and their Supermarine Spitfires, were subsequently integrated into the USAAF.

Capt. Donald Willis, an Eagle Squadron pilot alongside a Spitfire Mk V, late 1943. [National Archives]
The RAF’s 70th, 121, and 133 Eagle Squadrons would become the 334th, 335th, and 336th Squadrons of the 4th Fighter Group on the 12th of September 1942. These units flew Spitfire Mk IX’s and within the month were supporting the nascent bomber offensive which was targeting installations in France. The start of this effort went poorly, when only one aircraft out of a twelve plane flight returned, the rest having been lost to enemy fighters, harsh weather conditions, or having run out of fuel in the short range fighter. Thankfully for the Group, this would be their worst day of the war. Despite this setback, the unit saw its first major mission carried out on the 20th of October in the Calais area escorting B-17’s carrying out a high altitude raid. This would be the first major bomber operation carried out under escort and was met with success. Their Spitfires would prove a very capable fighter aircraft, but their short range rendered them unable to conduct escort missions far beyond the English Channel. In any case, this wouldn’t prove much of an issue, as for the rest of the year as they would mostly conduct fighter sweeps across the low countries and provide convoy cover (Bucholtz 9). However, with the changing of the year, the 4th would exchange their venerable Spitfires for new P-47’s.

The 4th FG flew their Spitfires in combat for the last time on April 1st, 1943, after which they completed the full transition to P-47C’s. This change was not viewed favorably, as most of the unit’s pilots disliked the considerably heavier Thunderbolt (Marshall & Ford 340). The changeover had little initial impact on operations, and the squadron was largely involved in the same missions as before. However, the group would later accompany bombers on deeper raids into Europe thanks to newly issued external fuel tanks for their P-47’s. They would use these new 200 gallon fuel tanks on an escort mission into Ghent on July 25th and soon after their first foray into Germany airspace over Westhoff-Emmerich. It should be noted that these fuel tanks were a rare piece of equipment at the time and the 4th only had them thanks to the efforts of Lt. Col. Cass Hough of the 8th Fighter Command’s technical section. They were, unfortunately, as troublesome as they were vital, often failing to transfer fuel above 20,000, and were later withdrawn as British made paper 108 gallon tanks became more available (Marshall & Ford 411).

Despite their complaints, the 4th FG’s veteran pilots would master their new planes and had put them to good use. In a battle defending a formation of B-17’s over the city of Utrecht, the 4th FG was credited for the destruction of nine enemy aircraft at the cost of one of their own, with the pilot having bailed out over the occupied Netherlands (Bucholtz 16). With their P-47s, the 4th would take up an important supporting role in the escalating bombing offensive, one which saw their longer ranged P-47s making more flights into the German frontier. This tempo and the 4th’s change in command under the more aggressive Lt. Col. Don Blakeslee would see the unit become among the most successful in the entire USAAF.

Col. James Matthew Blakeslee would lead the 4th FG from January to November 1944, after which he remained on the ground after several high profile pilots of the USAAF had been lost in a short period of time. He is pictured here receiving the Distinguished Service Cross from Supreme Allied Commander in Europe, Dwight Eisenhower. [National Archives]
Lt. Col. Blakeslee was made C.O. of the 4th with the turn of the year, and in addition to bringing new, more aggressive tactics to the table, he would work to ensure his unit was re-equipped with the new P-51. Blakelsee would meet personally with General William Kepner and argue that his squadron would be the best candidate for refamiliarization with the new plane as their experience with the similarly-engined Spitfire would make for an easier transition. Kepner was convinced, and subsequently put the 4th FG at the top of the list for P-51’s. The schedule for the transition was harsh as they continued to fly combat missions in their P-47’s while also familiarizing themselves with the new aircraft. The process was time consuming and they would not make their operational debut with their new planes until February 28, 1944 (Marshall & Ford 432). These Mustangs would nearly double the combat range of the unit, and the pilots favored them over their older P-47’s, but they experienced a variety of harsh teething issues and mechanical failures.

While the conversion was taking place, the 4th would be committed to Doolittle’s more aggressive strategy against the Luftwaffe, with the aim to achieve aerial supremacy over Western Europe before the invasion of France. As such their independent actions increased, and on January 31, 1944, they would join the 355th FG in bombing the Luftwaffe’s airfield at Gilze-Rijen (Marshall & Ford 425). In many ways this mission bore some similarity to the fighter sweeps they had conducted since they had flown with the RAF, but it would mark a first in that direct assaults on Luftwaffe airfields would then become more commonplace. Among the last major actions the unit would perform with its P-47s was its support of ‘Big Week’.

Their first combat mission in the new planes was fairly uneventful, on February 28, when flying as escorts for a formation of bombers attacking a V-1 launch site they encountered no enemy aircraft but strafed a Ju 88 on their way home. They would claim their first aerial kills two days later during a bomber withdrawal support mission near Frankfurt where they claimed two enemy fighters (Bucholtz 38). The following day the unit would help achieve a major milestone, the first fighter escort operation to Berlin and back. The operation would prove anything but easy, as deteriorating weather conditions saw most of the aircraft involved turn back. However, elements of the 3rd Bomb Division would press on, supported by the 4th, 55th, 354, and 363rd FG’s. The 4th would engage a formation of roughly 60 Fw 190’s and Bf 110’s northeast of Wittenberg in the day’s first encounter with the enemy. They claimed five victories but suffered one loss from enemy fire, and another as a result of a radio failure which made navigation across a storm in the English channel impossible. The pilot was later forced to ditch his aircraft in France after a failed attempt to reach neutral Spain (Marshall & Ford 439, Bucholtz 39).

Capt. Don Salvatore Gentile was among the leading aces in the 4th FG. He was credited with 21 aerial and 6 ground victories, though his combat service ended after a botched aerobatics stunt in front of assembled members of the press. He was grounded and went on a tour to raise war bonds, later becoming a test pilot. [National Archives]
Perhaps the most exciting encounter that day was experienced by Capt. Don Gentile and Lt. John Godfrey, both aces in the 4th. The two pilots were unable to join the rest of their flight as a result of extremely poor weather, but proceeded with their mission regardless. En route the weather would clear, and reveal a flight of roughly 50 Do 217 night fighters, pressed into service as daylight bomber destroyers, and dozens of Fw 190’s which were preparing to attack a nearby formation of USAAF bombers. The pair would decide to attack, in order to disrupt the enemy formation and prevent them from engaging the nearby bombers from an advantageous position. Gentile and Godfrey dove on the night fighters, damaging one and sending the group diving in an effort to escape. The engagement turned into utter chaos as the single-engined fighters joined in. In the confusion, the pair of aces would claim one enemy aircraft in a series of defensive fights that eventually saw them make their escape through the clouds. Flying on instruments and practically lost, they made their way back to England by their intuition, landing at RAF Hurn (Bucholtz 40).

The unit would return to Berlin on March 6 in support of a massive 8th Air Force operation. Favorable weather conditions would allow the 8th to dispatch a force of 730 bombers against a series of targets in and around the German capital, where they would meet the Luftwaffe in the largest air battle of the war up to that point. The 4th, led by Col. Blakeslee, would be tasked with escorting the bombers, which would prove a difficult undertaking, with the sheer number of opponents forcing the group to disperse into individual flights and sections to expand their coverage. The unit would be credited for the destruction of 15 enemy aircraft of the 45 claimed by P-51’s that day, in exchange for five losses. In comparison, P-47 units were credited with 37 kills for 5 losses, and P-38 units brought down three units at the cost of three of their own. It should also be noted the P-38’s comprised the minority of the fighters, while there were roughly twice as many P-47’s as there were P-51s. The USAAF would claim a total of 83 ‘confirmed’ enemy aircraft with the Luftwaffe having recorded the loss of 75 fighters (Marshall & Ford 439; Bucholtz 43, 45). The majority of these kills were twin engine and night fighters pressed into daylight service. This engagement, while not representing a distinct turning point, did demonstrate a noticeable shift in the war over Germany. Of the 672 bombers that proceeded with the mission, 69 failed to return, and 6 were written off. These were certainly heavy losses, but were a fraction of the nightmare that the Allies were facing in the summer and autumn of the previous year. Beyond that, Luftwaffe losses were mounting both in the sky and on the ground, and the use of its heavier, twin engined bomber destroyers had become untenable in the face of agile new opponents.

D-Day

During the first day of Operation Overlord, most fighter units were dedicated to countering a Luftwaffe response that never came. Several would go on to attack inland targets. [National Archives]
Over the coming weeks the 4th would continue to support the bombing campaign, but in June of 1944 they would participate in something far more decisive. The group would be among the many fighter units providing top cover for the invasion of Normandy. Throughout D-Day, each of the unit’s three squadrons would operate independently and continuously until nightfall. The day began with the 334th and 335th squadrons undertaking an offensive patrol under the command of the unit’s C.O., Col. Blakeslee, between 03:20 and 09:45 over Rouen, France. The patrol found no enemy fighters and sought out targets of opportunity, in their case a pair of locomotives that they strafed with their machine guns. Their only loss was 1st. Lt. Fraser, who had lost contact with the rest of the squadron and was subsequently downed by German fighters and taken prisoner. The 336th would sortie at 06:42 to provide cover for warships shelling the landing areas, which proved uneventful (Bucholtz 84).

At 11:20, the 334th would sortie again to Rouen with one section carrying bombs. They would attack a troop train to poor effect, though an encounter with a flight of 10 Fw-190 near their airfield at Evreux proved more successful. In the ensuing battle the 334th was credited with the destruction of four enemy fighters, with the only damaged P-51 making it back home. While this confrontation was happening, the 335th had attacked the marshaling yards at Fleury. The 336th would fly for the last time that day at 13:35 conducting a fighter bomber sweep near Evraux. They would find no targets and would lose an aircraft to ground fire, with 1st Lt. Freiderick being taken as a PoW. The last mission of the day would see the 334th and 335th conduct attacks against a radar station and a road convoy near Rouen. While successful in their mission, they incurred heavy losses when elements of the unit were attacked by around 15 fighters belonging to JG 2 and JG 26 as the US fighters attacked infantry positions.

Capt. Winslow Sobanski was a Polish infantryman at the outbreak of the war, eventually finding his way to the US where he then joined one of the Eagle Squadrons. He was among those killed in action during the group’s last sortie on D-Day. Pictured here in a P-47. [National Archives]
The day would prove exhausting, with pilots flying up to three missions from dawn to dusk. Between flights most of the 4th’s pilot’s would rest, usually either having coffee or trying to get some sleep in before their next mission. The different squadrons would also find themselves having vastly different experiences, with the 336th having spent most of the day covering the invasion force which the Luftwaffe hadn’t the strength to attack, and taking part in a fighter bomber sweep that found no worthwhile targets and saw one aircraft lost to flak. In comparison, the 334th and 335th spent the entire day conducting offensive sweeps which claimed a number of targets, but also saw them sustain higher casualties than any of the other US fighter squadrons over Normandy that day with ten fighters lost (Bucholtz 82, 83).

Shuttle Mission to VE-Day

Following the success of the landings, and subsequent breakout in Normandy, many of the USAAF fighter units would take on tactical missions in support of the armies in Western Europe, in addition to the ongoing strategic air campaign. However, some P-51 units were selected to participate in an escort mission in which the bombers would land at prepared airfields in the Soviet Union instead of returning to their home bases. A 45 aircraft detachment of fighters from the 4th would depart for the Soviet Union on June 20th. The mission would see them join a force of 1,000 bombers as they attacked targets in the Rhineland, and then on to Piryatin, Ukraine some seven hours away. 45 Mustangs of the 4th would make the trip, encountering some 25 enemy fighters over Siedlice, downing two, but losing one of their own. All but one of the remaining planes landed at their intended destination, with one 2nd Lt. Hofer being forced to land at Kiev after running low on fuel after pursuing enemy fighters (Bucholtz 88). However, unbeknownst to the assembled American aircraft, the formation had been trailed by a Ju 88. Soon after, a well coordinated attack by the Luftwaffe using its He 177 heavy bombers saw many of the US bombers hit, though their P-51’s were unscathed.

The P-51’s were subsequently dispersed and flew a variety of missions in the following weeks which brought them over Central Europe and the Mediterranean. They soon flew an escort mission against an oil refinery in Drohobycz, Poland on the 26th. The return leg of the mission took them to Lucera, Italy where they would support the bombing operations of the 15th Air Force. The largest of these missions would take them over Budapest to perform a fighter sweep ahead of the strike force. There they encountered 80 German and 18 Hungarian Bf 109G’s and a massive dogfight ensued. In the battle the 4th would claim eight Axis fighters at the cost of four of their own. This included 2nd Lt. Hofer who had died during a strafing attack against an airfield. (Bucholtz 89). The unit would be led back to England on the 3rd of July.

American and Soviet personnel during Operation Frantic. [National Archives]
Beyond Operation Frantic the 4th settled back into the ‘usual’ operations they’d had since most of the group had left for the Soviet Union. They continued to fly deep penetration and escort missions over Germany, though by the end of the summer, Luftwaffe activity in the air had been considerably reduced. The savage war of attrition over Germany had been decisively won by the USAAF, as the Luftwaffe began to feel ever more crippling shortages of experienced pilots and fuel. Ironically, the Luftwaffe’s supplies of fighter aircraft were secure, though transporting them to airfields would prove ever more troublesome through the remainder of the war. While they had the aircraft, a subsequent USAAF campaign against rail communications across Germany would make overland transportation difficult, and ever more frequent fighter sweeps made transiting by air a very dangerous prospect.

For the remainder of the war the 4th FG remained committed to supporting the strategic bombing campaign, especially as it pertained to offensive fighter sweeps and attacks against Luftwaffe airfields. Their last victory of the war was a probable destruction of an Me 262 that was damaged over the Prague/Ruzyne airfield, with the group credited for 1,058.5 total victories against enemy aircraft, both in the air and on the ground (Bucholtz 120). They would end the war among the most successful Fighter Groups in the USAAF, having come a long way from the overly boastful volunteers that had flown against the Luftwaffe before any other Americans.

The 99th Fighter Squadron ‘The Tuskegee Airmen’

As black aviators, the men of the Tuskegee-trained squadrons would face unique challenges, having to face prejudice from their own countrymen who sought to deny them the opportunity to fight. They were initially excluded from many of the pre-war programs that turned out many of the pilots who later joined the ranks of the USAAF and US Navy. Many who ran these programs espoused the belief that they were incapable of the judgment needed for leadership, and that they had lacked ‘the proper reflexes to make a first class fighter pilot’ in the words of General Edwin J. House (Moye 102).

Their chance came with the Civilian Pilot training program in 1939, having been excluded from the program the previous year. The program was extended to a series of predominantly black colleges and universities, with the most critical being the Tuskegee Institute in Alabama. The university would build a fledgling airfield that eventually grew into an Army Air Corps training base, which proved controversial even among hopeful applicants, as in their eyes they were clearly still segregated from the rest of the Army. While the controversies flowed in the small Alabama town, the Air Corps moved to create the first black pursuit squadron, the 99th.

Col. Benjamin O. Davis would lead the 99th fighter squadron and the later 332nd Fighter group. He would go on to become a Brig. General in the newly formed United States Air Force after the war. [San Diego Air and Space Museum]
The first cadets of the 99th would graduate March 7, 1942 under the command of Capt. Benjamin O. Davis. The squadron would subsequently fly within the US before its transfer to the Mediterranean in late March 1943, equipped with new P-40L’s (Moye 99). They occupied a former Luftwaffe airfield in Morocco and were to be attached to the 33rd Fighter Group after they had gotten some experience in theater. In May, the squadron would be deemed ready for service and would move to a field in Tunisia. They would see success there, but the leader of their fighter group, Col. William Moymer was immediately hostile to their presence. He failed to return the salutes of the 99th’s officers and he placed the squadron on patrol duties over secure air space. He would then openly criticize them for being ‘unaggressive’ for failing to claim victories over territory where they were unlikely to encounter enemy aircraft (Bucholtz 18, 19; Moye 101). In spite of this, the unit pushed on and would aid in the preparations for the invasion of Sicily.

The 99th’s first combat missions were fighter sweeps against enemy positions in Southern Italy, their first target being a German airfield on the island of Pantelleria on June 2, 1943. The airbase would be the site of many more attacks, including the unit’s first encounter with enemy fighters. On June 9th, six P-40’s from the 99th Squadron accompanied A-20’s to the airfield, encountering four enemy fighters. In the ensuing fight they successfully drove off the enemy aircraft, and damaged one, taking no losses of their own. A further effort was made to intercept a flight of Ju 88’s at high altitude but were unable to, as their P-40’s had their oxygen systems removed to save weight for the low altitude mission (Bucholtz 21). The pilots of the 99th were particularly enthused that in their first encounter with the enemy, they had managed to complete their mission and all returned home safely.

While they would eventually be known for their iconic red tailed P-51’s, members of the Tuskegee fighter squadrons would fly the P-40L, P-39Q, and P-47D before they were issued Mustangs. [National Archives]
The squadron would be redeployed days later, partially a result of Moymer who sought to see the squadron reduced to coastal patrol duties. Instead, the 99th was transferred to the 79th Fighter Group, who’s commander, Col. Earl E. Bates, did his best to integrate the unit into the group. While they remained formally segregated, they enjoyed a far more open and professional environment than what they endured with the 33rd (Moye 103). Their first mission with the unit was on July 2 and saw them escort a flight of 16 B-25’s to their target, a German airfield in Castelvetrano, Italy. It would prove less than ideal when the B-25’s failed to line up with their target on the initial approach and had to repeat the attack, giving Axis fighters stationed nearby the time they needed to scramble. Two of the 99th’s pilots were lost in the first pass from the German fighters, but the remaining members soon regained control of the situation. In the ensuing confrontation with enemy Fw 190’s, Bf 109’s, and a Macchi 202, the 99th would claim one confirmed destruction, one probable, and two damaged aircraft. Though perhaps most importantly, none of the B-25’s they were escorting came to harm (Bucholtz 21, 22).

The coming weeks saw them mostly fly ground attack missions in support of the ongoing invasion of Italy, and met very few enemy aircraft for the remainder of the year. It was during this time that they also discovered that the Tuskegee training center wasn’t large enough to supply a sufficient number pilots to the squadron, while also supporting the construction of three additional squadrons. Their pilots resultantly flew an abnormally high number of missions due to being short handed (Bucholtz 25). This period also saw them defeat a great deal of the unfair criticism leveled against them and had largely cemented a favorable reputation within the Army Air Force. Among the most notable victories on that front was an article in Time, which had previously published an article based on Moymer’s alleged grievances with the squadron. Maj. Roberts of the 99th would be quoted “people assumed we were not producing because we were negroes…but now that we have produced, things have changed.” The 99th had also succeeded in convincing most of the 79th FG of their worth, and had garnered a great deal of respect as they moved into 1944. Many white pilots of the 79th disobeyed an order from the commander of the Air Force commander in the MTO, and held a desegregated dinner party to celebrate the anniversary of the 99th’s combat debut (Moye 104, 105).

Forming the 332nd Fighter Group

While 99th gained valuable experience over the Mediterranean, they began to rotate pilots out to train the next pursuit squadrons to form a segregated fighter group. These squadrons were the 100th, 301st, and 302nd, all of which would be formed at Selfridge Field, Michigan. Selfridge would prove a particularly dreadful post for these men, as it was here that they would face intense discrimination both by the local populace and base staff, while being a stone’s throw from the racial powder keg of Detroit. However, this would not remain their home for long, and they would soon depart for their operational assignments by the end of the year. They would join the 99th in the Mediterranean Theater of Operations in January of 1944, being equipped with a set of used P-39s. These aircraft would prove troublesome in service due to their age and condition, and as such numerous accidental losses followed, so by the early summer of 1944, Col. Davis had managed the acquisition of new P-47Ds. However, the unit would soon transition again to the newer P-51 soon after the 99th joined the rest of the fighter group in July, something the group’s veterans would resent as they felt they had been segregated again after finding acceptance within the 79th FG.

Capt. Andrew Turner aboard a P-51. The group’s transition to this aircraft vastly expanded the range and variety of operations across the Mediterranean and Central Europe. [National Archives]
The group would fully transition to Mustangs by July of 1944, and would be reassigned to the 15th Air Force where they would support long range bombing operations. Their first mission in their new planes was on July 4th, where they took 40 aircraft to to escort two bomber wings, but they would encounter only a pair of Bf 109’s that made no attempt to attack the allied aircraft . Beyond this, their pace of escort missions rose and they would take part in supporting raids against Axis positions in Northern Italy and Southern France. Soon after, they would provide support for the amphibious invasion of Southern France. On August 12, All four of the 332nd’s squadrons were given specific targets, with the 99th striking radar stations in Montpelier and Sete, the 302nd attacking radar stations in Narbonne and Leucate, the 100th attacked the radar stations near Marseilles and Cape Couronne, and the 301st attacked four targets around Toulon. At the loss of three pilots, one captured and two killed, all of the targeted radar stations sustained considerable damage .

The remainder of the war saw the 332nd fly a considerable number of escort missions, including an earlier attack against the Ploesti oil fields in Romania on July 13th, 1944. It was during that mission that they had begun to cement their status as one of the most reliable escort units in the USAAF, after they dispersed a flight of eight German fighters that had attacked bombers of the 55th Bomb Wing. Their C.O., Col. Davis maintained an unwavering directive to his unit, on escort missions they were never to abandon their bombers. This didn’t sit well with some but it was accepted, in part because many felt that a failure to protect the bombers would come down harder on them than the other squadrons (Bucholtz 51, 105; Moye 102). As such, their record for defending bombers was exemplary, having lost only 27 bombers to enemy fighters from June of 1944 to April 1945. It should also be noted that 14 of these losses occurred during a single day when a failure in mission planning resulted in the bombers and their escorts failing to meet at the proper time. As the target that day was the Luftwaffe air base at Memmingen, Germany, losses were correspondingly high (Bucholtz 53, Haulman 2). This places the remaining 13 bomber losses among the other 178 escort missions they performed over ten months. This policy would however, result in the squadron having the lowest aircraft kill to loss ratio of any other P-51 squadron in the theater, however, they would still consistently outscore all of the veteran P-38 squadrons in the Mediterranean (Marshal & Ford 477).

Among their most impressive escort missions was in support of a bombing raid against the Daimler-Benz tank assembly plant in Berlin, on March 24, 1945. From the 332nd’s base in Ramitelli Italy, this was a 1600 mile round trip, the longest mission ever conducted by the 15th Air Force. 59 Mustangs of the 332nd would leave their base at 11:45 under the command of Col. Davis, though he would soon return after experiencing engine trouble and left the squadron in the command of Capt. Edwin Thomas. They would encounter some two dozen enemy fighters outside of the German capital, including a number of Me 262s. The jets would initially prove difficult to catch, and the aircraft, belonging to JG 7, would at first disengage from the bombers whenever the escorts drew close. However, several of the jets would later press their attack on the formation. In the ensuing battle 1st Lt. Earl R. Lane, Flt. Officer Joseph Chineworth, and 1st Lt. Roscoe Brown would each be credited with a confirmed kill on three downed Me 262s. On their return flight they engaged several targets of opportunity, including two trains. The success of this mission earned the unit one of their three Distinguished Unit Citations, and the personal thanks of Gen. Lawrence of the 5th Bomb Wing (Bucholtz 108, 109).

Beyond their role as escorts for the 15th Airforce’s bombers, the 332nd would be engaged in a number of fighter bomber missions across the Meditteranean and Central Europe. These missions were conducted whenever time permitted between bombing raids and would see the squadron engage a number of targets. These would include airfields and various transportation targets varying from trains to river barges. A raid on August 30, 1944 would mark the unit’s most successful day when the 332nd attacked poorly camouflaged aircraft at Grosswardein airfield, Romania. In the ensuing strafing attack, they would be credited with the destruction of 83 aircraft with a further 31 damaged, ranging from 30 Ju 88’s, to a pair of super heavy Me 323 transport aircraft (Bucholtz 66). They would mount similar attacks against Axis airfields from Romania to Hungary.


Pilot’s of the 332nd, Lt. Clarence ‘Lucky’ Lester on the right, leads the group with 3 credited victories, all claimed on the same day. [National Archives]
The 332nd would end their campaign at an airfield in Cattolica, Italy, and was credited for the destruction of 111 aircraft in the air, 150 on the ground, 57 locomotives, 600 rail cars, and had flown 15,533 sorties (Bucholtz 116). It was a common myth that the squadron had never lost a bomber to enemy fighters, this being a rumor circulated by the press near the end of the war. This was not the case, but even with the failure over Memmingen, their bomber losses to fighters were half of the average and they were a considerable morale booster for the bomber crews of the 15th Airforce.

Flight Characteristics and Pilot’s Remarks

 

[P-51B pilot training manual]
The P-51B would prove to be an excellent fighter, but one that could present some challenges to those unaware of its quirks. It shared most of its general flight and handling characteristics with its older Allison powered predecessors, though some alterations to the design would make themselves felt, and not always to the plane’s benefit or pilot’s wishes.

Overall, the Merlin Mustang’s would prove to be fast and highly maneuverable, but with more complex flight characteristics than the Allison powered models that came before. Under most flight conditions, the plane was positively stable and possessed controls that were light and responsive. This aspect had been improved from the previous models, as the P-51B would be equipped with improved internally sealed and balanced ailerons which kept control stick forces light. These were rated very well, though pilots would note they were still ‘mushy’ at low speeds. However, as the plane’s top speed increased, it was capable of pulling maneuvers that could prove hazardous to pilots. Above 4g turns where a pilot without a g-suit was partially blacked out, the stick reversal could be harsh, but the worst of its effects were eliminated by a 20lb bobweight that was incorporated into the control system later on (Dean 350, 349).

The plane’s stall characteristics were mixed, but mostly mild. A one g stall in a clean aircraft was characterized by a roll to the right which came on after rudder buffering and aileron snatching, and was easily recovered from. Pilots were generally positive about the stall warning and recovery characteristics. However, its accelerated stall behavior proved to be far less universally understood. Some pilots claimed an easy recovery after ample warning, and others claimed it came on suddenly and viciously. Its low drag wings would contribute partly to this, as with its lack of air flow disturbances, stalls could come on without much warning. In the event of a spin, recovery was achieved by throttling back and pulling up while directing the rudder in the opposite direction of the spin. A spin could be serious trouble as a typical recovery resulted in a loss of about 9,000 ft in altitude (Dean 351, 352; P-51B flight manual 80).

While the plane was certainly very capable in regards to its maneuverability, pilots would have to take great caution when performing maneuvers of any kind when the fuselage tank still contained fuel. When the 85 gallon tank still contained fuel, the plane’s center of gravity shifted considerably and induced severe longitudinal instability. Hard maneuvers with any considerable volume of fuel still in the tank would result in a stick reversal that would require the pilot to brace themselves against the movement of the stick. Failing to do so would result in a loss of control or a further tightening of the turn which could result in a high speed stall or even structural failure (Dean 347, 348). Both RAF and USAAF manuals would ban aerobatics with roughly forty or more gallons of fuel in the tank, and suggested caution once it had been reduced to 25 gallons (Pilot’s Training Manual 68, Pilots Notes 30). In service this issue was one that rarely affected the plane’s effectiveness in combat, as the long range tank was the first to be used on long patrols and escort missions and thus typically contained little or no fuel when contact with enemy aircraft was made.

On early and mid production P-51B’s, pilots would also have to be cautious of high speed snapping brought on by the aforementioned longitudinal instability while they were conducting rolls. Pilots caught unaware were often injured during this violent jolt, and rolls were restricted accordingly. The addition of a fin extension for the vertical stabilizer and reverse rudder boost tabs would largely solve this issue, and the restrictions were lifted on suitably modified aircraft (Dean 350).

Perhaps where the aircraft shined the brightest were its dive characteristics, which were achieved as a result of its low drag wings and fuselage. These granted it excellent acceleration and a higher critical mach number than most of its contemporaries. Due to the changes in air flow across an aircraft’s wings as a plane approaches the sound barrier, most aircraft would experience buffeting, and a loss of control along and total loss of lifting forces. This change in flight characteristics that results in this loss of control is known as compressibility, a phenomenon that occurs when an aircraft exceeds the speed of its critical mach number.

A visual explanation of compressibility from the P-51B’s pilot training manual, the disturbed airflow results in a loss of lifting forces on the wings and control surfaces. The P-51’s wings mitigated the worst of its effects until much higher speeds. [Pilot’s training manual]
Thanks to its laminar flow airfoil, the P-51 was almost unique in its ability to remain controllable at otherwise unheard of speeds. However, in a high speed dive the P-51 would eventually experience compressibility and a pilot needed to be aware of the changing characteristics of their aircraft. In the P-51 this would first be felt through a ‘nibbling’ at the controls, afterwards by the stick ‘walking’ back and forth, and lastly by the aircraft pitching up and down with motions that grew more violent as the aircraft picked up speed (Pilot’s Training Manual 74, 75). On earlier models that lacked the vertical stabilizer extension, there was also directional instability that occurred at high speed, which required rudder correction or the plane could be sent into a spin. However with the later modifications the plane was nothing less than astounding. In diving tests from 35,000ft, pilots were able to reach mach .83 while retaining control of the aircraft, and despite the violent shaking and buffeting of the aircraft, were able to recover from the dive. In more practical conditions, control characteristics would remain normal until the aircraft was between .72 and .74 mach, after which the plane would experience escalating tuck-under, or a tendency to pull downwards airspeed increased. The maximum permissible dive speed was set at 505 mph IAS below 9000 ft, and 300 mph IAS at 35,000 ft, TAS being 539 mph (Mach .81). The maximum permissible engine RPM in dives was 3300 (Dean 341, 342, 343). Overall, the P-51B proved to be phenomenal in a dive, with only the British Hawker Tempest gaining a slight lead in tests, it being another aircraft equipped with laminar flow airfoils (Ethell 62).

Its take-off procedure was fairly typical of contemporary US fighters and required a strong right rudder deflection during take off to counteract the powerful torque from its engine. Its best climb out speed was between 160 to 170 mph IAS, which was quickly achieved after its flaps and landing gear had been retracted (Dean 341). Landing was somewhat more challenging, as the 140 mph IAS glide slope offered poor forward visibility, and little was improved as the plane came in to land at about 90 mph. It was thus fairly common for combat pilots to make tail up, level landings in order to have a better view of the landing strip before touching down. Its widely spaced gear and wide tire tread otherwise made the landing fairly easy.

While the P-51B’s possessed some truly phenomenal flight characteristics, the same cannot be said for the canopy. In US Navy evaluations the ‘birdcage’ canopy was found to result in poor all-around vision, most notably fore and aft. It was also fairly restrictive and made turning to view behind the aircraft more difficult (Dean 353). The frame itself could also not be opened on take off or landing and thus proved to be of some annoyance to pilots. This would later be solved with the addition of the ‘Malcom Hood’ which provided excellent visibility and was far less confining. The rest of the cockpit was judged to be satisfactory and capable of accommodating pilots of varying stature.

The ‘birdcage’ was unpopular as it was quite restrictive in terms of visibility, and it could not be kept open on the landing approach or takeoff. [Pilot’s Training Manual]
Its armament however, was distinctly lacking and fairly unreliable. It’s armament of four .50 caliber AN/M2’s was considerably lighter than most US fighters of the time and were installed in such a way that the ammunition links were prone to deformation in high-g maneuvers. It was not uncommon for P-51B/C’s to return from their missions with several guns malfunctioning as a result of failures to feed or extract. As a gun platform, its qualities were judged as roughly the same as the P-40, and below those of the P-38 and P-47 (Dean 353).

Comparisons with American Fighter Aircraft: Early to Mid 1944

Entering service alongside the P-47 and P-38, the new P-51’s would compare very well. When it came to the P-47D, equipped with R-2800-63’s, these aircraft were in some ways complementary, and excelled in areas the other did not. Thanks to its powerful turbosupercharger, the P-47 would retain the power needed to outperform the P-51 above 25,000ft, but was significantly slower at lower altitudes. The P-47 was also less vulnerable to ground fire and thus better suited for ground attack missions. The P-51B however, outstripped the P-47D in rate of climb, linear speed, acceleration at altitudes below roughly 30,000ft, and dive performance (Ethell 70; Marshall & Ford 526). Ergonomically speaking, the P-51B was the superior aircraft, as the turbosupercharger controls of the P-47D added to the workload of the pilot.

The P-47’s Turbosupercharged R-2800 engine provided unparalleled performance above 30k feet, and it’s durability made it ideal for fighter bomber missions. It was fairly lacking in its rate of climb and acceleration at low to medium altitudes. [National Archive]
When it came to escorting bombers, the P-47D and P-51B were the most effective tools at the USAAF’s disposal. Both aircraft performed superbly at and above the altitudes the bombers typically flew at, though the P-51B would prove the more vital as it could travel significantly further. By late spring 1944, external fuel tanks had been introduced that extended the P-47’s escort radius across most of Germany, however, by this time the P-51B was capable of accompanying bombers beyond Poland (Marshal & Ford 516). While the shorter range of this aircraft was often used to excuse the high bomber losses during earlier campaigns, the fact is that had they been supplied with the proper external fuel tanks, they would have been capable of deep incursions into German airspace months before the P-51 entered service.

The P-38 experienced serious reliability and performance issues due to the extremely low temperatures encountered at high altitude over Northern Europe. Its poor high altitude dive performance was also widely known, and exploited by Luftwaffe pilots. [National Archives]
The older P-38J Lightning would not stack up quite as favorably against the new Mustang. While the P-38J possessed a better climb rate and acceleration, it was out-stripped in linear speed by the P-51B at all altitudes, and possessed a very low critical mach number which meant that virtually any opponent at high altitude could escape by diving away. To make matters worse, a number of technical and operational issues spelled trouble for these aircraft in the colder Northern European climate. These issues, compounded by the extremely poor cockpit and canopy of the P-38, saw Lightning squadrons fall behind Thunderbolt and Mustang squadrons in victory credits (Marshall & Ford 439, 516; Ethell 70).

While the P-38J would receive external fuel tanks that would allow it to travel to Berlin and back, it was held back by a number of factors that severely reduced its combat effectiveness. In the European Theater of Operations, the P-51B would present a clear and general improvement over the P-38s, which saw more success in other theaters with conditions that they were better suited to, namely the Mediterranean and Pacific.

German Fighter Comparison: Early to Mid 1944

Entering service near the end of 1943, the P-51B compared very well to the German Fw 190As and Bf 109Gs in service at that time. The typical Bf 109 encountered through the first half of 1944 was the Bf 109G-6 series, which possessed better firepower than those that preceded it, but was heavier, and initially slower for it. These planes were equipped with either the Daimler-Benz DB 605A, or the high altitude, DB 605AS engines, both of which were later equipped with MW-50 boost systems. In all cases the P-51B possessed the superior linear speed, but in the case of MW-50 equipped aircraft, the Mustang had a slightly lower climb rate at low to medium altitude (Marshall & Ford 526, 523; P-51 flight tests). Without the boost system, which came into widespread use in the summer of 1944, the Bf 109G-6 was considerably slower and had a clear disadvantage in top speed and climb rate at all altitudes. The disparity with the high altitude model was much narrower, though the P-51 still held an edge.

The Bf 109G-6 was the most common Luftwaffe fighter encountered by the P-51. Later versions boasted considerably higher engine power thanks to the MW50 boost system; they did not compare well to many western allied fighters prior to this. Here one prepares for a fighter bomber sortie. [Asisbiz]
When it came to maneuverability, both aircraft had their own advantages, with the Bf 109 having better low speed handling and the P-51 having the advantage at high speed. The dive performance of the P-51B was far superior even at lower altitudes as the Bf 109 experienced stiffening of the elevator at high speed.

Visibility the Bf 109 was more or less on the same level of the standard ‘birdcage’ P-51B, and this would largely remain the case, as both planes would be re-equipped with improved canopies that offered better visibility. However, the cockpit of the P-51 was considerably more spacious and was further improved by the Malcolm hood. The Bf 109’s greatest strength was that it was equipped with an automatic RPM governor and mixture control that took a great deal of work off the pilot.

In terms of armament, both aircraft were comparable, with an unmodified Bf 109G-6 possessing a pair of 13mm machine guns and either a 20 or 30mm cannon, which fired through the propeller hub. Of the two, the 30mm was far less common.

Overall, the Bf 109G-6 was a somewhat dated fighter, one that had its advantages, but was  generally outclassed by the new Mustang. However, upgrades like water-methanol injection, an improved vertical stabilizer, and a new canopy helped keep the aircraft competitive and staved off obsolescence. The much refined ‘Kurfurst’ series would match P-51 performance in a number of areas, but its introduction was well after the Luftwaffe had lost control of German airspace.

The P-51B would face several models of the Fw 190A, with the most up to date being the A-8. The P-51B would have considerable linear speed, climb, and high altitude dive advantages over the earlier models. The Fw 190A-8 would have the benefit of a significant boost in power to its BMW 801D-2 engine, first by means of a fuel injection system, and in the summer of 1944, they were judged robust enough to be run at higher manifold pressures and had their supercharger boost regulators overridden. These modifications allowed the engine to produce significantly more power and increased the aircraft’s top speed at all altitudes (Douglass 344). In terms of top speed, this put these two aircraft on closer footing at low altitude, and ahead of the other two American fighters. It was, however, nowhere close to offsetting the general disparities at higher altitudes. The excellent defensive characteristics of the aircraft helped to offset some of its disadvantages against the P-51, as the Fw 190A held the best roll rate in the theater, solid dive characteristics, and good rearward visibility.

The Fw 190A’s were completely outclassed at altitude by the P-51B, owing to their relatively low full throttle height. They would however, be on somewhat closer footing at lower altitudes and could hold their own against the other two American fighters. [Asisbiz]
In terms of armament it was no contest, as the earlier A-6’s and A-7’s possessed a pair of either 7.92mm or 13mm machineguns respectively, and a pair of 20mm cannons. This was increased to two pairs on the Fw 190A-8. In regards to ergonomics the Fw 190A was excellent, with good visibility, clean instrumentation, and an advanced engine control system which handled RPM, manifold pressure, and mixture through the use of a single, integrated electro-mechanical computer. Its controls too were tight and responsive, if a little heavy at speed, thanks to its push rod control system. However, as was also the case with the Bf 109, its cockpit was comparatively cramped compared to the P-51.

Subsequent models of both these aircraft, the most numerous being the Bf 109G-14 and the Fw 190D-9, would largely eliminate the performance disparity at low altitude. However, at medium to high altitudes, the P-51 would still enjoy a considerable edge in top speed, dive performance, and high speed maneuverability. Only later Bf 109G’s with enlarged superchargers and better high altitude performance were close to closing the gap, with the K-4 series finally achieving high altitude parity near the very end of the war.


The Bf 109G-14 and the Fw 190D-9 would enter service in the Autumn and Winter of 1944, though they would not entirely replace their predecessors by the end of the war. [Largescaleplanes, Asisbiz]
The Me 262 presented a much greater threat in the air for obvious reasons. The jet fighter possessed a top speed roughly 100 miles per hour faster than the P-51 and was the only Luftwaffe fighter capable of following it into a dive. It was, however, considerably lacking in acceleration, which presented itself most dangerously on take off and on the landing approach. While the high top speed of the jets meant that they could disengage safely from most confrontations, they were helpless if caught near taking off or landing. Thus the general strategy for defeating these aircraft was to catch them as they were returning to their bases, where Allied fighters would await them. This is not to say this was easy, as their airfields were well defended by some of the best flak units available to the Luftwaffe and they would eventually have their own dedicated fighter cover (Ethell 97, 98). Higher up the jet could prove a deadly opponent as when flown well, it was extremely difficult to catch and an experienced pilot had control over most engagements.

The Me 262 was a world first, and had many USAAF planners concerned. On paper it had the ability to wreak untold havoc on allied bomber formations, but its technical limitations and the general poor state of the Luftwaffe late in the war prevented it from operating in numbers large enough to make a major impact. [Asisbiz]
In any case, encounters with the new jet fighters were fairly uncommon as they were constrained by operational restrictions owing to the temperamental nature of the new turbojet engines and the lack of a dedicated trainer for the aircraft until late 1944. They would not be seen flying against the Allies in appreciable numbers until the late autumn of that year.

Building the P-51B & C

The P-51B’s and C’s were built at plants in Inglewood, California, and Dallas, Texas, respectively. The distinction exists due to the differences in manufacturing between these two facilities, but these are functionally the same aircraft. With the exception of the earliest model, the P-51B-1, which had a different aileron design, their components were interchangeable. The main production models were equipped with the Packard V-1650-7 engine. Deliveries of these models began in February of 1944 (Marshall & Ford 253)

Production of this aircraft was complicated greatly by the breakneck pace of its procurement, which saw massive orders placed before its prototype had completed testing. As such, the aircraft that left the factories differed considerably even when they were built mere weeks apart. While all WWII fighters underwent constant modification, the level and rate of changes made to the P-51B and C were extensive and rapid. In addition to minimal changes, like changing the pilot’s seat from a wooden one to a magnesium one, in a matter of weeks the P-51B would receive an additional fuselage fuel tank, an extension to its vertical stabilizer and a rudder anti-balance tab, and an elevator control system which made use of a 20lb bob weight (Dean 329). These features would constitute a considerable challenge to work into the design without compromising the pace of production for an aircraft that USAAF planners wanted in as great quantity in the shortest possible time.

The Inglewood P-51 production line. [North American Aviation]
This challenge would highlight both the greatest strengths and weaknesses in US aircraft manufacturing. Most aircraft factories in the US operated by building large batches where the design would be frozen to allow faster construction. Modifying the design meant changes to the production line, which meant slowing down or stopping. US factories operated at batch sizes of up to 1,500, compared to the British Supermarine Spitfire’s production lines which operated at or below 500. The compromise was the modification center, to which “finished” aircraft would be delivered to be fitted out to new modifications. In practice, this system was extremely inefficient and saw quality control drop significantly. It also proved to be a highly inefficient use of labor, and could represent between 25 to 50% of the total labor required to complete an aircraft. Quality control also dropped considerably as the modification center was primed to try and deliver aircraft as quickly as possible (Zeitlin 55, 59). Lastly, the centers saw a great deal of wastage of material, accumulating a much larger proportion of metal scrap from rushed fittings, and ruined parts than the production lines (O’Leary 142). The USAAF would have its Mustangs, but only at a considerable cost and of initial questionable quality.

In the end they were successful in that they delivered the P-51B in great quantities despite the rushed pace of procurement, development, and production. However, it certainly contributed to the severe teething issues experienced by the aircraft that would see it briefly grounded in March of 1944 and would trouble it for weeks later.

In all, 1,988 P-51Bs were built with the first leaving the production lines, at a very low initial rate, in the summer of 1943 with the first deliveries taking place in August, with a further 1,750 P-51C’s being built. Production of both types declined as the P-51D production began in January of 1944, with the last P-51B’s leaving Inglewood in March and P-51C production continuing for several more weeks (Dean 321).

Construction

Wings

The wing group of the P-51 was composed of each wing, bolted together at the centerline. Each wing was of a cantilever stressed skin construction and consisted of a main panel, the wingtip, the flap, and the aileron. The main panel was built up around a main forward spar and a rear spar, to which twenty one pressed ribs were attached. These spars were spliced together roughly around half their length. A self-sealing 90 gallon fuel tank was fitted at the inboard section and a bay for its .50 caliber machine guns and ammunition was found near the center. The ailerons were of a fairly heavy construction, being all metal and supported by two spars and twelve flanged ribs. They were aerodynamically balanced by a diaphragm attached to the forward edge of the aileron and sealed to the rear spar by a fabric strip. These were controlled by means of a cable, as were all of the control surfaces of this aircraft. These were equipped with trim tabs and were adjustable in flight. The flaps were all metal plain flaps that were hinged on three sealed ball bearings and were hydraulically actuated.

[Legends in their time]
The landing gear was hydraulically actuated with a fully retractable tail wheel. The main landing gear were fixed to the wings by a cast magnesium supports and were equipped with multiple disc brakes connected to the hydraulic cylinder by metal tubing. The wheels were 27 inches in diameter and possessed a fairly wide tread, which helped to give the P-51 excellent ground handling.

The wings of the P-51 were designed to achieve laminar flow and used a NAA/NACA 45-100 series airfoil. It would fall short of true laminar flow as even extremely minor surface imperfections resulted in airflow disruptions that made laminar flow impossible. However, these were among the most aerodynamically advanced wings used by any fighter during the Second World War, providing extremely low drag and excellent high altitude dive performance.

Fuselage

[Legends in their time]
The fuselage was composed of two main sections, both of which had a semi-monocoque construction. The main section was formed by four extruded longerons, around which the intermediate frames and stringers were connected. The upper longerons were extruded H-sections which extended from the sheet metal firewall and tapered into a T-section. The lower longerons, consisting of an H-section and U-channel, extended the full length of the main fuselage. This entire unit was made up of eight assemblies which were riveted and bolted together, these being the firewall, turnover, truss, upper deck, left and right side panels, radio shelf, web assembly, and the radiator air scoop.

The main fuselage section also contained the cockpit, the windshield being composed of a center pane of bullet resistant five pane laminated glass, with two Plexiglas windows to either side. The canopy was either a metal framed Plexiglas ‘bird cage’, or a Malcom Hood. The birdcage had panels that opened outward on the top and port side. The hood slid back across the rear of the canopy. Behind the pilot were lucite windows which enclosed the radio space. A relief tube was installed and stored beneath the seat, and proved quite useful considering the long flights that this aircraft commonly made.

Early P-51B instrumentation. [Legends in their own Time]
The rear section was comparatively simple, composed of two longerons, a shelf, five formers, and three solid bulkheads. The fuselage, as with the rest of the aircraft, was skinned in Alclad. This section was reinforced after structural failures during high speed rolls in early models.

Tail Section

The tail section was affixed to the rear fuselage and consisted of the horizontal stabilizer, elevators, fin, vertical stabilizer, and the rudder. The horizontal stabilizer was a one piece assembly supported by two spars, fixed to the fuselage by four bolts, and through which the vertical stabilizer was attached. The elevators consisted of a front spar with eighteen flanged ribs, and was initially fabric skinned with Alclad leading edges before it was later entirely metal skinned. These were fastened with five sealed ball bearing hinges and each had an adjustable trim tab.

The vertical stabilizer was supported by two spars along with four ribs and a detachable tip. Extensions to the vertical stabilizer by means of a fin were added to P-51B/C’s to correct for longitudinal stability issues with a full fuselage fuel tank, and to correct certain undesirable characteristics when the aircraft was put through a roll. The rudder was fitted at the rear of the stabilizer and was supported by a single spar to which twenty flanged ribs were attached. Much of the rudder was skinned with mercerized cotton, save for the reverse edge. The rudder was fitted with a trim tab and aerodynamically balanced by means of a 16.6 lb lead weight at the tip.

Engine Section

The engine section consisted of the engine mounting and external cowl components and was bolted to the firewall. The cowl consisted of a frame made of Alclad beams to which the cowl panels fastened. This frame acts as a cradle for the engine which is mounted by a bracket through anti-vibration units. The entire section is designed to facilitate easy access to the engine through panels, and the engine mount allows for the rapid removal of the Packard engine.

[Legends in their time]

Engine

The early models of the P-51B used a Packard V-1650-3, with this engine being replaced on the production line in February of 1944 with the Packard V-1650-7. These are largely the same engine, though their superchargers were geared for optimal performance at different altitudes and thus have different maximum outputs. The 1650-3 was designed specifically for high altitude use and gave the P-51B/C a full throttle height of 29,000 feet, the 1650-7 was geared to achieve a higher engine output at a FTH of 21,400 feet (Marshall & Ford 253).

These engines had a bore of 5.40 inches, a stroke of 6 inches, a displacement of 1,649 cubic inches, a compression ratio of 6.0:1, a width of 30 inches, a height of 41.6 inches, length 87.1 inches, a frontal area of 5.9 sq. ft, and a weight of 1690 lbs. They differed in that the -3 supercharger ratios of 6.391:1 and 8.095:1, and those of the -7 were 5.80:1 and 7.35:1 (Wilkinson 125, 127). They were both fitted with a four blade Hamilton-Standard 24D50-65 or -87 hydropneumatic propeller with aluminum blades of a diameter of 11 feet and 2 inches. These blades were either 6547-6, 6547A-6, or 6523A-24 types. The engine exhaust stacks were of a stainless steel construction which had a removable exhaust shroud to keep heat from the spark plugs and to reduce drag.

Packard V-1650-7 [Pilot’s training manual, Smithsonian]
Both engines used a two stage, two speed supercharger and was equipped with an aftercooler. The supercharger was automatically controlled and governed by the air pressure at the carburetor intake, which was found just below the prop spinner. The controls for the engine were conventional, requiring manual throttle and rpm adjustments.

Radiator and Cooling Systems

The engine was cooled by two separate systems, one dedicated to the engine, and the other cooled the supercharged fuel-air mixture. Both of these systems were connected through the main radiator matrix within the air scoop below the main fuselage, with the coolant flow maintained by an engine driven pump. A smaller radiator for the oil cooler was placed below and ahead of the radiator matrix for the engine and aftercooler. The radiator setup was designed to make use of the Meredith effect, which in practical terms meant that the heated air flow out of the radiator produced thrust which counteracted a large percentage of the drag incurred by the scoop. The outlet for the radiator was automatically controlled. This design was able to reduce net drag upwards of 90% and was one of the most important features which allowed the aircraft to achieve such a high top speed (Marshall & Ford 510).

[Legends in their time]
The hoses for the radiator which extended through roughly two thirds of the aircraft, and the unarmored radiator, which sat at the bottom center of the aircraft, constituted the most vulnerable part of the aircraft’s design. These made the aircraft fairly vulnerable to ground fire, as the high cooling requirements of the Packard Merlin engine meant that a failure of the cooling system wouldn’t take long to put the aircraft out of action.

Fuel System

The initial models of the P-51B possessed only two 92 gallon wing fuel tanks with an 85 gallon fuselage fuel tank being included later through modification kits and was eventually incorporated into the production line. The Mustang was also capable of carrying two external fuel tanks by means of wing mounts. Fuel was drawn only from individual fuel tanks, requiring the pilot to manage up to five individual sources of fuel throughout longer flights (Pilot’s Training Manual 26).

[Pilot’s Training Manual]
The inclusion of the 85 gallon fuselage tank would introduce new challenges, as the shift in weight caused by a full tank introduced severe longitudinal instability. For this reason this tank was the first to be consumed. The combined tankage was 269 gallons.

Armament and Armor

P-51B’s were equipped with four .50 caliber AN/M2 machine guns. Each inboard gun was supplied with up to 250 rounds, with the outboard weapons having 350 each. These guns were mounted at roughly 45 degree angles within the wing, which caused severe cycling issues when the guns were fired while the aircraft was pulling hard maneuvers. These issues were lessened with the addition of electric boost motors for the ammunition feed, but were not completely solved until the subsequent P-51D model. The guns were electrically heated to prevent them from locking up at high altitudes. These aircraft were typically equipped with the N-3B reflector gunsight, with later aircraft receiving K-14 gyroscopic gunsights.

[National Archives]
Wing pylons allowed the aircraft to carry a payload of up to 500 pounds at either side, being either external fuel tanks or bombs. These aircraft could be made to carry rockets by means of field modification kits. Armor plates were placed ahead of the radiator header tank, at the engine fire wall, and behind the pilot.

(Dean 355-376)

Conclusion

It would take a considerable effort to develop the P-51B from its Allison engined predecessors, and even greater hurdles would have to be overcome to produce them in the quantities needed. In the end, both were achieved and the P-51B would enter large-scale operation in the Spring of 1944. In spite of its harsh teething issues, it would become among the most decisive weapons of the Second World War. With its incredible range and medium and high altitude performance, the aircraft would prove instrumental in establishing air superiority over Western Europe prior to Operation Overlord, and contesting the skies over Germany itself.

P-51B production was switched over to the D model at Inglewood in March of 1944, but the aircraft would remain in service in large numbers through the end of the war. [National Archives]
Its design, while not revolutionary, was thoroughly advanced and represented a considerable leap in aerodynamics and airframe design. The P-51B would however, be only a starting point for the Packard Merlin Mustangs, as further refinements would result in the iconic, and much more widely produced P-51D.

Specifications

P-51B/C ( with Fuselage tank) Specification
Engine  Packard Merlin V-1650-3, V-1650-7
Engine Output [V-1650-7] 1630 hp [1720 hp]
Maximum Escort Fighter Weight 11,150 lbs (2x108gal external)
Gross Weight 9,681 lbs
Empty weight 6,988 lbs
Maximum Range [External Fuel] 1350 miles [2150 miles]
Combat radius [External Fuel] 375 miles [750 miles]
Maximum speed (V-1650-7) 444 mph (75″ Hg) at 20600ft
Armament  4x .50 cal M2 machine guns, 1200 rounds of ammunition
Crew Pilot
Length 32′ 2
Height (tail down) 12’8
Wingspan 37.03′
Wing Area 235.75 sq.ft

 

P-51B/C ( with Fuselage tank) Specification
Engine  Packard Merlin V-1650-3, V-1650-7
Engine Output [V-1650-7] 1630 hp [1720 hp]
Maximum Escort Fighter Weight 5058 kg (2×409 liters external)
Gross Weight 4391 kg
Empty weight 3169 kg 
Maximum Range [External Fuel] 2172 km [3460 km]
Combat radius [External Fuel] 603 km [1207 km]
Maximum speed (V-1650-7) 714 km/h (1905mm Hg) at 6279 m
Armament  4x 12.7mm M2 machine guns, 1200 rounds of ammunition
Crew Pilot
Length 9.80 m
Height (tail down) 3.86 m
Wingspan 11.29 m
Wing Area 21.9 sq.m

(Dean, Performance Tests on P-38J, P-47D and P-51B Airplanes Tested with 44-1 Fuel., Marshall & Ford)

Maximum Level Speed Speed at 67″ Hg, 3000 RPM 75″ Hg, 3000 RPM No wing racks, 75″ Hg, 3000 RPM
Sea level 364 mph 380 mph 388 mph
Critical altitude low blower 408 mph at 10400 ft 411 mph at 2300 ft 422 mph at 7400ft
Critical altitude high blower 426 mph at 23900 ft 431 mph at 20600ft 444 mph at 20600ft
Aircraft Specification Gross weight 9680lbs, P-51B-15  (V-1650-7)

*A note on fuels: The 75″ of manifold pressure figure represents the high end of performance using 150 octane fuels, these were typically only available to P-51 squadrons based in England.

Climb rate 67″, 3000 RPM 75″ Hg, 3000 RPM
Maximum at low blower 3,920 ft/min at 5600 ft 4,380 ft/min 2,300 ft
Maximum at high blower 3,170 ft/min at 19,200 ft 3,700 ft/min at 15,600 ft
Aircraft Specification Gross weight 9680lbs, P-51B-15

 

Maximum Level Speed Speed at 1701 mm Hg, 3000 RPM 1905mm Hg, 3000 RPM No wing racks, 1905mm Hg, 3000 RPM
Sea level 586 km/h 611 km/h 624 km/h
Critical altitude low blower 656 km/h at 3169 m 661 km/h at 701 m 679 km/h at 2255 m
Critical altitude high blower 685 km/h at 7284 m 693 km/h at 6278 m 714 km/h at 6278 m
Aircraft Specification Gross weight 4390 kg, P-51B-15 (V-1650-7)

 

Climb rate 1701 mm Hg, 3000 RPM 1905 Hg, 3000 RPM
Maximum at low blower 1194  meter/minute at 1707 m 1335 meter/minute 701 m
Maximum at high blower 966 meter/minute at 5852 m 1128 meter/minute at 4755 m
Aircraft Specification Gross weight 4390 kg, P-51B-15

(Performance Tests on P-38J, P-47D and P-51B Airplanes Tested with 44-1 Fuel.)

P-51 Variants through P-51D

North American USAAF RAF Engine Armament No. Built Additional Notes. First delivery
NA-73X Allison 1 Prototype. October 1940
NA-73, -83 XP-51 Mustang Mk I Allison 2x .50 cal MG, 4x .30 cal MG 622 RAF, export. August 1941
NA-91 P-51 Mustang Mk Ia Allison 4x 20mm cannons 150 ‘Plain P-51’. July 1942
NA-97 A-36A Allison 6x .50 cal MG, bombs 500 Dive Bomber. October 1942
NA-99 P-51A Mustang Mk II Allison 4x .50 cal MG 310 March 1943
NA-101 XP-51B Packard 4x .50 cal MG 2 (converted) P-51B prototype
NA-102, -104 P-51B Mustang Mk III Packard 4x .50 cal MG 1988 Inglewood production. Summer 1943
NA-101, -103 P-51C Mustang Mk IIIB Packard 4x .50 cal MG 1750 Dallas production. August 1943
NA-106 (through -124) P-51D Mustang Mk IV Packard 6x .50 Cal MG +8000 Bubble canopy. January 1944

(Dean 321)

P-51B & C Variants

P-51B & C Variants Notes Serial No.’s
P-51B-1-NA Earliest production model, steel aileron diaphragms, two point aileron attachment.  43-12093 to 12492.
P-51B-5-NA Three attachment points per aileron, non-magnetic diaphragm.  43-6313 to 6352, 43-6353 to 6752, 43-6753 to 7112.
P-51B-7-NA B-1s and 5s which received a new fuselage fuel tank carried this designation. Aircraft often carried prior designation in practice. Converted aircraft.
P-51B-10-NA Production model with fuselage tank.  43-7113 to 7202, 42-106429 to 106538, 42-106541 to 106738.
P-51B-15-NA Engine changed to Packard V-1650-7 (previous models were converted to this engine via supercharger kits).  42-106739 to 106908, 42-106909 to 106978, 43-24752 to 106738.
P-51C-1-NT Same as P-51B-5-NA.  42-102979 to 103328
P-51C-2-NT C-1s which received a new fuselage fuel tank carried this designation. Aircraft often used prior designation in practice. Converted aircraft.
P-51C-5-NT Same as P-51B-15-NA. 42-103329 to 103378, 42-103379 to 103778.
P-51-C-10-NT Production model with stabilizing fin extension. 42-10818 to 103978, 43-24902 to 25251, 44-10753 to 10782, 44-10818 to 10852, 44-10859 to 11036, 44-11123 to 11152.
P-51C-11-NT Production model. 44-10783 to 10817, 44-10853 to 10858,44-11037 to 11122.
F-6C Photoreconnaissance. Converted Aircraft.
TP-51C Dual control trainer. Converted Aircraft.

(Marshall & Ford, O’Leary)

Video

Gallery

Illustrations by Ed Jackson

XP-51B, 312093. The XP-51B’s were a pair of earlier Mustangs converted to use the Packard V-1650-3. Their cooling systems would prove the most troublesome, though the general teething issues these aircraft experienced were harsh and varied.
P-51B-7-NA 43-6913 ‘Shangri-La’. Debden, UK 1944. Debden ,UK 1944. This aircraft was flown by Capt. Don Gentile of the 4th Fighter Group, one of the unit’s leading aces.
P-51B. 325th Fighter Group. Poltava, USSR 1944. The 325th was among the units that participated in Operation Frantic, where they supported a series of USAAF raids launched from within the Soviet Union during the summer and fall of 1944.
P-51B-5-NA, 43-12214 ‘Rebel Queen’. Debden, UK 1944. This aircraft was flown by Col. Don Blakeslee, Commanding Officer of the 4th Fighter Group. This aircraft is an early production P-51B which had been equipped with a Malcolm Hood bubble canopy, this modification greatly improved visibility.
P-51C-10-NT ‘By Request’. Ramitelli, Italy 1944. This aircraft was flown by Col. Benjamin Davis, Commanding Officer of the 332nd Fighter Group. This is a late model which has been fitted with a fin fillet, extending from the vertical stabilizer. This addition greatly improved the aircraft’s stability in rolls and high speed dives.

B-17’s accompanied by a P-51B over England, March 1945.[National Archives]
A collection of P-51’s accompany a flight of B-24s of the 8th Air Force, near England. 1944. [National Archives]
The Malcom Hood bubble canopy would offer pilot’s great visibility compared to the ‘birdcage’. [National Archives]
The P-51A can be easily differentiated from its merlin powered counterpart by the tube shaped carburetor intake over the nose. [wikimedia]
Though most P-51B’s would be sent to Europe, some would serve in the China-Burma-India theater. Here a Mustang cruises alongside a C-47. [National Archives]
Ground crew pose alongside one of their planes. [National Archives]
A P-51B in the CBI theater is cleaned. This plane has had its exhaust fairing removed, a fairly common modification made in the field which some pilots believed cut down on drag. [National Archives]
A P-51B comes in to land, the wide tire tread and wheel base of these planes helped give these planes good landing and ground handling. [National Archives]
Ground crew pose with one of their planes, the tail fin extension as equipped to this plane helped alleviate some of the aircraft’s less desirable characteristics when it was rolled. [National Archives]
Among the challenges caused by segregation for the 332nd were personnel shortages. The only available training facility at Tuskegee struggled to turn out enough pilots and ground crew to support the segregated squadrons. Mechanics and armorers were among the most affected, especially when the fighter group rapidly transitioned from P-39’s, P-47’s, and P-40’s to P-51’s over the late spring and summer of 1944. [National Archive]
P-51B’s of the 325th Fighter group accompany bombers on their way to the Soviet Union during Operation Frantic. [National Archives]
The success of Operation Overlord saw the redeployment of many USAAF units to the continent. These P-51’s of the 9th AF were the first to be deployed to France. [National Archives]
The F-6C was a photo-reconnaissance variant that had a camera installed in the fuselage, the lens cover for which sits here just behind the radiator scoop. This model was credited with the last kill in the ETO, after downing a Fw 190 on May 8 1945 (Dean 339). [Wikimedia]

Credits

  • Written by Henry H.
  • Edited by  Ed Jackson & Henry H.
  • Illustrations by Ed Jackson

Sources

Primary:

  • Flight Tests On The North American P-51B-15 Airplane, AAF NO. 43-24777, 1944.
  • Preliminary Results of Performance Tests on a P-51B Airplane with 44-1 Fuel P-51B-5-NA, V-1650-7 Engine. 1944.
  • P-51B-15-NA 43-24777 (Packard Merlin V-1650-7) Performance Tests on P-38J, P-47D and P-51B Airplanes Tested with 44-1 Fuel. (GRADE 104/150). 15 May, 1944.
  • Matthews, H. F. Elimination Of Rumble From The Cooling Ducts Of A Single-Engine Pursuit Airplane. NATIONAL ADVISORY COMMITTEE FOR AERONAUTICS. 1943.
  • Messerschmitt A.G. Augsburg. (1944). Leistungen Me 109 G-14/U4 mit Db 605 Am u. Asm. 1944.
  • Pilot Training Manual for the Mustang. United States Army Air Force, 1943.
  • Pilot’s Flight Operating Instructions P-51B-1 Airplane. Evansville, Indiana: United States Army Air Force, 1943.
  • Pilot’s Notes for Mustang III Packard Merlin V-1650-3 Engine. Air Council, 1944.
  • Wilkinson, Paul. Aircraft Engines of the World. 1944.
  • The United States Strategic Bombing Survey: Over-All Report (European War). U.S. Govt. Printing Office, 1945.

Secondary:

  • Bucholtz, Chris. 332Nd Fighter Group: Tuskegee Airmen. Oxford: Osprey Publishing, 2007.
  • Bucholtz, Chris. 4Th Fighter Group “Debden Eagles”. Oxford: Osprey Publishing, 2008.
  • O’Leary, Micheal. Building the P-51 Mustang the Story of Manufacturing North American’s Legendary WWII Fighter in Original Photos. Specialty Pr Pub & Wholesalers, 2011.
  • Dean, Francis H. America’s Hundred Thousand: the US Production Fighter Aircraft of World War II. Schiffer Publ., 1997.
  • Douglas, Calum E. Secret Horsepower Race: Second World War Fighter Aircraft Engine Development on the Western Front. TEMPEST, 2020.
  • Ethell, Jeffrey L. Mustang: A Documentary History of the P-51. London: Jane’s, 1981.
  • Haulman, Daniel L. Nine Myths about the Tuskegee Airmen. October 21, 2011.
  • Marshall, James William; Ford, Lowell. P-51B Mustang: The Bastard Stepchild that saved the Eighth air force. Bloomsbury Publishing Plc. 2020. (Electronic)
  • Moye, J. Todd. Freedom Flyers: The Tuskegee Airmen of World War II. New York, NY: Oxford University Press, 2012.
  • Overy, Richard James. The Bombing War: Europe 1939-1945. London: Penguin Books, 2014.

Northrop P-61 Black Widow

sweden flag USA (1943)
Night Fighter – 706 Built

A P-61A of the 422nd NFS. (San Diego Air and Space Museum)

The Northrop P-61 was a night fighter designed to fulfill a largely overlooked gap in America’s air defenses in the years prior to its entry into the Second World War. Ambitious and groundbreaking, the P-61 would be the first fighter aircraft designed to carry a radar and was to be equipped with a state of the art remotely operated turret. However, the aircraft suffered numerous technical problems which led to many delays in its development. Despite its quirks, the plane proved to be popular with its pilots, effective in service, and far more capable in its mission than preceding American night fighters, while also proving itself effective in roles not envisioned at the time of its design.

Nascent Developments

The US Army’s night air defense services during the interwar years were perhaps their most neglected and least developed. This was in part due to the meager capabilities of the aircraft and detection systems of the time, but also general disinterest from senior leadership and, resultantly, poor funding. Despite advances in night flying instrumentation and training aids, most notably Edward Link’s ground trainer, efforts during the period to detect and intercept aircraft at night were largely futile. Success in testing was almost entirely based on the weather, as the search lights they coordinated with relied on acoustic detection, and their ability to find the enemy was largely based on luck whenever the skies were not clear. Attempts were even made to detect enemy aircraft by the weak electromagnetic waves emitted by their spark plugs, but these were met with predictably poor results. What methods they did develop were subsequently made useless by advancements in bomber design, as the new Martin B-10 was faster than most contemporary fighters, and the B-17, still in development at the time, showed even greater promise (McFarland 3, 4).

Andrew Link’s ground trainer allowed pilots to train for blind flying without actually taking to the air [American Society of Mechanical Engineers]
Night fighters would prove a largely unworkable concept during the interwar years due to the crude instruments employed to find the bombers, which themselves also stood a good chance of outrunning their pursuers, however, this would soon change. The development of radar and more capable fighters would prove to be the decisive factor that would transform the practice of intercepting aircraft at night from a clumsy mission dependent more on luck than anything else, to an essential service that would grow ever more precise in its ability to detect and bring down enemy aircraft.

 

Lessons Abroad

The night raids during the Blitz both proved the necessity of night fighter forces and would provide the first lessons needed to found the service [Encyclopedia Britannica]
Virtually all major new developments of the Air Corps’ night fighting capabilities in the years prior to the US entry into the Second World War were a result of two factors, new developments in radar and reports from observers sent to take note of the lessons the RAF were learning during the fall of France and the Blitz.

As the clouds of war drew over Europe during 1938 and 1939, it was clear that airpower would be a decisive component of any potential conflict. It was for this reason that president Franklin Delano Roosevelt massively built-up US military forces for the goal of defending both the mainland United States and its overseas military installations in 1939 and 1940. This build up had broad aims, but perhaps most importantly it saw the vast expansion of the US Army’s air power. This was to prove instrumental for those officers who wished the service to take on a much larger role in the US Armed Forces, and to finally cement their position in it, as the department had been reorganized several times during the interwar period.

They would soon see a massive leap in responsibilities as the Air Corps took up the bulk of air defense duties with the founding of the Air Defense Command in February of 1940. However, despite their eagerness to play such a major role, they also recognized their lack of experience and sought to understand the fundamentals of the modern air war in order to better fulfill this task. With the war waging in Europe, General Henry “Hap” Arnold was able to argue for the presence of US Air Corps observers overseas. In the spring of 1940, four officers were sent to London, Paris, and Berlin (Harrison&Pape 26). While these early postings were important for shaping foreign policy and building ties that would facilitate easier coordination with the RAF later in the war, they would soon become an essential source of information for Air Corps planners following the fall of France and throughout the Blitz.

It was during the Blitz that perhaps the largest gaps in US air cover would become evident, with various solutions being presented to help bridge them. Thankfully for the Air Corps observers, the Blitz would demonstrate exactly what they would need to develop to face any threat from the air. They recognized that they needed a modern air force, which could cooperate with sophisticated detection and communication networks to form a comprehensive air defense system that would leave any attacker badly mauled, day or night.

Brig. General Tooey Spaatz was the primary observer for RAF night fighting operations, and it was no coincidence that he later became Chief of the Air Corps Material Division at Wright Field. While the British night fighter services were still extremely crude at this point in the war, they presented a much better starting point for US planners than the virtually useless interwar experiments. Spaatz’s efforts largely shaped the requirements for the Air Corps’ night fighter, these being relayed to Northrop’s Chief of research, Vladimir Pavlecka, while he was at Wright Field working on another project. Alongside a set of specifications, he was told the plane would need to be a two-engined aircraft with a crew of two, a pilot and a radar operator, though the specifics of radar were not disclosed. At this time, Northrop was a new company and made for an obvious choice, as they had previously worked on an unbuilt night fighter design for the British, and were one of the only firms that were not at capacity at the time.

Soon, this new aircraft, designated the ‘Air Corps Night Interceptor Pursuit Airplane’, began to take shape. It would be powered by a pair of Whitney Double Wasp engines carried in nacelles that would be connected by a twin boom tail, and joined to the fuselage through the wings. It would carry a crew of three, a pilot, a gunner, and a radio operator who also doubled as a rear gunner. It would mount two turrets carrying four .50 caliber guns each and would be a large aircraft with a height of 13 feet and two inches (4.013 m), a length of 45 feet and six inches (13.87 m), a wingspan of 66 feet (20.12 m), it would weigh 22,654 pounds (10276 kg), and feature the new Zaparka flaps.  While this proposal bore many similarities to the later XP-61 prototype, much would change as the design was revised.

While the design was promising, work was slow, and though Northrop had a prototype designed in January of 1941, it would be many months until it was ready to fly, and years before it was ready for service (Harrison&Pape 30). As a result, the aircraft would not be ready for the war to come, leaving most of the night fighting duties to stop gap designs, such as the converted A-20 bomber designated the P-70, and the British supplied Bristol Beaufort.

Clean slate, Dull chisel: A history of early American night fighting

With the attack on Pearl Harbor, the night fighter force found itself entering the war with the lessons learned from the RAF, but without adequate training programs or equipment. Compared to what existed in England, the communications and detection infrastructure was very poor, as it relied on high frequency radio sets which proved troublesome, lacked sufficient identify-friend-or-foe capability, with early warning radar set up in poor positions, and worst of all, they lacked a dedicated night fighter force.

With the help of RAF advisors, they set out to correct these faults, with the Air Defense Operational Training unit being activated March 26th, 1942. The 81st Fighter Squadron (special) was chosen on May 28th, 1942 as the first official night fighter training unit and was staffed with officers who were enthusiastic about the promise of this new mission. This unit was later placed under the new Night Fighter Department, which itself was reorganized as the Night Fighter Division and made subordinate to the Fighter Department . Their curriculum was composed on July 4th, 1942, as pilots were to be trained to a high level of proficiency in instrument flying, blind take offs and landings, night formation flying, night gunnery, pilot-radar operator interception teamwork, Ground Control Intercept (GCI), and general air defense procedures.

While this unit was extremely useful in testing and building confidence in new equipment, like the SCR 540 radar, it was hit by numerous hurdles which prevented it from turning out the number of pilots needed. This was mostly the result of a shortage of aircraft, as the relatively small number of P-70’s, DB-7’s, AT-11’s, T-50’s, and B-80’s would prove a serious bottleneck, as would the delays in getting Airborne Intercept ground trainers. This problem remained late into 1941, as the 81st was deactivated and its personnel used to form the 348th and 349th Night Fighter squadrons in October of 1941. These two squadrons would be used to train new night fighter personnel with the hope that they could build 15 squadrons by 1943. However, this training schedule was overly ambitious and hampered by insufficient supplies of equipment. Sadly, in line with much of the troubled program, night fighter pilots graduated with no fanfare or any formal ceremony. They simply signed on a line and received their wings.

Night Fighter Squadron building would continue slowly until James Doolitle would push for its accelerated development in late 1942. Doolittle, after being so impressed by RAF night fighters over North Africa, called for four new night fighter squadrons to be formed, these being the 414th, 415th, 416th, and 417th. It was not until July of 1943 that real Night Fighter Squadron development began in earnest, as more aircraft and training material became available, and the new 481st Night Fighter Operational Training Group was formed under the command of Lt.Col. Winston Kratz (Harrison&Pape 104).

Trial by Fire: Pacific

The P-70 was the USAAF’s first true night fighter. Unfortunately, in practice, the plane proved to be totally inadequate for the task. [SDASM Archives]
The first night fighter deployments were to Panama and Hawaii, with the first P-70s becoming available in January of 1942. These planes were first used in improvised night fighting squadrons, like the 6th Night Fighter Squadron initially based out of Hawaii. They were, however, badly constrained by their inadequate support infrastructure and, as pilots would soon find out, the performance of their aircraft. Problems soon arose over the personnel shortages which required volunteers from signal corps officers and the enlisted maintenance crews to serve as radar operators. Problems improved very little following their move to Guadalcanal in February of 1943, where conditions were brutal.

Their objective was to try and stop the nightly raids by Japanese bombers which came over the islands to conduct nuisance raids. The P-70s were vectored onto these aircraft using the ground-based SCR 270 early warning radar without success, as the radar could provide only the azimuth to the target but not its altitude. Neither the radar crews nor pilots had much experience with GCI procedures and, combined with the meager capabilities of the P-70, the night fighters brought down few Japanese aircraft. While the night fighter crew’s living conditions improved thanks to new prefabricated shelters, their operational success did not. Their challenges were made far more difficult as the Japanese adapted to their tactics and their bombers began to fly above the P-70’s service ceiling, and went so far as to imitate American GCI operators in attempts to give faulty information to pilots (Harrison&Pape 68). P-70 crews did all they could to improve the speed and operational ceiling of the aircraft through serious modification which included installing propellers from B-17F’s and P-38 fuel pumps, though to no success.

Conditions at airfields across the Solomons were extremely poor, and made worse by occasional bombing raids [National Archives]
Frustration with the P-70 even managed to motivate the squadron to modify some of the P-38s they had been supplied with in an effort to replace the P-70. This would prove difficult, as the aircraft were not equipped with radar as they were to be used in conjunction with searchlights to find their targets. Two P-38Gs were modified by Lt. Melvin Richardson and a squadron radar mechanic by adding a second seat behind the pilot and building an avionics pod for the SCR-540 out of an external fuel tank. These modified planes were much faster than the P-70s and were capable of reaching high flying Japanese bombers, however, they could not convert enough fighters. Both would receive the Legion of Merit for their ingenuity, but apart from this small victory, the 6th would lack the means to conduct their missions.

Overall, the night fighter squadron’s experiences in the South Pacific proved dismal, having neither the properly trained personnel, support elements, or even aircraft needed to effectively complete their mission. In the end, what success they did have was a result of their ingenuity and perseverance rather than specialized training or the equipment they had been issued. Each victory over the Japanese bombers was a hard-won achievement equally celebrated by the aircrews and the Marines the enemy harassed on a nightly basis.

Trial by Fire: Mediterranean

While the 6th NFS was still deployed to Guadalcanal, the 414th and the 415th left for England in March of 1943, where they would soon be retraining on Bristol Beaufighters. The switch from the P-70 was a difficult one, as between its tendency for ground looping and engine fires resulting from landing gear failure, the Beaufighter proved an intimidating plane for the US pilots. While the Beaufighter proved to be significantly faster and more agile than their old P-70s, many pilots felt uneasy flying it, and even their RAF instructors would readily admit the aircraft was among the most difficult in British service. Unlearning the habits from the P-70 was difficult, but thanks to a comprehensive program from the RAF, the challenge was soon overcome. Now proficient, the 414th and 415th left for North Africa and went into action in July of 1943 (Harrison&Pape 80).

The Bristol Beaufighter would prove an effective weapon against the Luftwaffe and Regia Aeronautica, though its handling characteristics left much to be desired [National Archives]
Unlike their counterparts in the Pacific, the Mediterranean squadrons were largely successful thanks to their far superior Bristol Beaufighters, comprehensive training programs, and good technical support. Not only were they directed by far superior ground based radar systems, but these squadrons would later be the first to use the British AI Mk VIII centimetric radar sets, which, unlike the previous SCR-540, could operate at low altitudes. This radar was particularly useful, as it meant German bombers could no longer fly low to reduce the effective range of the aircraft’s radar. In November of 1943, the campaign proceeded and the two NFS’ would cover Allied convoys against the attacks of German bombers during the advance into Italy. The Germans would use a very different set of tactics compared to the Japanese, and made use of far more sophisticated equipment. As opposed to the single Japanese aircraft that often came in at around 30,000 ft (9144 m), the Germans tended to stay roughly between 10,000 and 15,000 ft (3048, 4572 m) in formations of various sizes. As opposed to the nuisance raids designed to keep the Marines from getting any rest, the Germans often sought to hit strategic targets, like harbor facilities and shipping vessels en masse.  The Germans would also later employ chaff, which cluttered up radar scopes, and tail warning radar on their bombers to warn them of the approach of night fighters.

The efforts of early American night fighter squadrons in the Mediterranean would thankfully prove to be the rule rather than the exception for the air crews still to come. In the future, they would expect well trained, specialized personnel, and effective ground control radar support. However, there were still strict limitations imposed by the equipment afforded to the night fighter squadrons, in particular their aircraft which, despite their greater speed, were anything but easy to fly.

XP-61 & YP-61: trouble, frustration, and promise

The XP-61 would prove to be a promising, but troubled design that would require a lengthy redesign [This Day in Aviation]
While American night fighter pilots had their first experiences in combat during 1942 and 1943, work on the new XP-61 continued. The program truly began in early 1942, after several contracts were issued. It was decided in February of 1942 that 410 aircraft would be procured with $7,136,689.56 being charged to the Defense Aide and $55,656,178.67 to Air Corps Appropriations. This contract stipulated the delivery of the first twelve to take place in April of 1943, with the final aircraft being handed over in January of 1944. However, contract negotiations saw the number of aircraft requested rise and fall significantly in the following months. In addition to deliveries to the USAAF, 50 P-61s were to be set aside for Lend Lease, though this was later dropped due to a lack of RAF interest in the aircraft (Harrison&Pape 89).

The XP-61 first flew on May 26th, 1943, but only after a long and troubled process that saw the prototype fly with different propellers than what were originally called for, a dummy turret, and without a radar. Despite these shortcomings, the prototype was initially quite promising, with the aircraft performing satisfactorily during its short preliminary flight, and its test pilot, Vance Breese, telling Jack Northrop “Jack, you’ve got a damn fine airplane!” (Harrison&Pape 89). The succeeding flights would, however, prove more troublesome ,as they soon revealed stability problems and reliability issues with the engines. These issues were tracked to the horizontal stabilizer and elevator assemblies, the short span Zap flaps, and the buildup of oil and gas in the engine crankcase. Engine failures and violent longitudinal instability soon grounded the XP-61. The stability problems were particularly troublesome, as they required redesigning much of the tail assembly of the aircraft and the addition of full span flaps in place of the Zap flaps.

Issues also arose with the use of the spoileron system on the second XP-61, which was designed to be used in conjunction with short ailerons as part of its lateral control system. The spoilers themselves were thin circular arcs that sat in grooves in the wings. These would deploy and allow the pilot remarkably good control at high speed, but their development would prove to be rather difficult. At first, they proved unstable as a result of their hinge movements, which was solved by reducing the size of the scoop. Later, serious vibration issues were found to destroy the spoilers within their wing slots. This, in turn, was found to be the result of high-speed airflow moving through the slot, and was remedied by installing plates to seal it off (Harrison&Pape 94).

In an effort to improve engine reliability, the initial Pratt & Whitney R-2800-25S engines were replaced with R-2800-10s, though this would fail to solve the problem. Despite the swap, engine failures were common in testing and it was found that cutting oil consumption resulted in oil starvation in the master cylinder, resulting in it freezing. With the rest of the articulating rods still going, the engine would work itself apart. After diagnosing the problem, the engineers at Pratt & Whitney soon resolved it with the installation of additional oil jets (Harrison&Pape 95).

These faults would see a major redesign of the XP-61, with the new model being designated the YP-61, which would act as both a prototype and pre-production model. This would incorporate a number of design modifications, including switching the tail boom’s construction from welded magnesium to aluminum alloy, the Zap Flaps being replaced by a full span trailing edge arrangement, and they would go on to incorporate the fix for the spoilerons (Dean 383). However, this would not represent an end to trouble, as the much-needed improvements in stability meant it could now be used for more demanding tests which would, in turn, uncover new faults.

As testing grew more strenuous, new problems began to arise, most notably with the fuselage. It was soon found that there were several structural weaknesses in the fuselage, with the vulnerable sections being subsequently strengthened, particularly around the canopy. The nose gear door also proved to be fairly fragile, which necessitated strengthening and having the aircraft’s 20 mm guns fitted with blast tubes to prevent the muzzle report from damaging it. It was also found that, at high speeds, the dome at the rear of the radar operator’s compartment could implode as a result of pressure difference. While this component was improved in prototyping, the problem would resurface in production models of the aircraft (Harrison&Pape 114).

At the nose of the aircraft, a new problem was found after a stopover in Arizona. In the sweltering 110-degree weather of the desert, it was found that the Plexiglas radome would deform and collapse once the aircraft was in the air, with the resulting shift in air flow causing the Lucite dome at the rear of the aircraft to blow out. While a new fiberglass dome was designed, it would not be ready well into serial production of the P-61A (Harrison&Pape 115).

The remote-controlled turret was also found to have several faults. The most immediate and concerning of these was that, while the guns were forward, the aircraft remained aerodynamically clean, but there would be intense buffeting as the turret traversed beyond 30 degrees in either direction. Following a redesign, the buffeting was mostly gone save for a far lighter effect that occurred when the turret was fully deflected to either side. In firing tests, it was found that use of the turret resulted in severe structural damage, with a similar problem being encountered with the aircraft’s 20 mm armament. In both cases, it was necessary to strengthen nearby structural elements with steel fittings, though owing to the complexity and size of the remote-controlled turret, a major redesign of the aircraft’s upper structure was required (Harrison&Pape 117, 118).

In spite of its many teething issues, the general flying characteristics earned the aircraft good marks from test pilots, exceeding Northrop’s guaranteed performance by one mile during speed tests, and was remarked upon by production project engineer Capt. Fred Jenks as follows: “The P-61 is an honest airplane. It has no mean tricks. In acrobatics such as loops, spins, Immelmanns, and fast turns, it behaves as a pursuit plane should. Its stalling gestures are near perfect.” (Harrison&Pape 121)

The YP-61, later redesignated P-61, would solve most of the issues of the troubled XP-61 and ease the transition for production models [NACA]

Britain and America on the same wavelength: Cooperation in developing the SCR-720

The Tizzard Mission

The P-61’s highly advanced air search radar was a device many years in the making, being a product of the interwar Tizard Mission, which sought to bring together US and British technical expertise for radar and radio development. It would officially become known as the British Technical and Scientific Mission, and it was not long into the war that approval was granted for the scientific material transfer to the US, which the British hoped would be reciprocated in kind by technical assistance and access to America’s electronics industries. The mission arrived in the US in September of 1940 with two gifts, a cavity magnetron, a device which allowed for the development of more advanced centimeter band radars, and an ASV Mark II surface search radar. Their audience was composed mostly of three groups, the Signal Corps, who had been struggling with the practical employment of radars for nearly a decade, microwave researchers, who were well versed on the technology but had yet to produce practical radar examples, and the US Army Air Corps, who were uninformed on technical matters but saw the promise of the technology (Brown 159, 160).

The mission got off to a good start, with the American audience suitably impressed. On the British side, they gained a great deal of information on the use of microwave techniques, and more importantly, access to larger electronics manufacturing industries and procedures. The US, on the other hand, gained access to the existing British radars and the cavity magnetron. While these advancements would have taken place eventually, the collaboration through the Tizard mission allowed rapid advancements in radar development and production in both countries. It was, of course, not without its negative consequences. Some trust in the American service-labs was lost when they gained the undeserved reputation of producing inferior equipment among the Armed Services when their interwar work was judged against the British (Brown 165, 166).

In the US, work on centimetric radar was mostly carried out through the MIT Radiation Lab, with their first goal being to produce an airborne centimeter band set (Brown 168). The benefits of a centimeter band radar over a meter band were considerable, as they would not require the use of drag inducing aerial antennas. The narrower band also meant less reflection from the ground, and while it was not yet known, they were less susceptible to jamming. In short, they represented a massive leap in capabilities over older radar sets (Brown 145).

The rooftop radar experiments above the Rad Lab were a major step in the development of practical centimetric radar [MIT Physics Department]
The Rad Lab’s first centimetric radar was a 10 cm band set operated from a roof in January of 1941, with an intensive development program to follow. It was hoped that, by February, it would be mounted in a B-18 for testing, and a month following that, they hoped to have it aboard an A-20. However, many issues plagued the rooftop experiment and it was not until March that the device was transferred over to the B-18. Work would continue, and after further collaboration with the British, a new series of technology transfers would benefit both programs, with the British gaining access to a better transmitter, and the US a better receiver. Soon, the lab would produce America’s first practical centimetric Aerial intercept radar, the SCR-520. Western Electric built a few of these sets but work soon transitioned to making a lighter version for the P-61, the SCR-720 (Brown 168, 169).

SCR 720

The SCR-720/AI Mk X was comprised of a number of components, some of which were encased within pressurized canisters [Mossie.org, thisdayinaviation]
The SCR-720 series was an advanced aerial intercept radar built by Western Electric and Bell Telephone Laboratories (Harrison&Pape 113). The radar operated on a wavelength of 9.1 centimeters at 3,300 megacycles, with a peak pulse power of approximately 70 kW. It had a maximum range of about 6 miles (9.65 km), which was later extended beyond 10 miles, at all azimuths between a search angle of 75 degrees to either port or starboard, with a minimum range of 300 feet (91 m). The system used a helical scan method and, in addition to the previously stated horizontal search angle, covered a total elevation of –30 degrees to +50 degrees. The device lacked provisions for IFF gear but could be used in conjunction with the SCR-729 transmitter, which was compatible with Mk III and Mk IIIG IFF sets along with beam approach beacons.

SCR-720 Radar operator’s display and controls [SCR 720 manual]
The display set up was composed of two indicator boxes, a two screen display to be used by the operator, and a far simpler one screen display which was for use by the pilot. The pilot’s indicator typically went unused, as it was less precise than the rear set and was generally redundant, as the pilot would be talked onto the target by the RO. The RO’s scope consisted of a range tube on the right, and an azimuth and elevation tube on the left. The display on the azimuth scope was dependent on the settings of the range scope, as only targets within certain set ranges would appear on the scope. This range was indicated by a marker line on the range scope and could be adjusted by the RO (Survey 28). The settings of these scopes were adjusted through the control box and synchronizer in the RO’s compartment.

In service, the SCR 720 offered many benefits over previous Allied centimeter band aerial intercept radars, in particular the slightly older British AI Mk VIII. A post war survey found that, while using the SCR-720, it was harder to lose maneuvering targets on the scopes thanks to the range/azimuth display which allowed the operator to follow the target’s course. The wide coverage meant it was unlikely a rapidly descending target would be lost, the range scope made course adjustment estimates to the target easier, and the range/azimuth display made intercepts across the flight path of the night fighter easier. Overall, the SCR-720 would prove to be the most precise and advanced AI radar set of the war and would see widespread use aboard the P-61 and DeHavilland Mosquito, which carried a British production of the device designated the AI Mk X (Survey 30).

However, while the radar was the best set in its day, it was also the most complex and was said to require the knowledge and experience on the level of a masters of electrical engineering just to make one’s way around the black boxes that made up the system (Harrison&Pape 113). In service, it would prove even harder to maintain where personnel and spare parts for the system would be sparse, and many of its components would prove vulnerable to the elements. This would generally prove an issue with P-61 squadrons, as they typically lacked personnel able to fix the boxes should problems arise, and often would not have enough spares to replace faulty components. Maintenance notwithstanding, the SCR 720 was a generation ahead of the previous SCR-540 and exceeded it in every capacity except ease of repair.

SCR 720 display diagrams [Radar Survey]

Enter the Black Widow: P-61A&B

An early production P-61A [National Archives]
The first P-61A rolled off the line in October of 1943 at Northrop’s plant in Hawthorne, California, with a public reveal later occurring at an Army-Navy show in Los Angeles in January of 1944. These aircraft were mostly unchanged from the last P-61 pre production aircraft, though this plane and the next 36 P-61As would be the only examples of the A model to be equipped with the remotely operated turret. The turret would be absent from the remaining 200 As and many of the succeeding B model, only to be reintroduced after a redesign (Pape 120,121).

Despite the revised model which would arrive much later in 1944, the P-61’s performance was roughly the same for its entire wartime service, with no major overall increases in horsepower or any major modifications to the airframe, apart from those to allow it to carry additional fuel and bombs and rockets for intruder and ground attack missions. Both the A and B models were powered by the Pratt and Whitney R-2800, with many early P-61As using the R-2800-10 and all aircraft beyond the P-61A-15 using the R-2800-65. Both engines produced 2000 hp, with the only major differences being their magnetos and ignition systems (Pilot’s training manual 11). Revisions to the design were gradual and often very minor between subtypes. As the P-61A matured, many new additions were made, including a new fuel system, underwing racks for bombs and fuel tanks, a water injection system, and additional oil capacity. The water injection system would boost the engine’s power about 15%, but only for brief periods and unsuitable for a lengthy climb or long-distance pursuit (P-61 training manual 12).

The P-61B would go on to extend the nose, revise the trim and hydraulic systems, alter certain instruments and displays in the cockpit, revise the heating system, alter the landing gear doors, and restore the turret (Dean 383). Much more work was done with regards the P-61C, which made use of significantly more powerful turbo-charged engines, though this aircraft did not see wartime use.

While the aircraft would mostly resemble the early prototypes, there had been more than a few major reworks of the airframe, most notably, the dropping of the ‘Zap Flaps’ for near full span types with an added lateral control system which made use of spoilerons. While these did prove troublesome in testing, the faults had been ironed out and the system worked to the satisfaction of pilots. The final configuration made use of slot covers and seals for the spoileron slots that solved the vibration problems and allowed for great lateral control for such a large aircraft while requiring little force on the part of the pilot. While this system was unconventional, pilots rapidly adapted to its use and were immediately appreciative of it, as it allowed for easily applied control at both very high and low speeds. It was particularly useful during landings, where they allowed for precise control on approach thanks to the automatic adjustment of the lateral control system with the flaps (Ashkenas 13, 14). This system was a major factor in making this fighter among the most maneuverable in the USAAF inventory, in spite of it also being the largest and heaviest.

Tough most of the aircraft’s worst issues were remedied in the prototypes, a few made it into the production models. The most glaring of these would be the plastic radome fitted to the nose of the aircraft, and the lucite tail cone at the rear, both of which would constitute fairly significant structural weaknesses. The plastic nose fitted to many of the early P-61As was weather sensitive and was prone to warping in the hot, tropical weather of the Pacific, or simply if left uncovered during a sunny and particularly hot day. The solution was painting the nose of the aircraft in a bright, reflective white paint, which raised obvious disadvantages when the aircraft was trying to stealthily pursue its targets in the dark. The lucite cone would prove more persistent and more dangerous (Harrison&Pape 115). On several occasions, these cones imploded during dives and high-speed maneuvers. While this presented little danger to the overall aircraft, the sound of rushing air through the radar operator’s position made communication between him and the pilot virtually impossible. This would be resolved later by the addition of metal reinforcement bands, though many older P-61’s would continue to fly without them.

While this was the first purpose built night fighter, in many ways, the aircraft fell short of the high hopes placed upon it, but would prove adequate for the purpose it was given. Responses to the P-61 were mixed though generally favorable, but complaints over speed and the difficulty of maintaining the SCR 720 radar persisted for all wartime models of the aircraft. Perhaps most unfortunate was that the top turret, which vastly complicated the aircraft’s design and added considerably to the aircraft’s frontal area and weight, was found to be completely unnecessary. The most immediate requests for improvement were for more powerful engines with better high-altitude performance, and for units to be supplied with more maintenance and test equipment for the SCR-720 radar, which the inadequately prepared ground crews struggled to keep in working order. While they were trained in basic maintenance and installation of the device, few had the technical skills necessary to actually repair faulty components. Neither of these would be provided in time to be of use during the war.

ETO: The 422nd and 425th Night Fighter Squadrons

French workers repair a runway at an airfield operated by the 9th AF [National Archive]
The first P-61 to leave the United States was a P-61A to be evaluated by the RAF in March of 1944. This aircraft was later returned in February of 1945, as the RAF were not particularly impressed with its performance and found its maximum range to be far too low. They needed night fighters for deep penetrations into German airspace in support of the ongoing strategic bombing campaign, and the P-61 simply did not fit the requirements. While the aircraft was by all metrics a poor fit for the RAF, the language and tone surrounding the growing competition between the P-61 and DeHavilland Mosquito would become increasingly petty and hostile within certain sectors of the US War Department and the Night Fighter division.

Beyond this evaluation aircraft, three Night Fighter squadrons would be deployed to England in anticipation of Operation Overlord. These were the 422nd, which departed on March 10th, the 423rd on April 1st, and the 425th which departed May 1st. During this period, only the 425th had P-61As slated to be shipped out with them amidst general concerns regarding the availability of the aircraft. The 422nd and 423rd were still equipped with the inadequate P-70 at the time of their departures (Overlord Build Up). Supplies would thankfully become more available, with the 422nd getting their first P-61As in late May and the 423rd becoming a photographic reconnaissance squadron and would not require the aircraft (Dean 285). Both squadrons would possess a small number of aircraft, with the 425th shipping out with only nineteen aircraft, and throughout their service in the European Theater it remained the case that replacement aircraft were in short supply.

None of these P-61As were equipped with the dorsal turret, and with copious time on their hands and the feeling that a second pair of eyes looking forward would be helpful, several crews in the 425th NFS had the bright idea to move the radar operator’s position up into the now vacant gunner’s seat. Along with technical representatives from Northrop, the chief radar and engineering officers, and a Capt. Russell Glasser, who possessed a graduate’s degree in mechanical engineering, they set out to modify the aircraft. The results were spectacular, with the pilot and R/O now able to communicate in the event of intercom failure, and the resulting shift in weight changing the slightly nose up to a nose down at cruise, increasing the cruising speed between 15 and 20 mph. This change was subsequently authorized for 9th Air Force’s P-61s (Harrison&Pape 205, 206).

Ground crews prepare a P-61 for action [National Archives]
It was not until July of 1944 that the European P-61s actually flew their first combat sorties, with the several months prior to this being taken up by training, including joint exercises with the RAF, and a race between the P-61A and a DeHavilland Mosquito NF Mk XVII. The latter was precipitated by a rumor that the USAAF was planning to replace the P-61 with the Mosquito.

In June, Lt. Col. Oris B. Johnson arranged for joint training with an RAF Halifax bomber squadron based at Croft, during which the Night Fighters would practice intercepting the bombers, who would in turn practice evasive maneuvers and other defensive tactics. The night fighters would be given an area to defend and would be vectored onto bombers by GCI. When they were in place to claim a ‘kill’, they flashed their navigation lights (Harrison&Pape 206). While this exercise was undoubtedly easier than what they would later be asked to do over France, it was important in building up the crew’s confidence in their abilities and equipment.

The ‘race-off’ was an event long in the making, with its roots in the War Department’s desire to purchase DeHavilland Mosquitoes for use as reconnaissance aircraft and night fighters. There were those in the department who wished to equip the Night Fighter Squadrons in the Mediterranean with Mosquitos, with the ensuing politics eventually driving a rumor that the War Department was planning to scrap the production of the P-61 in favor of the Mosquito, which were to be supplied by the UK and Canada. In any case, these proposals were impractical, as the British were extremely protective when it came to these aircraft. However, rumors soon filtered to the squadrons who were upset enough to propose a fly-off between the types. A demonstration was arranged on July 5th, 1944 at RAF station Hurn. The contenders were a P-61A and a Mosquito NF Mk XVII, with the results being that the P-61 out climbed and out turned the Mosquito between 5,000 and 20,000 feet.

The race was anything but clear cut, and it is extremely unlikely that it was just a fair competition that both sides took part in earnestly. Simply put, the RAF did not want to give the USAAF any more Mosquitoes than they absolutely had to, and were extremely motivated to throw the race. They had a great desire to ensure they were better supplied with the only night fighter in Allied service at the time that could fly long range missions into Germany. The results of the race are extremely suspicious given just how clear the P-61A’s win seemed to be in comparison to the years of evaluations which virtually always claimed that the Mosquito NF had the superior climb rate, and the P-61 had superior maneuverability. Members of the 481st NFTG who had flown planes came to the same conclusion, as did the AAF board, and even Col. Winston Kratz, director of night fighter training and a major proponent of the P-61 (Harrison&Pape 153, 156, 203). His words perhaps best sum up the event, “I’m absolutely sure the British were lying like troopers. I honestly believe the P-61 was not as fast as the Mosquito, which the British needed because by that time it was the one airplane that could get into Berlin and back without getting shot down. But come what may, the ‘61 was a good night fighter. In the combat game you’ve got to be pretty realistic about these things. (Harrison&Pape 209)”

The first real test of the P-61 in Europe came in July of 1944, when they were pressed into service against a new threat, the Fiesler 103 flying bomb or ‘buzz bomb’. The fast, unmanned weapon required the P-61 to enter a slight dive to catch them and, while they flew straight and level, they still proved a dangerous and challenging opponent. The bomb presented a small target but its massive warhead was capable of damaging a pursuer, something Capt. Tadas J. Spelis and F/O Eleutherios ‘Lefty’ Eleftherion would learn on the night of July 20th. Drawing in at 450 ft, Spelis detonated the bomb’s payload. which violently shook his plane and caused serious damage to the plane’s control surfaces and left much of the fuselage dented and perforated (Harrison&Pape 205).

Over the Channel: Autumn through Winter

P-61 crews grab lunch. Conditions for the crews of the 422nd and 425th grew basic as they operated from hastily prepared airstrips in the Autumn of 1944 [National Archives]
At the end of July, the 422nd and 425th would make the trip across the channel to provide afterhours protection for the US First and Third Armies, respectively. There, both squadrons would defend the Normandy beachhead as the Allies pushed forward into France. This period would largely inform the kind of fighting they would be doing for much of the campaign, intercepting lone German bombers and the occasional night fighter acting as an intruder, while also taking on alternate support missions. Shortly after the 422nd was deployed to the Cherbourg peninsula, they intercepted several Ju 88s, Do 217s, and Ju 188s as they attempted to harass Allied forces in the area, but kills were difficult to confirm owing to the contested areas these aircraft went down in.

This period also saw the P-61’s first encounter with a German night fighter when Lt. Paul Smith and Lt. Robert Tirney intercepted a Bf 110G-4 on August 7th, 1944. While Smith and Tirney approached the enemy, they were soon spotted and found themselves in a turn fight. While the maneuverability of the P-61 allowed them to keep up with the enemy, the two planes would end up colliding. Despite the impact, both planes would end up returning home, each carrying paint from their opponent. Records show elements of the German night fighter squadrons NJG 5 and 4 had been conducting ground attack operations that night without losses (Harrison&Pape 203; Part 4 Boiten 29).

This period also saw the first use of the P-61 in the ground attack role when the 425th NFS was called to assist an attack on German forces that had broken out of Lorient. Despite their early model P-61As lacking hard points for bombs, they were able to carry out the mission thanks to the powerful cannon armament of the P-61A. They conducted strafing runs on gun positions, truck convoys, and an artillery ammunition dump at the cost of one aircraft which struck a telephone pole in a low-level attack (Harrison&Pape 204).

Following the breakout in Normandy, there was a considerable lull in interceptions of enemy aircraft and the trickle of supplies to the unit meant much of the autumn of 1944 was characterized by inactivity. From September to November, GCI directed the 422nd’s P-61s into a total of 461 chases, resulting in 282 airborne radar contacts, 174 sightings, 20 of which were positively identified as enemy aircraft, and only 7 were shot down (McFarland 28). The use of Identify Friend or Foe (IFF) appeared to be limited, resulting in a high number of interceptions of friendly aircraft, and occasional friendly fire. Air crews in the 422nd NFS believed they had been fired on several times by RAF Mosquitos, and one Mosquito of the 305 Squadron, piloted by WO. Reg Everson, had been shot down by a P-61, with his aircraft being claimed as a ‘Ju 88’ (Harrison&Pape 302, peoples war).

As the Night Fighter Squadrons moved away from the beachhead and into airfields previously held by the enemy, their supply lines grew tighter and the enemy began to develop better tactics. A scarcity of fuel even threatened to keep the 442nd on the ground, but the crisis was avoided thanks to a little ingenuity. As fuel laden B-24’s came in for their deliveries in Florennes, Belgium, they would occasionally overrun the airstrip, whereupon the aviation gasoline would be siphoned out, and then stolen by the 422nd (Harrison&Pape 267).

A P-61B is prepared to sortie, Italy 1944. [National Archives]
While the Luftwaffe was less active at night during this period, their tactics had largely improved. Their typical after hours raiders became flights of bomb laden Fw 190s in the place of the lone medium bomber. The common types, the Ju 88, Do 217, Ju 87, and Ju 188, were still encountered, but were eclipsed by the more numerous 190s flying low altitude raids against Allied positions near the front line. The 190s would prove more difficult targets, as their small size made them hard to identify in the dark, and their speed and maneuverability meant they had a much better chance of slipping away from the larger night fighters. While they were harder to shoot down, the P-61 was still more than capable of breaking up their attacks and forcing them to return to base (Harrison&Pape 262).

With their superior speed and maneuverability, Fw 190 fighter bombers would prove a greater challenge for the P-61 than the usual medium bombers [Rod’s Warplanes]
The lull in Luftwaffe nightly activity in the autumn and winter of 1944 meant that both British and American night fighter squadrons could shift to offensive operations, and thanks to newer models of the P-61A and B mounting additional hardpoints for fuel and bombs, they would have an exceptional tool for this task. Both the 422nd and the 425th NFS would provide a vital service to the beleaguered 101st Airborne Division at Bastogne, Belgium, where they were able to provide air cover and ground attack support, day or night, in weather that kept most planes on the ground. The nightly air battles over the Ardennes took a similar, but intensified form as the Luftwaffe mounted a desperate offensive, sortieing aircraft to attack Allied positions, drop supplies, and mounted a score of night fighter intruder missions. These intruder missions had aircraft loiter around enemy airfields and attack any aircraft attempting to take off or land.

It was during this time that one of the greatest drawbacks of the P-61 made itself well known. It was a high maintenance aircraft and replacement parts and planes were scarce. During the Battle of the Bulge, only four of the 422nd NFS’s sixteen P-61s were operational, and keeping these four planes serviceable was a round the clock effort of the highest importance. Apart from the just as limited number of A-20s, the P-61s were the only aircraft capable of flying in the terrible weather conditions of the battle. Supplies had to be found outside of the regular channels, and crews were rotated out of these aircraft that each flew up to four missions per night. Combined, the 422nd and 425th NFS claimed a total of 115 trucks, 3 locomotives, 16 rail cars, sixteen aircraft, and had disrupted Luftwaffe activities in the area (McFarland 32, 33). The actions of the 422nd would go on to earn them another Distinguished Unit citation, and a commendation from the Commanding General of the 101st Airborne at Bastogne (Harrison&Pape 293).

However, this period was also considerably more dangerous as Luftwaffe’s night fighter squadrons were also performing similar missions in the same area. While they used comparatively obsolete radars, they could still present a threat. Of the scarce P-61’s active during the Battle of the Bulge, three were lost to unconfirmed causes (Dean 286).

Spring to VE-Day

The Battle of the Bulge would mark the apex of the NFS’ activity in the European Theater. The remainder of the European campaign would consist almost entirely of ground attack and intruder missions, as fuel shortages left most of the Luftwaffe grounded. Both the 422nd and 425th would commit themselves to ‘ground work’ against the usual targets; truck convoys and rail lines, as Tactical Air Command ordered a cessation of defensive air patrols, instead focusing on general offensive operations. In this role, the P-61 proved exceptional despite the design never being intended for such use, with the initial models not even possessing bomb racks.

Most ground attack missions would be conducted the same way, though some new tactics would be introduced to take advantage of the bomb racks added to the newer models of the P-61. During the beginning of 1945, P-61s would often carry napalm which they would drop in the vicinity of their targets. Using the fires for illumination, they carried out attacks with a combination of bombs, and in the case of the 425th’s modified P-61s, HVAR rockets. Rail yards, locomotives, and truck convoys were favored targets, as their drivers often felt it was safe to keep their lights on. While this may seem a ridiculous use of what were among the most expensive aircraft employed during the war, the 422nd was credited with damaging or destroying 448 trucks, 50 locomotives, and 476 rail cars for the duration of their service. Perhaps more impressive were the astoundingly low loss rates suffered on these intruder missions. From October of 1944 to May of 1945, the 425th NFS flew 1,162 intruder missions with the loss of only six aircraft. Despite the inherent dangers of flying at night, these missions actually proved to be far safer than daylight sorties (McFarland 29).

VE-Day [National Archives]

Luftwaffe Opponents

The typical encounters with the Luftwaffe were with its bomber, night attack, reconnaissance, transport, and occasionally night fighter forces. Their targets ranged from frontline positions, rolling stock, to airfields, and were typically attacked by lone aircraft or small formations of light attack aircraft, such as the Fw 190F&G or obsolescent Ju 87. P-61 crews would encounter virtually all medium bomber types in service with the Luftwaffe, including the dated He 111 and Ju 88A-4. Of all the aircraft encountered, only the Me 410 proved a serious challenge to intercept. They were employed as reconnaissance aircraft and their high speed meant they were only vulnerable to the P-61 when at a disadvantage. On roughly equal footing, the Me 410 could pull away (Harrison & Pape 275).

Encounters with enemy night fighters were fairly rare, as their squadrons generally only flew ground attack missions in August against the Normandy beachhead, and much later in December, in support of the Ardennes counteroffensive. They flew Bf 110G-4s, a few of the older Ju 88Cs, and the newest German night fighter at the time, the Ju 88G. While these aircraft flew with radar that had a much more limited range than the SCR 720 and were nearly useless at low altitude, their pilots were capable of putting up a much greater fight than those of the bomber and night attack squadrons. The first encounter between a P-61 and a Bf 110G-4 resulted in the latter being able to slip away after a collision, despite holding clear disadvantages in speed and maneuverability. While most of the new pilots the Luftwaffe were supplied with at the time possessed questionable proficiency at their tasks, most green crews remained on the ground as a result of chronic fuel shortages (Part 4,Boiten 33).

On the night of December 17th, several Luftwaffe night fighter squadrons would be committed to large scale ground attack operations in support of Von Rundstedt’s offensive. These missions were conducted by several dozen aircraft at a time that searched highways, rail lines, and known Allied positions for targets. These operations achieved a level of mixed success but at an extremely high cost, as the pilots were insufficiently trained for the mission and typically encountered accurately directed anti-aircraft fire (1944 Part 5, Boiten 68). The P-61s of the 422nd and 425th would find these night fighters significantly more challenging opponents than the medium bombers and transport aircraft they usually encountered. On several occasions, the German fighters slipped away from their pursuers and claimed two, later disproven, victories against P-61s, with the war diarist of Stab NJG6 commenting the “Black Widow inferior to Ju 88 and Bf 110 in dog fighting (Part 5, Boitens 77).” However, this confidence is likely due more to survivor bias than any major technical difference between these aircraft, as several German night fighters would be lost to P-61s. In all likelihood, it was the German night fighter pilot’s confidence in undertaking aggressive maneuvers in the dark that was the most probable reason for this assessment, as the P-61 was superbly maneuverable for its size.

Over the course of this offensive, the 425th would encounter a number of German night fighters and down several of them. Between the 25th and the 29th of December, three confirmed and two probable German night fighter losses can be attributed to this squadron’s P-61s, these being Bf 110 2Z+DH of NJG 6, Ju 88G-1s of NJG4 3C+RK and 3C+ZK, and two very likely Bf 110s, 2Z+DL and 2Z+CV (1944 part 5, Boitens 79, 84, 85). Given the short period and how few P-61s were serviceable, it is safe to say that the P-61 was certainly capable of taking on these opponents. However, it should also be noted that night fighters comprised a relatively small number of kills during this time, with many more being medium bombers and Ju 52 transport aircraft.

The 422nd NFS would later rebase in Langensalza, Germany. This posting allowed P-61 crews the opportunity to get a close look at the enemy’s standard night fighter, the Ju 88G. [National Archives]

CBI: The 426th and 427th

The 426th Night Fighter Squadron was called upon by General Henry H. Arnold for the defense of the B-29s based in Chengdu, China, as per the request of the Maj. Gen. Curtis LeMay (Harrison&Pape 222). They would also be joined by the 427th NFS following the end of Operation FRANTIC and the cancelation of any further deployments of USAAF bombers in the Soviet Union. While they were sent to defend the B-29 bases, they were soon found to be almost totally unnecessary, as there were little to no Japanese aircraft active after dark in the China-Burma-India Theater. Not long after their arrival in October, 1944, they would pivot almost entirely to an offensive role, and were mostly relieved of their defensive task to search for trains and truck convoys across the theater. Several aircraft were later modified to mount 4.5 inch rockets, as was the case for their counterparts in Europe (Harrison&Pape 215). They would be met with success, as the Japanese Army was reliant on a single network of roads that ran north to south, a single major rail line, and the Irrawaddy River to move men and material across the theater. Despite the massive patrol area, they could expect to find targets at these strategic bottlenecks (McFarland 40).

A pair of P-61s look for targets by daylight over Northern Burma. Note that the tail cone has been reinforced. [National Archives]
The challenges of operating in this theater largely mirrored those of the squadrons based in the Pacific, as supply lines were tight, the terrain proved difficult to construct airbases in, and the mountainous geography hampered the use of early warning radar. Fuel was particularly scarce and had to be shared with the B-29 squadrons based with them, which typically meant offensive operations were periodically called off when supplies of fuel ran out, as was a case for the 427th NFS’ detachment in China for the month of April 1945 (Harrison&Pape 236). The squadrons operated mostly dispersed across the theater as detachments, with the peak number of P-61s in the CBI being 53 in July of 1945. The number typically sat around 35 aircraft until June (Dean 386).

PTO: the 421st NFS

Many airfields across the PTO were built over coral beds which were difficult to clear and highly visible at night [National Archives]
The P-61 was a godsend to the Pacific night fighter squadrons who had long been forced to rely on the inadequate P-70, and with the exception of a few field modified aircraft, radar-less P-38s. Starting from early 1944, the various night fighter squadrons in the PTO would begin receiving P-61s and phasing out their long obsolete P-70s. Unlike Europe or the Mediterranean, the operations in the Pacific would not proceed at the pace of a gradual frontline that needed to be supported but rather saw the NFS deployed to newly constructed airstrips in support of larger amphibious operations which were targeted by raiders. Conditions were poor and extremely hard on airmen and planes alike, which brought unique challenges unknown to those in the ETO. In the words of S/Sgt. Harold Cobb of the 421st NFS: “Night fighting is not glamorous, but it is specialized in every degree, especially in the seven-league-boots, island-hopping war in the Pacific. Pilots must be able to take off and land without strip lights and on fields which are so new that construction is still in progress and the Seabees are still working (Kolln 51).”

The 421st got its hands on the P-61 in April of 1944, while it was based in Wakde, New Guinea, the planes having originally been shipped to Brisbane, Australia. The impact of receiving the new planes would prove considerable, both boosting the morale of the unit and giving it a long-needed replacement for its P-70s. The overall mission of the 421st was largely the same as it was for the 6th Night Fighter squadron in Guadalcanal two years earlier, to defend against nightly nuisance raids from Japanese bombers. The Japanese had also largely improved from their earlier campaigns, as they began to seek targets of greater strategic importance which they attacked with a far greater frequency. The men of the 421st were among them and their bases at Wakde and Owi were attacked regularly, often causing casualties and destroying aircraft. These airfields being built on lightly colored ancient coral beds made them both extremely visible at night and made it extremely difficult to dig shelters (Kolln 48). The effects of high explosives were also magnified, as they propelled sharp fragments of coral through the air with every bombardment.

The squadron also faced the same challenges posed by tropical environments, often with little improvement over the conditions almost two years ago. The prepared airfields were often built under difficult circumstances and challenging geography. The Seabees often had to work with coral beds, wetlands, and jungles that proved time consuming to develop into usable airstrips, often leaving little time and resources for improving the living conditions at these airfields.

Wakde Island before it had been captured by the US. The bright coral beds that were the foundations for the airstrip made an excellent aiming point for Japanese raiders [National Archives]
These conditions were also felt by the sensitive radar systems of the aircraft, especially the pressurized canisters which contained many of the system’s vital components. They had a tendency to depressurize, which resulted in electrical arcing at altitude, disabling the entire system. The 419th NFS had developed an improvised system where the electronics tanks were kept pressurized by engine-driven vacuum pumps, but it is unknown if this modification was ever taken up by any other squadrons (Harrison&Pape 149). Early models of the P-61A, which still had the plastic radome, also encountered trouble in the tropical heat and sun, as the nose of the aircraft would often soften and deform, which would impact the movement of the SCR 720’s scanner. In addition to the reflective white paint added to the nose, crews would fasten ‘sun shields’ while grounded to protect the radome in the tropical heat. As was the case with the European squadrons, supplies of replacement parts and aircraft were scarce, and in a unique twist in the Pacific, the improper packing of engines resulted in the loss of 400 R-2800s to corrosion. Combat readiness suffered as a result. The 421st considered it a ‘good day’ should six of their aircraft be operational during their operations from their later airbase at Tacloban (Harrison&Pape 241).

Conditions for the air and ground crews of the 421st were scarcely better. Harsh tropical weather, limited access to drinkable water, and disease were common in the South Pacific, with conditions only improving after redeployment to the Philippines. At Owi and Wakde, personnel had to overcome an outbreak of Typhus which claimed two, heatstroke which claimed one, and even Silicosis of the lungs which resulted in a single fatality. However, the base at Owi became perhaps the most livable thanks to the discovery of an artesian well near the unit’s bivouac area (Kolln 47,48).

In the PTO, P-61s often operated from small, busy airfields where accidents were not uncommon. This P-61 went off course on a blind landing through fog at Iwo Jima and crashed into another aircraft [National Archives]
Operations over New Guinea largely proceeded the same way as they had earlier, but with far greater success thanks to their new P-61s, which meant interceptions were comparatively easy, though new Japanese tactics would periodically disrupt their success. Perhaps the most surprising of these was the deployment of radar reflecting chaff from bombers as they made their way to and from their targets. The chaff were aluminum strips that reflected radar and presented on radar scopes as a single large ‘cloud’ that could obscure the positions of aircraft. In practice, the SCR 720 did not prove very vulnerable to chaff if the pilot had already been guided toward the target, as the air search radar proved powerful enough to burn through the interference. Far off ground-based radar stations would prove more vulnerable to it, especially older models (Kolln 55). The Japanese air force would also employ new tactics against the defenders. On August 5th, the Japanese sortied several fighters along with the typical high-altitude bombers. These aircraft were picked up on radar later than the bombers to which the night fighters had already been sent against. The Japanese fighters were, however, unable to inflict much damage and their tactics were soon understood by the defenders (Kolln 50). While the 421st NFS’ P-61s were largely part of the waning war in the South Pacific, their subsequent redeployment to the Philippines would place them in one of their most active theaters of the entire war.

Tacloban

The airfield at Tacloban was muddy, overcrowded, and under constant attack during the first weeks of the US Army’s return to the Philippines [National Archives]
The 421st was deployed to the airfield at Tacloban on October 31st, 1944 to provide nightly air cover for the amphibious operations in the Southern Philippines. The conditions were largely a repeat of those of the prior camps at Wakde and Owi, but the raids were far worse. What were once frequent occurrences became a daily fixture of the stay at Tacloban (Kolln 61). The airfield itself would also prove to be an extremely hazardous and ineffectual location to operate from, as would be the positions chosen for the GCI radars. Tacloban was extremely underdeveloped during the height of operations. It was without runway lights for fear they would attract Japanese bombers and the short, muddy airstrip was difficult for the heavy P-61 to operate from, with most landing attempts having to be repeated (Harrison&Pape 240). From their airfield, they were given a number of tasks which would include providing nightly air cover to invasion forces, escorting PT boats, convoy protection, and even conducting daylight patrols.

The Philippines would present a greater set of challenges to the 421st than New Guinea. For one, the aircraft and tactics used by the Japanese air forces were of a different nature entirely. While they previously worked mostly against occasional, small formations of bombers coming in at high altitudes, they now also fought against large numbers of fighters which flew continuous attacks, typically conducted at low altitudes. Coupled with poor GCI coverage of the area, the Ki-43s and A6Ms employed by the Japanese would prove an extremely difficult enemy to counter. The nightly attacks continued and many of the invasion planners became frustrated with the squadron’s inability to stop them entirely, eventually leading General Kenny to send much of the 421st to Peleliu, while the 541st Marine Air Squadron, equipped with the 565-5N, took their place at Tacloban. In the end, the 421st achieved seven kills during this time at Tacloban, and while this was a fairly significant level of success for the PTO, it was deemed unacceptable by the invasion planners (McFarland 37).

The reasons for this swap have long been debated, with claims ranging from the P-61 having insufficient range and loiter time, to the SCR-720 being unable to track more the more maneuverable Japanese fighters. In the end, however, the greatest problems faced by the squadron were its poor GCI support, its low number of serviceable aircraft which resulted from supply shortages, and the vulnerable, poorly suited airfield they had at Tacloban. The Marine night fighters who replaced them were credited with 23 kills, though most of these were during dawn or dusk missions. It does not appear that any technical failings of the P-61 were responsible for a perceived lack of performance, but rather, exceedingly poor operational conditions and biases held by the higher headquarters that placed the expectations of daylight fighters on the NFS 421st while not understanding how they would best be deployed or utilized (Kolln 72). Though most of the squadron had departed Tacloban, several planes and their associated personnel remained to ensure a smooth transition for the Marine aviators and to carry out their previous duties, though to a lesser degree.

The 421st would return at full strength to Tacloban in early January of 1945, after five weeks, and largely resumed the work they had been doing before they left. Most notably, this included the joint patrols with motor torpedo boats, especially the 7th PT boat squadron, which they had developed a good working relationship with. The P-61s would provide cover for the boats as they patrolled Surigao strait and the Ormac Bay area, with a squadron representative aboard to ensure smooth operation between the boats and their air cover (Kolln 75). Enemy air activity in the area had decreased significantly and once again took the form of periodic raids by bombers flying alone or in small formations.

The airfield the 421st returned to was vastly different from the one they had left just a few weeks prior. [National Archives]
The squadron would end the war at Ie Shima, Okinawa, in July of 1945. By this point, the Japanese armed forces were in a state of exhaustion but they were still capable of launching nuisance raids against front line positions and airfields, though the frequency of these raids was low and there were two other P-61 squadrons stationed in the area. This would also mark the beginning of the replacement period of the squadron’s P-61s for P-38Ms. The 421st would spend this time performing intruder missions against targets in Kyushu, Japan, with airfields tending to be the primary targets. In this role, they developed a bombing technique with their search radars, which would be used to measure the relative distance to the target, and in conjunction with the airspeed and altitude of the aircraft, a bomb release window could be worked out. Some pilots would even add marks on their windscreens as visual aids for the technique (Kolln 89). They were, however, unable to account for its effectiveness. There was little resistance to these raids as Japan had a comparatively underdeveloped night fighter service and their night fire control for their anti-aircraft artillery was little better.

With Japan facing famine and industrial breakdowns from the blockade, the prospect of a third atom bomb with more to follow, and their last hope for conditional surrender evaporating as the Soviet Union overran their mainland colonies, the war ended and the P-61’s wartime service came to an end.

Japanese Opponents

Despite being significantly less experienced with the use of ground based and airborne radars than the Germans, Japanese aviators and mission planners consistently demonstrated the ability to develop effective countermeasures and tactics to American night fighters. Japanese signals intelligence services would prove extremely effective and were able to determine the presence of enemy night fighters in areas without radar coverage by monitoring radio transmissions, and were even able to track the position of P-61s by their IFFs (Harrison&Pape 220, 319). They would also successfully employ chaff on a number of occasions, though to decreasing effect, as the US Army began to employ more advanced centimetric search radars that were less vulnerable to it. On Iwo Jima, for instance, raiders would typically use chaff roughly thirty miles out from the island and when they departed, which had the effect of blocking the older meter band SCR-270 and reducing the range of the centimetric SCR-527 (McFarland 39). In addition to this, they would also employ seaplanes to get the attention of night fighters, and once they had drawn them away from the raider’s target, they would land on the water’s surface or return home at low altitude (Thompson 71). This tactic appeared to have been used against the P-61s of the 421st while they were at Tacloban and to good effect, as the loiter time of the P-61 was rather low and they were often forced to return home after several of these non-encounters (Harrison&Pape 234).

Variants of the Mitsubishi G4M were among the most common raiders encountered by P-61s in the Pacific [Rod’s Warbirds]
In the Pacific, P-61s faced mostly medium and light bombers, though would face considerably more fighter aircraft as the war drew to a close. These aircraft employed a wider variety of tactics than those of the Luftwaffe, often to considerable success. However, they would still employ earlier tactics such as lone bomber, high altitude raids which were far less effective, as the P-61 did not have the difficulty the P-70 had in reaching high altitudes.

What they wanted but never got: The P-61C

The P-61C would be developed largely to fulfill the requests of most of the pilots who had flown P-61As and Bs. The design sought to add two major features, more powerful turbocharged engines to provide better high-altitude performance and a higher climb rate, and a set of air brakes. The air brakes would be designed by the AAF’s Wright Field staff in conjunction with Northrop. The design was first incorporated on a P-61A test aircraft, which was nearly lost after a portion of the air brake was sheared off the aircraft and nearly sent it out of control. The final design proved satisfactory and took the form of a two-part slotted panel with halves above and below either wing. These brakes also incorporated a novel system to reduce the asymmetric forces acting on the brakes. This worked by having the deployment of the lower set of brakes assisted by the raising of the top. The brake system exerted a counter force of roughly 1G when the aircraft was at high speed (Harrison&Pape 278, 281).

The engines would go through a considerably longer development period and were to be mounted on a new airframe. Initially, there was some debate on whether the engine should use either a two stage two speed supercharger, as the previous production models of the P-61 used, or a turbo supercharger. It was a new study under John M. Wild at Northrop that made the case for choosing the turbo charger, with his finding being agreed on by Wright Field’s Fighter Project Office. A CH5 turbo-supercharger was subsequently fitted to a P-61A for testing, the aircraft being redesignated the XP-61C. The XP-61C’s conversion was handled by Goodyear Aircraft out of Akron, Ohio, a firm that provided parts for Northrop. The aircraft was initially to be powered by the R-2800-77, though a production run could not be secured and a temporary installation of the R-2800-14’s were used in their place until the R-2800-73 was chosen for the production model. A parallel development that would later be designated the XP-61D made use of the R-2800-77. Cooling issues would bring an end to its development, with the P-61D being canceled as the P-61C entered production. The P-61C proved to be quite promising and a massive step above the previous models, with the aircraft’s service ceiling being raised to 41,000 ft (12497 m) and its maximum speed rising to 430 mph (692 km/h) (Harrison&Pape 279, 280). The P-61C would be the aircraft the test pilots had wanted from the outset, but would fail to make it into service fast enough to see combat.

Project Thunderstorm

Several Project Thunderstorm P-61Cs and an F-15A [NOAA]
While the P-61C arrived too late to take part in the Second World War, it would go on to make major contributions to meteorological research and aeronautical safety in the post-war Thunderstorm Project. The project began with the passing of the H.R. 164 bill in January of 1945, which authorized and directed the Weather Bureau to conduct a study on the causes and characteristics of thunderstorms for aviation safety. The bill would also authorize the appropriations needed for such a study and authorized the cooperation of other departments for assistance.

The finalized research plan called for a vertical stack of five aircraft to make passes through thunderstorms as they drifted over a network of meteorological recording stations in order to document the conditions within the storm. The Army Air Force would provide several P-61Cs and its derivative, the F-15A, for this purpose, as they were designed to withstand strong maneuvering loads and were judged strong enough to quite literally ‘weather the storm’ (Roscoe 26). These aircraft would be modified for the purpose, with wartime equipment being removed and meteorological research equipment installed in its place. The aircraft were prepared at NACA’s Langley Field with the equipment necessary to monitor turbulence and vertical air currents.

The Flight Plan [Roscoe]
For the tests, the planes entered thunderstorms at altitude differences of five thousand feet, with the highest aircraft being at 25,000 ft (7620 m). No storms were avoided, no matter how violent. The project first began around Orlando, Florida in 1946, before later moving to Wilmington, Ohio the following year. These locations were chosen on the basis of the frequency of thunderstorms and the nearby Air Force installations which had the radars needed to support the project. The project would see the P-61s fly through 76 storms for 1362 fly-throughs, during which they collected vital data which would help pave the way for safer air travel during the post war civil aviation boom and were used to build a foundational study for thunderstorm research (Roscoe).

The hail damaged radome of a P-61C [NOAA]

Pilot’s Remarks

Lt. Herman E. Ernst, an ace of the 422nd NFS, behind the controls of a P-61A. Note that the pilot’s radar indicator has been replaced by a compass below the gunsight [National Archives]
Exhaustive tests were performed on the P-61 to determine its flight characteristics, and they were largely found to be in line with the earlier prototypes. Pilots were highly appreciative of its easy handling on takeoff and landing, along with its favorable stall characteristics. Its controls were effective up until stall condition, which itself only occurred after ample warning. Stalls themselves were relatively predictable and virtually always resulted in the nose dropping, with no tendency for either wing to drop, and no corrections being needed to prevent a roll (Dean 391).

In addition to its great stall characteristics, the P-61 would prove to be exceptionally maneuverable for a plane of its size and weight. However, given its size and with two heavy engines on the wings, its roll rate was rather poor. Tests found the aircraft could be put through all hard maneuvers save for outside loops, continuous inverted flight, spins, snap rolls, and vertical reversements. Pilot’s praise was given mostly for its extremely light controls even at high speeds, which was largely thanks to its unorthodox spoileron based lateral control system. Even at speeds of 400 mph IAS, fast aileron and spoileron movement could be affected with one finger on the control wheel, though controls were found to be ‘sloppier’ around 100 mph IAS. One pilot was so confident in the P-61’s maneuverability that he felt he could turn with the best of fighters and, in the case of the F6F, he would “be on his tail so fast it was incredible.” In addition to its maneuverability, its trimming characteristics were also very good, such that it was possible to trim the aircraft out for cruising on a single engine at 130 mph IAS. It should, however, be noted that despite this praise, pilots rated its maneuverability fair to poor, as it was compared to far lighter single engine fighters. Overall, the plane was rated very stable on all axes with good rudder and elevator effectiveness (Dean 388, 390).

Despite its size, the P-61 was capable of pulling off some very impressive aerobatics. However, regulations stipulated that acrobatics were restricted under most combat conditions. [Pilot’s manual]
The P-61 also boasted excellent dive and recovery characteristics, with the book limits being set around 430 mph IAS depending on the arrangement of the aircraft, or around Mach .70. Beyond these limits, buffeting and tuck-under would occur, but the aircraft also demonstrated the ability to exceed these limits by a fair margin. One pilot would claim that he had no problems at speeds of 450 to 475 mph IAS at around 10,000 to 15,000 ft. In this case, he had achieved a true airspeed of 599 mph at Mach .83 and returned within the specified limit envelope at 512 mph TAS, Mach .70, as he reached 10,000 ft. Typically, if buffeting did occur, it was advised to exit the dive by means of a gradual pullout and with high caution should external loads be carried. Another pilot would claim that buffeting would occur far in excess of the recommended dive speed limits. The P-61 would be rated good in respects to its dive acceleration, control forces, recovery characteristics, and be ranked 8 out of 11 US fighter types in the category best stability and control in a dive (Dean 389, 390). The engine limits within dives were 30 seconds at 3090 rpm.

The P-61 was, however, not without its faults, and the most criticized and frequently voiced issues were concerning its acceleration and climb rate. These sentiments were echoed in many tests, and were most notable in rating the aircraft for takeoff, where pilots soon found themselves climbing at a disappointing rate after reducing power. Even at combat-power, the aircraft could at most manage 2500 ft/m at an altitude of 5000 ft (Dean 381). In the concluding remarks to exhaustive tests, it was the most frequently voiced complaint. While it was true many test pilots judged the aircraft somewhat unfairly against lighter single engine fighters, even its most enthusiastic testimonies were typically accompanied with remarks regarding its acceleration and climb performance (Dean 393).

The cockpit drew mixed reactions, with the general feeling being that the layout was adequate but not ideal. While most felt the layout was fair, eight of twenty-one pilots felt the arrangement was “cluttered”, with another ten remarking that they felt the pilot was seated too far from the instrument panel. This group was so displeased with the arrangement that they ended up rating P-61’s cockpit 3rd in the category “worst cockpit”. While the layout of the cockpit remained divisive, virtually all of them were displeased by the restricted visibility caused by the canopy frame (Dean 392).

The P-61B and several As would make use of the night binocular gunsight. These slid behind the pilot when not in use [P-61 Pilot’s training manual]
In terms of its weaponry and its stability as a firing platform, the P-61 was well rated. Equipped with four 20 mm AN/M2 cannons and up to four .50 cal AN/M2 machine guns, the P-61 was very well armed. Despite the lack of the turret on most P-61s, the aircraft’s armament was more or less equal to its contemporaries, the Mosquito NF and Ju 88G, which both carried an armament of four 20 mm cannons. Firing stability was also good, with only one out of fourteen test pilots finding it objectionable. However, problems with the turret would impact its usefulness, as early aircraft would experience intense buffeting on the tail surfaces when the guns were set to certain positions. This problem would later be solved and the turret reintroduced to the aircraft when it was redesigned and supply bottlenecks with the B-29 were resolved (Dean 393).

Overall, the P-61 would present an aircraft with mixed, but favorable characteristics. The aircraft would be superbly maneuverable and responsive for its size and presented excellent flight characteristics at high and low speeds. In contrast, pilots were not enthused over what they judged was a poor rate of climb and cockpit layout. The 481st Night Fighter Training Group would also go on to lodge complaints about poor cockpit visibility and short combat radius (Harrison&Pape 156). While the cockpit went unchanged, the relatively limited range, of only about 1000 miles, would be later brought above 1,800 miles with the use of external fuel tanks (Dean 382).

Construction of the P-61A and B

The wings of the P-61, except for the tips, used a fully cantilever riveted, stressed skin construction with two main spars. Each wing assembly was composed of an inner panel, an outer panel, and the wingtip. The inner panel contained the engine nacelle, two fuel tanks, and a section of the flaps. This portion was the largest wing section and was fitted to the crew nacelle by means of bathtub and lug type fittings at the end of the main spars. The front section also provided one of the main air intakes for the aircraft and an outboard for the oil tank. It was constructed in two parts, forward and aft sections.

A P-61 undergoing maintenance. Its outer wing panel has been detached from the nacelle group [National Archives]
The aft section was also built in two parts and it contained the wing flaps, spoilers, and ailerons. The section also contained an oil cooler and its associated exit shutter. There were six hydraulically actuated slotted flaps with a full deflection of 60 degrees. Relatively small ailerons were installed outboard of the flaps, which extended to the wing tips. To boost lateral control along with the ailerons, a series of spoilers were used and were found in trailing sections, ahead of the outer wing flaps. These were curved metal panels that extended from slots in the wings and were mechanically driven by the pilot’s control column along with the ailerons. Initially, there were also aileron and booster tabs fitted to the inboard end of the left ailerons, but these were removed on later models. The combined fuel capacity of the wing fuel tanks was 646 gallons (2445 liters).

The center fuselage was a semi-monocoque structure composed of transverse bulkheads and channel section frames, longerons at the upper and lower quarters, longitudinal bulb angle stringers, and stressed skin. It was attached to the wings by means of heavy forged fittings on both sides of the fuselage. This section contained the stations for the aircraft’s pilot, gunner, radio operator, the aircraft’s SCR-720 radar, fixed quadruple 20 mm armament, and the mechanically operated turret. The enclosures for the pilot and gunner positions were made from molded Lucite sheets and extruded metal framing, with forward sections protected by bullet resistant glass blocks. The radar nose cone was made of plexiglass on early models, before a switch to a less heat sensitive resin-impregnated fiberglass on later aircraft. The radio operator’s position was enclosed by a framework of Lucite in extruded metal frames, with a rear tail cone that was formed from two sheets of Lucite that had been cemented together and bolted to the rest of the framework. All positions had a seat with a metal pan, padded backs, safety belts, relief tubes, and hand fire extinguishers. The center fuselage was also fitted with armor plates to protect the crew and ammunition boxes. These were located behind the nose, ahead of the gunner, in front of the turret ammo boxes, and behind the radio operator. The standard crew layout on this aircraft was poor compared to contemporary night fighters. A failure of the intercom system left the aircraft combat ineffective, as each crewmember was isolated, the radar operator particularly so.

The P-61’s fuel system [Pilot’s notes]
The tail booms were of a monocoque structure and connected the nacelle group to the tail group. They also housed components for communication, identification equipment, the flight control cables to the rudders, elevators, and tabs. They were connected to the nacelle groups, which were composed of a semi-monocoque structure. These carried the engine mounts, main landing gear, and fuel tanks. The engines were mounted from a built-up welded steel tube frame that was bolted to this nacelle through vibration isolators and the engine cowling panels. The cowling sections were removable in large sections and were attached to the engine by quickly-detachable fasteners to facilitate easier access to the engines. The adjustable flap segments were controlled from the cockpit and were hydro-mechanically actuated.

The tail section consisted of the horizontal stabilizer, the elevator, and two vertical stabilizers and was connected to the tail booms. The two tail sections were supported by two spanwise spars that ran through the horizontal stabilizer and had the vertical stabilizers at either end. The rudder and elevators were fabric skinned and had trim and booster tabs built into their trailing edges.

The aircraft had a tricycle landing gear arrangement, with its nose wheel housed in the center fuselage and the main gear in the nacelle group. Each main gear was supported by two steel castings which were bolted to either side of the inside of the nacelle. Landing gear loads would be handled by a shock strut which was connected to these castings by a pair of trunnions. When the gear was retracted, it was hinged on these trunnions at the castings by lockbolts which would be held in either the extended or retracted position by a mechanical latching mechanism.

Engines

The mid production P-61A and the P-61B were powered by a pair of 2000 hp class of the R-2800 Double Wasp engines. This was an air cooled, two row, eighteen-cylinder radial engine with a 5.57 inch bore and a 6-inch stroke. These engines had a maximum RPM of 2700 and a compression ratio of 6.7:1. The early A models used the R-2800-10, with the remainder of the series and the B models using the R-2800-65, both of which produced a maximum output of 2000 hp. The later R-2800-65W boosted this to 2250 using water injection, and the C used the R-2800-73 which produced 2800 hp. The A and B models were equipped with two-stage, two-speed superchargers, but the C used turbosuperchargers.

T/Sgt. M. Stetson at work on a 9th AF P-61, 1944. [National Archives]
The exhaust system was a stainless-steel arrangement with the exhaust stacks distributed around the edges of the nacelle, making use of a flame dampening system used to reduce the visibility of the exhaust at night. The B model and the late A series aircraft were equipped with a water injection system. The first aircraft with this system carried 26 gallons (98 liters), good for 15 minutes, with later aircraft carrying 34 gallons (128 liters), which was enough for 20 minutes of use, though some aircraft would carry as much as 74 gallons (280 liters). Use of this system could boost engine power by up to 250 hp per engine, though only in short increments, with the suggested limit being five minutes at a time (Pilot’s manual 12). Engines with water injection were designated R-2800-65W.

The engines drove a pair of Curtiss Electric four blade constant speed, selective pitch, full-feathering propellers. The hubs were a pair of C642S with a set of 12-foot 2-inch diameter, 714-7C2-12 blades. Engine speeds between 1800 and 2300 RPM were restricted as a result of propeller vibration in that range. The props were capped by large metal spinners which enclosed the hubs and inboard prop sections.

Avionics

In addition to the SCR-720 search radar, the P-61 carried a well-developed electronics suite. This included an SCR 729 radio navigation system, an SCR 695 IFF, and an RC-36 intercom system. The P-61A and early B models were equipped with the SCR 718 radio altimeter, which was later replaced with the AN/AP1. Early models used a pair of SCR-522 radio sets which was simplified in later models by a single AN/ARC 3. In later aircraft, an AN/APS-13 tail warning radar set was also included. The navigation systems were also supplemented by a MN-26C radio compass with a MC-1206A range receiver.

The P-61’s SCR-720 air search radar was composed of six main units which were installed in a number of boxes throughout the aircraft. These were the modulator, the transmitter, the receiver, an indicator unit, the mixer, and the power supply unit. The modulator was a rotary spark gap, pressured type which produced a 4 kV pulse. The transmitter was magnetron regulated and installed in a pressurized unit. The mixer was a crystal mixer type with a soft rhumbatron switch valve. The receiver used a reflector klystron oscillator with automatic frequency control. The indicating unit used a two-tube range and azimuth elevation display set. The entire system was powered by a 1,200 watt, 115 volt, 1,600 cycle engine driven alternator (Survey 27).

Heaters

In addition to navigation, communication, and detection equipment, there were also considerable heating, cooling, and ventilation systems. On the P-61A, a series of fuel-air mixture heaters were used to provide heating for the cannons and to the crew through three ventilators. The B model decreased the number of heaters from four to just two heaters that were placed fore and aft.

Armament

Gunnery Equipment and Armament Cutaway Diagram [Pilot’s notes]
The standard armament of most P-61s was a set of four fixed 20 mm Hispano AN/M2 cannons that were set in a compartment at the bottom of the central fuselage. 200 rounds of ammunition could be carried for each gun. Sighting for the gun consisted of the L-1 type gunsight on the P-61A and the LY-3N on the P-61B, with both being a reflector type lit by a sight lamp. From the B model onward, the aircraft would also carry a set of night binoculars which were a specialized gunsight for use in low light conditions.

While this aircraft is often known for its remotely controlled, quadruple .50 caliber turret, only about half of P-61s actually carried one. The turret’s machine guns were each supplied with 560 rounds, were fired simultaneously at a rate of about 800 rounds per minute, had a 360-degree traverse, and a maximum elevation of 90 degree upward from the horizontal. While the guns could be fired from any of the aircraft’s three positions, only the gunner and radio operator could direct the turret. For the pilot’s use, the guns would be locked forward by latching the turret and flipping the switch labeled “pilot” from either of the other two positions, though in the B model, the turret would automatically return to the guns forward position when not in use. The .50 caliber guns were typically fitted with flash concealing tubes in the field after pilots found it could interfere with their vision adjusted for low visibility flight. In service, the turret was almost always used by the pilot and very sparingly by the gunner against targets ahead of the aircraft. Pilots often found it unnerving to see the turret firing forward without warning from the gunner, as it could both ruin their low light vision and sometimes they misidentified the gunfire as coming from an enemy behind their aircraft. As a defensive armament, it was of little practical use, as the radar operator’s illuminated instruments screens degraded his low light vision. The turret was directed by a sighting arm which sat atop a rotating column with firing controls in the grips and fitted with an N-6 reflector sight.

The aiming system for the remotely operated turret [P-61 Pilot flight operating instructions]
Several models included wing racks which were capable of carrying additional fuel tanks or bombs with a maximum weight of 1600 pounds (725.75 kg). Field modifications on some aircraft allowed for the use of rockets. (Dean 393-404)

Several P-61s of the 425th NFS were modified to carry rockets, in this case, HVARs. [Wikimedia]

Conclusion 

The P-61 was a somewhat troublesome, yet effective night fighter that proved to be a capable replacement for the useless P-70 and obsolescent Bristol Beaufighter. Most of its faults, apart from the poor layout of crew, were to be expected for such a sophisticated plane still in its ‘teething period’ and supported by a modest supply chain. In the space of roughly a year, which constituted its entire combat service, most of its faults were corrected or lessened.

The aircraft served admirably across the European, Mediterranean, China-Burma-India, and Pacific theaters. P-61 pilots would encounter a variety of opponents among the Japanese and German air forces, utilizing a variety of tactics and equipment. They would prove effective against all but a handful of these combinations. Surprisingly, despite never being designed with such a use in mind, the P-61 would prove exceptional in the ground attack role. It was among the few aircraft at the time capable of carrying out attacks at night, or in poor weather. In service its greatest danger was its limited material support. This scarcity of replacement aircraft and parts would hobble operations, but the resourcefulness of ground crews often kept their squadrons from being entirely grounded. In the end, the aircraft provided effective service during their somewhat short combat tour across much of the world, in the face of inadequate material support and, at times, extremely poor conditions.

While the P-61C would never see combat, it would perform a vital role in a foundational meteorological study. Despite never being used for its intended purpose, this variant’s legacy proved to be no less important.

Specifications and Production Numbers

Type Number Built First Delivery Description
XP-61 2 May-42 First prototype series
YP-61/P-61 13 Aug-43 Second prototype series, pre production
P-61A-1 45 Oct-43 Power turrets installed in first 37 planes, first production model
P-61A-5 35 Turret removed, R-2800-10 engine changed to R-2800-65
P-61A-10 100 Water injection system added
P-61A-11 20 Two underwing racks
P-61B-1,2,5,6,11 155 Jul-44 Extended nose, wing racks on 2,6, and 11
P-61B-10 45 Four underwing racks
P-61B-15,16,20,25 250 Turret revised and reintroduced with two and four gun versions, wing racks (two on the -16), radar gun laying on -25 with seven built
P-61C-1,5,10 41 Jul-45 Turbosupercharged R-2800-73 engines, air brakes
XP-61D 2 converted airframes Nov-44 Prototype, turbosupercharged R-2800-77’s, wing racks
XP-61E 2 converted airframes Apr-45 Prototype daylight fighter, 2 crew, bubble canopy, turret removed, increased fuel capacity, no radar, four nose mounted .50 caliber guns, developed from P-61B.
XP-61G 16 converted airframes 1945 P-61B-20 modified for weather recon, unarmed

All airframes were built at Northrop’s plant in Hawthorne, California

Specifications P-61A P-61B P-61C
Engine  R-2800-10, R-2800-65, R-2800-65W R-2800-65W  R-2800-73
Maximum Engine Output [boosted]  2000 hp [2250 hp]  2000 hp [2250 hp]  2800 hp
Maximum Weight  29249 lbs  39056 lbs 41138 lbs
Standard Fighter Weight  28202 lbs   29876 lbs  30068 lbs
Empty Weight  23158 lbs  24413 lbs  26418 lbs 
Range [maximum external fuel]  ~1000 miles [+1800 miles]  ~1000 miles [+1800 miles] 
Maximum Speed  366 mph at 20,000 ft 366 mph at 20,000 ft 430 mph at 30,000ft
Armament [turret]  4×20 mm AN/M2 [4x .50 cal AN/M2]  4×20 mm AN/M2 [4x or 2x .50 cal AN/M2]  4×20 mm AN/M2 [4x .50 cal AN/M2] 
Crew  Pilot, gunner, radar operator  Pilot, gunner, radar operator  Pilot, gunner, radar operator 
Length  48′ 11″ 49’7″ 49’7″
Wingspan  66′  66′  66′ 
Wing Area  664ft² 664ft²  664ft² 

 

Specification  P-61A P-61B P-61C
Engine {P-61}  R-2800-10, R-2800-65, R-2800-65W R-2800-65W R-2800-73
Maximum Engine Output [boosted]  2000hp [2250 hp]  2000hp [2250 hp]  2800 hp
Maximum weight  13267 kg   17715 kg 18660 kg
Standard fighter weight  12792 kg 13552 kg 13639 kg
Empty Weight  10504 kg  11074 kg 11983 kg
Range [maximum external fuel]  ~1609 km [~2897 km]  ~1609 km [~2897 km]
Maximum speed  590 km/h at 6 km 590 km/h at 6 km  692 km/h at 9144 m
Armament [turret]  4x20mm AN/M2 [4x or 2x 12.7mm AN/M2]  4x20mm AN/M2 [4x or 2x 12.7mm AN/M2]  4x20mm AN/M2 [4x 12.7mm AN/M2] 
Crew  Pilot, gunner, radar operator  Pilot, gunner, radar operator  Pilot, gunner, radar operator 
Length  14.91 m  15.11 m 15.11 m
Wingspan  20.12 m 20.12 m 20.12 m
Wing Area  61.69 m² 61.69 m² 61.69 m² 

Cruising Speeds for the P-61A&B at 28,500lbs (12927 kg)

Altitude Speed
9000ft : 2743m 253mph : 408 kph
16000ft : 4876m 256 mph : 412 kph

Video

Gallery

Illustrations by Ed Jackson

P-61A ‘25507’, a very early A model. Only the first 37 aircraft would carry turrets before it was removed in order to correct buffeting issues and as to not compete with the B-29 program for certain components.
P-61A-5 ‘Lady Gen’ 9th AF, 422nd NFS, Florennes, Belgium. 1st Lt. Paul A. Smith (Pilot) & 1st Lt. Robert E. Tierney. This particular P-61A was among the first to travel to the ETO and was flown by the first US night aces. P-61’s would eventually switch to a gloss black livery, which would eventually replace this particular plane’s olive drab and invasion stripes.
P-61B ‘Midnite Madness II’, 548th Night Fighter Squadron, Iwo Jima. Stationed on Iwo Jima, aircraft of this squadron defended the recently taken island from nightly visits by Japanese raiders and reconnaissance aircraft.
P-61B ‘23968’ 414th NFS, 12th Air Force, Italy with a detachment to the 422nd NFS.
While two of the four MTO night fighter units were re-equipped with P-61’s and British built Mosquito’s, the other two continued to use their old Beaufighters until the end of the war.
P-61C, Thunderstorm Project, USA. With their guns traded for meteorological equipment, a select number of P-61C’s would embark on a foundational meteorological and aviation safety project. While never designed for such use, the P-61 would provide data for a groundbreaking study that would reveal the effects of thunderstorms on aircraft and the behavior of the convection cells within storms, among other major scientific findings.

The R-2800-10 and R-2800-65 differed in regards to their magnetos and ignition systems. War Emergency Power with the R-2900-65W was rated for 2250 hp at 60 inches of Manifold pressure, 2700 RPM [Pilot’s operating instructions for the P-61A]
Pilot’s Instrument Panel. The Pilot’s radar indicator position below the gunsight is empty [P-61 training manual]
The AN/CPS-1 was a microwave early warning radar system used to great effect by the USAAF across Europe, and later the Pacific, providing P-61 crews with accurate target information. The set here was deployed in Luxembourg, where it provided support during the Battle of Bulge [National Air and Space Museum]
P-61Bs patrol over the Marianas [National Archives]
The addition of underwing racks would largely solve the P-61’s range issues and allow it to carry a considerable bomb load [National Archives]
Members of an aviation engineering battalion add extensions to an airstrip in order to accommodate the P-61 [National Archives]
A P-61 landing at Luzon, 1945 [National Archives]
The radar operator worked from an isolated position at the rear of the aircraft. This would prove a major drawback of the design, as the pilot would be without radar guidance should the intercom system fail, the R/O accidentally pull the IC cable out, or if the tail cone imploded in a dive. [National Archives]
Reinforcement kits were used to keep the tail cones of these aircraft from imploding in dives or high speed maneuvers [Wikimedia]
A very early production P-61A of the 6th NFS in the Marianas, 1944. [National Archives]
Brig. General Earl W. Barnes flew a P-61 modified for his personal use with an extra fuel tank installed in the place of its turret. Pictured here at the opening of an airfield at Middleburg Island, New Guinea. [National Archives]
At -35F (-37C), engine heaters were needed to start this P-61 stationed in Alaska, 1944. [National Archives]
The P-61 gave squadrons who formerly relied on the P-70 a massive boost in performance and confidence. Here, Major V. Mahr of the 6th NFS climbs into his P-61 in Saipan, July 1944. [National Archives]
The XP-61E was a prototype escort fighter based on the P-61B, two aircraft were converted, but further development was canceled after the war. [Wikimedia]
P-61s of the 548th NFS wait for their shift to start, Iwo Jima. [Wikimedia]
A P-61 sits at Clark Field, Philippines, in August of 1945. By this point, the P-61 was being phased out by the P-38M in several squadrons. It would continue to see some use in the post war years. [National Archives]
The F-15A reporter was a photo reconnaissance aircraft developed from the P-61 that went into production after the war. These aircraft would participate in the Thunderstorm Project and serve in the Korean war. [Murph’s Models]

Credits

  • Written by Henry H.
  • Edited by Stan L. and Ed J.
  • Illustrations by Ed Jackson

Primary Sources:

  • Ashkenas, I L. The Development of a Lateral Control System for Use with Large Span Flaps. No. 1015. NACA, 1946.
  • Pilot’s Flight Operating Instructions Army Model P-61A Airplane. (T. O. NO. 01-15FB-1). Commanding General, Army Air forces. January 15, 1944.
  • Pilot Training Manual for the Black Widow P-61. Office of Assistant Chief of Air Staff Training. 1944
  • Handbook of Operating Instructions for Radio Set SCR-720-A and Radio Set SCR-720-B. AN 08-10-181. Joint authority of the Commanding General, Army Air Forces, and the Commanding General, Army Service Force. (1943).
  • Northrop P61 Black Widow Pilot’s Flight Operating instructions. T.O No. AN 02-35VC-3. USAF, July 1945
  • Introduction Survey of Radar Part II. Air Publication 1093D Volume 1 First Edition. Air Ministry, June 1946.

Secondary Sources:

  • Boiten, Theo. Nachtjagd Combat Archive 1944 Part Four. Surrey: Red Kite, 2021.
  • Boiten, Theo. Nachtjagd Combat Archive 1944 Part Five. Surrey: Red Kite, 2021.
  • Braham, Roscoe R. “Thunderstorms and the Thunderstorm Project”
  • Brown, Louis. Technical and Military Imperatives: a Radar History of World War II. Taylor & Francis, 1999.
  • Dean, Francis H. America’s Hundred Thousand: the US Production Fighter Aircraft of World War II. Schiffer Publ., 1997.
  • Kolln, Jeff. The 421st Night Fighter Squadron in World War II. Schiffer Pub., 2001.
  • McFarland, Stephen Lee. The U.S. Army Air Forces in World War II: Conquering the Night: Army Air Forces Night Fighters at War. Air Force History and Museums Program, 1998.
  • Pape, Garry R., and Ronald C. Harrison. Queen of the Midnight Skies: the Story of America’s Air Force Night Fighters. Schiffer Publishing Ltd., 1992.
  • Price, Alfred. Instruments of Darkness: the History of Electronic Warfare. Greenhill, 2005.
  • Thompson, Warren E. P-61 Black Widow Units of World War 2. Osprey, 1998.
  • “WW2 People’s War – Reg EVERSON’S STORY.” BBC. BBC. Accessed August 1, 2021. https://www.bbc.co.uk/history/ww2peopleswar/stories/26/a3130426.shtml.

 

 

 

Edo XOSE-1

USA flag old United States of America (1945)
Observation Scout Floatplane – 10 Built

XOSE-1 taking off. Notice it is painted in the wartime colors. [axis-and-allies-paintworks.com]
The XOSE-1 was an observation float plane built by the Edo float company during World War II and was intended to be a possible replacement for the OS2U Kingfisher. Before being built, the type seemed promising and ten prototypes were ordered. Although development was slow, the aircraft would finally fly after the war had ended. Testing showed the design was riddled with flaws and, with the end of the war making the observation floatplane obsolete and unnecessary, the XOSE-1 program was cancelled.

History

Photo of the mockup XOSE-1.

Before America had entered the Second World War, it was realized that many assets in the United States arsenal were outdated to some degree. Many aircraft were unable to compete with their contemporaries around the world. One such piece of equipment would be the ship launched floatplane. A concept that originated in the 1920s and 1930s, it involved the use of small floatplanes that were carried aboard large warships and could be deployed via catapults for a number of tasks to assist their mothership. These missions included long range scouting, spotting for the warships’ main guns and also providing anti-submarine protection using depth charges or torpedoes. Most of America’s larger warships were equipped with catapults at the time for this purpose. The dedicated ship-based floatplanes the United States Navy (USN) operated at their entrance to the war was the aging Curtiss SOC biplane and the Vought OS2U Kingfisher. The latter would soon replace the former and would enter widespread service after the Attack on Pearl Harbor. Although the Kingfisher was just entering service, the search for a modern seaplane that would eventually replace the aircraft began. The new type was expected to carry out the same duties as its predecessor but also be able to effectively protect itself if needed. The OS2U only had one .50 caliber machine gun for offense, which wasn’t very helpful when against newer fighters. The first and most prominent aircraft that would rise to meet this role would be the Curtiss SC Seahawk, but it would not be the only type that would be built. In fact, a competitor would come from a little known company called Edo.

The Edo Aircraft Company is not a company often mentioned in history regarding the Second World War. The company was founded in 1925 by Earl Dodge Osborne, with the name being an acronym of his own name. Despite being rarely discussed among historians, Edo was immensely crucial to the war effort for the USN. Edo was a primary producer for aluminum floats before the war and would be the main producer for the floats on Navy floatplanes, like the OS2U. It was estimated that up to 95% of floats used on USN aircraft were built by Edo. Not only was Edo responsible for the production of the floats, they were also known for adapting said floats for use on the aircraft that would use them. Edo had become known for their work on floats, but they worked on a handful of their own floatplane designs in the years before WWII had started. However, this was around the time the company was created in 1925, and aircraft design had changed drastically since then. Given their background and knowledge with designing and fitting floats, the USN requested that the Edo company should attempt to design their own modern floatplane for the ship-based observation role. Eager to attempt building a modern aircraft, Edo eagerly accepted the request. On January 11th, 1944, they would begin work on their floatplane, which would be called the XS2E-1.

Frontal view of an XOSE-2 or XTE-1. The two were visually identical from the outside. [axis-and-allies-paintworks.com]
The preliminary design of the XS2E-1 was deemed acceptable by the Navy and an order for ten prototypes was made. The XS2E-1 would be a two seat design with a Ranger V-770-8 engine. The engine mount and cowling would also both be designed by Ranger (this company would become Fairchild after the war.) Additionally, a Westinghouse 19 turbojet was to be installed in the rear of the aircraft to offer increased thrust for evasion or to give chase to an enemy aircraft. This would make the aircraft a mixed powerplant type. Another order for eight more units was made some time after the first order, but an exact date is unknown. On March 16th, 1944, the USN opted to change the floatplane’s design. The Westinghouse 19 turbojet that was planned for the project was experiencing its own difficulties in development.

When the XS2E-1 was drafted, the turbojet, due to its development, had become much heavier than what Edo was expecting. Due to this weight increase and a high demand for the jet engine on other aircraft projects, it was removed from the XS2E-1. This caused a weight problem in the aircraft’s design, as it no longer had the additional thrust needed to operate with its then-current weight. Edo changed the aircraft’s design drastically to make the XS2E-1 lighter. A significant revision done was the removal of the second seat, making the aircraft a smaller, single-seater aircraft. This, however, meant all the work the 2nd crewmen was intended to do was now transferred to the pilot, which would include operating the radar system in addition to flying and observing.

A frontal shot of an XOSE-1 demonstrating its folding wings. [axis-and-allies-paintworks.com]
After the loss of the turbojet and the switch to a single seater design, it was decided to change the aircraft’s role to an Observation Scout floatplane. Another reason for the change was that, developing parallel to the XS2E-1, was the Curtiss SC-1 Seahawk mentioned earlier, an aircraft that was meant to fill the Scout role for the USN. Finding that developing two aircraft with the same role was redundant, the USN authorized the role change on the XS2E-1. With the new role, the XS2E-1 was redesignated as the XOSE-1. Not long after the role and design change, a full-scale wooden mockup of the new XOSE-1 was built and an inspection was held on November 24th, 1944. An early criticism of the design was linked to the removal of the second seat, as would-be operators complained the intense workload was too much to put onto the pilot. A variant was soon conceived, the XOSE-2, which would address this workload issue by reintroducing the second seat for another crewman. This second crewmen would be tasked with operating the onboard radar system and performing observation duties. An order for two XOSE-2s to be built, as well as for a derivative of the XOSE-2 that would be a dual-control training version, soon followed after conception of the two-seat variant. The trainer would be named the XTE-1. Progress on the program overall was slow up to this point, but Edo had added many innovative features to the design to improve its performance.

Side view of an XOSE-1 taking off. [axis-and-allies-paintworks.com]
The war came to an end before the XOSE-1 could take flight. The end of the war saw most of the projects the USN was working on be terminated immediately, as there was no purpose in developing them anymore. The XOSE-1, however, was saved from this fate, as the USN allowed the floatplane to continue development after the end of the war. The XOSE-1’s first flight took place on December 28th of 1945, only a few months after the war had ended. Since there was no urgency to press this new type of aircraft to the frontlines anymore, funding to the program was cut and work slowed down in accordance. The XOSE-2 version finally flew on September 24th, 1947, two years after the war was over. Two XOSE-2s were built. It is unknown exactly when the first XTE-1 was completed and flew, but two of this type were built as well. Originally, during its debut, the XOSE-1 was painted in the standard blue-on-the-top-white-on-the-bottom that mid war USN aircraft used, but would later be colored in the dark blue that late/post-war Navy aircraft were painted in.

Rear view of an XOSE-1 with the floats detached and the wheels attached instead. [axis-and-allies-paintworks.com]
Despite being a company that had only built a handful of planes two decades prior, the XOSE-1 was very promising from the outset, but problems soon began to arise during testing. The XOSE-1 experienced trouble with the Ranger built engines. The two seater XOSE-2 experienced many more problems and major changes had to be implemented in the design. Some remedies to the problems included increasing the height of the tailfin and the addition of a ventral strake below the tail to help with stability. Stability issues were found to be caused by the two seater’s larger canopy installed on the largely unmodified fuselage. By the time the stability issues were resolved, it was almost for naught, as the aircraft program was going nowhere.The shipborne floatplane type itself was beginning to show its obsolescence compared to newer technology. Exactly when the program ended or the whereabouts of the ten XOSE built are unknown, as details about the program during this time are sparse. It is unknown if the XOSE-1 was ever even tested from a ship, as many warships postwar would have their catapults removed. Most of the testing was done via land or sea takeoff, with wheels attached to the floats or a landplane conversion where the floats were replaced with a conventional landing gear. The type would be slowly replaced by ship-based helicopters, an idea that had begun during the Second World War and expanded upon thereafter. The era of the scout floatplane, especially shipboard ones, was over. It is most likely all of the XOSE-1s and its derivatives were scrapped before 1950, as all shipboard seaplane squadrons had been disbanded in 1949.

Design

An Edo XOSE-1 in flight [axis-and-allies-paintworks.com]
The Edo XOSE-1 was a single-seat floatplane design of all metal construction. It’s floatation was provided by one large aluminum float under the hull, and two smaller aluminum floats on the wingtips. This layout was used on almost every USN floatplane. In addition to floats, the aircraft was also able to be launched via catapult aboard a ship. An optional wheeled undercarriage was also available for ground based takeoffs.

The body of the aircraft would be constructed of metal and would contain 135 lbs (61 kg) of armor. The overall weight of the aircraft would be 5,316 Ib (2411.3 kg) standard and 3,973 Ib (1802 kg) empty. The fuselage would have a length of 31 ft 1 in (9.5 m) and a height of 14 ft 11 in (4.5 m). The XOSE and its variants had a unique construction that allowed many parts of the aircraft to be easily accessible for maintenance.

A rear view of an XOSE-1 with its wings folded back. [shu-aero.com]
The Edo XOSE-1 and all of its variants were equipped with the Ranger V-770-8 inline engine that gave it a top speed of 188 mph (302 km/h), a cruising speed of 111 mph (178.6 km/h) and a stall speed of 61 mph (98.2 km/h). The aircraft would have a climb rate of 1,350 ft/min (411.5 m/min) and a maximum service ceiling of 22,300 ft (6797 m). The XOSE-1 would also have a range of 600 mi (965.6 km).

The cockpit would allow protection for the pilot, as the canopy was bulletproof. The canopy was one piece and would slide down and behind the cockpit for easy movement in. On the XOSE-2/XTE-1, the cockpit would be lengthened to accommodate the additional crewman, who would do observation and radar tasks. The canopy on the two seater versions would be two parts and the forward segment would slide back over the rear section.

Fuel would be stored in the fuselage in self-sealing fuel tanks. The tail section of the aircraft would be of metal construction as well. The only differences between the two versions were on the two seaters, in which the tail of the aircraft had to be extended height-wise and a vertical strake beneath the tail was added. Both of these changes helped in the stabilization of the two seaters. The wings of the aircraft were also constructed of metal and would have a wingspan of 37 ft 11 in (11.6 m). The wings would utilize a unique feature for some of its control surfaces. The flaps, that extended outwards from the folding line, would retract automatically if enough water impacted them. This was put in place to prevent damage to these flaps. Additionally, there were retractable slats on the leading edge of the wings to increase drag. The wings themselves could be folded inward for easy storage aboard ships or hangars. Interesting to note, the wings had a manual folding system instead of a hydraulic system most aircraft at the time had.

For armament, the XOSE-1 was equipped with two M2 .50 caliber machine guns as standard. Two hardpoints were equipped on the wings that could allow the XOSE-1 to carry two 350 Ib depth charges or two 50 gallon drop tanks. Additionally, two emergency rescue racks could also be carried on the underside for air to sea rescue missions. A single hardpoint could also be used to carry a radar pod. There is also mention of the XOSE-1 having smoke projectors as well. The two-seat XOSE-2 would lose one of the M2 machine guns and only carry a single gun. The XTE-1 variant would be completely unarmed, given it was only a trainer.

Conclusion

With the Edo XOSE-1 program being terminated, this would be the last time Edo would build an aircraft all on their own. However, Edo would propose a very interesting concept to the US Navy in the 1950s for an amphibious fighter similar to the Convair F2Y Sea Dart. However, this type would never be built.

Variants

  • XS2E-1 – Initial design of the XOSE-1. The XS2E-1 was a two seater and mounted a larger engine as well as a Westinghouse J19 jet engine. This design was changed and became the XOSE-1.
  • XOSE-1 – Single seat reconnaissance floatplane. The XOSE-1 had two .50 Cal M2 machine guns mounted in the wings and two hardpoints for depth charges. 6 were built.
  • XOSE-2 – Two seat version of the XOSE-1. The E-2 version would have a radar operator, a lengthened canopy, and only a single .50 cal for defense. Two were built.
  • XTE-1– Tandem control version of the XOSE-1. This version would be unarmed and would be used for training purposes. Two were built.

Operators

  • United States of America – The XOSE-1 and its variants were only tested by the United States Navy.

Edo XOSE-1 Floatplane Specifications

Wingspan 37 ft 11 in / 11.6 m
Length 31 ft 1 in / 9.5 m
Height 14 ft 11 in / 4.5 m
Wing Area 237 ft² / 22 m²
Engine 520 hp (387.7 kW) Ranger V-770-8 Inline Engine
Propeller 2-blade Hamilton Standard constant-speed propeller (9ft / 2.7m diameter)
Powerplant Ratings
Horsepower output Altitude
Take Off 550 hp Sea Level
Normal

(Approx. 84% Throttle)

500 hp 800 ft / 244 m
Fuel Capacity 120+58 US Gal / 454+219 L
Weights
Empty 3973 lb / 1802 kg
Gross 5316 lb / 2411.3 kg
Maximum 6064 lb / 2750.6 kg
Climb Rate (at sea level) 1,350 ft / 411.5 m per minute
Maximum Speed 188 mph / 302.6 kmh
Cruising Speed 111 mph / 178.6 kmh
Stalling Speed 61 mph / 98.2 kmh
Range 600 mi / 965.6 km
Maximum Service Ceiling 22,300 ft / 6797 m
Crew 1 pilot
Armament
  • 2x 12.7x99mm / .50 cal Browning AN/M2 machine guns
  • 2x 350 Ib / 158.8 kg Depth Charges

Gallery

Illustrations by Ed JacksonEdo,d

Edo XOSE-1 in Standard Wartime Colors
Edo XOSE-1 with the additional ventral stabilizers added
A view showcasing the retractable flaps on the engine.

Two Edo XOSE-1s in flight together [shu-aero.com]
A side view of the XOSE-1 in flight A side view of the XOSE-1 in flight. [axis-and-allies-paintworks.com]
Rear view of an XOSE-2 or XTE-1. [axis-and-allies-paintworks.com]

Credits

  • Article written by Medicman11
  • Edited by Stan L. and Ed J.
  • Illustrated by Ed Jackson
  • Jane’s All the World’s Aircraft 1947
  • Norton, Bill. American aircraft development of WWII : special types, 1939-1945. Manchester: Crécy Publishing Ltd, 2016. Print.
  • Wagner, Ray. American combat planes of the 20th Century : a comprehensive reference. Reno, NV: Jack Bacon & Co, 2004. Print.
  • Buttler, Tony. American secret projects : fighters & interceptors, 1945-1978. Hinckley: Midland, 2007. Print.x

Vultee XA-41

USA flag United States of America (1944)
Prototype Ground Attack Aircraft – 1 Built

XA-41 in flight

The Vultee XA-41 was a single-engine aircraft that began life as a dive bomber. Months later, its role was changed to a low-level attack aircraft. The XA-41 performed admirably in flight tests, but the United States Army Air Corps (USAAC) eventually decided that the fighter aircraft then in service were already performing well enough in the attack role. Despite its job being erased, the XA-41 continued development as a testbed, showing off the powerful XR-4360 engine it mounted and how much it could carry. The aircraft itself would have been deadly had it been produced in large numbers, as it boasted four 37mm cannons. As the war went on, the XA-41 was still being tested. Throughout the trials, the aircraft had extremely good performance, even being able to outturn a P-51, but its speed wasn’t quite enough for its role. At one point, it was given to the Navy for testing and eventually it would wind up at Pratt & Whitney (PR). At PR, it served as a testbed through the war and was eventually scrapped in 1950.

History

Cutaway Concept for the XA-41

The XA-41 began as part of a United States Army Air Corps (USAAC) requirement in October 1941 for a new single-seat attack aircraft. The Douglas A-20 Havoc (and soon to be produced North American A-36 Apache) was performing well at the time, but the USAAC wanted something new. The aircraft requested had to be able to reach at least 300mph (482.8kph) at sea level, have a service ceiling of at least 30,000ft (9,144m), and a range of 1,200 miles (1,932km). For the attack role, the aircraft was to have either 37mm, 20mm, or 50. cal guns mounted in the wings. Given this imposing armament, it is likely the aircraft would have attacked soft targets or even been used for tank-busting.

The USAAC commissioned Vultee Aircraft Corporation, Kaiser Fleetwings, and Curtiss to design a new aircraft for the role. Kaiser Fleetwings developed the XA-39, which would have mounted the R-2800-27 engine. Their aircraft didn’t progress beyond the mockup stage. Curtiss reused their naval XTBC-1 prototype for their part, renaming it the XA-40. This also didn’t go beyond the mockup stage. Vultee’s answer was the V-90, a ground attack aircraft mounting the fairly new and powerful R-4360 engine. Interestingly, the XA-41 started off as a dive bomber, despite it being commissioned as an attack aircraft. It isn’t often stated, but the Army had been interested in dive bombers since 1940, going as far as purchasing several Navy designs. The Army bought several SB2D-1 Helldivers in December of 1940 and renamed them the A-25 Shrike. They also had a troubled history with one of Vultee’s own aircraft, the A-35 Vengeance, which they tried numerous times, but to no avail. The XA-41 was most likely a chance for the Army to have a successful dive bomber or attack aircraft. The Army was satisfied with Vultee’s V-90 design and awarded a contract for two prototypes on November 10, 1942. Shortly after a mockup inspection, the Army interestingly switched the role from a dive bomber to a dedicated attack aircraft. The switch was rather abrupt and caused a delay in the development.

XA-41 Prototype

On April 30th, a new contract was signed which included a static mockup. Vultee continued construction on the project until the prototype was halfway completed, at which point the Army decided that the most current aircraft, the Republic P-47 Thunderbolt and North American P-51 Mustang, were already quite capable in the attack role. But this wouldn’t be the end of the XA-41. Maj.Gen Oliver P. Echols, Chief of the Material Division, opted not to cancel the program and instead complete it in order to show the true potential of a new attack aircraft with the new R-4360 engine. This decision went through on November 20th, with the prototype ¾ the way through completion. The 2nd prototype was cut at this point and only one would be built (Serial No. 43-35124/5).

On February 11th, 1944, the XA-41 flew for the first time from Vultee Field, California, with test pilot Frank Davis at the controls, and landed at March Field, California. Several flights after this were conducted at the nearby army base. The aircraft was flown by both Vultee and Army pilots, and both agreed it handled well. There were some problems here and there, which Vultee quickly fixed with some additions to the airframe. On June 25, the Army accepted the XA-41. On July 16th, on its 60th flight, the aircraft was ferried to Eglin Field, Florida. Testing proved the XA-41 was an exceptional aircraft, with many great features. The craft had an excellent turn rate, being able to outturn the P-51. For its size, it carried an impressive arsenal of weapons. But the Army wanted an attack aircraft that could also defend itself if the need would arise, and the 350mph (563.2 km/h) of the XA-41 wasn’t that impressive compared to other aircraft in service. The United States Navy became interested in the XA-41 at some point and the prototype was given to them for testing at Patuxent River, Maryland. The Navy wanted to see if the aircraft could be flown from aircraft carriers. After the Navy briefly tested it, the XA-41 was given to Pratt & Whitney (PR) on August 22, 1944. It was obvious at this point that the XA-41 would never see combat, but would remain in the US as a testbed aircraft. Serving with PR, it was used as a flying testbed for their R-4360 engine, as well as having a supercharger mounted. As testing continued, the aircraft was purchased by PR on October 9 and re-registered as NX6037N. There are few documents that reference the XA-41 post-war. The only thing mentioned is that the sole XA-41 was finally scrapped in 1950, having served for many years at PR.

Design

The XA-41 was a conventional single-engine aircraft. It had a slight gull wing and a tail sitter configuration. The landing gear in the wings would retract inboard and was placed widely to allow better landing performance. During development, the tail wheel had doors installed to completely cover it in-flight. The cockpit was placed far forward and raised to allow the pilot to see over the engine, giving him better visibility when attacking ground targets. The ventral tail had an extension that spanned most of the length of the aircraft. This was added during development. A spinner was also added at some point. The XA-41 mounted the PR XR-4360 Wasp Major engine, which was the main reason the Army and PR were so interested in the project.

For armament, the XA-41 had four M2 Browning .50 cal machine-guns mounted in the wings. For the attack role, it was meant to mount four more 37mm cannons (sources don’t mention what particular kind of gun) in the wings. All armament in the wings was placed outside of the propeller’s range. For bombing, the XA-41 had a bomb bay that could carry four 500Ibs bombs, a torpedo, additional fuel, or two 1,600Ibs weapons. In total, it had up to 6,500Ibs of ordnance. Documents mention that up to 1,100Ibs of additional bombs could be mounted to the wings. The aforementioned competing XA-39 only sported the four Brownings, two 37mms, as well as a predicted carrying capacity of six 500Ibs bombs.

Variants

  • XA-41 – [The sole prototype built, used as a testbed for the XR-4360 engine.]

Operators

  • United States of America – The United States Army Air Corps would have operated it had it entered production. After serving as a testbed for the Army, the Navy and Pratt & Whitney also did tests with the aircraft.

XA-41 Specifications

Wingspan 54 ft / 16.4 m
Length 48 ft 7 in / 14.8 m
Height 14 ft 5 in / 4.4 m
Wing Area 540 ft² / 164.5 m²
Engine 1x 3,000 hp ( 2240 kW ) XR-4360-9
Propeller 1x 4-bladed Hamilton Standard propeller
Weights
Empty 13,400 lb / 6078.1 kg
Gross 18,800 lb / 8527.5 kg
Maximum 23,359 lb / 10595.4 kg
Climb Rate
Rate of Climb at Sea Level 2,326 ft / 708.9m per minute
Maximum Speed 353 mph / 568 kmh at 15,000 ft / 4572 m
Cruising Speed 270 mph / 434.5 kmh
Range 800 mi / 1287.4 km
Maximum Service Ceiling 27,000 ft / 8229.6 m
Crew 1 pilot
Armament
  • 4 Browning M2 machine guns (400rpg)
  • 4 37mm cannons (30rpg)
  • Up to 6,500 Ibs of weapons

Gallery

llustrations by Haryo Panji https://www.deviantart.com/haryopanji

Vultee-XA-41 Original Prototype Colors
Vultee XA-41 with possible service markings (artist interpretation)
XA-41 at Patuxent River
XA-41 on the runway
XA-41 in a slight climb
XA-41 parked on the ramp

Sources

 

Douglas XTB2D-1 Skypirate

usa flag USA (1945)
Prototype Torpedo Bomber – 2 Built

XTB2D-1 on the Runway
First prototype of the Skypirate on the runway.

The Douglas XTB2D-1 “Skypirate” was a large, single-engine torpedo bomber built for use on the Midway class carriers during World War 2. At the time, it was the largest aircraft to be used aboard a carrier, dwarfing even two-engine designs. Unfortunately for the Skypirate, engine troubles, little support from the US Navy (USN), and numerous setbacks with the construction of Midway-class carriers nearly doomed it from the start. By the time it was airworthy, it was trying to fill an obsolete role which other aircraft, such as the TBF/TBM Avenger, already filled adequately. Work continued after the war, with several attempts to revive the program but it proved to be too costly and the Skypirate program was finally cancelled in 1947, with the two prototypes being scrapped in 1948.

History

With engagements such as the Battle of the Coral Sea and the hunt for the Bismarck, the effectiveness of torpedo bombers, such as the TBF/TBM Avenger and Fairey Swordfish, was clear. With the announcement of the large Midway-class carriers, the possibility of a new torpedo-bomber/scout bomber came about. In February 1942, a competition was put forward by the Navy for this role. The Douglas Aircraft Company, based in Southern California, proposed the Skypirate. The single-engine Skypirate was picked from eight different designs, most of which were two-engined. The Bureau of Aeronautics (BuAer) wasn’t expecting a single engine design to be submitted, assuming the specified massive carrying capacity would require a two-engine design. The program was being headed by Ed Heinemann as lead designer and Bob Donovan as the chief engineer, who would be on the project until the end.

XTB2D-1 Frontal View
An impressive look at the massive Skypirate from the front.

In November of 1942, Douglas was given permission to begin production of two prototypes and a mockup of the XTB2D-1 (then called the Devastator II, before being changed to Skypirate). Delays in the development of the Midway class would continue to hamper the Skypirate throughout its life. The finished product was a formidable aircraft, capable of carrying four torpedoes from land or two torpedoes from a carrier, the former being four times the carrying capacity of the TBM Avenger. In March and May of 1943, the mockup was inspected and an order for 23 production aircraft was put in. This was enough for a single squadron to operate from a Midway carrier. Problems began about this time, with the delivery of engines and propellers being delayed. By 1944, the Skypirate was still not airborne and it was obvious it wouldn’t be operational anytime soon. With earlier torpedo bombers performing adequately, a lack of support from the Navy, most of the Japanese fleet in shambles and continued delays with the Midway class (which would eventually sail after the war), the 23 production planes were cancelled. On February 18th, 1945, the first Skypirate was rolled out of the production facility, being completed on March 13th and finally going airborne on May 8th. Neither of the prototypes had any defensive armaments, but they were tested with torpedoes and drop tanks. Although no production was to ever start, the Skypirates would continue flying until the end of the war. During one such flight in June of 1945, a Skypirate was damaged mid-flight, but the craft was brought down safely. Engine problems were a frequent issue with the Skypirate and propeller problems would ground it in August of 1945, not flying again until after the war.

Skypirate Landing
Perhaps the most well known photo of the aircraft, the Skypirate prepares to land.

Postwar, the aircraft industry changed with the introduction of jet aircraft, thus eliminating the need for many prototypes being developed during the war. The Skypirate was no exception. With the torpedo bomber role now fading, the Douglas firm looked at other options to revive their Skypirate. Some ideas included using the Skypirate for an electronic warfare role or even as an anti-submarine aircraft (a role overtaken by another piston engine aircraft, the Grumman AF Guardian), but none of these propositions ever managed to become reality. As the Cold War was just beginning, the Skypirate program ended in 1947 and the 2 prototypes were scrapped in 1948.

Design

The Skypirate is most likely the largest single-engine aircraft to ever be designed for carrier operations. In comparison, the twin-engined B-25 Mitchell medium bomber measured around the same in length and width.

In flight
The 2nd prototype in flight, notice how the tail is shorter in comparison to the first prototype.

The initial Skypirate design had an internal bomb bay, which the prototypes dropped in favor of four external Mark 51 Mod 7 bomb hardpoints. These hardpoints could carry a range of weapons including 500Ib-2000Ib bombs, torpedoes, depth charges, mines or even incendiary bombs. The use of up to 4 Mk.13 Torpedoes (from land) were planned had it entered production. The Skypirate could alternatively carry up to 8,400Ibs of bombs. For offensive armament, the Skypirate had 4 M2 Browning machine guns in the wings. For defense, it had a Firestone model 250CH-3 remote turret behind the cockpit which carried 2 M2 Brownings and a turret in the back of the lower fuselage which carried a single M2 Browning. The lower turret was remotely fired through electronic control and powered hydraulically. Drawings indicate that Mark 2 Gun Containers could be added for extra forward firepower but none were ever attached during testing. 300 gallon drop tanks were also fitted during testing and could have been used had the craft been operational.

The Mockup
The sole mockup made alongside the 2 prototypes.

Along with such an impressive weapons payload, the Skypirate was full of advancements which would have improved its performance. To get such a large aircraft off the ground, the Skypirate was powered by a single Pratt & Whitney XR-4360-8, the largest radial engine ever built up to that time. The engine had a unique exhaust style that combined the exhausts in alternating rows to lower the effects of backpressure. Being a carrier-based aircraft, the Skypirate had folding wings as well as a catapult hook. The inclusion of a tricycle landing gear was also interesting, as it helped with bomb loading and carrier space. Most single engine aircraft of the time preferred using a tailwheel. The Skypirate had large flaps that extended the length of wingspan. The outer flaps served as ailerons while the midsection flaps were used as dive flaps. The dive flaps could also be lowered to help the aircraft cruise or assist in turning to help ease the stress off the aircraft when fully loaded. To assist with bombing or flight in general, a Type 3 Sperry vacuum-controlled, hydraulic autopilot was also to be added. A de-icing system was also added that pumped hot air over the wings and tail section.

The planned modifications of the prototypes are interesting to note. The 2nd prototype (Bu.36934) differed from the first, having a shorter tail of 8.6 ft, compared to the regular 10.5 ft tail of the original design. This was done most likely to conserve valuable space when inside a carrier. Along with these differences, plans to fit a jet engine in the fuselage of the 2nd prototype were made, but nothing ever came to fruition. The first prototype (Bu.36933) had a larger tail and was planned to be converted for the scout bomber role. These plans included adding cameras onboard. As with the jet engine designs, these also never came to be.

Variants

  • XTB2D-1 Bu.36933 – Prototype version, lacked any armament
  • XTB2D-1 Bu.36934 – The 2nd prototype. The tail was shortened to 8.6 ft. Also lacked any armament.
  • TB2D-1 – Proposed production version, 23 were ordered and planned production was to be 100 built every month. These versions were to be fitted with four .50 caliber machine guns in the wings, two in a Firestone power turret and one remotely controlled in the ventral hull. Eventually, the production versions were cancelled in favor of higher priority projects.

Operators

  • United States of America – Slated to be used aboard the Navy’s Midway-class carriers, with the end of the war and other setbacks, the XTB2D-1 was never used operationally.
XTB2D-1 Rollout
The design team poses with the first Skypirate on rollout day.

TB2D-1 Specifications

Wingspan 70 ft / 21.3 m
Length 46 ft / 14 m
Height 22 ft 6 in / 6.9 m
Wing Area 605 ft² / 184.4 m²
Engine 1x 3,000 hp ( 2240 kW ) XR-4360-8
Propeller 1x 8 bladed Hamilton Standard contra-rotating propeller
Fuel Capacity 501 US gal / 1896 L
Oil Capacity 28 US gal / 106 L
Empty Weight 18,405 lbs / 8350 kg
Gross Weight 28,545 lbs / 12950 kg
Maximum Weight 34,760 lbs / 15765 kg
Rate of Climb at Sea Level 1,390 ft / 425 m per minute
Time to 10,000 ft / 3048 m 8.2 minutes (Normal) 10.2 minutes (Military)
Time to 20,000 ft / 6096 m 22.3 minutes (Normal) 26.5 minutes (Military)
Maximum Speed 340 mph / 550 km/h at 15,600 ft / 4755 m
Cruising Speed 168 mph / 270 km/h

312 mph / 500 km/h (with torpedoes)

Range 1,250 mi / 2010 km (Torpedoes)

2,880 mi / 4635 km (Maximum)

Maximum Service Ceiling 24,500 ft / 7470 m
Crew 1 pilot

2 gunners

Armament 4 Browning M2 machine guns mounted in the wings (1600rds)

2 Browning M2 machine guns mounted in turret (1200 rds, incl remote 50.)

1 remote Browning M2 machine guns mounted in ventral hull

4 x Mk 13 Torpedoes (from land)

2 x Mk 13 Torpedoes (from carrier)

2 x 2,100 lbs Bombs

Total of 8,400 lbs payload capacity

2 x Mark 2 Gun Containers

Gallery

llustrations by Haryo Panji https://www.deviantart.com/haryopanji

Douglas XTB2D Skypirate Side View
Douglas TB2D Skypirate Side View – With the Defensive Gun Pod

 

XTB2D End
The last known photos of the Skypirates before being scrapped.
XTB2D Loaded
One of the prototypes with mounted Mk-13 torpedoes.
XTB2D Backside
A back view of one of the prototypes
XTB2D-1 on the Runway
First prototype of the Skypirate on the runway.
XTB2D-1 Frontal View
An impressive look at the massive Skypirate from the front.
Skypirate Landing
Perhaps the most well known photo of the aircraft, the Skypirate prepares to land.
In flight
The 2nd prototype in flight, notice how the tail is shorter in comparison to the first prototype.
The Mockup
The sole mockup made alongside the 2 prototypes.
XTB2D-1 Rollout
The design team poses with the first Skypirate on rollout day.

Sources

 

Plane side view

Consolidated Vultee XP-81

USA flag United States of America (1944)
Prototype Escort Fighter – 2 Built

The first XP-81 powered by the TG-100 engines (Convair)

The Consolidated Vultee XP-81 was a prototype mixed power fighter developed in late 1943 by the Consolidated Vultee Aircraft Corporation in order to meet an Army Air Force requirement calling for a high altitude escort fighter. Plagued by slow development and engine problems, the XP-81 would never see active service and development would be terminated in 1947. Despite this, the XP-81 still holds a distinct place in history as America’s first turboprop engine plane to fly and the world’s first plane to fly with a turboprop engine and a jet engine together.

History

With the formal introduction of the Boeing B-29 Superfortress on May 8th of 1943, it would be clear that a high altitude escort fighter would soon be needed to accompany the Superfortress on its bombing missions over the Pacific. In the summer of 1943, this need was realized and the United States Army Air Force (USAAF) issued a list of design requirements that consists of the following:

  1. 1,250 mile (2,012 km) operating radius
  2. Fuel for 20 minutes of combat plus reserve fuel supply for landing
  3. Cruising speed of 250 mph (402 km/h) at 25,000 ft (7,620 m)
  4. Maximum speed over 500 mph (804 km/h)
  5. Combat ceiling of 37,500 ft (11,430 m)
  6. Climb rate of 2500 fpm (feet per minute) / 762 mpm (meters per minute) while at 27000 ft (8230 m)
  7. Two engines*
  8. 12 ° angle of vision over the nose

* – The USAAF recommended that the designers use a two engine setup consisting of a propeller engine for long range flights while complemented by a jet engine for high speed combat situations.

Promotional illustration of the XP-81 showing a diagram of the XP-81. (Consolidated Vultee)

Interested in this proposal, the Consolidated Vultee Aircraft Corporation, later known as Convair, began work on an aircraft which would meet the specifications, appointing Charles R. Irving, who was a chief engineer of the Vultee Field Division and Frank W. Davis, the assistant engineer, who was also the chief test pilot, as the leaders of the design team. The project was known as the “Model 102” within Consolidated Vultee. In the early stages of development, the designers faced a dilemma of engine selection. The Pratt & Whitney R-2800 Double Wasp radial engine was considered, as was the General Electric TG-100 turboprop engine. After some evaluating and testing however, the TG-100 was selected as it was deemed to have superior performance for combat and cruising situations. As for the jet engine in the rear, a relatively straightforward choice to mount a General Electric J33-GE-5 (also known as I-40) jet engine was made. After a couple of months of development, Consolidated Vultee submitted a preliminary design proposal to the United States Army Air Force in September of 1943. Relatively interested in this design, the plane was given the greenlight for further development and received the designation “XP-81” by the Air Material Command.  

Detailed work on the XP-81 began in January 5th of 1944 and on January 18th, Consolidated Vultee was given the contract (no. W33-038-ac-1887) by the USAAF worth about $4.6 million to construct two flying XP-81 prototypes and one airframe for ground testing under the USAAF project name “MX-480”. Another contract followed on June 20th of 1944 worth $3,744,000 for the two flying examples, the airframe and the testing data. The contract was later modified to include 13 YP-81 under the project name “MX-796”. The construction of the first XP-81 prototype would begin on January 20th at the formerly independent Vultee aircraft factory in Downey, California but problems soon surfaced. Some time in April, the Air Material Command was notified that there would be a delay in the delivery of the TG-100 due to a couple of technical difficulties. As such, construction of the first prototype was delayed as the designers sought out an alternative engine to replace the TG-100 in June.

Consolidated Vultee flight test crew poses with the first XP-81 prototype. (XP-81)

The Packard V-1650-3 (some sources state V-1650-7), which was the American copy of the British Rolls-Royce Merlin engine, was selected to fill in the gap and the USAAF promptly provided Consolidated Vultee with such an engine taken from a North American P-51D Mustang. Within a week of receiving the engine, Consolidated Vultee engineers were able to install it after making considerable structural modifications to the first prototype’s airframe. A radiator similar to that of the Lockheed P-38J’s “beard” radiator would also be mounted on the XP-81, under the propeller spinner. Unfortunately for the designers however, the change of powerplant would add 950 lb (431 kg) to the plane while taking away 960 hp at takeoff and 1720 hp at top speed. With the relatively slow development, the first XP-81 prototype would finally be completed in January of 1945 bearing the serial number of “44-91000”.

Although the aforementioned issues with weight gain and horsepower loss were present, the Packard engine powered XP-81 was still deemed safe for flight tests, and as such, the first XP-81 prototype was prepared for test flights at Muroc Dry Lake in California and finally took to the skies on February 7th of 1945 with Frank W. Davis in the cockpit. Amazingly enough, 46 test flights were made with the Packard engine and it accumulated a total of 47.75 flight hours. In the testing phase it was noted that with the Packard engine installed, the XP-81 had poor directional stability at low speeds and the occasional splatter of oil on the windscreen by the propellers. Plans to replace the Packard engine were brought up on May 18th of 1945 when the TG-100 turboprop was finally available. The conversion was completed and the first prototype was returned back to Muroc for more tests on June 11th. Due to the new engine installation, extensive ground work had to be accomplished before flight tests were to continue. Throughout June 23rd to December 20th of 1945, numerous ground tests were conducted and a few problems surfaced. For one, the TG-100 was difficult to start and once it did, the pilot would have difficulty controlling the propeller. As this was an early turboprop engine,

reliability was low and the turbine wheels had to be replaced constantly, sometimes only after half an hour of use. The 10 inch (25 cm) oil cooler for the TG-100 was also deemed a problem, and it was thus increased to a 12 inch (30 cm) system instead. Perhaps the biggest problem however, was the throttle lag the XP-81 suffered. Frank W. Davis describes the problem by stating “The pilot had about a 10 second lag when he wanted to go and about 2 seconds lag when he wanted to stop, with both thrust and drag being powerful and non-adjustable when they did occur.” (Consolidated Vultee XP-81, by Steve Ginter). The ground personnel concluded in these ground tests that the current Aeroproduct A542 propeller and drive shafts were incompatible with the TG-100, and that new propellers should be developed. An emergency engine feathering system was also recommended.

The first flight of the XP-81 with the TG-100 engine occured on December 21st of 1945. This was the 47th test flight the first XP-81 underwent. Performance was rather satisfactory, and the flight concluded after a mere 5 minutes. Excessive oil consumption was noted however. Test flights with the TG-100 proved disappointing as the turboprop did not perform as it was advertised, delivering less horsepower than was expected. Out of the estimated 2,300 hp the TG-100 was suppose to achieve, only 1,400 hp was achieved. The I-40 engine was no help either, as it developed nearly 250 lb (113 kg) less thrust than advertised as well. The estimated performance of 478 mph (769 kmh) at sea level was not achieved with only a mere 400 mph (643 kmh) achieved. Due to these factors, the performance achieved was similar to that of the Packard engine installation. Despite these problems, the XP-81 still did well in some aspects. The relatively decent handling and decent climb rate was complemented, as was the light controls. The second prototype (serial no. 44-91001) was produced some time before November of 1946, and was ready for flights by February of 1947. It featured a longer ventral fin than that of the XP-81 and had a four blade Hamilton Hydromatic propeller replacing the Aeroproducts propeller used on the first prototype. Unfortunately, it is unknown what date the second prototype made its maiden flight, but it is speculated that it first flew some time in February of 1947.

In total, 116 flights were made by both of the XP-81 prototypes, 22 of which were done by the second XP-81 prototype. More tests were planned, as on January 14th of 1947, Consolidated Vultee called for the following areas to be studied and tested:

  1. Firearms testing of the Browning AN/M2 and the Hispano T31. Bombs and rockets tests will also be included.
  2. Anti-icing equipment efficiency.
  3. Control characteristics and lateral stability.
  4. Cabin pressurization experiments.
  5. Power plant operations.
The first XP-81 prototype taxiing on the Muroc airstip in preparation of a flight on January 22nd of 1946 (SDAM)

However due to the previously mentioned issues of the XP-81 underperforming, the USAAF gradually lost interest in the XP-81 program. Consolidated Vultee was well aware of this, and they had been trying since December of 1946 to improve their design. A proposal was made in December 31st to the Air Material Command to fix the underperforming prototypes. This proposal suggested that an improved TG-110 (the ones that would have been used on the YP-81) should replace the TG-100 and a J33-19 jet engine should replace the J33-GE-5. The Air Material Command however was not impressed by the proposal due to the amount of redesigning and time needed and in early 1947, their engineering department ceased work on the TG-100 turboprop engine. Things would look even more grim for the XP-81 when on January 27th of 1947, the contract for the 13 YP-81 pre-

The first XP-81 prototype flies over the Mojave Desert. (Convair)

production fighters were cancelled. Finally on May 9th, the XP-81 program reached its end when the government decided to cancel the contract on its development. The two prototypes were then taken in by the USAAF on June 24th and 25th. Finalization of the cancellation was conducted on June 23rd of 1948 after the USAAF was reorganized into the United States Air Force (USAF) when Consolidated Vultee was reimbursed with $4,578,231 for their work on the program.

Though development stopped for the XP-81 program, the two prototype’s story did not end there. At the time when the USAAF took in the prototypes, the engine and propeller development branches of the Air Material Command was in the middle of developing more advanced propeller control techniques and a suitable machine was needed to perform tests on as wind tunnels and models were not available. The USAAF promptly provided the two XP-81s which were redesignated as “ZXF-81” for this new role. The two planes were then stored in Edwards AFB (previously known as Muroc AAF) for future use. Unfortunately, they were never used and on April 29th of 1949, all useful parts and gadgets were stripped from the two planes by order of the USAF. The two empty airframes were then dragged onto the photography & bombing range of the Edwards AFB.

Despite the XP-81s now sitting in the desert, Consolidated Vultee was still not willing to yield completely. The company tried proposing reviving the XP-81 program using different power plants and repurposing the role. The proposal called for the use of the British Armstrong-Siddeley Double Mamba turboprop producing 4,000 hp and a Rolls-Royce R.B 41 jet engine producing 6,250 lbf (2,835 kgf) of thrust replacing the original engines. The idea behind this was to create a ground attack aircraft which could be exported to other countries. However, this idea was understandably met with skepticism by the Air Force, but an investigation to see the feasibility of this proposal was made. On September 14th of 1950, a report was finalized stating that at least ⅔ of the airframe would need to be modified in order to mount the new engines.  New drop tanks, rocket rails, hardpoints and various other parts would also need to be redesigned. Another investigation was done on this proposal by comparing the hypothetical performance to the all-turboprop Douglas A2D Skyshark, a ground attacker aircraft in service with the USAF. It was determined that the Skyshark would outperform the XP-81 with British engines in all aspects, so there was no point in developing an inferior aircraft. Another factor that was noted was the excessive amount of maintenance, training and logistics needed to service the ground attacker. With all these factors in mind, the proposal was discarded by the USAF and Consolidated Vultee finally gave up on the XP-81.

Frank Davis sits in the cockpit of the XP-81. (Convair)

The two XP-81 airframes would remain in the desert exposed to the elements for decades until August of 1994 when Air Force Flight Test Center Museum curator Doug Nelson retrieved them. They were in derelict condition, with the second XP-81 prototype being more damaged than the first. As of 2018, the two airframes remain in the National Museum of the United States Air Force in Dayton, Ohio awaiting future restoration. Although never seeing service, the XP-81 still holds a distinct spot in history as America’s first turboprop engine powered plane to fly and the world’s first plane to fly with a turboprop engine and a jet engine together.

The first XP-81 prototype at the Edwards AFB shortly before recovery. (AFFTCHO)

Design

Airframe:
The XP-81’s semi-monocoque fuselage was constructed using age hardened 24-SRT aluminum alloy, followed by the exterior surfaces being flush riveted. The entire fuselage is made from metal. The wing design was a NACA laminar flow type, made from aluminum-alloy. The design allowed for a stressed-skin wing which was flush riveted as well, with the rivet heads being milled. Due to the relatively heavy materials used in the wings, the surface was relatively smooth thus allowing for good aerodynamics. The majority of the heavy plating was mounted in the frontal 34.5% of the wings, and thus allowed a decent mount for aerial weapons and permitted ordinance to be mounted. There were spoilers present on each wing which automatically operated in accordance to the ailerons. Another interesting feature was a thermal anti-ice system derived from the hot hair emitted from the TG-100 turboprop and the exhaust. Within the fuselage two fuel tanks were installed directly behind the cockpit, making for a total 811 gallons (3670 L) of fuel. The fuselage also housed the XP-81’s tricycle landing gear which was electrically operated. The main gear was fitted with disc brakes, also doubling as a parking brake.

The first XP-81 prototype’s Packard engine installation being finalized at the Vultee plant. (Convair)

The canopy on the cockpit was based off of the British bubble design, which allowed for a relatively clean 360° view. This type of canopy was used on many planes in service with the United States and Britain. The canopy would be controlled by the pilot via a hand crank on the left hand side of the cockpit. For fatal combat situations, an emergency canopy jettison system was provided allowing for the pilot to bail out quickly. The pilot’s seat was an ordinary World War II styled seat, but this was eventually replaced with an ejection seat modelled after the one used on the Convair XP-54. As the XP-81 was a long range fighter, an automatic piloting system was also installed. The cockpit would also be pressurized using the air from the TG-100 engine. For pilot comfort, a temperature system was installed allowing for optimal temperatures in all climate and altitudes.

For communication, the XP-81 was fitted with a VHF (Very High Frequency) SCR 522-A radio set. The cockpit also had room for a BC-1206 beacon receiver and an SCR 695 identification friendly-or-foe system, but these were never installed. The pilot would operate the SCR 522-A radio from the right side of the cockpit, where the radio controls were based.

It is also interesting to note that the second YP-81 prototype had a longer ventral fin than the first prototype.

Powerplant:

The Packard-powered XP-81 prototype idle on the Muroc airstrip. (Convair)

The XP-81’s design called for a General Electric TG-100 (also known as XTG-31-GE-1) turboprop and General Electric/Allison J33-GE-5 (I-40) jet engine as its power plants. The first prototype had a four blade Aeroproducts A542 brand propeller driving the TG-100 while the second prototype had a Hamilton Standard Hydromatic 4260 propeller instead. The TG-100 had a capacity for 8 gallons (30 L) of oil while the I-40 had 3.5 gallons (13 L). In terms of fuel, 811 gallons (3,670 L) was available in the XP-81’s two standard fuel tanks in the fuselage, but could go up to 1,511 gallons (5,720 L) with the installation of drop tanks.

The TG-100 Turbo Prop Engine

Armament:
The standard armament envisioned for the production P-81 would consist of either six 12.7x99mm Browning AN/M2 machine guns with 400 rounds each or six 20x110mm Hispano T31 cannons with 200 rounds each. The loadout of these guns would be in groups of three in each wing. For ordinance, a single hard point was mounted under each wing, allowing the plane to carry two bombs size ranging from 100 lb (45 kg) to 1,600 lb (725 kg), allowing for a maximum of 3,200 lb (1,451 kg). Chemical tanks, drop tanks, depth charges could also be equipped. Alternatively, 14 High velocity Aircraft Rockets (HVAR) could be carried.

Variants

  • XP-81 – Prototype fighter variant powered by a TG-100 turboprop and I-40 jet engine. Two examples were produced and extensively tested up until the cancellation of the project. Both prototypes were redesignated as “ZXF-81” in 1948 and stored in Edwards AFB. They would be stripped of useful parts and towed to the photography/bombing range near Edwards AFB and left there in derelict condition until August of 1994 when they were retrieved by Doug Nelson. The two airframes still survive to this day and are currently awaiting restoration at the National Museum of the United States Air Force in Dayton, Ohio.
  • YP-81 – Planned batch of 13 pre-production fighter variant powered by a lighter but more powerful TG-110 turboprop engine and an uprated General Electric J33-GE-5 turbojet engine. These planes would have been armed with either the Browning AN/M2 machine guns or the Hispano T-31 cannons. It would have differed from the XP-81 by having the wings moved back 10 inches (2.54 cm). No YP-81s were produced.
  • ZXF-81 – Post development termination designation for the two XP-81 prototypes. This designation signified that the prototypes were now flying test beds. However, no use of the prototypes after its termination was noted.
  • XP-81 (British Engines) – Unofficial variant proposed by Consolidated Vultee some time in 1949/1950 calling for the revival of the XP-81 project using British Armstrong-Siddeley Double Mamba turboprop producing 4,000 hp and a Rolls-Royce R.B 41 jet engine producing 6,250 lb (2,835 kg) of thrust replacing original engines. This new variant would be used as a ground attacker that would be solely used for export. This proposal never saw any development and was thus discarded.

Operators

  • United States of America – The XP-81 was intended to be used by the USAAF, but development carried over to the USAF. The project was eventually cancelled.
Consolidated Vultee Aircraft Corporation XP-81
(Taken from “Consolidated Vultee XP-81 by Steve Ginter”)
Wingspan 50 ft 6 in / 15.39 m
Length 44 ft 8 in / 13.61 m
Height 13 ft 10 in / 4.21 m
Wing Area 425 ft² / 39.48 m²
Wing Loading 45.9 ft² /  4.26 m²
Tire Loading 80 in² / 516.12 cm²
Wings Sweep Back
Wing Dihedral
Root Cord 13-1
Engines 1x General Electric XTG-31-GE-1 (TG-100) turboprop
1x General Electric / Allison J33-GE-5 (I-40) jet
Oil Capacity TG-100: 8 gallons / 30 L
I-40: 3.5 gallons / 13 L
Empty Weight 12,755 lb / 3,887 kg
Normal Weight 19,500 lb / 8,845 kg
(Maximum internal fuel with reduced armaments)
Maximum Combat Weight 24,650 lb / 11,181 kg
Fuel Capacity 811 gallons / 3,070 L – Internal Fuel Tanks
1511 gallons / 5,720 L – Internal Fuel Tanks + Drop Tanks
350 gallons / 1,325 L – Individual Drop Tank
Center of Gravity Max Forward – 17%
Max Aft – 27%
Rate of Climb 0 to 5,000 ft / 0 to 1,524 m – 5,200 fpm / 39.31 mps
Time of Climb 30,000 ft / 9,144 m in 9.6 minutes
Speed 299 mph / 481 kmh at Sea Level
253 mph / 407 kmh at 15,000 ft / 4,572 m
224 mph / 360 kmh at 30,000 ft / 9,144 m
Maximum Speed 546 mph / 877 kmh
Diving tests were never finalized due to propeller and engine problems. Flight #90 on September 4th of 1946 achieved the highest speed as mentioned above.
Range Conditions under maximum combat weight

Ferry Range – 2,393 mi / 3,851 km
Speed at 247 mph / 397 kmh – 2,002 mi / 3,222 km
Speed at 274 mph / 441 kmh – 1,622 mi / 2,610 km

Service Ceiling 47,000 ft / 14,000 m
Crew 1x Pilot
Radio Equipment SCR 522-A VHF Radio
Armament 6x 12.7x99mm Browning AN/M2 (400 rpg, 2,400 total)
or
6x 20x119mm Hispano T31 (200 rpg, 1,200 total)
Never Fitted on Prototypes, Intended Armament
Gunsight 1x K-14 Gyro Gunsight
Ordinance 2x hardpoints capable of carrying 3,200 lb / 1,452 kg of either bombs, depth charges, chemical tanks or drop tanks
or
14x 5 inch / 12.7 cm High Velocity Aircraft Rockets (HVAR)

Gallery

Plane side view
Sideart of the XP-81 by Escodrion

 

XP-81 parked on a ramp.
Side view of the XP-81 on the ground.
Rear view of the XP-81 on the ground. Note the position of the exhaust pipe relative to the two fuselage mounted air intakes.
XP-81 in flight.
Instrument panel of the XP-81. (USAF)
Controls of the XP-81. (USAF)
The second XP-81 prototype at the Edwards AFB shortly before recovery. (AFFTCHO)
The second XP-81 prototype at the Edwards AFB shortly before recovery. The helicopter shadow visible is possibly an H-21 ‘Flying Banana’ (AFFTCHO)
The second XP-81 prototype preparing to be transported. (AFFTCHO)
The second XP-81 prototype preparing to be transported. (AFFTCHO)

Sources:
Ginter, S. (2006). Consolidated Vultee XP-81. Simi Valley: Steve Ginter.Jenkins, D. R., & Landis, T. R. (2008). Experimental and Prototype: U.S. Air Force Jet Fighters. North Branch: Specialty Press.Wegg, J. (1990). General Dynamics Aircraft: And their Predecessors. London: Putnam Aeronautical Books.Baugher, J. (1991). Convair XP-81. National Museum of the US Air Force (n.d.). Convair XP-81. Images: Side Profile Views by Escodrion – https://escodrion.deviantart.com

 

Beechcraft XA-38 Grizzly

usa flag USA (1944)
Prototype Attack Plane – 2 Built

Colorized photos by Michael J.

The Beechcraft XA-38 Grizzly was an experimental attack aircraft stemming from a USAAF requirement for a two seated attack bomber. Two prototypes were constructed between 1944 and 1945, and saw extensive testing within the US. The Grizzly showed promising performance, but was ultimately cancelled due to the engines intended for use was given priority to the Boeing B-29 Superfortress and the inevitable victory of the Allies.

History

Bottom view of Beechcraft XA-38 (S/N 43-14407) in flight. (U.S. Air Force photo)

In 1942, the United States Army Air Force (USAAF) issued a requirement for a two seater attack bomber. Beechcraft was quick to respond, and proposed their design to the USAAF. The USAAF was very interested in the design, and ordered two prototypes to be constructed in December of the same year after granting the contract to Beechcraft. In anticipation of the two prototypes, the USAAF assigned serial numbers to them, being “43-14406” and “43-14407”.

 

Beechcraft specifically designed the Model 28 to be able to destroy gun emplacements, ships, armored vehicles and bunkers while keeping great maneuverability and able to remain airborne after being damaged. All of this would be done by the addition of a powerful 75mm T15E1. The task of developing the Grizzly was given to a team led by Bill Cassidy with Jess Vint and Alex Odevseff in charge of designing the armaments, Bill Irig in charge of the control surfaces, Gus Ericson in charge of the design of wings, Mervin Meyers in charge of hydraulics, Ralph Harmon in charge of the landing gear structure and Noel Naidenoff in charge of the engine compartment. The Grizzly is common thought to be a modified Model 18 design, but this is untrue. The Grizzly did take inspirations from the Model 18, though.

Beechcraft XA-38 during ground vibration tests. Tests were set-up to determine natural frequencies excited during engine operation. (U.S. Air Force photo)

The first Grizzly (43-14406) was delivered to the Army Air Force and flown on May 7th of 1944 by test pilot Vern L. Carstens. The first test flight went relatively well except for an unplanned touch-and-go during landing. This was due to Carsten’s inexperience with landing such a large plane. The first prototype had no armaments installed, but had a wooden mockup of the 75mm T15E1 cannon. In the next few test flights, the Grizzly proved itself to be very aerodynamically stable, and made a good impression with the designers. A memorable flight test includes a performance comparison between the Grizzly and a recently manufactured North American P-51B. The Grizzly and P-51B were put in a mock pursuit, and the P-51B was reported to have been unable to keep up. Afterwards on July 7th of 1945, the first Grizzly was transferred to Wright Field to be used by the USAAF. It then participated in 38 test flights from between October 13 and October 24 of 1944 flown by Captain Jack W. Williams. Williams affectionately noted that the Grizzly was a great aircraft and “very maneuverable” for an aircraft of its size. It is also interesting to note that the turrets on the first Grizzly were dummies.

Front view of Beechcraft XA-38 (S/N 43-14407) in flight. (U.S. Air Force photo)

The second Grizzly (43-14407) had its maiden flight on September 22nd of 1945, once again piloted by Carstens. The second prototype was intended for armament testing, so it had all weapons fitted. The 75mm T15E1 prototype cannon was already successfully tested the year before on July 1st. The second Grizzly flew a total of 38 hours afterwards in tests at Eglin Field in Florida.

As successful as the Grizzly was, it would never reach mass production status because of two reasons. The first reason was that the R-3350 engines were in short supply, as the Boeing B-29 Superfortress had top priority for the engines. Second, the war situation was already in America’s favor, thus cancelling out the need for such an aircraft. As a result, both of the Grizzlies were retired from active service. One was scrapped while the other one was transferred to the Davis-Monthan Airfield in Arizona, meeting an unknown fate.

Design

The XA-38 Grizzly was an all metal, two seat, twin engined semi-monocoque plane with cantilever wings, conventional landing gears, oleo-pneumatic shock absorbers and hydraulic brakes. It was powered by two Wright R-3350-43 Duplex Cyclone engines, the same engine that powered the Boeing B-29 Superfortress. The propellers measured at 4.32m (170in) each in diameter.

The plane used flush riveting and butted skin joints to give the plane its pristine, shiny look. The foil used to construct the wings were derived from the NACA 23000 series which was good for high and low speeds. The engine hub was made from stainless steel and aluminum alloy. The oil coolers were placed in the wings. Four fuel tanks were installed in the wings with a capacity of 2,422 litres (640 US Gallons). Two self-sealing fuel tanks were also placed behind the cockpit which can carry 681 litres (185 US Gallons) if needed to. There were pumps and connectors installed onto the tanks, which would stop fuel flow if the tanks took damage.

Beechcraft XA-38 (S/N 43-14407, the second and last XA-38 built). (U.S. Air Force photo)

As for armaments, the XA-38’s 75mm T15E1 was the defining feature. Could carry 20 rounds fed by a Type T-13 feeding system. There would be a Type N-6 reflector sight to help the pilot aim. The cannon would fire every 1.2 seconds if the pilot pressed the trigger button. Two .50cal (12.7mm) M2 Browning machine guns were installed under the cannon with 500 rounds each. The entire nose section of the XA-38 could be unhinged, where mechanics can easily access the guns for maintenance.

The Grizzly had two turrets, one located on the top of the fuselage and one on the bottom. The turrets were manufactured by General Electric and had two M2 Brownings each with 500 rounds. These two turrets were controlled by a single gunner seated in the rear fuselage. He would aim the guns with a periscope to control the turrets. Interestingly enough, the turrets can also be fixed to fire forward to accompany the T15E1. Ordinance wise, the XA-38 was designed to carry a wide range of things. It could have carried bombs, napalm, torpedoes, fuel tanks, smoke tanks and depth charges.

Operators

  • United States of America – The USAAF was the sole operator of the XA-38. The USAAF extensively tested the XA-38, and concluded it was a success. However, the XA-38 never reached mass production status.

Beechcraft XA-38 Grizzly

Wingspan 64 ft 4 in / 20.52 m
Length 51 ft 9 in / 15.77 m
Height 15 ft 6 in / 4.72 m
Wing Area 626 ft² / 58.15 m²
Engine 2x Wright R-3350-43 Duplex Cyclone (2,300hp)
Fuel Capacity 640 US Gallons / 2,422 L

Additional Tanks: 185 US Gallons / 681 L

Maximum Takeoff Weight 35,265 lbs / 15,996 kg
Empty Weight 22,480lbs / 10,197 kg
Wing Loading 56.3 lb/ft² / 275.1 kg/m²
Power Loading 7.67 lb/hp / 3.48 kg/hp
Climb Rate 2,600 ft/min / 792 m/min
Maximum Speed 330 mph / 531 km/h at Sea Level

348 mph / 560 km/h – at 5,000 ft / 1,525 m

370 mph / 595 km/h – at 17,000ft / 5,180m

Cruising Speed (75% Throttle)

289 mph / 465 km/h – at Sea Level

350 mph / 563 km/h – at 16,000 ft / 4,875 m

Range 1,625 mi / 2,615 km – at 225 mph / 362km/h

745 mi / 1,200 km – at 289 mph / 465 km/h

Service Ceiling 29,000 ft / 8,840 m
Crew 1x Pilot

1x Gunner/Observer

Forward Firing Armaments 1x 75mm T15E1 (20 rounds)

2x 12.7mm M2 Browning (500 rpg)

Defensive Armaments 2x 12.7mm M2 Browning (Upper Turret) (500 rpg)

2x 12.7mm M2 Browning (Belly Turret) (500 rpg)

Ordinance Planned Options:

  • Bombs
  • Fuel tanks
  • Napalm
  • Torpedoes
  • Depth Charges
  • Smoke Tanks

Gallery

Grizzly Prototype 314407 – armed with two 500 lb bombs

Sources

Beechcraft XA-38. (2008). National Museum of the US Air ForcePearce, W. (2013). Beech Aircraft Company XA-38 Grizzly. Old Machine Press.Pelletier, Alain J. (2005). Beech aircraft and their predecessors. Putnam, Images:  Side Profile Views by Ed Jackson – Artbyedo.comColorized Images by Michael J.

Douglas XB-19

usa flag USA (1941)
Prototype Heavy Bomber – 1 Built

The XB-19 parked on the ground.

The XB-19 was a heavy bomber designed in 1935 to fulfill a request made by the United States Army Air Corps (USAAC) to develop an experimental heavy bomber with extreme range. Although slow in its development and obsolete by the time it was produced, it served as a test vehicle to evaluate plane and engine performances. The sole XB-19 was converted to a cargo transport plane and was eventually scrapped in 1949. The XB-19 was the largest plane operated by the USAAC and USAAF until the Convair B-36 came into service.

History

The roots of the XB-19 can be traced to 1935 on February 5th when the United States Army Air Corps (USAAC) commenced “Project D”. The purpose of Project D was to experiment with the maximum distances achievable with bombers. The USAAC contacted and discussed the project with Douglas Aircraft Company and Sikorsky. Douglas representatives agreed to the terms of the design and plans were made during a conference on June 5th, 1935. The initial plan was to begin the basic design on July 31st of 1935, detailed designs on January 31st of 1936, and have the plane physically produced by March 31st, 1938. The plan however was soon found out to be too ambitious, with the designers underestimating the work required. The designers would be plagued with a lack of proper funding and the sheer enormity of the task. The project would finally be completed in May of 1941, nearly four years after the original deadline.

Douglas XB-19 under construction. (U.S. Air Force photo)

Douglas Aircraft Company received a contract to the project in October of 1935 which required Douglas to create a general and detailed design of the plane, create a mockup of the plane and test the wing centre section, undercarriage, and engine nacelles of the plane. Douglas accepted the contract on October 18th. Later that year, the USAAC would evaluate the mockups provided by Douglas and Sikorsky. Douglas’s design was ultimately chosen, and was given the task of further developing the plane.

The XB-19 under construction at the Douglas Aircraft Factory located in Santa Monica, California. 1940.

The plane would be known as the “XBLR-2” (Experimental Bomber Long Range 2) in its early stages of development. The progress of developing the bomber proved to be tedious and slow. Lack of funding would severely hinder work on the plane. During that time the USAAC made a change to the requirements, the plane was suppose to be powered by four Allison XV-3420-1 engines producing 1,600 horsepower each, but was ordered to be replaced by four Wright R-3350 engines producing 2,000 horsepower each instead. This would also hinder work as the plane had to be slightly redesigned. As time went on, Douglas had to loan a Douglas OA-4A from the USAAC to test an experimental tricycle landing gear configuration intended for the XBLR-2. The tests proved to be a success. Later, the XBLR-2 would be redesignated as “XB-19” (Experimental Bomber 19). Douglas eventually managed to scrape together enough funds to produce a prototype, and the production was authorized on March 8th of 1938.

XB-19_38-471_at_Mines_Airfield_Colorized copy
The XB-19 parked at Mines Airfield. (Colorized by Michael J.)
Washing XB-19 at March Field 1941. (Colorized by Michael J.)

During its development, the Douglas company had many problems with the XB-19. They were forced to allocate more funds than initially expected, and needed design staff to work on other aircraft which had a more promising production future. They claimed the XB-19’s design was obsolete due to the production delays it suffered over the past three years and the fact that the plane’s weight was far heavier than expected. The Douglas company officially made a recommendation to cancel the XB-19 project on August 30th, 1938. This recommendation was denied by the USAAC. Interestingly enough, two years later, the USAAC would suggest that the slow development of the XB-19 already rendered the project obsolete when they removed the plane from the top secret classified list. The XB-19 would finally be completed in May of 1941.

The XB-19 parked on the ground next to a P-40 Kittyhawk.

Shortly after completion, the XB-19 was used in taxiing tests on May 6th, 1941. The flight test was scheduled to be on May 17th, but was postponed three times due to critical mechanical errors. The landing gear brakes were found to have defects, its engines had backfiring issues, and the propeller pitch control system had to be worked on. On June 27th however, the XB-19 would finally have its maiden flight. In the maiden flight, seven crewmembers were on board with Major Stanley M. Umstead in charge. The flight lasted 55 minutes from Clover Field in Santa Monica to March Field. The flight went by smoothly without any problems and was successful. Shortly afterwards, Donald Douglas would receive a congratulatory telegram from President Roosevelt. The USAAC unofficially accepted the XB-19 in October of 1941.

Eager observers watch the XB-19 preparing for its maiden flight. Clover Field, 1941.

After the Japanese attack on Pearl Harbour on December 7th of 1941, the United States was on high alert. The XB-19’s turrets were armed and a new layer of olive camouflage paint was applied, replacing its bare metal USAAC livery. It would make 4 more tests flights in California before being transferred to Wright Field on January 23rd, 1942 as another safety measure. By then, the XB-19 had over 70 hours of flight time.

The XB-19 was finally accepted officially by the USAAF in June of 1942 after minor modifications were made to the plane’s brake system. The contract cost to the United States government was $1,400,064. The Douglas Aircraft Company also spent $4,000,000 in personal company funds. The XB-19 was extensively tested by the USAAF for eighteen months to see the engine performances and different altitudes and the maneuverability of the aircraft.  The results of these tests would later go on to influence the design of the Boeing B-29 Superfortress and the Convair B-36. The XB-19 performed well in all aspects and was generally free of problems. The only problem noted however was the inefficient engine cooling process. Due to this, the cooling gills on the plane had to be open the whole time in longer flights, thus reducing the effective speed of the XB-19.

The XB-19 in flight over Santa Monica with an AT-6 following it.

After the XB-19 was thoroughly tested and experimented with, the USAAF no longer had a need for it. It was brought to the Wright Field and modified to be a cargo transport aircraft. It was refitted with Allison V-3420-11 engines and had its armaments removed. The new aircraft would be designated “XB-19A”. For the next two and a half years, the XB-19A would fly to numerous airfields within Ohio. It was documented to have been stationed at Wright Field, Patterson Field, Lockbourne Air Base, and Clinton Country Air Base. The XB-19A would make its last flight on August 17th, 1946, where it flew to the Davis-Monthan airfield in Arizona from Wright Field to be stored. It stayed in storage for three years before finally being scrapped in 1949, thus ending the legacy of the XB-19.

XB-19_in_flight_1942_Colorized copy
The XB-19 in flight some time in 1942. (Colorized by Michael J.)

To this day, only two wheels of the XB-19’s landing gear survives. One can be seen in the Hill Aerospace Museum in Oregon, and the other can be seen in the National Museum of the United States Air Force in Ohio.

The wheel of the XB-19 with a car and person for comparison.

Design

The XB-19 is described as a colossal, all metal low wing monoplane installed with a conventional tricycle landing gear. The two main wheels of the landing gears measured at 2.44 m (8 ft) in diameter, which was impressive for the time. The original design specifications ordered wanted the engines to be four Allison XV-3420-1, but was swapped for four Wright R-3350-5 engines instead with a three blade metal propeller with a 5.18 m (17 ft) diameter. The engines would be switched once again to Allison V-3420-11 after the plane was repurposed as a cargo transport aircraft. The plane could carry an impressive amount of fuel, at 38,178 L (10,350 US Gallons) in its auxiliary fuel tanks, with an optional 3,210 L (824 US Gallons) that could be stored in the bomb bay.

A shot of the underside of the XB-19 with the gear down.

The XB-19 carried 8,480 kg (18,700 lbs) of ordinance usually, but could be overloaded to 16,828 kg (37,100 lbs) if fuel was reduced significantly. As for armaments, the initial prototype was unarmed. Later though, two 37mm Oldsmobile T9 autocannons, five 12.7mm M2 Brownings and six M1919 Brownings were fitted to the plane. One T9 was fitted to the nose while the other was fitted to the upper front turret, each accompanied by a single M1919 machine gun. There would be one M1919 on each side of the bombardier’s position, and a M1919 on each side of the stabilizer. A single M2 Browning was fitted in the tail of the XB-19, two M2 Brownings on each side of the galley compartment, one in the bottom turret, and one in the upper powered turret.

In the crew compartment, there was eight seats and six bunks. The compartment could accommodate two flight engineers, and six relief crew members. The normal combat crew consisted of sixteen people. (Refer to Specifications Table).

The cockpit of the XB-19.

Variants

  • XB-19 – The original model and design. Initially developed as a long range heavy bomber for the USAAC, but was outdated by the time it entered service. It served as a “flying laboratory”, testing engine performances and plane handling. It was converted to the XB-19A after the USAAF no longer had use for it.
  • XB-19A – The XB-19A was a converted XB-19 using improved Allison V-3420-11 engines. It was used as a cargo transport aircraft after the air force was done experimenting with it. All armaments were removed. It was scrapped in 1949.

Operators

  • United States of America – The XB-19 and XB-19A was operated by the USAAC and USAAF throughout its service life.

 

Douglas XB-19

Wingspan 212 ft / 64.62 m
Length 132 ft & 4 in / 40.34 m
Height 42 ft / 12.8 m
Wing Area 4,285 ft² / 398.091m²
Wing Loading 32.6 lb/sq ft / 159.5 kg/sq m
Power Loading 17.5 lb/hp / 7.9 kg/hp
Engine 4x Wright R-3350-5 Duplex Cyclone (2,000 hp)
Fuel Capacity 10,350 US Gallons / 38,178 L – in auxiliary fuel tanks + 824 US Gallons / 3,120 L – in bombay (Optional)
Maximum Weight 140,000 lbs / 63,503 kg
Empty Weight 86,000 lbs / 39,009 kg
Climb Rate 650 ft/min / 198 m/min
Speeds Cruising: 135 mph / 217 km/h – Sea Level

Operational: 186 mph / 299 km/h – @ 15,700 ft / 4,785 m

Maximum Speed: 224 mph / 360 km/h – @ 15,700 ft / 4,785 m

Normal Range 5,200 mi / 8,369 km
Maximum Range 7,710 mi / 12,408 km
Service Ceiling 23,000 ft / 7,010 m
Crew 2x Pilots

1x Commander

1x Navigator

1x Engineer

1x Radio Operator

1x Bombardier

2x Flight Mechanics

1x Turret Operator

8x Gunners

6x Relief Crew

(24 Crew – 16 Active, 2 Emergency Stations, 6 Relief Crew)

Defensive Armament 2x 37mm Oldsmobile T9 Autocannon

5x 12.7mm M2 Browning

6x 7.62mm M1919 Browning

Normal Ordinance 18,700 lbs / 8,480 kg
Maximum Ordinance 37,100 lbs / 16,828 kg *

* – with reduced fuel load

 

Douglas XB-19A

Wingspan 212 ft  / 64.62 m
Length 132 ft 4 in / 40.34 m
Height 42 ft / 12.8 m
Wing Area 4,285 ft² / 398.091m²
Wing Loading 32.71 lb/sq ft / 159.8 kg/sq m
Power Loading 13.51 lb/hp / 6.1 kg/hp
Engine 4x Allison V-3420-11 (2,600 hp)
Loaded Weight 140,230 lbs / 63,607 kg
Empty Weight 92,400 lbs / 41,912 kg
Maximum Speed 265 mph / 426 km/h
Cruising Speed 185 mph / 298 km/h
Normal Range 4,200 mi / 6,759 km
Service Ceiling 39,000 ft / 11,885 m

Gallery

A spectacular shot of the XB-19 flying low. 1942.
XB-19A on the ground with Allison V-3420-11 engines.
Crewmen washing the XB-19 at March Field, some time in 1941.
The crew of the XB-19 operating in the cockpit.
XB-19A on the ground with Allison V-3420-11 engines.
XB-19 Before Scrapped
A photo of the XB-19 post-war before it was scrapped.

 

Sources

Bunker, Howard G. Development, Test and Acceptance of Douglas XB-19 Airplane, AAF No. 38-471. 1942, pp. 18–25, Development, Test and Acceptance of Douglas XB-19 Airplane, AAF No. 38-471., Francillon, René J. McDonnell Douglas aircraft since 1920. Putnam, 1988., Images:  Side Profile Views by Ed Jackson – Artbyedo.com, Colorized Images by Michael of PE

 

Vought F4U Corsair

sweden flag USA (1942)
Naval Fighter Plane – 12,571 Built

black-and-white-f4u

The F4U Corsair is another most famous fighter and fighter-bomber of WWII, although it saw action mostly against the Japanese in the Theatre of the Pacific, therefore being primarily used by the US Navy and the Marines. This airplane in particular was specifically designed for aircraft carriers, being a naval aircraft in essence, although initial doubts over its performance on-board an aircraft carriers made it to serve initially as a land-based asset. It saw also action during the Korea War as a ground attack and Close Air Support (CAS) aircraft, and with the French in the Indochina, Algeria, and Suez Canal crisis. It also saw some service in the Atlantic during WWII, mainly with the British Fleet Air Arm, where reportedly provided air cover to the airplanes attacking the battleship Tirpitz, and served in the Indian and Pacific Oceans. The Corsair contributed to change the balance over the skies of the Pacific by shooting down many Mitsubishi A6M Zeros, although not as much as the Grumman F6F Hellcat.

The Corsair is single-seat and single engine fighter/fighter-bomber for day and night-time, featuring a characteristic inverted gull wing (Similar to that of the Junkers Ju-87 Stuka and the Loire-Nieuport 40) and a very long propeller-blade. The development of the Corsair began following a request by the US Navy for twin and single-engine fighters in 1938, with the single-engine required to obtain the maximum speed possible and a stalling speed of no more than 110 km/h (70 mph), and a long range. Interestingly, the initial requirements comprised the aircraft to carry anti-aircraft bombs to be dropped on enemy formations. That same year, Vought – the builder company – was awarded a contract to start with the development of the Corsair.

The Corsair was a pretty advanced aircraft for the times, and this characteristic meant that its development would find several problems that required solution, which in turn, were quite remarkable. Even so, the Corsair required improvements while in service, which does not deny the fact that it was one of the greatest and unique airplanes of the war, let alone a good complement to other aircraft carrier-based fighters and among the best naval fighters in the war.

One of the main features during development was the incorporation of the largest engine available, the Pratt & Whitney R-2800 V-18 Double Wasp of 2250 hp, requiring the installation of a wide three-blade Hamilton propeller. This installation had two visible effects on the design: First, the characteristic shape of the airframe, where the bow is basically the area where the big and long engine is located, almost displacing the cockpit further aft. Second, it yielded speeds of up to 652 km/h (405 mph), making it the first single-engine American design to reach such speed. But the first problems emerged, especially in regards to diving speed that, although achievable, meant considerable damage to control surfaces and access panels, as well as problems with the engine. Spin recovery standards also needed to be revisited.

The wing itself, along with the longitudinal shape, were both a challenge when designing the frame. In regards to the inverted gull wing, it was purposed to make the width and the landing gear as short as possible, benefiting also the minimization of drag, as the anhedral of the center section gave an optimal meeting angle between the wing and the fuselage. Yet the weight of the wing alone neutralized those effects. But it also had the problems when recovering from developed spins, as the shape of the wing interfered with the elevator. It also had problems with the starboard strip, that used drop without warning, requiring the installation of small stall strips on the leading edges. The port wing also had the potential of stalling and dropping in failed landings, which was further dangerous if throttle was abruptly increased in such cases. The inverted gull wing was also a product of solving the problem of the landing gear, as they needed to be tall enough to keep the propeller away from the ground (the same problem the Saab J-21 had). It simply shortened the length of the legs, while the landing gear was able to retract and rote 90° into an enclosed wheel well, maintaining the streamline of the wings.

The Corsair, however, was benefited during its development thanks to the experiences of other air forces when the war sparked in Europe. As a result, the set of 2 X 7.62mm synchronized engine cowling-mount machine guns, and the 2 X 12.7mm wing machine guns was deemed unsuitable, prompting the armament scheme to be modified. Three 12.7mm machine guns were fitted on each wing, increasing the firepower of the Corsair.

redbull-f4u-corsair-and-b25-mitchell
A Corsair and Mitchell bomber, fly together at an airshow.

As it was abovementioned, other problems prevented the Corsair to serve as a carrier-based fighter until 1944, mainly those related to the type of landing required in that type of vessel. Not only the wing-related problems when performing this manoeuvre, but also the location of the cockpit plus the long bow made landings particularly dangerous for new pilots. Furthermore, during landing approaches manoeuvres, the oil from the hydraulic cowl flaps had the tendency to spatter onto the windscreen, compromising visibility, and the oleo struts had bad rebound when landing, making the entire aircraft to bounce upon landing. The top cowl flap down was sealed, while a valve was fitted to the landing gear legs in order to solve the issues, solution that were, on the other hand, implemented by the British firstly. It had its first flight in 1940, entering in service in December 1942 intended as a naval fighter, but these problems delayed its utilization as carrier-borne fighter and the US Navy initially preferring the F6F Hellcat, but it also meant that the Marines would use the Corsair as their main air assets, and it was with this branch that the Corsair began to carve its reputation. It entered in service in the late 1942, where the Marines began to make use of it at the Battle of Guadalcanal and the Solomon Islands, where its first debut was rather disappointing. But once the Marines learned how to maximize the advantages of the Corsair, they began to contest the air supremacy the Japanese had. It also saw extensive action as a fighter-bomber/attacker in the Marshall Islands, Palaus, Iwo Jima and Okinawa.

It was the British the ones that solved the operational problems of the Corsair for naval use, as they began to operate with the Corsair in 1944, on-board the HMS Victorious. Those Corsairs saw action as carrier-borne aircraft by supressing Flaks and providing escorts to aircraft performing raids against the Kriegsmarine battleship DKM Tirpitz in three raid operations: Operation Tungsten, Operation Mascot and Operation Goodwood. Later on the British Corsairs were deployed in the Indian Ocean and the Pacific, attacking Japanese targets on April 1944.

The Corsair saw action in post-WWII conflicts such as the Korean War, the Indochina War, among others. Many served with other air forces as surplus or donated aircraft, where it served more than 30 years after WWII was over, when it scored its last air victories and gave an honourable closure to an era past gone. 10 F2G ‘Super’ Corsair series also served as civilian racers after the war.  A total of 12571 Corsairs were built, being in service with the US Navy, the Marines and other air forces from 1942 to 1979, attesting the good quality of the aircraft and its endurance, being produced until 1953. A total of 15,386 Mustangs were built.

Design

The Corsair is a low inverted gull wing fighter, with a single tail and a single engine: Pratt & Whitney R-2800 V-18 Double Wasp of 2250 hp, with a wide propeller fitted as to maximize the power yield. As a result of the size of the engine, the bow or nose of the Corsair is particularly long, which made the cockpit to be located further aft. The relocation and reconfiguration of the armament – which was placed at the wings – and the resulting relocation of the fuel tank in front of the cockpit contributed to its location in the airframe, which in turn had to be elongated.

The wings with their characteristic shape were the result of the need for shortening the legs of the landing gear and for accommodating also a folding wing, while being located also well ahead the pilots’ cockpit, making the Corsair to have a cross shape. This wing design also resulted in the Corsair having remarkable aerodynamics over similar airplanes of its type. The shape of the wing was also beneficial in the sense that the meeting angle between the wing as the fuselage reduced drag and saved the utilization of wing root fairings, although the bent wing tended to neutralize such benefits given its weight. On a similar way to the Saab J 21, the supercharger air intakes, alongside the oil coolers, were placed at the wings, this case on the anhedraled center section of the wings. The combination of the propeller diameter, the engine and the wing’s shape and length – alongside the resulting aerodynamics – made the Corsair the fastest naval aircraft the US had at its disposal. The flaps were changed to a NACA slotted type while the ailerons were increased in span.

The fuselage, mainly the large panels, were made of aluminium and attached to the frames by spot welding, which eliminated the use of rivets. The top and the bottom areas of the outer wings were made out of fabric, as well as the ailerons, the elevators – which were also made of plywood – and the rudder. At the rear an IFF (Identification Friend or Foe) transponder device was installed.

The landing gear consisted typically of two ‘legs’ at the wings and a rear small wheel, with the carrier-based version having also a tail hook for the arresting cable. All of the set was retractable, only that the ‘legs’ at the wings rotated 90° and then swivelled backwards, a trait that common among many US fighters. Noteworthy to remark that the landing gear was hydraulically operated, alongside the cooling flaps, the wing flaps, the wing folding and locking, the arresting gear, the gun charging, and the dive breaks.

The aft cockpit had some interesting features and modifications resulting from the assessed hazards while landing on an aircraft carrier. As this problem was the result of the nose and the location of the same cockpit, a rectangular plexiglass panel was fitted in the lower center section, so to allow the pilot to see below and perform carrier landings with more safety. In addition, armour plates were applied to the canopy area, with the windscreen being a 38mm bullet-proof installed internally and the behind the curved windscreen. To aid the pilot’s rear view, half-elliptical planform transparent panels were placed at each side of the structure right behind the cockpit, yet the view provided was rather limited.

The aft section of the Corsair is also full of noticeable characteristics, with a projecting fuselage tip where the vertical stabilizer is placed, which is large. The horizontal stabilizer is, in turn, placed ‘aft’ of the tail.

The Corsair’s armament was originally a set of two 7.62mm machine guns at the frontal section of the nose, and two 12.7mm machine guns, one at each wing. But as the abovementioned reports from the war in Europe obliged the armament to be modified, the final disposition was of 6 X 12,7mm machineguns at the wings, three on each side.

Death has bent wings.

F4U Climb

The Corsair was the most effective fighter the US Navy and the USMC had from the moment it was introduced and entered combat in the Solomon Islands in 1943. It was appraised by the pilots due to its performance and its capacity to remove the threat posed by the Mitsubishis A6M Zeros, as well as to break Japanese bombing raids. It was also capable of outfling and outfighting any land-based aircraft. It was capable of performing interception, bombing, ground-attack and fighter missions. The Corsair was a fighter that was also an ace-maker, with Kenneth Walsh (21 kills), Gregory “Pappy” Boyington (28 kills) and Joe Foss (26 kills). It was under Boyington lead that his squadron, the “Black Sheep” were the most effective squadron, scoring 97 kills and 103 damaged airplanes on the ground. Noteworthy to remark, the Corsair was also appraised by Admiral Nimitz giving its performance.

As the Corsair was cleared for carrier use, it began to operate on-board USS Essex and USS Bunker Hill. The Corsair also performed dive bombing missions in the Marshal Islands as it dropped more than 90718 kg (200000 lbs) of bombs against Japanese installations. It also took part in combats at China Sea, Okinawa, Iwo Jima, Formosa and the Philippines. It also took part on the Saigon and Tokyo Raids, which were diversionary attacks prior to Okinawa. It was also during Okinawa where they had to operate as fleet air defence against the Kamikaze attacks in the earlier stages of the battle, performing CAS with bombs, rockets and Napalm once the threat was neutralized. They reportedly achieved remarkable feats, like keep flying after ramming an enemy. The Corsairs scored 2140 Japanese airplanes with only 189 Corsairs lost, along with 14 warships and 33 merchants sunk (Saigon raid). These scores earned the Corsair Nimitz’s appraisal and a US government citation, and the builder granted an “E” after the War.

The Corsair was among the few WWII-era aircraft to serve right into the earlier days of the Cold War, as it took part in low altitude attack fighter-bombing and CAS missions in Korea, as well as heckling the enemy in night missions. It also attacked enemy installations It dropped bombs, Napalms, rockets and cannons the same way as in WWII, being both aircraft and pilots both veterans of that conflict, and operating from WWII aircraft carriers (USS Essex and USS Bon Homme Richard). As tough as it was, it was able to cary alarge payload and remain more time in the combat zone for CAS missions, and even the Corsair even managed to kill a North Korean Mig-15. The Corsair also had a high rate of availability and hard resistance against enemy fire.

One last dogfight over the jungle

F4U in flight

When the 1969 ‘Soccer War’ sparked between Honduras and El Salvador, both nations were having among their air forces inventories some WWII-era fighters, namely F4U/FG-1 and P-51D/TF—51 fighters. These airplanes were to perform the last dogfight between WWII-era (or piston-propelled engine) airplanes, like two medieval knights clad in armour, ready to joust for a last opportunity as to write the last chapter of an era. The morning of the 17th of July, 1969, the encounter was bound to take place. As Honduran Captain Fernando Soto was leading a group of three F4U-5 to strafing missions at the border, one of the Corsairs was attacked by two Salvadorian P-51, with Capt. Soto shooting it down. But there was to be a second encounter between the veteran aircraft, as late on the same day, during a bombing mission alongside another F4U-5, they encountered Salvadorian FG-1. The result was that both FG-1 were shot down, making of Capt. Soto the only “Ace” of the War.

P-51 of the Salvadorian Air Force, piloted by US mercenaries, patrolled the Salvadorian skies and border, looking also for the Honduras Corsairs, with no avail.

Variants

  • F4U-1 (Corsair Mk I)/FG-1 – This was the first production series of the Corsair, being characterized by a ‘bird cage’ canopy and a low seating position, featuring also the definitive abovementioned modifications for the series-production models, including the 6 X 12,7mm machine guns’ configuration. An additional pair of auxiliary fuel tanks were installed in each wing edge A two-seat trainer was built but was not accepted by the US Navy. The Corsairs in service with the Marine Corps did not had folding wing capacity neither they were fitted with an arrester hook but a pneumatic tail wheel, as they were land-based, receiving the designation FG-1 and being built by the Goodyear. Those with the British Fleet Air Arm were denominated Corsair Mk I.
  • F4U-1A (Corsair Mk II) – A post-war denomination introduced to differentiate the mid-to-late production batch. This version – which would be the second production version – would have a new type of canopy, similar to a Malcolm hood type – like that of the Spitfire – and with only two frames. It had a simplified windscreen, which improved visibility overall along with the canopy being taller. That the pilot’s seat was raised 180mm (7 in), in combination with a lengthened tailwheel strut, meant that visibility was also improved, solving the problems posed by the long nose. This is the version that, along canopy modifications, also introduced wing and undercarriage oleo struts modifications, becoming in the US Navy carrier-based version. This version also received a new power plant, the R-2800-8W water-injection engine, and the capacity to carry a center-section fuel drop tank. Goodyear also built a variant of this version, land-based and without folding wing capacities. Those in service with the British had their wings modified – shortened by 2cms/8 in – for use in their carriers, denominated FG-1A.
  • F3A-1 (Corsair Mk III) – Denomination for those built by the Brewster, which none of them reached front-line units as the building both production and quality control were poor, noticeable after having speed restrictions and broken wings (due to poor quality wing fittings).
  • F4U-1B – Unofficial post-war denomination to identify Corsairs modified for Fleet Air Arm use.
  • F4U-1C – Ground attack and fighter version, with the 6 X 12,7mm guns replaced by a set of 4 X 20mm AN/M2 (Hispano-Suiza) cannons thus providing considerable firepower for ground attack missions. Based on the F-4U-1. This version had a remarkable performance in the Battle of Okinawa, as it was introduced in 1945.
  • F4U-1D/FG-1D/F3A-1D (Corsair Mk IV and Mk III) – Ground attack and fighter version, developed and built in parallel to the F4U-1C. It had the new engine fitted in the F4U-1A, yielding speeds of up to 684 km/h (417 mph). It also carried an increased payload of rockets and a twin-rack plumbing for an additional belly drop fuel tank, which increased firepower but also drag. The range was also increased, meaning it could perform long missions. A single piece – Malcolm hood type – canopy was adopted firstly as a standard for this version, then for the following Corsairs. Goodyear and Brewster also produced this version, under denominations FG-1D and F3A-1D, respectively.
  • F4U-1P – Photo-reconnaissance version.
  • XF4U-2 – Nigh-time fighter version fitted with two auxiliary fuel tanks.
  • F4U-2 – Experimental carrier-based night-time fighter. Armed with 5 X 12,7mm guns, with the starboard gun being replaced by an Airborne Intercept radome containing a radar. 32 were modified by Naval Aircraft Factory, ant two more were modified in the front-line. It saw action in the Solomon Islands and in Tarawa.
  • XF4U-3 – Experimental version used to test different engines that never entered into combat. Goodyear also produced some units of this version, denominated FG-3. A single XF4U-3B was produced with some modification, intended to be issued to the British Fleet Air Arm.
  • XF4U-4 – Version with new engine and cowling.
  • F4U-4 – A naval fighter/fighter bomber version, being the last one taking part in WWII, as it was introduced by late 1944. It was powered by a 2100 hp dual-stage-supercharged V18 cylinder engine, with its power boosted to 2450 hp when the cylinders were injected with a water/alcohol mixture. An air scoop was fitted under the nose, while the wing fuel tanks were removed. The propeller was also changed from a three blade to a four blade type. The new engine, the mixture and the new propeller blades allowed the F4U-4 to reach speeds of up to 721 km/h (448 mph) and a better climbing rate (4500 ft/min / 1180 m/min). A flat bulletproof windscreen was also installed, avoiding optical distortions. Versions with wingtip tanks and a six-blade contra-rotating propeller were proposed but ultimately rejected by the US Navy.
  • F4U-4B – Corsair that were set to be delivered for the British Fleet Air Arm, but were confiscated by the US.
  • F4U-4C – A version with an alternate weapons set of 4 X 20mm AN/M2 (Hispano-Suiza) cannon. 300 delivered.
  • F4U-4E/F4U-4N – Night fighters with the starboard wing radar radome. The F4U-4E was equipped with an APS-4 search radar, and the F4U-4N was equipped with an APS-6 search radar. These Corsairs would have an armament of 4 X 20mm AN/M2 (Hispano-Suiza) cannons. These Corsairs served in the Korean War.
  • F4U-4K – Experimental drone version
  • F4U-4P – A photo-reconnaissance version.
  • XF4U-5 – Version with new engine cowling.
  • F4U-5 – A modified version of the F4U-4, introduced in 1945 and aimed at increasing the Corsair’s performance and introduce many of the suggestions issued by the pilots. It was powered with a Pratt & Whitney R-2800-32(E) engine with a two-stage supercharger of 2850 hp. Automatic blower controls, cowl flaps, intercooler doors and oil cooler for the engine were fitted. Spring tabs for the elevators and rudder, a modernized cockpit, a retractable tailwheel, heated cannon bays and pitot head were also fitted. The cowling was lowered two degrees, and the wings were all-metal. 223 units delivered.
  • F4U-5N – A radar equipped version. 214 units delivered.
  • F4U-5NL – A winterized version equipped with rubber de-icing boots on the leading edge of both wings and tail. 72 units delivered and 29 units modified from F4U-5N.
  • F4U-5P – A long range photo-reconnaissance version. 30 units delivered.
  • F4U-6/AU-1 – A re-designated AU-1 (which in turn, was based on a modified F4U-6), which was the ground-attack version in use by the Marine Corps. The AU-1 had extra armour protecting both pilot and fuel tank, as well as extra racks, and the oil coolers relocated inboard to reduce changes of ground fire damage. The supercharger was redesigned for low-altitude operations. Capable of carrying up to 3720kg (8,200lbs) of bombs and of reaching speeds of 383 Km/h (238mph) or 479 Km/h (298mph) when armed with bombs or rockets and with one or two fuel tanks. At empty payload this version could reach speeds of 626 Km/h (389mph). produced in 1952 and retired in 1957, seeing action in the Korea War.
  • F4U-7 – Version based on the AU-1 for service with the French Navy.
  • FG-1E – Goodyear-made Corsairs FG-1 with radar equipment.
  • FG-1K – Goodyear-made Corsairs FG-1 used as drones.
  • FG-3 – A turbosupercharger version from modified FG-1D airframes.
  • FG-4 – Goodyear-made Corsairs F4U-4 that were never delivered.
  • Super Corsairs (F2G-1 / F2G-2) – Versions developed after the war, powered by a Pratt & Whitney R-4360 Was major with 4-row 28-cylinder radial engine and a teardrop/bubble canopy. The F2G-1 had a manual folding wing and a 4,3m (14ft) propellers, the F2G-2 had hydraulic operated folding wings, 4m (13ft) propellers and carrier arresting hooks. Development problems delayed and finally ended further developments, with the F2G-2 becoming racing planes.

Operators

  • f4u-4-vmf-124-13-kenneth-a-walsh-okinawa-06-45_03
    U.S. Navy F4U-4 – VMF 124 No 13 – June 1945
  • United States of America – The Corsair was primarily used by the US Navy and the United States Marine Corps in most of the campaigns of the Pacific War. It started its service at Solomon Island in 1943 as fighter in the hands of the USMC, where three famous Pacific War American Aces marked their scores with Corsairs. It also took part of dive bombing operations in the Marshal Islands, seeing also action in the China Sea, Okinawa, Iwo Jima, Formosa, the Philippines and also in the Tokyo and Saigon Raids. In Okinawa, it became the main defence against Kamikaze attacks. The Korean War brought the Corsairs back given its capacity to carry large and heavy amounts of payload/ordnance, performing ground-attack and CAS missions, used by the USMC. Many were also sold as surplus aircraft, serving in the air forces of Argentina, El Salvador and Honduras.
  • f4u-1f-mk-1-1855-nas-faa-5f-jt150-10-43_03
    Royal Navy F4U-1 (F.Mk.1) 1855 NAS FAA – Oct 1943
  • United Kingdom – 2,012 Corsairs were issued to the British Royal Navy Fleet Air Arm in 1943, where the wings were clipped 8 inches in order to increase storage in the lower carrier decks, being the British Corsairs the first ones to be used in on an aircraft carrier. The Corsair also took part as escort fighter and anti-air defences in three operations – Operation Tungsten, Operation Mascot and Operation Goodwood – against German battleship DKM Tirpitz. In 1944, British Corsairs took part in operations at the Indian and the Pacific Ocean, remarkably used in Java as bombers. It was during Corsair service with the British, that enhancements for carrier operation were made.
  • French Navy F4U-7 - 14F Aeronavale No.133704 - Circa 1956
    French Navy F4U-7 – 14F Aeronavale No.133704 – Circa 1956
  • France – France and its naval air branch or Aéronavale operated with 69 AU-1 and 94 F4U-7s from 1954 to 1964. It was introduced to replace the Supermarine Seafires, Grumman Hellcats, Curtiss Helldivers and SBD Dauntless that equipped the naval air service. They operated from 4 carriers – Arromanches, Dixmude, La Fayette and Bois Belleau – that were part of the French Navy. 4 squadrons – the 14F, 12F, 15F and 10F – were operating with the Corsair, alongside two training squadrons – 10S and 57S. French Corsairs intervened firstly in Indochina, as they were handed by the US (AU-1 Korean War veterans) and where they were well received by French troops and pilots. In Indochina 6 Corsairs lost and 2 pilots dead.
  • The Corsairs also operated in Africa, namely in Algeria, Suez and Tunisia. In Algeria, they provided fire support, bombing, reconnaissance and protection of airborne troops. There were some considerable losses due to accidents and AA fire took place. In Suez, they took part in operations from carriers Arromanches and La Fayette, attacking the Cairo-Almanza airfield with only one loss against 12 planes damaged and 1 damaged of the Egyptians. The last action the French Corsairs saw was in Tunisia, where they provided support to besieged troops at a French airbase after Tunisian independence, attacking also Tunisian troops and vehicles. 3 Corsairs were lost due to the AA. The French reportedly used the Corsairs to experiment with anti-tank missiles, but they were never used. As new carriers and new air naval assets were introduced, the Aéronavale withdrew its Corsairs.
  • f4u-1a-rnzaf-22sqn-49944-1944_03
    Royal New Zealand AF F4U-1A – 22 Sqn 49944 – July 1944
  • New Zealand – The New Zealand air force shifted from the P-40 to the Corsair in 1944, receiving in total 424 airframes as a lend-lease, with 13 squadrons operating it. The RNZAF operated with F4U-1A, F4U-1D and FG-1D, concentrating on attacking the bypassed islands with ground support, escort and air patrols. Only 17 Corsairs were lost, as the Japanese air superiority was, by the time the Corsairs were received, almost neutralized. A squadron equipped with Corsairs served an occupation duty for two years once the Pacific War was over.
  • Argentina – Argentine acquired the day-time and night-time fighter versions of the Corsairs (26 F4U-5/5N/5NL) in 1957, being incorporated to the Argentinian aircraft carrier ARA Independencia. As the abovementioned versions were fitted with Radar, Argentina became the first nation in the region to operate aircraft with radars. They intervened during the 1958 border incidents with Chile, and in the period of 1959-1960, the Corsairs were used as submarine chasers – equipped with depth charges – following the detection of unidentified submarines. They also took part during the political revolt of 1963, being 1964 their last year of operational service during another set of border incidents with Chile. They were withdrawn from service in 1968.
  • Honduras – The Honduras Air Force operated the Corsair from 1956 to 1979, with 19 units. The Honduran Corsairs also took part in the 1969 ‘Soccer War’, where a single Corsair scored three victories against two Salvadorian Corsairs and one Mustang, piloted by Capt. Fernando Soto. These were the only air-to-air victories of the war. The Honduras Corsairs also performed strafing missions at the border. The Corsair that scored those victories is now a war memorial.
  • fg-1d-el-salvador-fas-201-67087-1958_03
    El Salvador Air Force FG-1D – 67087 – Circa 1958
  • El Salvador – The Salvadorian Air Force operated the Corsair from 1957 to 1976, with 25 F4U-/FG-1D. They took part in combats during the 1969 ‘Soccer War’, where took some losses in the hands of the Honduras Air Force operating similar F4U-4 and F4U-5 fighters.
  • Germany – Germany captured only one British Corsair that was forced to land in Norway due to technical issues while taking part in Operation Mascot.
  • Japan – Japan also captured two Corsairs after emergency landings, with one possible tested in flight.

 

F4U-4 Specifications 

Wingspan  12,49 m / 41 ft 0 in
Length  10,27 m / 33 ft 8 in
Height  4,5 m / 14 ft 9 in
Wing Area  29,17 m² / 314 ft²
Engine  1 Pratt & Whitney R-2800-18W 18 cylinder radial engine of 2,250 hp
Propeller Diameter  4,06 m/ 13 ft 4 in
Maximum Take-Off Weight  6149 Kg / 13,556 lb
Empty Weight  4174 kg / 9,202 lb
Loaded Weight  5626 kg / 12,405 lb
Maximum Speed  718 km/h / 446 mph
Range  2511 Km / 1,560 miles
Maximum Service Ceiling  12650 m /41,500 ft
Climb Rate  3050m in 5,1 minutes (22.1 m/s; 4,360 ft/min)
Crew  1 (pilot)
Armament
  • 6 X 12,7mm (0.50 caliber) M2Browning machine guns or 4 X 20mm AN/M2 cannons.
  • Up to 1000 kg (4000 lbs) of bombs.
  • 8 X 127 mm high velocity aircraft rockets.

Gallery

f4u-4-vmf-124-13-kenneth-a-walsh-okinawa-06-45_03
U.S. Navy F4U-4 – VMF 124 No 13 – June 1945
U.S. Navy F4U-1 VMF 123 no. 15 ‘Daphne C’ – July 1943
f4u-1a-vf-17-jolly-rogers-17640-bighog-11-43_03
U.S. Navy F4U-1A – VF-17 17640 ‘Big Hog’ – Nov 1943
U.S. Navy F4U-1D - VMF-451 20141 - Apr 1945
U.S. Navy F4U-1D – VMF-451 20141 – Apr 1945
Royal Canadian Navy FG-1D - 1841 Sqn BuNo 76236 - Aug 1945
Royal Canadian Navy FG-1D – 1841 Sqn BuNo 76236 – Aug 1945
f4u-1a-rnzaf-22sqn-49944-1944_03
Royal New Zealand AF F4U-1A – 22 Sqn 49944 – July 1944
Royal New Zealand AF F4U-1 - No. 21 Sq NZ5315 BuNo 49909 - Jun 1944
Royal New Zealand AF F4U-1 – No. 21 Sq NZ5315 BuNo 49909 – Jun 1944
f4u-1f-mk-1-1855-nas-faa-5f-jt150-10-43_03
Royal Navy F4U-1 (F.Mk.1) 1855 NAS FAA – Oct 1943
French Navy F4U-7 - 14F Aeronavale No.133704 - Circa 1956
French Navy F4U-7 – 14F Aeronavale No.133704 – Circa 1956
fg-1d-el-salvador-fas-201-67087-1958_03
El Salvador Air Force FG-1D – 67087 – Circa 1958

 

 

Sources

Aviation Publications (1977). Pilots Manual for F4U Corsair., Berger, R (Ed.). Aviones [Flugzeuge, Vicenç Prat, trans.]. Colonia, Alemania: Naumann & Göbel Verlagsgessellschaft mbH., Chant, C (2001). Aviones de la Segunda Guerra Mundial [Aircraft of World War II, Fabian Remo Tamayo & Fernando Tamayo, trans.]. Madrid, Spain: Editorial LIBSAD’Angina, J. (2014). Vought F4U Corsair. Oxford, UK: Osprey Publishing., Gunston, B. (1995). Guía Ilustrada de los Cazas y Aviones de Ataque Aliados de la Segunda Guerra Mundial (II). [An Illustrated Guide to Allied Fighters of World War II, Gearco, trans.]. Barcelona, Spain: Ediciones Folio (Original work published in 1981)., Gustin, E. (n.d.). Chance Vought F4U Corsair. F4UCorsair.com.HISTARMAR (n.d.). Chance-Vought F4U-5 Corsair. Fundación HISTARMAR.Lyford, C., & Tillman, B. (2014). Corsairs vs. Mustangs: The last dogfight. The Flight Journal, 16-24.Marsaly, F., & Bocognano, J. (2002). Les Corsairs de la Royale. Aérostories.org.Rochotte, L. (1999). Les Corsair Français. Net-Marine.Sherman, S. (2012). Vought F4U Corsair. Acepilots.com.The National WWII Museum (n.d.). Airplanes in the Us Freedom Pavilion: The Boeing Center.UTDallas (n.d.). Chance Vought/LTV History.Vought.org. (n.d.). F4U Corsair – WWII Record. Vought.org.Werbaneth, J. P. (2009). His Majesty’s Corsairs. Avalanche Press.Vought F4U Corsair. (2016, September 11). In Wikipedia, The Free Encyclopedia. Images: Black and White F4U by Brian Struble / CC BY 2.0,   F4U in FlightF4U Climb by Airwolfhound / CC BY-SA 2.0, Royal Navy F4U by Tony Hisgett / CC BY 2.0,  Redbull F4U Corsair and B25J Mitchell by Ronnie Macdonald / CC BY 2.0,   Side Profile Views by Ed Jackson – Artbyedo.com, Additional Side Profile Views by Brendan Matsuyama