He 112 in Hungarian Service

Hungarian Flag Kingdom of Hungary (1938)
Fighter Aircraft – 4 aircraft operated

Despite being not adopted for service by the German Luftwaffe, the He 112 had great potential as an export aircraft. Spain, Romania, and Japan were some of the countries that got their hands on fighter aircraft. Hungary, with its close ties to Germany, also wanted this fighter in its inventory, though it was not to be. Unfortunately for them, despite their efforts, only a few of these aircraft would ever see service with their Air Force. This was mainly due to the reluctance of Germany to provide the necessary parts and licenses, and the start of the Second World War. The few aircraft that did reach Hungary were mainly used for crew training and even saw limited combat use.

One of few He 112 in Hungarian service. Source: www.destinationsjourney.com

A brief He 112 history

Prior to the Second World War, the Luftwaffe was in need of a new and modern fighter to replace the older biplanes that were in service, such as the  Arado Ar 68 and Heinkel He 51.  For this reason, in May 1934, the RLM issued a competition for a new, modern fighter plane. While four companies responded to this request, only the designs from Heinkel and Messerschmitt were deemed sufficient. The Heinkel He 112 was a good design that offered generally acceptable flight characteristics and possessed a good foundation for further improvements. The Bf 109 on the other hand, had slightly better overall flight performance and was much simpler and cheaper to build. Given the fact that the Germans were attempting to accelerate the production of the new fighter, that alone was seen as a huge advantage over the He 112. Ultimately it would not be accepted for service, and only 100 or so aircraft would be built. These would be mainly sold abroad, with those remaining in Germany being used for various testing and evaluation purposes.

He 112 the unsuccessful competitor of the Bf 109. Source: www.luftwaffephotos.com

While the He 112 project was canceled by the RLM, to compensate for the huge investment in resources, Heinkel was permitted to export this aircraft.  A number of countries such as Austria, Japan, Romania, and Finland showed interest, but only a few actually managed to procure this aircraft, and even then, only in limited numbers.

Hungarian Interest in the He 112

Being that it was on the losing side of the First World War, the Hungarians were in a similar situation to Germany in regard to military restrictions under the Treaty of Versailles. Crucially, it prohibited the Hungarians from developing their air forces. In time though, the Allies became less and less involved in maintaining the Treaty, and the Hungarians began slowly rebuilding their air force.  By 1938 the  Magyar Királyi Honvéd Légierő MKHL (English: Royal Hungarian Home Defence Air Force) was openly presented to the world. At that time, the Hungarians undertook steps to rebuild their armed forces in the hope of reclaiming some of their lost territories. For a modern air force, they needed better fighter designs, as their aged biplanes would not be sufficient. By 1938, they had improved their relations with Germany, and it was then possible to acquire new equipment from them.

The Hungarian military delegation that was in Spain during the civil war observed the relatively new Heinkel He 112 fighter in action and immediately became interested in it. In June 1938, a military group disguised as a civilian delegation visited Heinkel’s company. Three Hungarian pilots had the chance to flight test the He 112V9 aircraft. They were highly impressed and urged the Hungarian Army officials to adopt this aircraft. Unsurprisingly, based on the glowing report, the Hadügyminisztérium (Ministry of War Affairs) asked Heinkel for 36 such aircraft.

Unfortunately for them, Heinkel never actually put the He 112 into mass production, given the fact that it was not adopted for service with the German Air Force. It did, however, build a small series that was intended for Spain and Japan. The Hungarian offer was not considered as important, and thus no aircraft would be delivered to them. The Reichsluftfahrtministerium RLM (English: German Ministry of Aviation) also intentionally delayed the delivery of weapons to Hungary. This was done to politically and economically pressure the Hungarians and Romanians who were on the brink of war at that time, in an attempt to reduce tensions.

Still, the Hungarians persisted, and at the start of 1939, they requested again for the 36 aircraft, and once again, the Germans denied this request. However, a single He 112 V9 was given to Hungary and was used for flight testing near Budapest. On the 5th of February 1939, it crashed during a test flight against a CR-32 biplane fighter.  In March 1939, another aircraft was sent to Hungary, this one being a He 112 B-1. It was extensively tested by the Hungarians who generally liked its design.

The He 112 V9 was lost in an accident during its first test flight. Source: D. Bernard Heinkel He 112 in Action

As the Romanians acquired a batch of 24 He 112 In 1939, the Hungarians were concerned over their neighbor’s growing military strength. Realizing that the Germans would not deliver the promised aircraft, they decided to ask for a production license instead. This was granted, and Heinkel also delivered two more He 112 B-1 with the Jumo 210E engine. When the license document arrived in Hungary in May 1939, a production order for the 12 first aircraft was given to the Weiss Manfréd aircraft manufacturer. Several changes were made, including the installation of 8 mm 39.M machine guns and the addition of bombing racks. In addition, the original 2 cm cannons were to be replaced by the Hungarian, domestically built, Danuvla 39, though it is unclear if any were actually installed. As the preparation for the production was underway the three available He 112 were adopted to service. This received coded designation V.301 to 303 where the V stands for Vadász (English: Fighter).

The B-series was in many aspects a complete redesign of the previous series. Including the introduction of a new tail unit, and part of the fuselage, to name a few. Source: www.luftwaffephotos.com

Despite the best Hungarian attempts to put the He 112 in production, the situation was made impossible by the coming war between Poland and Germany. The RLM would officially prohibit the export of any German aircraft engines and equipment at the start of the war. This meant that the vital delivery of the Jumo 210 and DB 601 engines could not be made. Based on this fact, all work on the Hungarian He 112 was canceled. Instead,  Weiss Manfréd investigated to see if it could reuse most of the He 112 production line to produce a new domestic design named WM–23 Ezüst Nyíl (English: Silver arrow). While one prototype was built it was lost in an accident which ended the project.

The WM–23 Ezüst Nyíl prototype. Source: www.destinationsjourney.com
The V.303 during pilot training in 1940. Source: www.destinationsjourney.com
Despite their best efforts, the Hungarians managed to operate only three He 112 (not including the single aircraft last way back in Budapest) Source: www.destinationsjourney.com

In Combat

In the Summer of 1940, the rising tension between Romania and Hungary over Transylvania reached a critical point. Transylvania was once part of Hungary but was lost after the First World War when it was given to Romania. By 1940, the Hungarian Army began preparing for a possible war with Romania over the territory. As neither side was willing to enter a hastily prepared war, negotiations began to find a possible solution. But despite this, there were some minor skirmishes, and Hungarian aircraft made several reconnaissance flights over Romania. The three Hungarian He 112 were stationed near the border, and the Romanians also had some He 112 in their inventory. While the Hungarian He 112’s did take up to the sky, no combat action by them was reported. Ultimately, at the end of August, Romania asked Germany to arbitrate the issue regarding the disputed territory, With Hungary being given the northern part of Transylvania in the settlement.

During the Axis invasion of Yugoslavia in April 1940, Hungary once again mobilized its He 112s. These were stationed near the border with Yugoslavia but they were not used in any combat operations.

By the time the Axis attacked the Soviet Union in June 1941 all three He 112 were used as training aircraft, with their secondary role being to protect the Weiss Manfréd factory. Due to a lack of spare parts, there was no point in sending this aircraft to the frontline. Two aircraft were involved in a landing accident where they were damaged. While their final fate is not completely clear, they may have been destroyed in 1944 when the Allies intensified their bombing campaign against Hungary. It is unlikely that the He 112s were operational at this point.

The V.301 had an accident where the pilot forgot to activate the landing gear. This is not surprising given that most pilots at that time mainly flaw on the older biplanes that had fixed landing gear. Source: www.destinationsjourney.com
In the Summer of 1941 the V.303 was damaged during a landing where the left landing gear wheel simply broke off. Source: www.destinationsjourney.com

Technical Characteristics

The He 112 was an all-metal, single-engine fighter. The monocoque fuselage consisted of a metal base covered by riveted stress metal sheets. The wing was slightly gulled, with the wingtips bending upward, but otherwise had a conventional construction.

During its development life, a great number of different engines were tested on the He 112. For the main production version, the He 112 B-2, it carried a 700 hp Jumo 210G liquid-cooled engine, with some others being equipped with the  680 hp Jumo 210E engine. The He 112 had a fuel capacity of 101 liters in two wing-mounted tanks, with a third 115-liter tank placed under the pilot’s seat.

The landing gear was more or less standard in design. It consisted of two larger landing wheels that retracted into the wings and one semi-retractable tail wheel. The He 112 landing gear was wide enough to provide good ground handling and stability during take-off or landing.

The cockpit received a number of modifications. Initially, it was open with a simple windshield placed in front of the pilot, with Later models having a sliding canopy.

The armament was changed throughout the He 112’s production, and the last series was equipped with two 7.92 mm MG 17 machine guns and two 2 cm MG FF cannons. The ammunition load for each machine gun was 500, with 60 rounds for each of the cannons. If needed, two bomb racks could be placed under the wings.

Conclusion

The He 112, although few in number, provided the Hungarian Air Force with one of its first modern fighter aircraft. Despite the Hungarian attempts to acquire over 30 aircraft from Germany, this was never achieved. In the end, the Hungarians only had three operational He 112, and one  was lost in an accident during testing. While these were stationed on the front line on two occasions they never saw actual combat action. By 1941 due to a lack of spare parts, they were allocated for training purposes. The Hungarians eventually got a production license for the Messerschmitt Bf 109G making the few available He 112 unnecessary.

 

He 112B-1 Specifications

Wingspans 29  ft 10  in / 9.1 m
Length 30  ft 2 in / 9.22 m
Height 12 ft 7 in  / 3.82  m
Wing Area 180  ft² / 17 m²
Engine One r 680 hp Jumo 210E  liquid-cooled engine
Empty Weight 3,570  lbs / 1,620 kg
Maximum Take-off Weight 4,960 lbs / 2,250 kg
Climb Rate to 6 km In 10 minutes
Maximum Speed 317 mph / 510 km/h
Cruising speed 300 mph / 484 km/h
Range 715 miles / 1,150 km
Maximum Service Ceiling 31,170 ft / 9,500 m
Crew 1 pilot
Armament
  • Two 20 mm (1.8 in) cannons and two machine guns 8 mm (0.31 in) machine guns and 60 kg bombs

Credits

  • Article written by Marko P.
  • Edited by  Henry H.
  • Ported by Marko P.
  • Illustrated By Godzilla

Illustrations

Source:

  • Duško N. (2008)  Naoružanje Drugog Svetsko Rata-Nemаčaka. Beograd
  • G. Punka (1994) Hungarian Air Force, Squadron Publication
  • J. R. Smith and A. L. Kay (1990) German Aircraft of the Second World War, Putnam
  • D. Monday (2006) The Hamlyn Concise Guide To Axis Aircraft OF World War II, Bounty Books
  • D. Bernard (1996) Heinkel He 112 in Action, Signal Publication
  • R.S. Hirsch, U, Feist and H. J. Nowarra (1967) Heinkel 100, 112, Aero Publisher
  • C. Chants (2007) Aircraft of World War II, Grange Books
  • S. Renner. (2016) Broken Wings The Hungarian Air Force, 1918-45, Indiana University Press

 

Lippisch P 13a

Nazi flag Nazi Germany 

Ramjet powered aircraft

None built

In the later stages of the Second World War, it was becoming apparent to both the Luftwaffe  (English German Air Force) and the German Government that the Allied air forces were gaining air superiority. This realization saw them turn to new and fantastical ideas in a desperate attempt to turn the tide of the war. Some of these represented new improvements to existing designs, the introduction of the newly developed turbojet engine, and even more esoteric and experimental methods. In many cases, these were pure fantasies, unrealistic or desperate designs with no hope of success. Few of them reached any significant development, and among them were the works of Alexander Martin Lippisch. While Lippisch helped develop the Me 163, the first rocket-powered interceptor, his other work remained mostly theoretical. One such project was the unusual P 13a, ramjet-powered aircraft that was to use coal as its main fuel source. While some work was carried out late in the war and soon faced insurmountable technical problems, thus nothing came of the project.

Artistic presentation of how the P 13a may have looked. Source:  Luftwaffe Secret Jets of the Third Reich

History

Before the start of the Second World War, aviation enthusiast and engineer Alexander Martin Lippisch, was fascinated with tailless delta wing designs. Lippisch’s early work primarily involved the development of experimental gliders. Eventually, he made a breakthrough at the Deutsche Forschungsinstitut, where he worked as an engineer.  His work at DFS would lead to the creation of the rocket-powered glider known as the DFS 194. As this design was a promising experiment in a new field, it was moved to Messerschmitt’s facility at Augsburg. After some time spent refining this design,  it eventually led to the development of the Me 163 rocket-powered interceptor.  While it was a relatively cheap aircraft, it could never be mass-produced, mostly due to difficulties associated with its highly volatile fuel. In 1942, Lippisch left Messerschmitt and ceased work on  the Me 163 project. Instead, he joined the Luftfahrtforschungsanstalt Wien (English: Aeronautic Research Institute in Vienna) where he continued working on his delta-wing aircraft designs. In May 1943 he became director of this institution, and at that time the work on a supersonic aircraft was initiated.

In the later war years, among the many issues facing the Luftwaffe, was a chronic fuel shortage. Lippisch and his team wanted to overcome this problem by introducing alternative fuels for their aircraft. Luckily for his team, DFS was testing a new ramjet engine. They were designed to compress air which would be mixed with fuel to create thrust but without a mechanical compressor. While this is, at least in theory, much simpler to build than a standard jet engine, it can not function during take-off as it requires a high airflow through it to function. Thus, an auxiliary power plant was needed. It should, however, be noted that this was not new technology and had existed since 1913, when a French engineer by the name of Rene Lorin patented such an engine. Due to a lack of necessary materials, it was not possible to build a fully operational prototype at that time, and it would take decades before a proper ramjet could be completed. In Germany, work on such engines was mostly carried out by Hellmuth Walter during the 1930s. While his initial work was promising, he eventually gave up on its development and switched to a rocket engine instead. The first working prototype was built and tested by the German Research Center for Gliding in 1942. It was later tested by mounting the engine on a Dornier Do 17 and, later, a Dornier Do 217.

The Dornier Do 217 was equipped with experimental ramjets during trials. Source: tanks45.tripod.com

In October 1943, Lippisch won a contract to develop the experimental P 11 delta-wing aircraft. While developing this aircraft, Lippisch became interested in merging his new work with a ramjet engine. This would lead to the creation of a new project named the P 12. In the early stage of the project, Lippisch and his team were not completely sure what to use as fuel for their aircraft, but ramjets could be adapted to use other types of fuel beyond aviation gasoline.

Unfortunately for them, LFW’s facilities were heavily damaged in the Allied bombing raids in June 1944. In addition to the damage to the project itself, over 45 team members died during this raid. To further complicate matters, the scarcity of gasoline meant that Lippisch’s team was forced to seek other available resources, such as different forms of coal. This led to the creation of the slightly modified project named P 13. In contrast to the P 12, the cockpit was relocated from the fuselage into a large fin. This design provided better stability but also increased the aircraft’s aerodynamic properties.  The overall designs of the P 12 and P 13 would change several times and were never fully finalized.

The P 12 and 13 small-scale models, in both configurations, were successfully tested at Spitzerberg Airfield near Vienna in May 1944. The project even received a green light from the Ministry of Armaments. In the early stages of the project, there were some concerns that the radical new design would require extensive retraining of pilots. However, the wind tunnel test showed that the design was aerodynamically feasible and that the aircraft controls had no major issues. Based on these tests,  work on an experimental aircraft was ordered to begin as soon as possible.

A proposed P 12 aircraft. Its designs changed greatly over time, before being finally discarded in favor of the letter P 13. Source: The Delta Wing History and Development

The DM-1 Life Saver 

While working on the P 12 and P 13, Lippish was approached with a request from a group of students from Darmstadt and Munich universities. They asked Lippisch to be somehow involved in the P 12 and 13 projects. Lippisch agreed to this and dispatched one of his assistants under the excuse that for his own project, a wooden glider was to be built and tested. The previously mentioned student’s and Lippisch’s assistant moved to a small warehouse in Prier and began working on the Darmstadt 33 (D 33) project. The name would be changed to DM 1 which stands for Darmstadt and Munich.

At this point of the war, all available manpower was recruited to serve the German war effort. For young people, this often meant mobilization into the Army. One way to avoid this was to be involved in some miracle project that offered the Army a potentially war-winning weapon. It is from this, that numerous aircraft designs with futuristic, and in most cases unrealistic, features were proposed. Many young engineers would go on  to avoid military service by proposing projects that on paper offered extraordinary performance in combat.

The students and Lippisch managed to nearly complete their DM1 test glider when the war ended. Source: airandspace.si.edu

While it was under construction, preparations were made to prepare for its first test flight. As it was a glider it needed a towing aircraft that was to take it to the sky. A Sibel Si 204  twin-engine aircraft was chosen for the job. However, this was not to be done like any other glider, being towed behind the larger aircraft. Instead, the DM-1 was to be placed above the Si 201 in a frame, in a similar combination as the Mistel project. The estimated theoretical speeds that were to be reached were 560 km/h (350 mph).

Allegedly, there were four different proposals for the DM’s that were to be fully operational. The DM 2 version was estimated to be able to reach a speed of  800-1,200 km/h (500 – 745 mph). The DM 3’s theoretical maximum speed was to be 2,000 km/h (1,240 mph) while the fate of the DM 4 is unknown. Here it is important to note that these figures were purely theoretical, as there were no supersonic testing facilities to trial such a design. It is unclear in the sources if these additional DM projects even existed, even if in only written form. We must remember that the whole DM 1 glider idea was made to help the students avoid military conscription and that Lippisch himself never saw the DM 1 as any vital part of the P 13.

In any case, the glider was almost completed by the time the war ended and was later captured by the Western Allies. Under the US Army’s supervision, the glider was fully completed and sent to America for future evaluation. It would then be given to the Smithsonian Institution.

 

A DM 1 test glider being under construction. Source: hushkit.net
The Siebel Si 204 was to be used as a carrier for the DM 1 glider for the expected first-flight tests. Due to the end of the war, this was never achieved. Source: www.silverhawkauthor.com

Work on the P 13

As the work on the P 13 went on, the name was slightly changed. This was necessary as different variations of the P 13 were proposed. The original  P 13 received the prefix ‘a’ while the later project’s designation continued alphabetically for example P 13b. After a brief period of examination of the best options, the P 12 project was discarded in favor of P 13. The decision was based on the fuel that the aircraft should use. What followed was a period of testing and evaluation of the most suitable forms of coal that could be used as fuel. Initial laboratory test runs were made using solid brown Bohemian coal in combination with oxygen to increase the burn rate. The fuel coal was tube-shaped, with an estimated weight of 1 kg, and encased in a mesh container through which the granulated coal could be ejected. The testing showed serious problems with this concept. While a fuel tube could provide a thrust that on average lasted 4 to 5 minutes, its output was totally unpredictable. During the testing, it was noted that due to the mineral inconsistency of the coal fuel, it was impossible to achieve even burning. Additionally, larger pieces of the coal fuel would be torn off and ejected into the jet stream. The final results of these tests are unknown but seem to have led nowhere, with the concept being abandoned. Given that Germany in the last few months of the war was in complete chaos, not much could be done regarding the Lippish projects including the P 13a.

As more alterations to the original design were proposed its name was charged to P 13a. Here is a drawing of a P 13b that was briefly considered but quickly discarded. Source: The Delta Wing History and Development

In May 1945, Lippish and his team had to flee toward the West to avoid being captured by the advancing Soviets. They went to Strobl in Western Austria, where they encountered the Western Allies. Lippisch was later transported to Paris in late May 1945 to be questioned about his delta wing designs. He was then moved to England, and then to America in 1946. The following year,  American engineers tested the DM 1 glider at the wind tunnel facility of the Langley Field Aeronautical Laboratory. The test seems promising and it was suggested to begin preparation for a real flight. A redesign of the large rudder was requested. It was to be replaced with a much smaller one, where the cockpit would be separated from the fin and placed in the fuselage. Ironically Lippish was not mentioned in this report, as technically speaking he was not involved in the DM 1 project. Nevertheless, he was invited for further testing and evaluation of this glider. If this glider and the Lippish work had any real impact on the US designs is not quite clear.

Despite no aircraft being ever completed, one full-size replica of this unusual aircraft was built after the war. It was built by Holger Bull who is known for building other such aircraft.  The replica can now be seen at the American Military Aviation Museum located in Virginia Beach.

An interesting full-size replica of the P 13 located at the American Military Aviation Museum. Source: Wiki

Technical characteristics

DM 1

The DM 1 glider was built using wooden materials. Given that it was constructed by a group of young students, its overall design was quite simple. It did not have a traditional fuselage, instead, its base consisted of a delta wing. On top, a large fin was placed. The cockpit was positioned in front of the aircraft within the large vertical stabilizer. To provide a better view of the lower parts of the nose, it was glazed. The landing gear consisted of three small landing wheels which retracted up into the wing fuselage. Given that it was to be used as a  test glider, no operational engine was ever to be used on it.

The DM 1 side view. In contrast to the later P 13a design, the pilot’s cockpit position was placed above the wings. This was necessary as the engine was to be added. Source: airandspace.si.edu
A DM 1 was captured by the Allies after the war. Its unique shape is quite evident in this photograph. Source: Wiki
A good example of DM 1 (to the right) and P 13a models that showed the difference between these two. The P 13a could be easily distinguished by its engine intake and the different position of the pilot cockpit. Source: Wiki

A good example of DM 1 (to the right) and P 13a models that showed the difference between these two. The P 13a could be easily distinguished by its engine intake and the different position of the pilot cockpit. Source: Wiki https://imgur.com/a/QW7XuO5

P 13a

The P 13 is visually similar but with some differences. The most obvious was the use of a ramjet. This means that the front, with its glazed nose, was replaced with an engine intake. Here, it is important to note, that much of the P 13a’s design is generally unknown, and much of the available information is sometimes wrongly portrayed in the sources. The P 13a never reached the prototype stage where an aircraft was fully completed. Even as the war ended, much of the aircraft’s design was still theoretical. Thus all the mentioned information and photographs may not fully represent how the P 13 may have looked or its precise characteristics, should it have been finished and built.

The exact ram engine type was never specified. It was positioned in the central fuselage with the air intake to the front and the exhaust to the back. As the main fuel, it was chosen to use small pieces of brown coal which were carried inside a cylindrical wire mesh container. The total fuel load was to be around 800 kg (1,760 lbs). Combustion was to be initiated by using small quintiles of liquid fuel or gas flames.  The overall engine design was changed several times during the work on the P 13 without any real solution to the issues of output consistency. Given that the ramjets could not work without an air thrust, an auxiliary engine had to be used during take-off, though a more practical use would be to tow the P 13 until it could start its engine. A rocket takeoff ran the risk of the engine failing to ignite, leaving the pilot little time to search for a landing spot for his unpowered aircraft.

 

An illustration of the proposed P 13a engine interior. The use of coal as fuel may seem like a cheap alternative but given that this kind of technology was never employed may be an indication of its effectiveness. Source: theaviationgeekclub.com

The wing construction was to be quite robust and provided with deflectors that would prevent any potential damage to the rudders. The wing design also incorporated a sharp metal plate similar to those used for cutting enemy balloons cables. These proposed properties of the wings are another indicator that the P 13 was to be used as an aircraft rammer. Another plausible reason for this design was the fact that given it had no landing gear the aircraft design had to be robust enough as not to be torn apart during landing. The wings were swept back at an angle of 60 degrees. The precise construction method of the wings (and the whole P 13 a on that matter) are not much specified in the sources. Given the scarcity of resources in late 1944 it is likely that it would use a combination of metal and wood.

A drawing of the P 13a interior. Its overall construction was to be more or less standard in nature. This could not be said for the aircraft’s overall shape design. Source: D. Sharp Luftwaffe Secret Jets of the Third Reich

The fin had to be enlarged to provide good flight command characteristics. In addition, given that the position of the cockpit was in the fin, it had to be large. The fin was more or less a direct copy of one of the wings. So it is assumed that it too would share the overall design.  The fin was connected to the aircraft by using four fittings.

The cockpit design was to be simple and cheap to build. The pilot was to have plenty of room inside the large fin. The cockpit was provided with a large glazed canopy that provided a good view of the front and sides. The seat and the instrument panel were bolted to the cockpit floor and walls. These could be easily detached for repairs. The instrument panel was to include an artificial horizon indicator, altimeter, compass, and radio equipment, Given that it was to operate at a high altitude oxygen tanks were to be provided too. Despite being intended to fly at high altitudes the cockpit was not to be pressurized. Another unusual fact was that initially the P 13 was to have a crew of two, but this was quickly discarded.

A possible example of how the inside of the pilot cockpit may have looked. Source: D. Sharp Luftwaffe Secret Jets of the Third Reich

Here it is important to note that the version of the P 13 with the large fin is often portrayed as the final version of this aircraft. However, Lippisch never fully decided whether he should go for this version or the second that used a smaller fin with the pilot cockpit placed above the engine intake. Depending on the proposed version they are drastically different from each other. Lippisch, for unknown reasons, presented the British intelligence officer with the version that used the smaller fin and the American with the second version.

During its development phase, many different alterations of the P 13 were proposed. Isource: D. Sharp Luftwaffe Secret Jets of the Third Reich

Landing operations were a bit unusual. To save weight no standard landing gear was to be used. Instead, Lippisch reused the Me 163 landing procedure.  As the  P 13 was immobile on its own, a small dolly would be used to move the aircraft. Once sufficient height was reached the dolly was to be jettisoned. In theory, this was an easy process, but in practice, this operation offered a good chance of failure and was much less safe than conventional landing gear. Sometimes the dolly either failed to eject or it bounced off the ground hitting the Me 163 in the process, with often fatal consequences.

The Me 163 which did not have traditional landing gear, had to be prior to the flight, transported to the airfield before launching into the sky. Source: warbirdphotographs.com

The aircraft was to land with the nose raised up from the ground. This limited the pilot’s view of the ground. In addition due to its small size and in order to save weight, nontraditional landing gear was provided, instead, it carried a landing blade skid. To help absorb the landing impact, additional torsion springs were to be used. This bar had to be activated prior to the landing, it would emerge from beneath the aircraft fuselage, with the rotation point located at the front. Once released it was to guide the aircraft toward the ground. After that, the torsion springs were to soften the landing. This whole contraption seems like a disaster just waiting to happen and it’s questionable how practical it would be.

A drawing that showed how the P 13a was to land using a guiding landing blade skid. Source: D. Sharp Luftwaffe Secret Jets of the Third Reich

One interesting feature of the P 13 was that it could be easily disassembled into smaller parts which would enable effortless transport. Another reason was that due to the engine’s position in order to make some repairs or replacement of the engine, the remaining parts of the wing and the large fin had to be removed.

Was it an aircraft rammer? 

The precise purpose of the P 13a is not quite clear, even to this day. Despite being briefly considered for mass production, no official offensive armament is mentioned in the sources. So how would the P 13a engage the enemy? A possible solution was that it would be used as a ram aircraft that was supposed to hit enemy aircraft damaging them in the process. In an after-the-war interrogation by British officers, Lippisch was asked if the P 13 was to be used as an aerial ram aircraft. Lippisch responded the following “

“.. The possibilities of using the P.13 as a ramming aircraft had been considered but Dr Lippisch did not think that athodyd propulsion was very suitable for this purpose owing to the risk of pieces of the rammed aircraft entering the intake. This would be avoided with a rocket-propelled rammer…”

This statement contradicts the building description issued by the LFW issued in late 1944. In it was stated the following about this potential use. “…Due to tactical considerations, among other things, the speed difference of fighters and bombers, preferably when attacking from behind, though the thought was given to the installation of brakes ..  and although ample room for weaponry is present, the task of ram fighter has been taken into account – so that the ramming attack will not lead to the loss of the aircraft, thanks to its shape and static structure.”

This meant that this concept may have been considered by Lippisch at some point of the project’s development. The P 13 overall shape resembles closely to aircraft that was intentionally designed for this role. That said, it does not necessarily mean that the P 13 was to ram enemy aircraft. The use of such tactics was considered but their use was discarded, as it was seen as a futile and flawed concept. The project itself never got far enough to have an armament decided for it.

The precise method of how to engage the enemy aircraft is not clear as the P13a was not provided with any armament. It is sometimes referred to in the sources as it was to be used as a ram aircraft. Source: theaviationgeekclub.com

Conclusion

The Lippisch P 13 is  an unusual aircraft project in nearly all aspects. Starting from its shape, which proved, at least during wind tunnel tests, that the concept was feasible. On the other hand, its engine seems to have simply been abandoned after discouraging test results. It is unlikely that such a combination would have worked to the extent that the P 13 designer hoped it would. During the testing, they could not find a proper solution to providing a constant thrust with sufficient force to reach a speed that was expected of it. So the whole concept was likely to be doomed from the start.

The DM 1 however, while it was never seriously worked on by Lippisch himself, managed to save a group of young students who used the project to avoid being sent into combat.

DM-1 Specifications

Wingspans 5.92 m / 19  ft 5 in
Length 6.6 m / 21  ft 7 in
Height 3.18 m / 10 ft 5 in
Wing Area 20 m² / 215 ft²
Engine None
Empty Weight 300 kg / 655 lbs
Maximum Takeoff Weight 460 kg / 1,015 lbs
Maximum Speed 560 km/h / 350 mph (gliding)
Landing speed 72 km/h / 45 mph
Release altitude 8,000 m  (26,240 ft)
Crew 1 pilot
Armament
  • None

 

Theoretical Estimated Lippisch P 13 Specifications

Wingspans 5.92 m / 19  ft 5 in
Length 6.7 m / 21  ft 11 in
Height 3.18 m / 10 ft 5 in
Wing Area 20 m² / 215 ft²
Engine Unspecified ramjet
Maximum Takeoff Weight 2,300 kg / 5,070 lbs
Maximum Speed 1,650 km/h / 1,025 mph
Flight endurance 45 minutes
Fuel load 800 kg / 1,760 lb
Crew 1 pilot
Armament
  • None mentioned

Illustrations

The Lippisch DM-1, unnecessary to the overall project, it none the less allowed a group of students to escape military service.

 

A possible silhouette of the P13.

Credits

  • Article written by Marko P.
  • Edited by  Henry H.
  • Ported by Marko P.
  • Illustrated By Medicman11

Source:

  • A. Lippisch (1981) The Delta Wing History and Development, Iowa State University Press
  • D. Nesić  (2008)  Naoružanje Drugog Svetsko Rata-Nemačka. Beograd.
  • D. Monday (2006) The Hamlyn Concise Guide To Axis Aircraft OF World War II, Bounty Books.
  • J. R. Smith and A. L. Kay (1972) German Aircraft of the WW2, Putham
  • B. Rose (2010) Secret Projects Flying Wings and Tailless Aircraft, Midland
  • D. Sharp (2015) Luftwaffe Secret Jets of the Third Reich, Mortons

 

 

WM–23 Ezüst Nyíl

Hungarian Flag Kingdom of Hungary (1939)
Fighter Aircraft – One prototype

In their search for a new fighter, the Magyar Királyi Honvéd Légierő MKHL (English: Royal Hungarian Home Defence Air Force), approached the Germans for help. Initially, a deal was made with the German Heinkel company for the delivery of new He 112 fighters and a production license. However, nothing came of this deal, which led to the Hungarians attempting to develop their own fighter, partially based on the He 112.

The WM–23 Ezüst Nyíl prototype. Source: www.destinationsjourney.com

The He 112 In Hungary

In the late 1930s, the Hungarian Air Force was slowly in the process of rebuilding its combat strength by the acquisition of new aircraft. For a modern air force, they needed better fighter designs, which they were then seriously lacking. Luckily for them, they began to improve their relations with Germany, so it was possible to acquire new equipment from them. In June 1938, a  Hungarian delegation was sent to the Heinkel company, and the pilots that accompanied this delegation had a chance to fly the He 112 fighter. This aircraft was  Heinkel’s response to the Reichsluftfahrtministerium’s (English: German Ministry of Aviation) request for a new fighter. While generally a good design, it ultimately lost to Messerschmitt Bf 109. While the He 112 project was canceled by the RLM, to compensate for the huge investment in resources and time to it, Heinkel was permitted to export this aircraft to foreign buyers.  Several countries such as Austria, Japan, Romania, and Finland showed interest, but only a few actually managed to procure this aircraft, and even then, only in limited numbers.

He 112 the unsuccessful competitor of the Bf 109. Source: www.luftwaffephotos.com

The Hungarians were impressed with the He 112 and placed an order for 36 such aircraft. For a number of logistical and political reasons, the decision to sell these aircraft to Hungary was delayed.  A single He 112 was given to Hungary in February for evaluation but was lost on its first flight. Realizing that the Germans would not deliver the promised aircraft, the Hungarians instead decided to ask for a license. This was granted and Heinkel also delivered two more He 112 B-1s. When the license arrived in Hungary in May 1939, a production order for the 12 first aircraft was given to Weiss Manfréd aircraft manufacturer.

The first He 112 to reach Hungary, it was lost in an accident during its maiden test flight. Source: D. Bernard Heinkel He 112 in Action

A New Fighter

Despite the best Hungarian attempts to put the He 112 in production, this was prevented by the war between Poland and Germany. At the start of the Second World War, RLM officially prohibited the export of any German aircraft engines and equipment. This meant that the vital Jumo 210 and DB 601 engines would not be available. Based on this fact, all work on the Hungarian He 112 had to be canceled.

The Hungarian Air Force only operated a few He 112, which saw limited service before being reused as training aircraft. Source: www.destinationsjourney.com

As the Hungarians had the license for the He 112, some parts could still be domestically produced. In essence, this offered the Hungarians the chance to develop a new fighter, based on the He 112 blueprints. Not wanting to waste this opportunity, the Hungarian Ministry of War Affairs issued a directive to commence developing a new domestic fighter by reusing some components from the He 112. The whole project was undertaken by WM’s own chief designer Bela Samu, who began development in early 1939. To speed up development, the He 112 wing design was copied, but given the comparatively underdeveloped Hungarian aircraft industry, the wing was to be built of wooden materials instead of metal, as it was on the He 112. Other differences included using an oval-section fuselage, different armament, a new engine, and a cockpit redesign.

The first prototype was completed quickly by the end of 1939. In its prototype stage, the aircraft was painted in a light gray livery, earning it the nickname Ezüst Nyíl (English: Silver arrow)  from the personnel that worked on it. Once it was issued to the Air Force for testing, it received the standard Hungarian camouflage scheme, and the designation V/501 was also allocated to it. The maiden test flight was undertaken close to Budapest on the 23rd of February 1940. The flight proved successful and a maximum speed of 530 km/h (330 mph) at a height of 5 km (16.400 ft) was achieved. Some issues were detected, the most problematic proved to be the strong vibration caused by the exhaust system. Despite this, the project development pressed on.

The WM 23, possibly at an early stage of development. It was powered by a 1,030 hp WM K-14B engine which gave it comparable power to the He 112. Source: www.destinationsjourney.com

Short Service Life

Despite the time and effort put into the project, it all went for nothing as the prototype was lost in an accident in February, or April, depending on the source, 1942. During a test flight at high speeds, one of the ailerons simply broke off. The pilot lost control of the aircraft and had to bail out. The uncontrolled plane hit the ground and was utterly destroyed, and with it, the whole project was canceled.

Beyond this major setback, another reason why this project was canceled was the start of the license production of the German Bf 109G fighter. It was much easier, and faster, to commence production of this aircraft, thanks to German technical support, than to completely develop new tooling and equipment for the WM 23.

Technical Characteristics

The WM 23 was a mixed-construction single-engine fighter heavily inspired by the German He 112. Given its somewhat obscure nature, not much is mentioned in the sources about its overall construction. Given the urgency of the project, instead of the monocoque fuselage, the Hungarian engineers decided to use a simpler oval-section fuselage which consisted of welded steel tubes and then covered with plywood. The wings, as mentioned, were taken from the He 112, but had one huge difference, being made of wood, including its control surfaces.

The landing gear was another part more or less taken directly from the He 112. They consisted of two larger landing wheels that retracted into the wings, and one semi-retractable tail wheel. But based on the photographic evidence, their overall design changed during the prototype’s development. On the prototype, possibly at an early stage, a V-shaped front landing gear strut was used. This was later replaced by a large single-leg landing gear. The cockpit was equipped with a sliding canopy that slid to the rear.

The WM 23 was powered by a 1,030 hp WM K-14B (sometimes marked as 14/B) engine. This engine was developed based on the license of the  French Gnome and Rhone 14K engine, a fourteen-cylinder radial engine equipped with a single-stage, single-speed supercharger. As mentioned, during the fifth test maximum achieved speed was 530 km/h (330 mph).

While the prototype was never fitted with an offensive armament, the Hungarians had plans for a potential armament In the wing, two 8 mm (0.33 in) machine guns were to be installed. In addition, two 12.7 mm (0.5 in) heavy machine guns were to be added atop the engine compartment. Lastly it was to have a  payload of two 20 kg bombs  (44 lbs).

 

There are very few surviving photographs of the WM 23. While showing promising performance, the destruction of the only prototype and the commencement of the Bf 109G’s production in Hungary ultimately lead to the cancelation of this project. Source: D. Bernard Heinkel He 112 in Action

Conclusion

The WM 23 was an interesting Hungarian attempt to domestically develop and build a fighter aircraft that was greatly influenced by the He 112. It showed to be a promising design, with the prospect of entering serial production. However, the loss of the single prototype put an end to this project. By 1942, the Hungarians simply did not have the time to start over again with the WM 23, so they abandoned it in favor of the license production of the German Bf 109G.

WM 23 prototype Specifications

Wingspans 31  ft 5  in / 9.6 m
Length 29  ft 10 in / 9.1 m
Height 10 ft 9 in  / 3.3  m
Wing Area 199  ft² / 18.5 m²
Engine One 1,030 hp strong WM K-14B
Empty Weight 4,850  lbs / 2,200 kg
Maximum Take-off Weight 5,733  lbs / 2,600 kg
Maximum Speed 330 mph / 530 km/h
Crew 1 pilot
Proposed Armament
  • Two 12.7 mm (0.5 in) heavy machine guns and two machine guns  8 mm (0.33 in) machine guns plus a bomb load of 20 kg (44 lbs)

Credits

  • Article written by Marko P.
  • Edited by  Henry H.
  • Ported by Marko P.
  • Illustrated By Carpaticus

Illustrations

The WM-23’s factory test colors
WM-23 with Hungarian Airforce livery

Source:

  • D. Monday (2006) The Hamlyn Concise Guide To Axis Aircraft OF World War II, Bounty Books
  • D. Bernard (1996) Heinkel He 112 in Action, Signal Publication
  • G. Punka, Hungarian Air Force, Signal Publication
  • R.S. Hirsch, U, Feist and H. J. Nowarra (1967) Heinkel 100, 112, Aero Publisher
  • C. Chants (2007) Aircraft of World War II, Grange Books
  • J. R. Smith and A. L. Kay (1990) German Aircraft of the Second World War, Putnam

 

 

 

 

 

Douglas F3D, F-10 Skyknight

United States of America (1951)

Nightfighter & Electronic Warfare Aircraft: 265 built

Skyknights on the muddy ground of K-6, Korea. (USAG Humphreys)

Designed after the Second World War, the Douglas Skyknight was meant to be the defender of the American carrier group after dark. The ambitious design sought to use all of the lessons learned from night fighter design and tactics in the Second World War, and produced the first specialized all-weather jet fighter. While it would prove too cumbersome for use on the smaller WWII era fleet carriers, the Skyknight would prove to be an exceptional nightfighter when disembarked during its combat debut in the Korean war. Yet beyond the conflict, and entering rapid obsolescence as a fighter in the 1950s, the Skyknight would prove itself to be an able electronic warfare aircraft and a pioneering aircraft in its field.

American Night Fighter Experiences in WWII

The more advanced, radar equipped night fighter of the Second World War was an ad-hoc creation which combined state of the art airborne detection systems with often pre-existing fighter designs. The resultant creation was an interceptor capable of bringing down enemy aircraft at night, or in very poor weather. While they carried their own radars, the limits of their range required they be directed by ground controllers into the path of their target. Alternatively, they could be tasked with offensive patrols to harass the enemy over their own airfields or carry out ground attack missions. In either case, sorties were demanding, with the crews of these aircraft having to endure patrols of several hours, flying almost entirely by instruments, and occasionally in extreme weather conditions. Even without the dangers of aerial combat in the dark, flying these aircraft was often exhausting and dangerous work. Vertigo was as deadly an opponent as any enemy they might encounter. Unsurprisingly, the best qualities that these aircraft could possess were straightforward flying characteristics, like good handling and stability. A good night fighter needed to be uncomplicated, and forgiving in how it flew.

The P-61 was a massive, but fairly short ranged night fighter with a mixed service record between the European and Pacific theaters. (san diego air and space museum)

The American nightfighters of the Second World War can largely be broken down into two groups. These were the heavier, twin engine, land based types in use with the Army Air Force, and the lighter carrier-embarked forces of the Navy. The Air Force’s principle nightfighter was the Northrop P-61 Black Widow, a purpose built, if somewhat over-engineered design that saw use across Western Europe and the Pacific. The design was cutting edge, featuring a state of the art airborne intercept radar system, and an impressive, if totally unnecessary, remote controlled gun turret. It flew exceptionally well, was nimble beyond what its size would suggest, and was fast enough to catch all but high flying, fast recon aircraft. Yet the design had two serious limitations. Foremost was its disappointing endurance, as in spite of its size, its limited fuel capacity and massive Pratt and Whitney R-2800 engines meant it had a range comparable to many single engine fighters. This was partially resolved by the installation of wing pylons which could fit either fuel tanks or bombs, though having to choose between ordinance or range imposed significant mission limitations. Less serious was its poor crew layout. As designed, the pilot and radar operator sat at opposite ends of the fuselage, hampering communication and, in the event of an accident, the loss of the intercom completely isolated the radar operator from the other two crewmembers. This limitation was overcome by the crews of the 425th Night Fighter Squadron, who moved the radar operator’s equipment to the gunner’s position. However, this modification was almost entirely limited to the European theater.

While the P-61 proved a capable night fighter, and an excellent all weather ground attack aircraft, there was much to be learned from the Mosquito Night Fighter Mk. 30’s that were made available to American crews near the end of the war. The Mosquito featured a side-by-side pilot and radar operator arrangement, and a large internal fuel capacity that gave it excellent range without having to install external fuel tanks. While it was less maneuverable, it was arguably the best night fighter of the war, capable of pursuing targets over long distances and attacking enemy rear line airfields at night, without having to sacrifice ordinance for range.

In all, the experiences of the P-61 crews were mixed. In Europe, they provided good night cover for the Army Corps they were assigned to, and during the siege of Bastogne, they were among the only fighters providing protection and support to the beleaguered Airborne forces in the city, when poor weather kept all but a handful of aircraft grounded. In the Pacific, they were less successful, particularly towards the end of the war. During the battle for the Philippines, they often struggled to deal with the swarms of Japanese fighter bombers that flew dawn and dusk attack missions. The P-61’s were never designed or intended to defend against such forces, and found them a challenge to bring down. Where in Europe they gained the personal thanks of the commanding officer of the 101st Airborne Division within Bastogne, they were the target of General MacArthur’s personal frustrations as his beachheads were continually harassed by Japanese forces.

Small composite units of F6F and F4U Corsairs operated aboard US carriers for night defense. They proved just as capable when deployed ashore. (wikimedia)

In contrast to purpose built P-61, the night fighters of the US Navy were fairly simple conversions of existing fighter planes converted to serve as a defense for carriers at night. Variants of the F5F Hellcat and F4U Corsair were fitted with small, wing mounted radar sets to allow them to track and engage targets at night or in poor weather at the direction of ground controllers. They were far simpler aircraft, and generally were tied down more heavily to their ground controllers as a result of their shorter range, and simpler radar systems. Within the fleet, the duty of these night fighters was to contend with enemy aircraft that attempted to attack naval vessels under the cover of darkness. When assigned to land based Marine corps aviation, they were typically charged with the protection of amphibious operations and providing air cover for important installations. In both cases, these light night fighters proved very successful, and in the case of the Philippines, F6F Hellcat night fighters ended up replacing P-61’s as the defenders of the beach head. However, the limitations of the single engine fighters left the navy wanting something more. The Hellcat and Corsair night fighters were fast, but they had a fairly short range, and lacked a dedicated radar operator. The benefit of the heavier night fighter was its ability to more easily re-acquire targets which may have evaded the first attack and longer endurance, which allowed it to pursue and catch fleeing targets over an extended chase.

The new F7F-N was hoped to be the ultimate carrier based nightfighter of the Second World War, carrying both a second crewman to operate radar and navigation systems, and having a significantly better range. However, it was not to see use during the war, it was too large, and it was soon to be obsolete. The piston engine was being superseded by the jet turbine, and the carrier air wing of the future would soon need an aircraft to contend with threats far faster than any of their existing fighters.

Douglas’ Dark Knight

In August of 1945, at the very end of the Second World War, the Navy’s aviation bureau set its requirements for a new carrier based, jet night fighter. It called for a top speed of 500 mph (805 km/h), a service ceiling of up to an altitude of 40,000 ft (12,192 m), and a 125 mile (201 km) radar intercept range. Beyond its performance requirements, it was to also carry a pressurized cabin with temperature controls, and a robust set of de-icing equipment. Four companies presented bids, these being Douglas, Curtiss, Grumman, and Fleetwing, with preliminary design work beginning in October. By April of the following year, the contest was over, with only Douglas’ proposal receiving a letter of intent, the resources to construct three prototypes, and ground testing materials. Its daylight counterpart was to be the Grumman F9F Panther, with the two fighters being poised to propel the US carrier forces into the jet age.

Designer Ed Heinemann with his 88 Oldsmobile alongside the Skyknight. (smithsonian)

The design dubbed ‘Skyknight’ incorporated many of the lessons learned from the Second World War. The aircraft’s chief designer was the prolific Ed Heinemann, who designed some 20 aircraft through a very productive career. Some of his most notable creations were the venerable SBD Dauntless, AD Skyraider, and A-4 Skyhawk. His Skyknight featured a cutting edge search radar which was operated by a crewmember seated beside the pilot, allowing for easier communication. The radar was also to feature a fire control system which gave the pilot an accurate lead on his target, allowing him to engage maneuvering targets and those that were not visible to him. They attempted to fully resolve the drawbacks of the Black Widow by installing large internal fuel stores, but the high fuel consumption of the turbojet engines meant the aircraft had comparable range to most other jet fighters of the era. Beyond its combat ability, it was to be a very straightforward aircraft to fly, with stability at both extremes of its speed limits. Its only eccentricity was that it had an ejection chute as a means of crewmen to escape the aircraft in an emergency. Altogether, it was a conventional, honest aircraft that flew well.

The first XF3D-1 prototype was flown by test pilot Russ Thaw on March 23, 1948, with the second flight following in June, and the final in October. Apart from basic safety and performance tests, the aircraft was flown in mock intercepts against single seat jet fighters. Even with their World War II era SCR 720 radar, they easily managed 85 mile (136 km) intercepts with GCI support. The Air Force also conducted tests on the aircraft, and inquired about installing the fighters with afterburning engines, but declined and instead developed the all-weather F-89 Scorpion.

The Skyknight was a remarkably stable and maneuverable aircraft. (aerocorner)

After satisfactory land based flight testing, the Skyknight received a production contract. The F3D-1 would replace the prototype Westinghouse J-34-WE-22 engines with more powerful WE-38’s, and the WWII era SCR-720 radar system was replaced by the new AN/APQ-35, a change made in the third prototype. The new radar boasted a much longer effective range, and was the first airborne radar with a lock-on feature, which allowed for the continuous, automatic tracking of a radar contact. Modifications continued to be made on the F3D-1’s as the engines were again changed to the J34-WE-34’s and their plastic-glass nosecones were switched to fiberglass.

The Skyknight was built around the concept of the heavy night fighter, and was thus at the limits of how large a carrier-borne aircraft could be. In an age where carriers were originally designed for single-engined, piston aircraft, the F3D-1 would prove rather troublesome. The comparatively massive Skyknight was difficult for carrier deck crews to maneuver about the ship, and prepare it for launch. The bridle, which connected the nosewheel to the catapult, needed to be significantly stronger than those used for other navy fighters, and the proximity of the wheel to the engine intakes required a greater level of safety, and these precautions lengthened launch procedures. The landing gear shock absorbers too were also judged inadequate, as the plane bounced excessively during arrested recovery, and high vibrations were noted. This was particularly worrying, as the Skyknight’s radar system used vacuum tubes, and was quite fragile. These, and other problems, saw the F3D rated for marginal daylight use and was prohibited from launching and recovering at night.

Another major challenge was to be found in training programs, both for pilots and radar operators. The training program for radar operators was notably lacking, and provided no specialized training for the complex radar systems aboard the Skyknight. A result of an underappreciation for how sophisticated the job was, and a lack of funding.

The Skyknight’s were significantly larger than all previous carrier borne aircraft. This often proved challenging on carriers designed for single engine piston aircraft. (Naval Air Museum)

Landing issues were resolved in the subsequent, and final production model, the F3D-2. The new model was designed primarily to get larger, more powerful engines into the aircraft, though the J46 engine they were slated to receive never materialized. However, they still represented a serious improvement over the first model, as they were equipped with an improved version of the J34 engine, an autopilot, gun laying radar, tail warning radar, wing spoilers to increase the aircraft’s roll rate, and they received the modifications that would get them cleared for their full use aboard aircraft carriers. They soon succeeded the small number of F3D-1’s built, with the first aircraft being flown in February 1951.

The first of the new F3D-1 Skyknights took flight on February 13th of 1950, with the Navy accepting the first deliveries, which were then turned over to Composite Squadron VC-3 in December of the same year. One of Douglas’s test pilots, LaVerne Brown, would give the Naval Aviators an introduction to the aircraft. The Skyknight would fully enter service in February, with the aviators familiarizing themselves with the new aircraft, and being bolstered by another combat squadron, VC-4. The new squadron would be the only one to actually be deployed aboard carriers.

They would not prove ideal. The dimensions of the aircraft proved problematic, being far larger than what the deck crews were accustomed to, and they were occasionally mishandled, resulting in minor damage. The weight of the aircraft also complicated the use of the carrier’s catapults. The H-8 hydraulic launching system needed to be used close to its maximum power setting to launch the Skyknight, and if the bridle was not well connected to the aircraft, the catapult hook could break free, and be sent hurtling toward the end of the track at the bow of the ship. The collision would necessitate a lengthy repair, and during their deployment aboard the USS Lake Champlain, it happened twice, much to the frustration of all aboard. Lastly, the low mounting position and slight downward cant of the aircraft’s engines baked the wooden deck of the carriers and had a habit of setting alight any flammable materials which may have leaked from any of the planes or machinery present. These were never large, or particularly dangerous, but any conflagration on the deck was met with an alarm and the entire ship was sent to fire quarters. The Skyknight’s were seen as extremely inconvenient, and frustrating to the carrier’s commanders, who were also very unhappy that the plane had only one, very specialized use. The night fighters and their crews quickly became the black sheep of the air group, even to the other pilots.

The size of the new Skyknight earned it the nickname “Willie the Whale”, with ‘Whale’ slowly overtaking Skyknight as the crew’s preferred moniker. (US Navy)

Pilot’s views of the new aircraft were mixed. The Skyknight was like nothing naval pilots had flown before, and not only because it dwarfed every other plane on the flight deck. It lacked all the familiar trappings of a navy fighter, and if anything it reminded many of them more of a transport aircraft than any fighter they had ever flown. Beyond that, the tandem seating arrangement proved unique, as did the spacious, carpeted, air conditioned canopy equipped with a cigarette lighter and ashtrays. While the Skyknight was not proving to be the answer to after hours protection the US Navy was looking for, many of those assigned to the new jet could not help but be fascinated. They also soon found they could not help but be frustrated when they were asked countless questions about what the plane could not do and the capabilities it did not have. Unlike the other fighters of the US Navy, the Skyknight was not a fighter bomber, but that never stopped the press from asking questions about how many bombs it could carry, or their commanders asking if they could perform daylight strike and patrol missions.

The Skyknight was a bad fit for the carriers of its day. It was far too large and did not have the versatility that might have justified its many inconveniences. It was the bane of the carrier air group, and left deck crews and other pilots irritated, as it meant more work for them. It was a massive, catapult breaking, deck burning, fire starting annoyance that did only one mission. In spite of this, the Skyknight was to become one of the most exceptional aircraft of its day.

While Naval aviation found the Skyknight totally unsuitable for their purposes, the Marine Corps were eager to get ahold of them. The Marines still flew their piston engined Tigercats and Corsairs, and believed wholeheartedly in the two-man nightfighter concept the Navy still was not entirely sold on. As opposed to the Navy, with jet fighters aplenty, the Skyknight represented a massive upgrade for the Marines, who already flew a fair share of oddball planes. It was thus in the Corps that the Skyknight found its new home, and would soon demonstrate itself as an incredible nightfighter.

The Forgotten War

Following the end of the Second World War, the Korean Peninsula was controlled by a combined US-Soviet commission, which eventually saw the creation of two governments on either side of the 38th parallel, the formal boundary for Soviet-American jurisdiction. The American supported Republic of Korea was founded in the South, and the Soviet aligned Democratic People’s Republic of Korea was founded in the North. In spite of what the names might suggest, both regimes were dictatorships, and neither recognized each other’s legitimacy. Clashes occurred at the border, and the DPRK was emboldened when the US and Soviet forces withdrew in 1948. North Korean leader Kim Il Sung was confident his forces could reunite the country in a decisive military campaign, and received permission from Soviet leader Joseph Stalin to launch the invasion. On June 25, 1950, an artillery barrage heralded the start of the war as the DPRK’s forces pushed South, and their marines made landings along the Eastern Coast.

The war would escalate into an international conflict that brought in the Soviet Union, China, and the United States with its many allies under the banner of the newly formed United Nations. The air war over the peninsula was an odd affair, with several of the air forces involved having only recently been formed, and flying a mix of World War II era and modern jet aircraft. The People’s Republic of China had only been founded in 1949, with an airforce so new it had not even been fully organized by the start of the war. With Soviet support, they received training and aircraft, becoming a fully realized military force by the end of the war. The DPRK was likewise supported, and possessed a force of WWII era fighters and ground attack aircraft of Soviet make. The Soviets themselves sent pilots and aircraft, seeing it vital that they gain some experience in what was becoming the first modern air war. They would, however, maintain that they were never directly involved, with their pilots officially flying with the Chinese air force.

The UN Forces would operate an eclectic mix of aircraft. Here an outdated F-80 Starfighter shares the ramp with a modern F-86 Sabre. (National Archive)

The UN forces were backed by the largest airforce in the region, the US Far East Air Force, stationed in Japan. The force mostly operated the then obsolete B-29 and B-26 bombers, and F-80C jet fighters. It was by far the most powerful air force in the region, but unsuited for tactical support missions. The Air Force was supplemented by the US and Royal Navies with their carriers, and later, disembarked Marine aviation forces.

Technologically, the forces involved used both the crudest and most cutting edge equipment available. The Soviet aligned forces were, initially, almost entirely dependent on older WWII era stock, their main fighters being Yak-9Ps and La-9’s until attrition ground them away after several months of fighting. After roughly a year, they began to be replaced by the cutting edge Soviet MiG-15, which allowed North Korean and Chinese pilots to claim a level of parity, even as they were largely overwhelmed over much of the peninsula. The disparity in numbers would force them into a defensive strategy which involved a great deal of night operations, and basing nearly all of their new MiG’s in China, where their airfields could not be stalked by American fighters.

The UN forces flew a bewildering variety of propeller and jet aircraft, especially when compared to Communist forces, who by the middle of the war were flying little else than MiG-15s and light ground attack planes into combat. Once their forces were better established in the theater, American air forces pursued an offensive anti-air campaign over the northern half of the peninsula using their own cutting edge F-86 Sabre, while swarms of piston engined F-51 Mustangs, F4U Corsairs, and new Skyraiders were used for close air support, and massed B-29’s were flown against strategic and tactical targets.

The Soviet MiG-15 proved an incredible upgrade over the outdated Yakovlev piston engined fighters, matched in performance only by the Air Force’s F-86 Sabre. (Museum of the US Air Force)

These strategic raids were much the same as those of the Second World War. The Superfortresses targeted factories, power generating infrastructure, and bridges, though the inaccuracy of their methods left much of the northern half of the peninsula in ruins. In an effort to stop the raids, the Communist forces used their new MiG 15’s as interceptors, and could comfortably attack these formations with their combination of heavy cannons, and near unapproachable speed. Only the less common American F-86 Sabres were fast enough to catch them, and thus any real hope of keeping the B-29’s safe during daylight hours was gone. Their solution was to transition to night bombing, which would eliminate all but a few very specialized Soviet MiG 15 crews from being able to intercept them. This nocturnal shift in the war over the peninsula saw night fighting transitioning from a mostly tactical affair, involving aircraft raiding or defending frontline positions at night, to a strategic one that pitted each side’s most advanced aircraft against one another over control of the northern half of the peninsula.

Sallying Forth

Marine Nightfighter squadron VMF(N)-513 arrived in Korea in August of 1950 with a dozen Corsair night fighters, and a very difficult job to do. The pilots of the ‘Flying Nightmares’ flew night ground attack sorties in their WWII era fighters. After shuffling from airfield-to-airfield as the Chinese army began its southern march, the unit was reinforced by VMF(N)-542, most of which was returning stateside. The Nightmares received new pilots from the retiring squadron, and some twelve heavy F7F Tigercat night fighters. With them came a new job, night interdiction, which proved to be more dangerous, but much more important. The UN forces had air superiority over much of the peninsula, and thus the Communist forces took to moving most of their supplies at night, often in well armed, well protected convoys. For two years, the Nightmares flew some of the most dangerous missions of the war, with 54 aircraft being lost to all causes. It was in June of 1952 that the squadron was resupplied, again given a new aircraft, and a new mission.

The F4U-5N was the most common American night fighter of the early Korean war. They were primarily tasked with night ground attack missions. (Wikimedia)

While they had received planes and pilots from the 542 in-theater, a cadre of that squadron had retrained on the new F3D Skyknight. They would join the Flying Nightmares in June, bringing fifteen new night fighters, and shortly after, retiring the squadron’s Corsairs. As the Skyknight was virtually useless in an air to ground role, their task was to be the escorts for the air force’s B-29 raids over Northern Korea. They would, however, not enter combat for some time, as the cadre had not been equipped with the blast tubes for their 20 mm cannons. Lt. Col. Lambrecht would take charge of the deployed unit, now with 12 aircraft, 3 having been retained in Japan.

The unit would quickly install the tubes after they arrived on August 5th, with the first combat sortie penned for the 7th. It was to be flown by a joint Royal Air Force-USMC crew, with RAF pilot Squadron Leader John Gardener, and Marine radar operator Staff Sergeant Kropp taking the Skyknight up on its first patrol. It was a local patrol mission, and apart from investigating an unidentified IFF emergency code, not much occurred. Over the next several days, more missions were flown, though no enemy aircraft were intercepted. As it was in the Second World War, night intercepts were difficult, and any failures on the part of the ground based radar director, or the RO on the plane, could result in a botched intercept. Even with the new radar, closing with the target was still a challenging affair that tested the pilot and radar operator alike. It was clear that even with new technical advances, bringing down enemies at night would require a mastery of the equipment, and excellent coordination between all parties.

Having been mostly discarded by the Navy, the Skyknight soon found itself among other oddball aircraft in the inventory of the Marine Corps. Unlike the Navy, the Marines were not ones to turn away an offer for new jet fighters. (Smithsonian)

The enemy they chased was typically in one of two groups, either cutting edge MiG 15’s that were usually flown by Soviet pilots, and rarely encountered outside of the North, or very light trainer aircraft flown in a ground attack role. These were usually Yakovlev 18’s, or the exceedingly obsolete wood and fabric Po-2 biplane. Rarely were these attacks very serious, though their frequency earned them the moniker “bed check Charlie”, a title formerly held by Japanese night raiders of the Second World War. The Skyknights rarely encountered the light piston engined planes, and the MiG’s were their primary opponents.

They lacked radar, but the Soviet pilots were well trained in instrument flying and were proficient in ground directed radar intercepts. They were thus reliant on a local radar, and the tell-tale glow of the Skyknight’s turbojet engines to attack them. Their most effective tactic was a trap in which one MiG flew a straight and level course, while a second trailed it at a lower altitude. Should the first plane find itself pursued, the ground radar would warn them to speed up, and direct the second aircraft to climb and attack the pursuing Skyknight. As the American night fighter had a tail radar, it was often forewarned of the approach of the trailing MiG, but on one occasion, the Soviet pilot claimed a victory. The other threat to the Skyknight were radar directed searchlight traps, which disoriented the crew while AAA batteries attempted to bring them down. This proved far less dangerous than the MiGs.

The Skyknight would prove to be one of the only two aircraft to challenge the MiG-15, though unlike the faster Saber, it relied on its sophisticated radar systems to ambush the mostly blind MiGs. (USAG Humphreys)

While the Skyknights of the 513th were working themselves into combat, a pair of incidents would leave a dark mark on some of the unit’s early service. On August the 15th, the Squadron’s CO, Col. Lambrecht disappeared while on patrol from Kusan, and the Corsair sent to search for him failed to identify any wreckage. On the 1st of September, a catastrophic engine failure brought down another Skyknight. Flown by pilot Maj. Harrold Eiland with his RO, MSgt. Alois Motil, the plane’s starboard engine experienced power fluctuations before breaking down. A clanging noise alerted the crew, as the RPM gauge and fire alarms remained steady. Then the port engine failed, and the plane lost all thrust. As the plane was flying out from the airbase, it fell into the sea, and only Motil escaped the crash. A two month investigation grounded the planes until the culprit was found. It proved to be a turbine compressor failure, which sent shattered turbine blades through the fuselage and into the second engine. While local flights were still carried out, combat patrols would not be flown again until October 17th, when armor plates were installed aboard the aircraft.

The Nightmares wasted no time, and once they were airborne again, they took on the job of escorting the Air Force’s bombers under the leadership of Lt. Col. Homer Hutchinson, who succeeded the late commander in early September. He was notably a much more aggressive commander, who tasked his pilots with seeking out enemy road traffic on their return from their escort missions. Their first escort mission was conducted on November 3rd, to cover the areas where the Air Force’s night fighters were not permitted. Their F-94 Starfire carried sensitive equipment, and could not be directed over Communist held territory.

The Skyknight’s escort strategy mirrored the RAF’s nightfighter tactics of the Second World War. Upwards of six fighters were flown on separate tracks to find and bring down the enemy. One group flew barrier patrols between the bombers and known enemy fighter bases, a second group flew with the bomber formation, and the final group flew over the bomber’s target area. A typical escort operation involved nine Skyknights.

The first victory was soon claimed, with Maj William Stratton and RO MSgt Hans Hoglind catching an enemy at 14,000 ft (4267 m). They struck the aircraft with 20 mm cannon fire, hitting the port wing, fuselage, and tail pipe, with the burning plane shortly descending rapidly out of sight. It was claimed as a Yak 15, but declassified Soviet records identify the aircraft as a MiG 15 flown by Capt. V. Vishnyak, who survived and brought the wrecked MiG home. The squadron’s second victory came on the 8th, when Capt. Oliver Davis and his RO Dramus Fessler were vectored on to a target. The enemy noticed them and attempted to evade, though Davis turned with them and fired. Several shots struck the rear of the enemy aircraft, which set fire to its engine, and they saw it lose control, before plummeting to the earth. The plane belonged to Soviet pilot Lt. Ivan Kovalov.

These new victories inspired great confidence after the incidents of the previous Autumn. Now going into winter, the Skyknight crews of the 513th settled into a routine of escort, and offensive patrols. Between November and January, they claimed four enemy jets and were getting a better handle for the ordeal that the escort mission was soon proving to be. The massive number of aircraft airborne, and the limited number of ground directors meant that communications with GCI operators were heavily strained. Coupling that with the task of navigating the predetermined patrol areas for about two hours, it all added up for a demanding job for pilot and RO alike.

In spite of all that, they proved extremely successful. The Skyknight was proving to be an exceptional night fighter, and was conducting patrols over Northern Korea with impunity. The only real threat were MiG traps, which could only be conducted in clear weather, and depended on perfect coordination between radarless planes and their ground controller.

In those first three months, bomber losses fell, and between February and July, no B-29’s would be lost to enemy fighters. The ungainly Skyknight, once considered almost useless by the Navy, was now proving itself indispensable to night operations over Korea.

The Long Haul

The Nightmares would be joined by another Skyknight unit in the Summer of 1953. VC-4, detachment 44N, the ‘Nightcappers’, arrived in Korea aboard the USS Lake Champlain in early June of 1953. The four planes were proving an absolute nuisance to the operation of the carrier, and they were prevented from flying as much as possible. The detachment’s officer, Lt. O’Rourke, would try his best to argue for more flight hours, in order to simply retain their proficiency. They would fruitlessly attempt to fly daylight missions, after the Carrier Air Group commander did everything possible to prevent them from flying at night. The commanders of the carrier wished to be rid of the planes. The Communist air forces lacked the strength to attack an American carrier in daylight hours, much less at night.

K-6 was a perpetually rainy airfield that hosted a mix of Naval Aircraft. (USAG Humphreys)

After being effectively grounded aboard the USS Champlain, O’Rourke successfully petitioned for the unit to be sent ashore to join the Marine aviators. They settled into airfield K-6, alongside the 513th, and were quickly worked into their schedules. Settling in proved a challenge, as they traded their carrier berths for quonsets, at the rainy, muddy airfield outside of Pyeongtaek. They drew Marine fatigues, though rain gear was in high demand and low supply. It rained constantly and the airfield had a permanent muggy atmosphere, which made landing more difficult, and keeping dry an impossibility. The two were combined during the frequent airstrip overruns, when the planes rolled off the tarmac and into the mud. It was rarely a dangerous affair, though it was always a cold and dirty job dragging out the stuck aircraft. The Navy aviators would also soon learn that the Skyknight had been banned from most airfields in Korea, with the exception of emergencies, as its low mounted engines gouged holes in asphalt as easy as it had baked the wooden decks of the carriers.

The culture shock would also require a good deal of adjustment. Whereas the carrier was well regimented and ran with a clean and ordered efficiency, the Marine Corps was a force which took the odds and ends it was given, and made due. Perhaps the best example came down to how to cut the engine tail pipe to size, so as to have the exhaust be the right temperature. If it was too hot, the turbine blades could overheat and break, if it was too cool, it would not produce anywhere near the amount of thrust it should. For the adjustment, the Navy used a prescribed manual for the process, they turned the engine on, checked the temperature with a specialized gauge, turned it off and let it cool, cut a section, and repeated the process for several hours until the numbers matched the manual.

The Marines turned on the engines to full power, and checked the temperature using the cockpit instruments. If it was low, they shut off the engine, and took a pair of large pliers and bent crimps into the hot tailpipe to shrink the diameter of the outlet. They then turned it up again to full power until the temperature was at least 40 degrees above the book’s absolute maximum allowable temperature. With that, they marked a red line at the max and told the pilots not to exceed that unless they absolutely had to. Doing this made their planes some 20 to 30 knots faster than their navy counterparts, regardless of how much rougher they were. It was harder on the blades and tended to scorch the pipes, but O’Rourke felt the extra performance could make the difference when trying to catch MiGs that held a confident speed advantage. This kind of resourcefulness would prove a necessity from operating from K-6, as spare parts were scarce, especially for the aircraft’s fragile, complicated radar systems.

In the end, they came together, and the Navy and Marine Aviators would be fully integrated and billeted together, in the words of O’Rourke with “no bitching”.

Toward Armistice

Reinforced, the 513th continued its job of MiG chasing. Their job remained the same, and they still had the same issues. GCI services were overburdened, and the radar station on Cho-do island missed a good deal of contacts. While on patrol, the long search range of the AN/APQ-35 was particularly useful, and crews reported spotting numerous contacts that the GCI stations never called out. In comparison, the MiG pilots enjoyed excellent radar direction owing to good training, a larger number of ground stations, and a defensive operations which made for easier planning. Their ability to react to the American night fighters led a number of aircrews to believe the MiGs began sporting radar sets, though this was never more than a rumor.

The number of intercepts of MiG 15s declined, though the aggressiveness of the Soviet, and by then some Chinese, pilots remained. They were learning the strengths of the aircraft, primarily its much higher top speed, and the tried and true tactic of diving and staying low so that the plane would become lost in the haze of radar ground returns. While they were getting better at escaping the Skyknights, they lost their chance to chase the bombers. B-29 ‘Double or Nuthin’ was the last to be shot down on the night of January the 29th, with all but one crewman surviving the war.

The final MiG kill likely belonged to a Lt.Jg Bob Bick, who had been determined to claim a MiG since arriving at K-6. He did so with CPO Linton Smith on July 2, after pursuing a contact, firing, and setting it ablaze. His next message to his GCI director was that his aircraft had taken several cannon hits, and Bick’s plane fell off the radar screens at Cho-do. Bick had fallen into a MiG trap, and though he had claimed the bait plane, the trailing MiG had him. In a unit as small as detachment 44N, his loss was felt hard. A second Skyknight failed to return from that patrol area two nights later, though there were no radio communications, or Soviet records, that might suggest a cause.

There was a superstition amongst the 513th with aircraft that bore a 13. Note the mud emblematic of runway overrun. (smithsonian).

Having successfully shot down one of the Skyknights, the Soviet crews felt a burst of enthusiasm and doubled down on the bait tactic. Three or more trailing aircraft replaced the lone tracker, and they flew out more frequently. The bait plane too embraced their role and made themselves as visible as possible. O’Rourke claims to have chased a MiG flashing its navigation lights, and as he closed to ID the plane, his tail radar alerted him to six pursuing fighters. He promptly broke off the engagement. In the last months, they failed to claim any MiGs, but they had completely stopped them from intercepting the B-29’s. In the final tally, the Skyknights claimed six MiG 15’s, and lost one of their own in combat, with another possibly sharing its fate. There were another four losses, attributed to accidents.

A less dangerous, but much more frustrating threat came in the form of the harassment attacks from the so-called ‘night hecklers’. By 1953, these were training aircraft, usually Yak 18’s and the comically outdated Po 2 biplane. Unlike the Yak 9’s and Lavochkin fighters that Chinese and Korean aviators flew earlier in the war, these light aircraft could be flown from pastures, hidden with ease, and could be flown so low that long range radar stations could not detect them. Apart from raising alarms, a number of them carried out a successful strike against UN force fuel reserves at Inchon. They were otherwise a threat only to a good night’s sleep, as their bomb loads were extremely light, and they were not putting their best pilots in these disposable aircraft. In addition to the AAA gun crews who had not had any targets for months, the Navy’s night fighter squadrons were called in to deal with the ‘hecklers’. There was some excitement among the aircrews, as the prospect of a defensive intercept was a new mission.

Excitement soon turned to disappointment. The pilots of the 513th expected calls to scramble and chase down contacts, but what they got were long nights playing cards in their full flight suits in the summer heat. The ‘hecklers’ were undetectable by radar, and there was rarely a forewarning of their attack. The Skyknight itself was also unsuitable for it, as the disparity in speed between the two aircraft meant the pursuer rarely had a chance to fire before they had to break off the attack to avoid collision. The first heckler was shot down in December, brought down by 1Lt. Joseph Corvi and MSgt. Dan George. It was even more notable, as the conditions were blind, and the crew downed the Yak 18 non-visually, using their radar. Apart from another probable kill, there was little luck to be found against ‘bed check Charlie’.

To better deal with them, Corsair and Skyraider night fighters were brought into K-6 from aircraft carrier dettatchements. These aircraft were handier at low speeds and had much better loiter time, so they could stay airborne and search for much longer. When they did pick up the enemy, they could stay on them as they stuck close to the terrain.

As the war came to a close, and an armistice was fast approaching, both sides fought tooth and nail for where the final North-South demarcation would lie. While diplomatic talks were underway at Panmunjom, the Skyknight’s mission soon changed. B-29 escort missions were over, as were patrols over the Yellow sea. They were to patrol the frontline, which proved extremely disappointing to the crews who were accustomed to owning the night skies over Northern Korea. Oddly enough, in the last week of the war, they were also tasked with ground attack missions, a job once reserved for the squadron’s now retired F7F Tigercats.

513th Squadron members alongside an F7F Tigercat and an F3D Skyknight. For a time, the squadron was flying Corsairs, Tigercats, and Skyknights, but by the war’s end, they were a purely jet aircraft operation (Smithsonian)

For the members of the 513th, the war ended at 2200 hours July 27, 1953. They soon transitioned to training operations, and DMZ no-fly line enforcement. This marked the end of the Skyknight’s surprisingly exceptional role as a night fighter. As an aircraft that had failed miserably in its planned purpose, the air crews of the 513th found in it something that could take them deep into enemy territory, and hunt the most dangerous opponent the war had to offer. Of the two aircraft that posed a threat to the MiG-15, one was a brand new, cutting edge interceptor in the form of the F-86 Sabre, and the other was an underpowered night fighter designed weeks after the end of the Second World War. It was a remarkable tool to a squadron that proved itself extremely flexible, flying three aircraft it had no pre-war training with in a damp and unforgiving environment.

Obsolescence and Testing

The Skyknight’s would remain in limited use as fighters after the Korean War, retiring from the role in the mid 1950’s. (US Navy)

The Skyknight was a dated plane even before it saw use in Korea, and by the end of the war, it was totally obsolete. Aeronautics was progressing in leaps and bounds, new fighters were breaking the sound barrier and mounting much more sophisticated radar systems, far better than the then archaic APQ 35. The squadron that had made a name with the Skyknight, VMA-513, dropped its ‘Night’ suffix when it traded its Skyknights for the Douglas Skyray, a supersonic, all weather interceptor. As it was slowly brought out of combat service, some 200 Skyknights were available for new jobs.

This saw an expansion of the Skyknight’s secondary role, flight testing. In addition to general aerodynamic and safety studies, the cavernous sections of the aircraft once occupied by its radar systems could be repurposed to carry equipment for any number of tests. Perhaps the most important of these was for the carrier automatic landing system. The Skyknight was the first aircraft to carry the Bell ALS, and in 1957, one was used to test the system aboard the USS Antietam. The system was designed to help guide an aircraft on the approach within plus or minus thirty feet (9 m) longitudinally, and twenty feet (6 m) vertically, to the arresting wires on the carrier deck. Ironically, an aircraft that proved so terrible for carrier service played a major role in developing one of the most important systems for modern carrier aircraft.

A lone Skyknight prepares to test Bell’s automatic landing system. (Wikimedia)

Another major, if not quite as groundbreaking task the plane received was in testing early air to air missiles. Throughout the fifties, the first practical air to air missiles were introduced, and while they were not mature enough to totally replace guns, the Air Force and Navy pursued them, believing that they soon might. The Skyknight was chosen as the test aircraft for the Sparrow missile program, and while the weapon did mature into one of the most effective air to air weapons of its day, its first iterations were extremely crude. Sparrow I was a beam riding missile which was directed by the aircraft’s radar into the target. To test it, 28 F3D’s, both 1’s and 2’s, were converted into missile carriers, receiving between two and four wing pylons to accommodate the new weapons, and their 20 mm cannons were removed. These aircraft entered limited service with Marine fighter squadrons, and a proposal for an updated design to carry six missiles was introduced. Nothing came of the program, as the missile was incapable of hitting maneuvering targets and was generally unfit for use in combat.

The Skyknight’s tested, and very briefly employed the Sparrow I missile. They proved to be totally unsatisfactory, and the planes were soon relieved of the weapons (tail spin topics)

As important as missiles was the ever evolving field of electronic warfare. It was becoming ever more vital to know the positions of enemy radar installations, communications infrastructure, and, as would become vital later on, surface to air missile systems. The Skyknight was recognized as an ideal candidate for this kind of reconnaissance mission, as the removal of its radar systems left ample space for electronic surveillance equipment and radar jammers. One F3D-2 would be modified in 1955 and equipped with a panoramic surveillance radar, direction finding and analysis equipment, and a pair of 200 watt noise jammers. Two of its cannons were removed, with two remaining to give the aircraft some form of defense and to avoid weight distribution issues. The plane was modified at MCAS El Toro by two electronic warfare veterans, WO Joe Bauher and MSgt. ER Grimes.

While the Skyknight was far too obsolete to be a fighter, its forgiving handling and large electronics bays allowed it to shift effortlessly into the realm of electronic warfare. (aerocorner)

The prototype was soon joined by a second test aircraft and the pair were evaluated and refined at the Naval Ordnance Test Station China Lake, and the White Sands Missile Range. They proved satisfactory and soon orders to convert 35 Skyknights to F3D-2Q, later redesignated EF-10B, electronic surveillance aircraft were approved. The first of the modified aircraft were received at the very end of 1956 and delivered to the Marine squadron VMCJ-3, with additional deliveries being made to VMCJ-1 through -3 in the following years. With its conversion complete, the Skyknight was to begin its second career.

Back in the Saddle

The first new deployment of the EW Skyknight began in July 1958, with VMCJ-3 rebasing to MCAS Iwakuni, Japan. While its original mission was to help with defensive electronic warfare training, it was not long until they were recruited into the Peacetime Aerial Reconnaissance Program, and used as a surveillance tool against the Soviets, Chinese, and DPRK in East Asia. Under the code name ‘Shark Fin’, the Skyknights flew offshore patrols to gather electronic data on radar stations and communication networks. Among the most crucial patrols were those around the Soviet Far East, though their patrols ranged all over the region, with forward deployments spanning from Tainan, Taiwan, to Misawa, in Northern Japan. Their first major find came in 1959, when they were the first to detect a modern Soviet P-12 ‘Spoon Rest A’ early warning radar which was based near Vladivostok.

Most of these patrols were well within international waters, though patrolling aircraft were still sent out to meet them. With the Soviets, it was a nearly carefree affair. While Skyknight and MiG pilots were engaged in a deadly cat and mouse game nearly a decade earlier, they made peaceful, routine intercepts over the Sea of Japan and its neighboring waters. One Captain Chuck Houseman would remember monitoring communications between a MiG pilot and his ground controller. When the fighter pilot asked what to do when the Skyknight’s ECMO began to take photos, his ground controller suggested he smile. On another occasion, in 1965, he flew with external fuel tanks which bore a message which, in Russian, read “JOIN THE US MARINE CORPS”. The joke was not appreciated in the higher echelons of his command, and the tank was soon painted over after they received complaints from the NSA. Soviet encounters were usually without issue, though the Skyknights would often try to avoid them by flying out over open waters and, with their twin 300 gallon (1135 liter) external fuel tanks, wait until the MiG’s ran low on fuel, before resuming their patrol.

Marine EF-10’s on the tarmac. They would convert 35 Whales for EW work. (aerocorner)

Flights near China and Korea proved far more challenging. They were met by pilots that flew far more aggressively, and on occasion, attacked patrols over international waters. While the Skyknight’s were never attacked, they always needed to be wary and tried to limit contact whenever possible. When avoidance was not an option, they were often escorted. Beyond this, the Chinese and North Korean air forces set up fake navigation beacons to try and throw off patrolling aircraft and lure them into their national airspace, where they could then be brought down. These dangers aside, no Skyknights were ever lost during these missions, and they recorded important data on Soviet radar systems.

With VMCJ-3 engaged in its Shark Fin operations in Asia, VMCJ-2 was working closer to home. Their job was to monitor Cuban military expansion flying patrols dubbed ‘Smoke Rings’, beginning in 1960. Unlike the relatively easy job of monitoring early warning radars in East Asia, Soviet technicians in Cuba were keen to keep their work under wraps, and shut down their systems if they thought they were being surveilled. This was soon noticed by the patrolling Marine aviators, who soon learned to fly under radio silence, and operate from less conspicuous airfields, particularly those in the Bahamas and Jamaica. Their work soon paid off, as in the next year, they detected the operation of a P-20 ‘Token’ radar system, used as a ground control radar for MiGs.

The Smoke Rings patrol work built up considerably as the situation in Cuba escalated after the failed invasion in the Bay of Pigs, which would lead to a significant Soviet military build up, culminating in the deployment of ballistic missiles to the island. During this period, the Skyknight’s would prove vital in uncovering, and confirming, the locations of SA-2 surface to air missile sites that would eventually prevent U-2 overflights of Cuba. When the missile crisis arrived, the Marine’s job would be to act as radar jamming support should the crisis turn into a conflict. Thankfully, they were never called upon for that task, though in the years to follow, they still patrolled the island to keep a picture of the situation, and to give new crews practical experience before they were deployed to Vietnam.

Vietnam

While the Marine Skyknight pilots were snooping along the seas of East Asia, and flying rings around Cuba, the US had become embroiled in the brutal civil war which followed the end of French control over Vietnam. Much the same as Korea, this war between two heavily militarized states would see widespread destruction, and a massive technological arms race. Airpower would be a major component to the US strategy, both seeing its traditional use, and a way to offset the considerable numerical disadvantage on the ground. It also proved a way to get around the DMZ between the North and South, which was created to prevent a direct invasion from either side. The Democratic Republic of Vietnam would weather a brutal air campaign with help from their patrons, the Soviet Union and the People’s Republic of China. At first, they had little more than anti-aircraft artillery and various light anti-aircraft armaments, but as was the case in Korea, the Soviet Union would step in and deliver the tools and training needed to build a formidable defense against American air power.

Much like the MiG-15 over Korea, the Soviet S-75 (SA-2) would prove a game changer that shifted the strategies for both sides. (Smithsonian)

The Soviets would provide aircraft, radar systems, and training personnel to build the Vietnamese People’s Air Force an effective GCI system to intercept American bombers. They began modestly, with a small force of MiG 17’s, a subsonic fighter with a gun armament. However, in 1966, the Soviet Union would begin to supply the more advanced, supersonic MiG 19, and the much more modern MiG 21. As impressive as the MiG 21 was, it did not cause the shake up that the deployment of the SA-2 surface to air missile did, which Vietnamese anti-air troops began training on in 1965. The system was robust, easily transportable by truck, and very effective for its day, with the Soviet Union supplying 95 batteries, and over 7000 missiles during the war. The triple layer of defenses, in which lower altitudes were covered by flak, higher altitudes by SAMs, and the MiG’s which were effective in both areas, proved to be a serious danger to American aircraft over northern Vietnam. However, there was a unified weakness that all of these systems shared and could be exploited. They all needed radar support to function.

The Skyknight’s would join a staggering number of aircraft involved with signals reconnaissance and jamming efforts, but among those providing direct support during Operation Rolling Thunder, it was the only major electronic warfare aircraft. Its partner was the EB-66, a faster, sleeker aircraft that boasted a more modern suite of jammers and signals intercept equipment, and was capable of airborne refueling. It was by far the more capable aircraft, but by the start of 1965, they were in short supply. Against them was a fledgling, but quickly growing network of North Vietnamese SAM batteries and ground control stations for MiGs.

The growing EW requirements of the US air strategy would put an intense workload on the Marine EF-10 pilots, while more advanced aircraft prepared to enter service. (Marty Lachow)

VMCJ-1 were deployed to the airfield at Da Nang in April of 1965 under the command of Lt. Col. Otis Corman, with six EF-10B’s and a complement of 93 men. While it might seem odd that so old an aircraft was being brought in for such an important job, both the Navy and Air Force lacked an aircraft that could fully replace it. While the SA-2 missile was known to them in the late 50’s, and had gone on to down U-2 spy planes, budget constraints and a lack of concern over the weapon stifled the timely development of a tactical jamming and signals intelligence planes that could work closely with strike aircraft. While the Navy was to receive the new EA-6A Electric Intruder, technical delays would see it deployed at the end of 1966. As it was, the Skyknight, now known almost universally as the ‘Whale’, was to play an important role in plugging the gap until more advanced aircraft became available.

The Whale’s flew their first missions on April the 19th, flying radar reconnaissance flights throughout Indochina. Their findings allowed them to plot the network of North Vietnamese early warning and fire control radars near their side of the DMZ. As the month came to a close, the air war took a turn when MiGs downed two F-105s. In response, the Whale’s were sent to suppress North Vietnamese ground control radars. Equipped with ALQ-39 jammers configured to counter the enemy’s early warning and flak directing radars, they flew ahead of strike groups, jamming and dropping chaff to confuse MiG ground directors. The EF-10’s were soon in high demand to support Navy and Airforce operations, and it was not long until they were working at a 300 percent higher rate than they were in peacetime. This was later decreased to 200%, but the crews and planes were still operating near their limits.

VMCJ-1 flew both the Skyknight, and the Mach 2 capable F-4R reconnaissance aircraft, representing some of the oldest, and most modern aircraft in US service. (Sam Gill)

DMZ patrols and jamming support continued routinely until the 24th of July, 1965, when an F-4C Phantom was shot down by an SA-2. While, previously, the sites were off limits out of concerns that killing a Soviet advisor might escalate the conflict, strikes against two SAM sites were authorized three days after the Phantom was downed. The mission, Spring High, involved the use of all six EF-10’s acting in support of a strike force of 48 F-105s. The Whales flew as screens for the F-105’s, jamming the radars used by the flak, SAMs, and MiGs. None of the strike aircraft were lost to radar guided assets, but six were lost to low level anti-aircraft fire.

One of the early challenges faced in these missions was the lack of a dedicated escort, which proved concerning, as the defenseless Whales were typically the first in and last out. While none were ever lost to MiG’s, aircrews were often concerned enough to set up informal escort flights with other Navy units. Such was the case with Chuck Houseman, who organized an escort flight with a squadron of Marine aviators who flew F-8 Crusaders from the carrier USS Oriskany. MiG’s aside, the greatest concerns were typically fuel related, as the planes were operating at the limits of their range and carried jammers and chaff dispensers on their wing pylons, where they could otherwise carry additional fuel.

By the end of the summer of ‘65, the SA-2 threat continued to evolve. Batteries sprouted up around the North, and their operators were honing their expertise on this new weapon. Facing the SAMs would require a new set of tactics that blended a mix of electronic deception and fast, aggressive flying. Named Wild Weasel, these strike aircraft were given the dangerous task of venturing into defenses designed to kill them, and tear them down. In this, the Whale was to play an early, vital role. While it had no real offensive capability, it could jam the radars of the anti-aircraft guns and SAMs, and use its signal analyzing capabilities to get a fix on their locations. They would accompany the flights of F-100’s, and later F-105’s, destined to attack the site directly and provide them with vital support. At this early chapter in anti-SAM tactics, most of the strike aircraft lacked the radar warning equipment that gave them an alert when they were being targeted. Until the devices began arriving in mass next year, one of the Whale’s most important jobs was simply to tell them when they were being targeted, and when they needed to go evasive.

In the fall, SA-2 networks and radar guided flak batteries had encompassed much of the North, and the job of the Whale’s grew more complex, and dangerous. The Vietnamese crews too were learning, often setting up several radar stations, while only using one of them to guide their weapons. The Whale’s performed well, but the limits of the aircraft and its equipment became apparent when they suffered their first and only loss to the SAM batteries. In March of 1966, the SAMs would finally catch one of the jamming aircraft, forcing a change in tactics.

With the prohibitions on their use near SA-2 sites, the Whale’s transitioned to supporting Naval operations along the coast. (Jerry Parks)

Following the incident, the EF-10 was no longer permitted within 20 miles (32 km) of a SAM site, and its mission area was effectively pushed out over the coast. This new patrol area would see them mostly supporting Naval operations, as the Navy would not possess their own jamming aircraft, the EA-6A, until the end of the year. The Whale’s new task was to fly in Navy strike aircraft toward the coast to screen their approach with jammers, while also taking note of the active air defenses over Northern Vietnam. They would prove essential, to the point that missions could be called off if no supporting EF-10’s were available.

The Whales would fly a much less conventional mission over Laos and Cambodia, where they aided in the project known as ‘Blind Bat’. In an effort to curtail the supply line known as the Ho Chi Mihn trail, the USAF outfitted several C-130H cargo aircraft with massive night vision devices. Using these, they hoped to spot the faint lights emitted by trucks and bicycle lamps as they made their nocturnal journey south to deliver supplies to the forces of the Viet Cong guerrilla fighters in the south. When the C-130 spotted something, it dropped illumination flares over it, and whatever was exposed would be attacked by the pair of B-57 Canberras which trailed the spotter. The fourth aircraft of the troupe was an EF-10, there to protect the others from radar guided AAA.

They flew off the wing of the C-130, with the two bombers following behind them. While they were never exposed to much of a threat from the ground, the Whale crews who flew these missions considered them the most dangerous during their entire combat tour. It is understandable, considering all four aircraft often flew in blacked out liveries, with a single navigation light atop the C-130 to provide reference between them. Poor weather and moonless nights were also common, as the porters along the trail knew they’d be hardest to spot in such conditions. Little was improved during a successful mission, as the flares and the exploding ammunition along the trail ruined the pilots’ night vision, leaving them to readjust as they turned for home.

End of Watch

By the end of 1966, the Whale’s replacement began to arrive in growing numbers. The EA-6A Electric Intruder was superior in every regard, but it proved unreliable as it went through a rough teething period as it was deployed to the theater. The first arrived at the end of October, ahead of a series of strike operations toward the end of the year. However, the new planes would not be able to do the job alone, and they were joined by the squadron’s venerable EF-10’s. A massive number of strikes were launched starting December 2nd, 1966, under the largest EW umbrella so far, consisting of six Intruders and ten Whales. While the Intruders could handle some of the more dangerous work, the Whales could cover transiting strike aircraft, and monitoring and jamming the growing number of radars for enemy AAA batteries.

A Super Whale departs the airfield at Da Nang. They are distinguished by the lack of any dorsal antennas. (Jerry Parks)

Even into the Autumn of the following year, the EA-6’s were still proving challenging to keep serviceable. It proved frustrating enough that the Corps decided to upgrade its Whale’s to bridge the gap until the Intruder’s readiness rates improved. The ‘Super Whale’ would feature a new broadband radar receiver, an ALR 27 radar warning receiver, and an improved panoramic display for detecting and classifying hostile radar systems. The new suite radically improved the crew’s ability to classify enemy systems and gave instantaneous missile launch warnings. The first of the modified planes was delivered in March, and crews soon flew them on their now familiar missions.

The eight Super Whales of VMCJ-1 continued to fly until September 1969, having been fully replaced by more modern aircraft. By the end of its service it was almost unique in its age, and its pilots often remarked on the fact that few airmen were assigned to something so eccentric. It was an aircraft designed with WWII era technology, and it made its pilots well aware of that fact. There were so few of them that the training materials for the aircraft were sparse, and no formal training program existed, so learning to fly and use the aircraft’s systems was an on-the-job affair. In a sense, each crewman familiarized themselves in their own way. A NATOPS manual was produced, but only near the very end of the plane’s combat tour in Vietnam. Many airmen felt pride in having mastered such an unconventional plane, especially one that flew quite well. Sentiments aside, they all knew it was an extremely obsolete plane kept flying by the kind of resourcefulness the Marine Corps is known for. The newest planes were almost twenty years old and had seen constant use in that time. The stockpile of parts was low, and there was not a single aircraft that was not completely wrung out. As fond of them as some pilots were, they were all happy with their new Intruders, and the Skyknight was finally retired in 1970.

Crew Remarks and Flying Characteristics

From the ground up, the Skyknight was designed to be stable, maneuverable, and to have no quirks in flying that might surprise the pilot. The designers were extremely successful in this regard, with pilots praising solid flying characteristics that some went as far as to call immaculate. Most contemporary jet fighters were known for being a handful, if not outright dangerous to fly, but with its hydraulically boosted controls, spoilerons, and positive longitudinal stability, it was an easy handling aircraft. Even over Vietnam, pilots found it a very forgiving, comfortable aircraft to fly. This being said, no one was ever much impressed by the look of the aircraft, with some pilots remarking that with its broad, flat wings, wide fuselage, and deep set cockpit, that Skyknight was a transport aircraft masquerading as a fighter.

A Skyknight crew prepares to depart. (National Archives)

It was, however, underpowered, having never received engines much more powerful than those on the prototypes. Its top speed was poor, and a fully loaded plane had a downright sluggish climb rate. While it was slow, this did not prevent it from scoring 6 victories against MiGs over Korea, and preventing the rest from chasing B-29’s. However, speed was not the primary issue, but rather reliability. These engines were fairly primitive turbojets developed from the first combat models. Engine failure brought down a number of these aircraft, and exploding turbines would prompt the fitting of armor plating to prevent shrapnel from traveling through the fuselage.

In its intended use as a carrier based night fighter, it was an almost total failure. The plane was simply too large and prone to mishandling by deck crews familiar with much smaller aircraft. It was also almost beyond the capabilities of the hydraulic catapults in use at the time, and accidents would result in serious damage being done to the system. The low cant of the engines scorched the wooden flight decks and ignited any flammable materials, resulting in fire alarms and a shut down of the flight deck. Thus, they could only be kept idling if positioned off the side of the ship. When they were later modified to correct for landing issues, they were controllable on the approach, and generally had good landing characteristics. However, the large flat windscreen was easily obscured by rain, and seaspray in poor weather conditions. The windshield wiper did little to improve visibility on the approach.

While the Skyknight might have been at home on the supercarriers which entered service in the 50’s, they were nothing but trouble on the WWII era carriers of their day. (smithsonian)

When deployed ashore, the issue of the low slung engines remained, and the exhaust was capable of warping, or boring holes, in tarmac. The position of the engines also made them vulnerable to foreign objects and debris on airfields, though this was amended with the use of intake covers which were removed when the aircraft was lined up on the runway. The aircraft was otherwise very capable when operating from airfields.

The Skyknight’s range was also rather short, given the high rate of fuel consumption from its crude turbojet engines. This was mostly resolved through the use of 150, and later 300 gallon (567, 1135 liter) wing mounted fuel tanks. However, in later electronic warfare missions, a jammer or chaff dispenser was often carried on one of the two wing pylons, shrinking the total fuel capacity of the aircraft. Over Vietnam, aircrews would occasionally fly with one engine off while they were transiting to stretch the endurance of the aircraft on longer missions.

The radar suite of the Skyknight was advanced, though its complexity did not lend itself well to ease of use or repair. The AN/APQ-35 featured a gun laying radar, which directed the pilot where to fire, a search radar, which the RO used to find targets, and a tail radar which warned the crew of pursuers. The radar presented information through three scopes.

The AN/APG 26 gun laying radar was the first with lock-on capability, automatically tracking a selected target. This feature was engaged by the RO, who centered the radar on the target and pushed a button. It was useful, but its position ahead of the search radar created a small blindspot. Some crews opted to remove the smaller device entirely to clear the blindspot, and many simply felt it was unnecessary given the high performance of the search radar which could be used to guide the pilot onto target.

The AN/APS 21 was a massive, cutting edge, and very complicated, search radar. It was a truly excellent piece of equipment, but was described by O’Rourke as being designed ‘by engineers that had never flown a plane’. The radar itself was not stabilized, and the plane of the scanning radar shifted with the positions of the aircraft. An RO also had to operate it within certain limitations. If set to a fast sweep at the widest angle, the dish would swing rapidly back and forth, destabilizing the plane and breaking itself. Broken radars were not uncommon given the fragility of vacuum tube electronics, with radar serviceability capping squadron readiness near 60% in Korea.

However, for all its quirks, it was an extremely impressive piece of engineering for its day. The radar was able to detect large contacts at 125 miles (201 km) and had an adjustable search angle that could be set as wide as 170 degrees. For all its trouble, it was worth it.

The massive tandem radar array of the APQ-35. (Jay Miller)

Lastly, the system had an AN/APS-28 tail warning radar which sat at the very end of the aircraft. It was very similar to the older APS 19 radar system found on Corsair night fighters, and thus very familiar to those who retrained for the Skyknight during the Korean War. It employed a small scope on the AN/APQ-35 console and displayed the position of contacts behind the aircraft up to four miles, with a crude approximation of their altitude, with the contact being noted as being as level, above, or below the aircraft. It also had a secondary display consisting of a quadrant of warning lights that would warn the crew of pursuers and their relative distance and position. In service, it was almost useless, as it was set off at lower altitudes and any friendly aircraft nearby, for instance, any B-29’s they might be escorting. The quadrant lights were thus typically removed and the RO would refer back to the dedicated AN/APS-28 radar scope every few minutes.

The AN/APQ-35 suite was later removed when the Skyknights were converted to electronic warfare aircraft, replaced with an EW kit comprised of a panoramic surveillance receiver that displayed the direction of radar and communications systems. It was fed information by a radio direction finder, and a pulse analyzer for identifying radar emissions. The original system used a once quite sophisticated APR-13 receiver, which displayed information on an oscilloscope, and a direct audio output of the radar transmission, to classify and give the direction of ground based radar systems. This allowed the ECMO to identify and locate any number of radar stations, though this took a good deal of work. The Skyknight proved to be a groundbreaking EW platform, but for its crews, this meant dealing with cutting edge yet crude equipment.

Being from the 1950’s, the system was a cumbersome affair that required its operators to manually set operating frequencies and offered no automation of any kind. It could surveil enemy radar systems and provide missile launch warning for friendly aircraft, however, it was nearly impossible to do both at once. Thankfully, in roughly its last year of service, the Marine Corps replaced the analogue EW suite with an APR-33 threat receiver, an ALR-27 missile launch warning system, and a new panoramic display which displayed the entire range of Soviet and Chinese radar frequencies. The new system took a lot of work off the ECMO and automatically warned them of a missile launch, representing a comprehensive upgrade.

Operating the aircraft’s jammers, both on-board and those on the pylons was a relatively straightforward affair. In Vietnam, this virtually always meant using wing mounted ALQ-31 pods that could carry two jammers that were configured to counter different radar systems. While the Skyknight was pulled from the SA-2 jamming mission, it proved very helpful in jamming AAA fire control radars. Crews often remarked that anti-aircraft fire became totally inaccurate once the jammers were in range of the enemy radar system.

Construction

The Skyknight was a solidly built aircraft with a conventional construction. It featured a very wide, monocoque fuselage with folding, mid level, two spar wings. The front of the aircraft contained the search radar or electronic warfare equipment, and the armament, all enclosed in a fiberglass cone. The engines were contained in low slung nacelles within the fuselage, behind which were the airbrakes. At the rear of the aircraft was a conventional stabilizer configuration, with a tail warning radar at the very back of the aircraft. The aircraft had a fully retractable tricycle landing gear arrangement with a deployable tailwheel to prevent tail strikes.

Armament and fuel stores (Standard Aircraft Characteristics)

The wings of the aircraft were of conventional construction, though it saw an early use of hydraulically boosted control surfaces. Combined with a set of spoilerons added to the production series of the aircraft, its roll rate was excellent and maneuverability was retained at high speeds. With the fuselage air brakes allowing for the pilot to prevent overshooting a targeted aircraft, or avoiding overspeed, the plane was remarkably controllable in all aspects of flight. They were designed to fold just beyond the outboard pylons.

The final engine of the aircraft was the Westinghouse J34-WE-36. It produced up to 3,400 lbs (1542 kg) of thrust, leaving the aircraft fairly underpowered. Attempts to re-engineer the aircraft were canceled when the J46 never became available. The J34 was a development of the J30, a WWII era jet engine, and was largely obsolete before entering service due to the rapid strides in turbine development after the Second World War. The engine was an axial compressed turbo jet, with 11 compressor stages, and two turbines. In its early service, it was fairly unreliable and dangerous, as the turbines could break and send their blades through the fuselage, into the second engine. Armored deflectors were thus installed early in its military service. The engine ran on 115/145 octane AVGAS, and not jet fuel, a feature which seriously highlighted its obsolescence in later years.

The console for the APQ-35 contained all of the controls and displays for the three radar systems; it proved compact, but complicated to use. (Pilot’s Handbook)

The Westinghouse AN/APQ-35 radar suite comprised three self contained radar units, being the X-band AN/APS-21 Search radar, AN/APG-26 gun laying radar, and AN/APS-28 tail warning radar. The search radar was by far the largest unit and presented a maximum instrumented range of 200 nautical miles for ground contacts. In airborne use, it could detect targets out to a range of about 120 nautical miles. Its scan area could be adjusted in terms of elevation, and had an adjustable horizontal search angle between 30 and 170 degrees. The search radar sat in tandem with the gun laying radar, with the smaller system ahead of the main unit. The gun laying radar had a maximum range of 4000 yards (3657 m) and was activated by aiming the search radar onto the target and engaging the lock feature. The smaller radar would then automatically track the locked target and adjust the aircraft’s gunsight to give an accurate lead. The tail warning radar had a range of 3 nautical miles and was fixed. The console had three scopes, being a plan position indicator scope, an azimuth scope that gave directional guidance toward the target, and a tail warning scope. Long range target information was displayed on a large plan position indicator scope which was used exclusively by the search radar, while the azimuth scope was shared with the gun laying radar and used to guide the pilot on the final approach to the target.

On the electronic warfare model, the radar suite was replaced by a collection of radio emission monitoring equipment, jammers, and countermeasure dispensers. The original EW suite consisted of an APR-13 panoramic surveillance receiver, replaced in the early sixties with the ALR-8, which included a APA-69A direction finder, and an ALA-3 pulse analyzer. The direction finder and the receiver each had their own console and were used to track and classify actively emitting radar systems. The ALR-8 could monitor most of the Soviet, and Soviet derived, radar systems of its day. This was done with a pair of oscilloscopes, one circular in the case of the panoramic indicator, the other linear in the case of the pulse analyzer, and a direct audio output of the radar emission. These gave the direction and pulse width of the radar system respectively, while the audio output could also be used to identify the pulse width and type of radar system. Each had their own distinctive tone, occasionally allowing for easy classification. A constant pulse rate indicated a fire control type system either directing a SAM or anti-aircraft gun batteries.

EW suite changes, the Super Whale setup is on the right. (EF-10B NATOPS)

These systems were seriously overhauled with the Super Whale upgrade under AFC 199. This included a panoramic ULA-2 indicator console which displayed the directions of all emitting radars, and no longer required the ECMO to manually search frequencies. Defensive upgrades included an APR-33 fire control monitor receiver and an ALR-2 missile launch warning receiver. These upgrades automated much of the ECMO’s workload, and allowed for the aircraft to perform missile warning duties while also investigating radar emissions.

Originally, the aircraft was only equipped with a pair of ALR-2 200 watt jammers, which were acceptable through the 1950’s, but totally inadequate for use over Vietnam. They were typically supplemented by outboard jammer pods and countermeasure dispensers. They often carried an ALQ-31 pod that could fit two jammers, which were typically configured to jam the early warning radars used to provide GCI for MiGs, and fire control radars for anti-aircraft batteries. The other major EW tool was the ALE-2 chaff dispenser, which could be used to create metallic, radar reflecting clouds of aluminum strips. Other avionics included VHF radio communication systems, a UHF radio, a VHF beacon homing receiver, a radio altimeter, and a radio compass. These systems were upgraded throughout the Skyknight’s long career, and new systems, like equipment to use the Tactical Air Navigation System, were added.

The Skyknight’s armament consisted originally of four Hispano Suiza M2 20mm cannons with 200 rounds carried for each weapon. The pilot was provided with a radar directed Mk. 20 Mod. 0 gunsight which could provide automatic targeting for a locked target. On electronic warfare variants of the aircraft, the armament was reduced to two weapons, retained for balance and self defense purposes.

A variety of ordnance could be carried on the outer pylons, being unguided bombs up to a weight of 2000 lbs (907 kg) per pylon. The 11.75” ‘Tiny Tim” rocket could also be mounted, though it is unlikely they were ever used, as only a handful of ground attack missions were carried out with this aircraft near the end of the Korean war, and only with unguided bombs. The pylons were otherwise used only for carrying 150, or 300 gallon fuel tanks (567, 1135 liters), in addition to the electronic warfare equipment described above.

The Skyknight featured more creature comforts than most other navy fighters by the time of its design. It was air-conditioned, its floors were carpeted, and an electric cigarette lighter was installed in the instrument panel, with ashtrays at the crew’s elbows. It was perhaps the only fighter aircraft to be equipped with a built-in cigarette lighter. However, it was not retained on the EF-10 and the engineers at Douglas removed it when they were updating the instrument panels. Both crewmembers were provided with urinals in the form of relief tubes for use on long flights.

F3D escape diagram. (Pilot’s Handbook)

The escape system consisted of a chute positioned between the crewmen, and at its end was a panel which would be ejected by means of explosive bolts. The crew would then use a bar over the opening, at the rear of the cockpit, to hurdle themselves down the chute and clear the plane. It was effective, though it meant that one could not safely bail out of the aircraft below 2000 ft. Crewmen who ditched the aircraft were to escape via the roof panel, which also doubled as the means to enter and exit the plane.

Conclusion

A preserved Skyknight, ironically stored aboard the USS Intrepid. (Tony Inkster)

In spite of failing in its original goal almost completely, the Skyknight’s career in the Marine Corps saw it become the unsung hero of two wars and earn the respect of its crews. It seems almost impossible that an ungainly nightfighter rejected for its original use could have ever brought down cutting edge MiG’s, made the Air Force’s B-29’s untouchable, or claim the night skies over North Korea for its own. Yet in the end, the Marine’s made due, and ‘Willy the Whale’ became one of the most successful fighters in the theater. Beyond this shocking combat debut, it almost effortlessly transitioned into an entirely different role, surveilling radar systems and providing electronic support for US forces in the technological cat-and-mouse race over Vietnam. Developed in an age when planes had their operational life spans measured in months, the Marine Corps flew the Skyknight for almost twenty years, a testament to its ruggedness and versatility.

F3D-2

Specification

Engine Westinghouse J36-WE-36
Engine Maximum thrust 3,400 lbs (1542 kg)
Fighter weight internal stores only 24,614 lbs (11,164 kg)
Fighter with 2 x 150 gal tanks 26,731 lbs (12129 kg)
empty weight 14,898 lbs (6757 kg)
Combat Range [with external fuel] 1,195 nmi
Combat Range [internal stores] 995 nmi
Maximum Speed 426 kts @ 15,000 ft (4572 m)
Cruising Speed 395 kts
Combat Ceiling 35,000 ft (10,668 m)
Armament 4x 20 mm Hispano Suiza M2 cannon
Crew 1x Pilot

1x Radar Operator

Length 45′ 5″ (13.84 m)
Height 16′ 1″ (4.9 m)
Wingspan 50′ (15.24 m)
Wing Area 400 sq.ft (37.16 m2)

Variants

F3D-1 (F-10A): First production version, J34-WE-34 engines. 28 built.

F3D-2 (F-10B): Improved, final production model. J34-WE-36 engines, lock on capability, General Electric G-3 autopilot, wing spoilers. 237 Built, final aircraft built in March of 1953.

F3D-1M: Sparrow missile testbed.

F3D-2M (MF-10B): Four missile hardpoints, no cannons. Brief service life. 16 converted from F3D-2s.

F3D-2Q (EF-10B): Electronic Warfare Aircraft. 35 Converted from F3D-2.

F3D-2T: Night fighter trainer. 5 converted.

F3D-3: Proposal, swept wing night fighter with J46-WE-3 engines.

Illustrations

 

VMF(N)-513, Korean War. During their combat tour ,The Flying Nightmares developed a superstition around aircraft numbered 13. The crew of this aircraft improvised an alternate number.
VF-14, USS Intrepid, 1954. The Top Hatters were among the last squadrons to fly the Skyknight as a night fighter. They quickly transitioned to more modern fighters.
VMCJ-1, Da Nang, Vietnam War. in 1967, all of the in-theater EF-10B’s were upgraded to ‘Super Whales’. They continued to serve for two more years before being phased out by more modern aircraft.

 

Credits

  • Article written by Henry H.
  • Edited by  Stan L.
  • Ported by Henry H.
  • Illustrated by Hansclaw

 

Sources:

Primary:

“Eyes In the Night”. Naval Aviation News. V33-34 1952-1953.

Pilot’s Handbook Navy Model F3D-2 Aircraft. Secretary of the Air Force and the Chief Bureau of Aeronautics. 15, July 1952.

NATOPS Flight Manual Navy Model EF-10B Aircraft. Chief of Naval Operations. 1 April 1969.

Night Fighters Over Korea. G.G. O’Rourke with E.T. Woolridge. Naval Institute Press.

Standard Aircraft Characteristics F3D-2 “Skyknight”. 15 February 1952.

Secondary Sources

F-3D/EF-10 Skyknight Units of the Korean and Vietnam Wars. Joe Copalman. Osprey Publishing. 2022.

F-105 Wild Weasel vs SA-2 “Guideline” SAM. Peter Davies. Osprey. 2011.

All Hands. No. 648-658. 1971.

Naval Fighters Number Four Douglas F3D Skyknight. S. Ginter.

Korean Air War Sabres, MiGs and Meteors 1950-53. Michael Napier. Osprey. 2021.

Naval Aviation News, Obituary Heidemann, Jan-Feb. 1992.

Skyknight. R.E. Williams. Naval Aviation News. 1983.

Into the Jet Age: Conflict and Change in Naval Aviation , 1945-1975. E.T. Wooldridge. Naval Institute Press. 1995.

A History of Marine Fighter Attack Squadron 531. Colonel Charles J. Quilter II and Captain John C. Chapin. History and Museums Division Headquarters, US Marine Corps. 2001.

US Marines in Vietnam High Mobility and Standdown 1969. Charles R. Smith. 1988.

Sparks over Vietnam The EB-66 and the Early Struggle of Tactical Electronic Warfare. Captain Gilles Van Nederveen. College of Aerospace Doctrine, Research and education. 2000.

Aircraft Carriers a History of Carrier Aviation and its Influence on World Events Volume II 1946-2005. Norman Polmar. Potomac Books. 1969.

Boulton Paul P.105 & P.107

Great Britain (1944)

Strike Fighter Concept

Side view of the Boulton Paul P.105C. This was the single-seat fighter version of the aircraft, armed with four 20mm cannons. (Boulton Paul Archive Photos)

The Boulton Paul P.105 was a concept for a multi-purpose, single-engine aircraft that was designed to fill a number of carrier based roles. To do so, the P.105 would utilize a unique and innovative method that would use interchangeable fuselage sections and cockpit modules that would allow the aircraft to perform different missions. These modules could be changed quickly to fill a needed role aboard carriers or airbases. The aircraft would not be chosen for production, and The P.105 would be developed further into the P.107, a land-based escort version. The P.107 would have a rear-facing turret and a twin boom tail design to allow greater traverse of the gun. This design wouldn’t be adopted either, and the program would conclude before the war’s end.

History

Late in the Second World War, the Royal Naval Air Arm began seeking out a new aircraft design that would be able to fill both the fighter and bomber roles aboard their carriers. Having one aircraft perform multiple roles would eliminate the need for specialized carrier-borne aircraft to fill the fighter, dive bomber, and torpedo bomber roles that were currently in operation. No official requirements were ever put out to build such an aircraft, but several companies had begun developing aircraft that would fit this role, which had become known as the “Strike Fighter”. Westland, Blackburn, Fairey and Boulton Paul would all develop designs that correspond to the strike fighter role. Boulton Paul’s aircraft design would be known as the P.105.

After the production of their Defiant turret fighter was finished, Boulton Paul began producing the Fairey Barracuda carrier bomber under license. After working extensively with a naval aircraft of this type, lead aircraft designer of Boulton Paul, John North, began to show interest in developing new aircraft to serve the Royal Navy’s carriers. The timing for this interest was beneficial too, as the Royal Air Arm began showing interest in new aircraft that were to be used in the Pacific Theater. He would first design a single engine fighter, dubbed the P.103 which would compete for the Navy’s Specification N.7/43 aircraft project. The P.103 was a heavily reworked Defiant with the turret removed and the design heavily cleaned up to make for a more effective fighter. Two designs existed for the P.103; the A and B, with the A using a Rolls Royce Griffon engine and the B using a Bristol Centaurus engine. The P.103 would utilize a number of innovative features, such as contra-rotating propellers, a low drag wing, specialized landing gear that became shorter when stowed, and elevators with automatic trim tabs. In addition, a more radical design was also submitted, the P.104, which was a twin-boom pusher. Despite both the P.103 and P.104 satisfying the specification, the Navy ultimately would find that a Hawker Tempest variant that was to be produced could easily be adapted to this role. This aircraft would become the Hawker Fury, and naval-ized into the Sea Fury.

While the P.103 wouldn’t be built, there were plans to test many of its design features on an existing aircraft. A Defiant was chosen to be extensively modified with most of the features found on the P.103, including the contra-rotating “dive-brake” propellers driven by a Centaurus engine, electric trim tabs, specialized shortening landing gear, and automatically closing landing gear doors. This aircraft, known as the Special Features Defiant, would also go unbuilt, with only a Defiant being modified with the elevator trim tabs. Boulton Paul wouldn’t yield any aircraft from this specification, but a new design would soon come from John North, who would continue working on Naval aircraft projects, looking to create an aircraft that would replace the Fairey Barracuda. Using design aspects intended for the P.103, and newer features found on the Special Features Defiant, he would design the P.105.

Static model of the standard P.105A. (British Secret Projects 1935-1950)

The P.105 was a small, high-performing aircraft that was meant to perform a number of duties aboard carriers. To achieve this the P.105 would have a unique design feature. To fill the variety of carrier-borne roles, the P.105 would have modular cockpit and bomb bay sections. Each of these modules would pertain to a particular role and would include necessary equipment to operate for the given task. The interchangeable modules included a two-seat torpedo-bomber with the necessary modifications to carry a torpedo (P.105A), a two-seat reconnaissance aircraft with an extended cockpit with changes to improve visibility (P.105B), a single-seat fighter armed with four 20mm cannons (P.105C) and a dive-bomber (P.105D). All aircraft aside from the C would be armed with four 12.7mm machine guns. With this system, it was thought more P.105 airframes could be stored inside hangars and carriers, while the unused modules could easily be stored and would take up less space, compared to having a number of different aircraft specified for specific roles, in theory, increasing the combat capacity of the carrier the P.105 would be stationed on. Boulton Paul expected the aircraft to be very high performance, and the P.105C fighter version, would be thought to serve as an excellent penetration fighter. Like its predecessors, the P.105 was originally going to utilize a Griffon 61 engine, but before performance predictions were done on the design, it would change to a Centaurus with counter-rotating propellers. The brochure on the details of the aircraft was submitted to the RNAA, but no order for production came about.While no particular reason was given for the design not being chosen, the modularity concept may have been less convenient in practice then on paper. Another reason could be that current aircraft at the time were deemed to have been performing adequately and didn’t need such an extensive replacement.

A side view plan drawing showing the layout of the Boulton Paul P.107. (Boulton Paul Archive Photos)

Although the P.105 wasn’t granted production, the design was further reworked into the Boulton Paul P.107. The P.107 was a return to basics for Boulton Paul, being a single-engine two-seat fighter with a turret. It can be assumed the P.107 began development during or shortly after the P.105 had been created. John North expressed many concerns with aircraft meant to operate in the Pacific War, with the biggest issue being the extreme range an aircraft would need in order to operate efficiently in this conflict. While details are sparse on its development, the P.107 extended range escort fighter appears to be his own attempt to create an aircraft meant to amend this issue. Overall, the P.107 shared many aspects of the P.105C, continuing to use the same overall design, Centaurus engine with contra-rotating propellers, and the same armament of four 20mm cannons. However, the P.107 wasn’t meant to operate from carriers, instead being designed as a land-based aircraft. Changes done to the design for this reason include the lack of folding wings and the removal of the torpedo blister. The aircraft would also benefit with the addition of a turret housing two 12.7mm machine guns. To improve the firing efficiency of the turret, the single fin of the P.105 was changed in favor of a twin fin design, which improved the firing range of the guns. The P.107 could also be configured for different roles, such as a dive bomber and for reconnaissance, but it is unknown if it used the same modular system the P.105 used. As was the case with his earlier designs, the P.107 wasn’t selected for production either.

Design

3-Way drawing of the P.105B. This was the reconnaissance version. (British Secret Projects 1935-1950)

The Boulton Paul P.105 had a conventional monoplane fighter layout. In the front, it would utilize a 6-bladed contra-rotating propeller that had reversible pitch. Originally, the design would have mounted a Griffon 61 V-12 inline engine but was changed in favor of the Centaurus 18-cylinder radial CE.12.SM engine instead. The wings on the P.105 were inverted gull wings, much like those on the Vought F4U Corsair or Junkers Ju 87 Stuka, which allowed the mounting of a larger propeller. To allow for easy storage aboard carriers, the wings were able to fold inwards. The fuselage had the most interesting aspect of the design overall, and that was its interchangeable cockpit and lower fuselage modules. Each variant of the P.105 would use different modules that would pertain to the intended role it served. The P.105A was a torpedo bomber and would use the torpedo blister present under the tail, and provisions for carrying another crewmember. The P.105B was a reconnaissance aircraft, and its cockpit would be lengthened to sit a pilot and observer. It would use a glass hull beneath the observer to assist in spotting. The P.105C was an escort fighter and would be a one-man aircraft. The last was a dive-bomber version, which only has very sparse details available. The dive bomber would carry up to two 1,000 lb (450 kg) bombs, most likely in an internal bomb bay module. The tail of the aircraft would be a conventional single rudder and tailplane arrangement. The armament of the P.105 was a standard two to four 12.7mm machine-guns in the wings of the aircraft, with the only deviation being the P.105C, which would use four 20mm cannons instead.

3-Way view of the P.107. Notice the turret and twin tail. (British Secret Projects 1935-1950)

The P.107 borrowed many aspects of the P.105 design, but changed some details to better fit its role. The engine and front sections would stay the same, keeping the contra-rotating propellers and Centaurus engine. Reference materials refer to the aircraft as being able to convert from an escort fighter to either a fighter-bomber, or photo reconnaissance aircraft. However, whether it was a conventional conversion, or via the module system the P.105 used is unknown, the latter being most likely. The wing design would stay the same, with the inverted gull wing style. Given its land-based nature, the wings no longer needed to be folded to conserve space, and the torpedo blister under the tail was removed. Behind the pilot, a gunner would sit and remotely control two 12.7mm machine guns. The machine-guns would be housed within the aircraft, with only the ends of the barrel protruding out. To give the gunner a better firing arc, the single tailfin was switched to a double tailfin. The turret and twin tail design are the most obvious differences between the P.107 and P.105. The aircraft’s fuel would be stored in a main tank beneath the crew members and two smaller drop tanks. The fuel amount was expected to give the aircraft a 3,000 mi (4,827 km) range, with up to 30 minutes of combat. The drop tanks could be switched for 2,000 Ib (900 Kg) of bombs. For offensive armament, the P.107 would use four 20m cannons mounted in the wings.

Conclusion

While no P.105 or P.107 would be constructed, the designs do attempt to amend issues that were present at the time. The Strike Fighter designation would eventually become a standard type of aircraft aboard carriers, and aircraft meant to fulfill multiple roles would also eventually be developed, but none would ever use such a unique system as the interchangeable fuselage of the P.105. It is interesting to note that the P.105 and P.107 appear to be the last military propeller aircraft that Boulton Paul would design before their switch to trainers and jet powered research aircraft, the aircraft themselves being distantly related to their Defiant fighter that they became known for during the war.

Variants

 

  • Boulton Paul P.105A– Two-seat torpedo bomber version of the P.105.
  • Boulton Paul P.105B– Two-seat reconnaissance version of the P.105. This version would have a glazed hull for the observer.
  • Boulton Paul P.105C– Single-seat Fighter version of the P.105.
  • Boulton Paul P.105 Dive bomber– Dive bomber version of the P.105. No designation was given to this design. (P.105D?)
  • Boulton Paul P.107– Land-based escort fighter derived from the P.105. The P.107 shared many design aspects with the P.105 but would remove features that would be needed for carrier use, such as the lack of folding wings. The P.107 would also have a turret and the tailplane would be switched to a double rudder design to accommodate the turret’s firing arc. Photo reconnaissance and fighter bomber versions of the P.107 are also mentioned.

Operators

 

  • Great Britain – Had they been built, the P.105 and P.107 would have been used by the Royal Fleet Air Arm, with a focus of being used in the Pacific Theatre aboard carriers and from land.

Boulton Paul P.105 Specifications

Wingspan 38 ft / 11.6 m
Length 34 ft 5 in / 10.5 m
Folded Width 15 ft 4 in / 4.67 m
Wing Area 250 ft² / 23.3 m²
Engine 3,000 hp ( 2,200 kW ) Centaurus CE.12.SM engine
Fuel Capacity 260 gal (1,180 lit)
Weights 12,285 Ib / 5,572 kg with torpedo

12,509 Ib / 5,674 kg with bombs

Climb Rate 3,660 ft/min / 1,110 m/min
Maximum Speed 469 mph / 755 km/h at 20,000 ft / 6,000 m
Cruising Speed 407 mph / 655 km/h
Range 1,300 mi / 2100 km – 3,320 mi / 5340 km
Crew Pilot

Other crew member (Depending on the variant)

Armament
  • 2-4 12.7mm machine guns (All versions)
  • 1x Torpedo (P.105A)
  • 2x 1,000 Ib (454 kg) bombs (Dive Bomber)
  • 4x 20mm cannons (P.105C)

Boulton Paul P.107 Specifications

Wingspan 38 ft / 11.6 m
Length 34 ft 8 in / 10.6 m
Wing Area 250 ft² / 23.3 m²
Engine 3,000 hp ( 2,200 kW ) Centaurus CE.12.SM engine
Fuel Capacity Main: 495 gal (2,250 lit)

Drop Tanks: 140 gal (640 lit)

Weight 15,900 Ib / 7,200 kg
Max Speed 470 mph / 755 km/h at 22,000 ft / 6,700 m
Range With Drop Tanks: 3,000 mi / 4,800 km

Without: 2,200 mi / 3,540 km

Fighter-Bomber: 700 mi / 1,120 km

Crew 1 Pilot

1 Gunner

Armament
  • 4x 20 mm guns
  • 2 x 12.7mm machine guns in rear facing turret
  • 2,000 Ib (907 kg) of bombs

Illustrations

Boulton-Paul P.107
Boulton-Paul P.105 Reconnaissance Variant

Credits

  • Article written by Medicman11
  • Edited by  Henry H.
  • Ported by Henry H.
  • Illustrated by Haryo Panji

Sources

Si-204E

Siebel 204

Nazi flag Nazi Germany (1940)
Light Transport and Trainer – Number built: 1,175

While often seen as less exciting than their combat counterparts, transport and auxiliary aircraft provided vital services in moving cargo, and training new pilots. Light transports which could combine both duties were thus extremely desirable during the war as theaters stretched across continents and pilot attrition was high. Luckily for the Luftwaffe, the Siebel company provided them with a simple but effective aircraft that could easily fulfill both roles. This was the Si 204, which saw wide-scale use both during, and after, the conflict.

The Si 204. Source: www.airwar.ru/

Siebel company history

The story of  Siebel began back in 1936 when Hans Klemm opened a new aircraft factory the,  Flugzeugbau Halle GmbH.  This company would go on to produce license-built aircraft, including the Focke-Wulf Fw 44, and Heinkel He 46. Between 1936 and 1937, a new project led by Hans Klemm was initiated. This was a light twin-engined transport aircraft designated as Fh 104. While the work was going on, Klemm decided to hand over the factory to well-known aircraft enthusiast Fritz W. Siebel. The same year the name was changed to Siebel Flugzeugwerke Halle GmbH. Under new management, the work on the renamed Siebel Fh 104 continued. The Siebel Fh 104 would prove to be a solid design and was pressed into Luftwaffe service as a communication and liaison aircraft. In 1942 the production of this aircraft was terminated, by which time only some 46 were built. The Siebel factory would survive the war and even produce a few new aircraft designs. It would continue to exist up to 1968 when it was merged with Messerschmitt-Bolkow GmbH.

The first aircraft to come out of the o Siebel Flugzeugwerke production was the Siebel Fh 104 Source: hwww.armedconflicts.com

The Siebel 204 

Following the success of the Fh 104, Siebel received a request from the Luftwaffe officials in 1939 to design and build a new twin-engine, 8-passenger transport aircraft.  So Siebel and his team of engineers began working on such a design. While they may have used the experience gained while working on the Fh 104, their next project was a completely new design. The first prototype Si 204 V1 (D-AEFR) was completed in early 1940, and was flight tested on the 25th of May the same year. Sources disagree about the year when the maiden flight was made. For example, D. Nešić and M. Fratzke mentioned that it happened in 1941 while M. Griel placed it in 1940. The test flight proceeded without any major issues, so the development of this aircraft carried on. In October 1940 the Si 204 V2 (D-IMCH) was flight tested. Both of these would serve as bases for the pre-production A-0 series which were to be operated by the German Lufthansa airline. The first prototype was scrapped in 1942 while the second remained in use up to early 1944 when it was lost in an accident.

Following its successful testing, the first production version known as Si 204A was built. It was powered by two 360 hp, or 465 hp depending on the source, Argus As 410 engines. The Si 204A-0 and A-1 were put into production in 1941, the precise numbers are not clear but were likely limited. As the war dragged on these were mainly used for crew training, a role to which they proved well suited.

The Luftwaffe was generally satisfied with the Si 204A’s performance as a trainer but requested that a new version of it be built. This version was dedicated to various crew training tasks including; radio navigation, instrument flying, bombing, and communication. Other requests were made regarding its front canopy design and stronger power units. For this reason, the engines were replaced with two  600-hp Argus As 411 12-cylinder engines. Additionally, the original stepped canopy was replaced with a fully glazed canopy.

The new version was to be designated Si 204D. The fate of the skipped B and C versions is unclear, but these were likely only paper projects.  The Si 204V3 and V4 served as bases for the Si 204D aircraft. Both were flight tested in early 1941, withhe V3 being lost in an accident during mid-1942 while the fate of the V4 is not known.

Technical characteristics

The Si 204 was designed as a low-wing, twin-engine, all-metal transport, and training aircraft. Its fuselage was made of round-shaped formers each connected with a series of metal bars. These were covered with sheet metal plating. On the fuselage sides, there were four rectangular windows.

The wings and tail units were also of an all-metal construction. The wings were built using only a single spar. The dihedral tailplane was divided into two fins and rudders, which were located on their tips.

In the last months of the war, due to shortages of resources, Siebel attempted to replace some metal components using wooden materials. The end of the war prevented any of these wooden components from ever being used.

The pilot and his assistant were positioned in the front. As many German bombers had a fully glazed canopy, to help with the training and adaptation of new pilots, the Si 204 was also equipped with such a designed canopy. It largely resembled the one used on the He 111. Thanks to it the pilot had an excellent view during the flight.

As mentioned earlier, Si 204D was powered by two 600 hp Argus As 411 12-cylinder engines, these used two variable pitch blade propellers. The maximum speed achieved with these engines was around 364 km/h. With a fuel load of 1.090 liters, the maximum operational range was around 1.800 km.

The landing gear was more or less a standard design. It consisted of three wheels. The landing gear retracted back into the engine nacelles. These were not fully enclosed and part of the wheels was exposed. The tail wheel was not retractable.

While initially designed as a passenger transport aircraft, the Si 204 would be primarily used for crew training. For this reason, its interior compartment could be equipped with different training equipment depending on the need. Including radio, radar, or navigation equipment.

The Siebel 204D side view. Its overall design is quite similar to the German he 111 bombers. Source: www.airwar.ru
Siebel pilot cockpit interior. The pilot and his assistant had an excellent view of the surrounding thank to the large glazed cockpit. Source: www.airwar.ru
The Siebel 204D had standard landing gear. The two front wheels retracted back into the engine nacelles. These were not fully enclosed and part of the wheels was exposed. Source: www.airwar.ru

Production

Despite being Siebel’s own design, the factory itself lacked production capabilities as it was already heavily involved in the manufacturing of other designs including the Ju 88. The actual production was redistributed to two occupied foreign factories. The first were the SNCAC factories located in Fourchambault and Bourges in France, which came under German control after the successful end of the Western Campaign in 1940. The second production center was located at the  Czechoslovakian Aero factory, which was also occupied by the Germans even before the war started. Other companies like BMM and Walter were also involved in the production of this aircraft.

The production numbers were initially low, for example, the SNCAC only managed to build five aircraft per month during 1942. From 1942 to 1944 this company produced some 150 Si 204D aircraft. Czechoslovakian production capabilities proved to be better, managing to manufacture some 1007 such aircraft by the end of the war. The total production of all versions during the war is around 1.175 aircraft according to H. A. Skaarup. This number, as is the case with many German production numbers, may be different in other sources.

Service

As mentioned earlier the Si 204 was mainly used for crew training for various roles, transportation, and glider towing. While there is quite limited information on their precise service life, it appears to be quite a successful design and was praised by the Luftwaffe pilots. By the end of the war, some were even equipped with various radar equipment including FuG 217R and FuG 218V2R tail warning radars to train night fighter pilots. Interestingly the Si 204 was employed for the training of further Me 262 pilots.

It is often mentioned that the Si 204 was the last Luftwaffe aircraft to be shot down. Near Rodach in Bavaria, just a day before the Germans capitulated to the Allies. That kill is accredited to Lieutenant K. L. Smith, a pilot of a P-38 Lightning from the 474th Fighter Group. How valid this claim is difficult to know precisely due to the general chaotic state in Germany at that time.

During its service life, the Si 204 proved to be an effective aircraft, completely suited for its designated role. Source:www.airwar.ru

Combat adaptation attempts

For fighting against Partisan movements in occupied Europe, older or modified aircraft were often reused, preserving the more modern aircraft for the front line use.  The Si 204 was seen as tempting for such a  modification, so the Siebel engineers tried to develop a fully armed combat version of this aircraft. To fulfill this role some extensive modifications were needed.

Inside its front fuselage, two 13 mm MG 131 heavy machine guns were placed. Each was supplied with 500 rounds of ammunition, stored in a metal ammunition bin. These were to be operated by the pilot. For this reason, he was provided with a Revi 16A-type gun sight. For protection against enemy aircraft, on top of the fuselage, a fully glazed turret armed with one 13 mm MG 131 was added. The turret movement was electrically controlled. Elevation was -10 to +80 while it could achieve a full 360 rotation.

The interior of the Si 204 received a bombing bay that could carry 12 70 kg bombs. External bomb racks with a capacity ranging from 50 to 500 kg were added. The pilot seat received armor plates for his protection from enemy fire on the Si 204E.  Due to its relatively slow speed, using this aircraft against a well equipped enemy was dangerous, so it was to be restricted to night bombing action only.

In 1944 two prototypes were completed and tested. Besides these two, the number of Si 204E’s built is unknown. Given its experimental nature, possibly only a few prototypes were ever completed. Allegedly these saw limited action fighting the Belarusian Partisans. The extent to which they were used in this role if used at all, remains unknown.

The Siebel 204E could be easily distinguished by its glazed turret, located on the fuselage top. This version is somewhat obscure as it is not known how many were built and if they ever saw action in combat. Source: www.silverhawkauthor.com

Carrier proposal

With the Allies slowly getting the upper hand in the air over Europe, the Luftwaffe became ever more desperate to find a solution to this problem. Mass production of cheap fighters was seen as a possible solution. One such project was proposed by Professor Alexander Lippisch, best known for designing a series of glider fly-wing designs. He was also involved in designing various bizarre aircraft projects, including the unusual  P 13a  aircraft.

A drawing of Professor Alexander Lippisch P 13a fighter. Source: D. Sharp Luftwaffe Secret Jets of the Third Reich

While working on the P 13, Lippish was approached with a request from a group of students from Darmstadt and Munich universities who wanted to avoid conscription to join his work. Lippisch agreed to this and dispatched one of his assistants under the excuse that for his own project, a wooden glider was to be built and tested. They together managed to build an experimental  DM-1 glider.. However, this aircraft was not to be towed like any other glider. Instead, the DM-1 was to be placed above the Si 201 on brackets and carried.  However, nothing came of this project, and no such attempt at deploying the glider was made as the war ended.

Professor Alexander Lippisch’s work involved designing unusual and unorthodox aircraft designs including the Li DM 1. Source: Professor Alexander Lippisch’s work involved designing unusual and unorthodox aircraft designs including the Li DM 1. Source: www.fiddlersgreen.net

After the war

When the war ended, the Si 204 would see more service in the hands of many other nations. The advancing Allies managed to capture a number of fully operational aircraft. These were immediately put to use either as transport, liaison, and evaluation purposes. At least one Si 204D was extensively used by the British pilot Captain Eric Brown, who was the chief test pilot of the Royal Aircraft Establishment at Farnborough. He was involved in a British project tasked with taking over German war research installations and interrogating technical personnel after the war.

He was generally impressed with the Si 204D’s overall performance, performing many flights on it. He later wrote about its performance. “The Si 204D was really a viceless airplane to handle, with inherently good stability about all three axes and good harmony of control. It was very well equipped for its tasks, and the later model I flew had an autopilot fitted. Like all German aircraft of that era, it was a mass of electrics, with extensive circuit breaker panels, and all very reliable.  However, the one thing the Germans never got right was wheel brakes, and the Sievel was no exception..”

A group of six or more Si 204 was captured by the Allies. Source: www.asisbiz.com

The Siebels that were moved to Farnborough were extensively used during 1945 for various roles, like communication, providing navigational guidance, and transporting pilots to various captured Luftwaffe airfields. The last operational flight of the Si 204D at this base was recorded at the start of 1946.

After the war, the Si 204 saw the most common use in French and Czechoslovakia, which actually continued to produce this aircraft. In French service, these were known t as NC 700, powered with As 411 engines, NC 701 ‘Martinet’, powered by two Renault 12S  engines, and NC 702,  a modified version of the Si 204A. In total the French constructed over 300 aircraft of this type. Some would see service in French Asian and African colonies. The last operational flight was carried out in 1964. Two NC 702’s would be given to Maroko in 1960, but their use and fate is unknown.

After the war, the French sold 7 NC 701 to Poland. They were used mainly for mapping photography. These were operated until the mid-1950s’ before being put out of service.

By mid-1960 some 5 French-built Siebels were given to the Swedish National Geographic Institut. These were mainly used for taking meteorological photographs.

The second country that produced the Si 204 was Czechoslovakia. They were built in two versions, the C-3 for the army and C-103 for civilian use. Both were mainly operated in their original transport roles. From 1945 to 1950 some 179 would be built.

The Soviets also managed to capture an unknown number of operational Si 204. These were briefly pressed into service before being replaced by domestic-built designs.

Switzerland also operated at least one Si 204D. This aircraft and its crew escaped from Germany on the 7th of May 1945 and landed at Belp near Bern. The Si 204D  would remain in Switz use under the B-3 designation.

Soviets operated an unknown number of Si 204. Their use was brief as it was replaced with new Soviet-built designs. Source: www.armedconflicts.com
During late 1945 and early 1946 the Si 204 were used by the Western Allies for transport and evaluation. Source: www.airwar.ru

Production Versions

  • Si 204 –  Prototype series
  • Si 204A – Transport and training version built in small numbers
  • Si 204B and C – Unknown fate, but likely paper projects only
  • Si 204D – Model with a new glazed cockpit and powered with a stronger engine
  • 204E – Experimental modification for combat operational use
  • Flying carrier – One Si 204 was to be modified as a carrier for the Doctor Alexander Lippisch experimental all-wing fighter, but was never fully implemented

Operators

  • Germany – Most produced planes were used by the Luftwaffe primarily used for crew training
  • Czechoslovakia – Produced some 179 additional aircraft for military and civilian use
  • France –  Over 300 modified aircraft (with French engines) were produced in France and saw wide service up to 1964.
  • Soviet Union –Operated some captured  Si 204
  • Poland – Brought 7  NC.701 from France after the war
  • Macoro – Operated two French NC 702
  • Sweden –  Operated five French-built Siebels
  • Switzerland – Used at least one Si 204 under the designation B-3
  • American and Great Britain – Both briefly operated a number of captured Si 204 after the war

Surviving aircraft

Today there are a number of partially or wholly survived aircraft Si 204. For example, the French Aviation Museum in Paris had one Si 204A and another located in the Escadrille du Souvenir close to Paris. One Si 204 is located at Sweden Lygvapen Museum.

Conclusion

While Germany in the Second World is better known for designing and producing a series of combat aircraft, their auxiliary aircraft are often overlooked. The Si 204 was one such case, despite its successful design, it is rather poorly documented in the sources. Its design was a success which can be seen in its after-war use, most notably by the French up to the mid-1960.

 

Si 204 D Specifications
Wingspans 21.33 m / 70 ft
Length 12 m / 39 ft 3 in
Height 4.25 m / 14 ft
Wing Area 46 m² / 495 ft²
Engines Two Argus As 411 engines
Empty Weight 1.500 kg / 3.300 lbs
Maximum Takeoff Weight 3950 kg / 8,710 lbs
Climb Rate to 1 km In  3 minute  30 seconds
Maximum Speed 364 km/h / 226 mph
Cruising speed 340 km/h / 210 mph
Range 1,800 km / 1,120 miles
Maximum Service Ceiling 7,500 m /  24,600 ft
Crew Pilot and his assistants plus eight-passenger
Armament
  • None

Illustrations

Si-204D
Si-204D
Si-204E
Si-204E

Credits

  • Article written by Marko P.
  • Edited by  Henry H. & Stan L.
  • Ported by Marko P.
  • Illustrated By Ed Jackson

 

Sources

  • D. Nešić (2008), Naoružanje Drugog Svetskog Rata Nemačka Beograd
  • H. A. Skaarup (2012) Axis Warplane Survivors
  • D. Mondey (2006). The Hamlyn Concise Guide To Axis Aircraft OF World War II, Bounty Books.
  • D. Donald (1998) German Aircraft Of World War II, Blitz Publisher
  • J. R Smith and A. L. Kay (1972) German Aircraft of the Second World War, Putnam
  • Jean-Denis G.G. Lepage (2009), Aircraft Of The Luftwaffe 1935-1945, McFarland & Company Inc
  • Captain E. ‘Winkle’ Brown (2010) Wings of the Luftwaffe, Hikoki Publication
  • M. Griehl (2012) X-Planes German Luftwaffe Prototypes 1930-1945, Frontline books
  • T. H. Hitchcock (1998) Jet Planes Of The Reich The Secret projects, Monogram Aviation Publication
  • M. Frazke Siebel Fh 104/Si 204, Flugzeug Profile

 

Lockheed C-130 in Argentine Service

Argentina (11 Aircraft)
Transport Aircraft

TC-63 and its crew on Puerto Argentino during the war for the disputed island. The pilots are holding a captured British Flag. (zona-militar)

Introduction

The pages of history are rich in aircraft that have a worthy place in the hall of fame of aviation. From fighters like the McDonnell Douglas F-4 Phantom, to ground attack aircraft, such as the Ju-87 Stuka or the Su-25 Frogfoot, or to civilian aircraft like the Queen of the skies, the world renowned Boeing 747 series.

One of these famous planes is the legendary Lockheed Martin C-130. The Hercules, as it is also known, has a place in the aviation hall of fame. From its service during the Vietnam War, to Operation Desert Shield and Desert Storm, and, more recently, the evacuation from Afghanistan of American forces, the C-130 proved to be a valuable asset to its operators. However, it is an aircraft with just as excellent a reputation in foreign service.

By the mid 1960s, Argentina did not have a viable Airlifter, the latest one in use being the severely outdated Ju-52. The Argentine Air Force was looking to satisfy this necessity, and it did so by acquiring the C-130 mid-weight air transport.

About the Hercules

After the Korean War, planes like the C-47, C-119, and C-46 were already starting to show their age and were no longer adequate for the needs of the time. Improvements had to be made, and fast.

On the 2nd of February 1951, the USAF issued a GOR (General Operating Requirement) for a new transport to Boeing, Fairchild, Lockheed, Douglass, Martin, Chase Aircraft, North American, Northrop, and Airlifts Inc.

This new transport plane was required to carry at least 92 passengers, 72 combat troops or 64 paratroopers in a cargo compartment that should be at least 41 ft. (12 m) long, 9 ft. high (2.7 m), and 10 ft. wide (3 m).

The USAF was very emphatic that this plane had to be specially designed as a military aircraft, and not a converted passenger plane. The use of a hinged loading ramp at the rear was also strongly suggested and newer technologies, like turboprop engines, were also available if the company chose to incorporate them. This specific engine could produce structural damage if it suffered a mid-flight failure, so security measures and structural reinforcement had to be taken.

The first prototype, the YC-130, first flew on the 23rd of August 1954. Since then, 29 versions have been developed, and it is still being operated by over 70+ countries, with the H variant, one of the combat cargo versions, being the most widespread of them all.

The variants of the C-130 range from combat cargo, tanker (KC-130), maritime patrol (PC-130), a psychological warfare version(EC-130J Commando Solo), a recon variant created for the then Iranian Imperial Air Force, and a ground attack (AC-130 Specter/Spooky) model equipped with automatic high caliber cannons and a 105 mm howitzer.

It is praised by its crews as a gentle and comfortable aircraft to fly, with very responsive controls. It is also considered a very maneuverable aircraft for its size. Designed to be structurally strong and able to land in even the harshest of terrains, it was a natural choice for the Argentine Air Force considering it had unpaved runways in Antarctica, and improvised landing strips in northern Argentina. The capability of this plane to also perform medical extraction missions was also crucial, as Argentina had no available aircraft to perform this kind of mission, let alone an aircraft capable to operate in rugged terrain or in the middle of a combat situation. This aircraft is optimized to perform combat transport duties, as it is sturdy, being able to resist small arms and even missile fragments. It also requires a very short take-off and landing distance, meaning it does not need a paved runway for take-offs or landings. All of these characteristics made it the perfect choice for the Argentines, as with this plane, they would have a choice besides ships for transporting personnel and supplies to their Antarctic bases, performing aeromedical evacuations, and supplying troops on the frontlines via airdrops or with a regular landing near the area of operations, without the need of a specialized or regular airstrip.

Layout of an aeromedical C-130 H (brookside press)

Its cargo bay is wide and tall, making it possible for its operators to load a variety of vehicles, personnel, and regular cargo. Humvees, paratroopers, APCs, and light tanks are many examples of the type of cargo the Hercules can carry. Its size goes according to the requirements stipulated in the GOR issued by the USAF.

Crew layout of a regular C-130. The pilot sits at the right of the picture, the copilot to the left, navigator center, the flight engineer at the far left, and cargo bay master sits opposite him. This layout is common for all variants. (picryl)

The first three, TC-61, TC-62, and TC-63

On the 15th of September 1967, a contract was signed between the FAA (Argentine Air Force, shortened in Spanish) and the USAF for the purchase of three C-130Es. Later that year, a training course was authorized and the instruction of the future crews for these aircraft began in late October. In July 1968, the first crew departed from Argentina for Sewart AFB, where a thorough training took place, and the pilots were introduced to the aircraft, its components, parts, manuals, and operation instructions. A crew of mechanics, as well as then Vice commodore (NATO equivalent Lt. Colonel) Ricardo Francisco Degano arrived at Sewart AFB. The mechanics were also assigned to an instruction course on the maintenance and repairing of the aircraft. Degano was assigned later that year as the first Squadron Chief once all training operations were concluded.

All personnel were then transported by the USAF to Marietta, Georgia, where the Lockheed construction and design facilities were located, and later accommodated at USAF Dobbins Air Reserve Base. There, the Argentine officers and petty officers received and inspected the three units bought by the FAA, their respective spare parts, technical support team, as well as operative and logistical teams, as signed in the contract. Even though the airplanes were received by the Air Force envoys in Marietta, the official reception took place in the Bahamas due to budget restrictions. This was a legal tax loop orchestrated by the USAF and FAA so that Argentina paid less in taxes and managed to afford 3 units of the C-130. Later that week, they flew back to Dobbins ARB, the planes then being piloted by Argentine crews.

In early December 1968, the three Hercules took off from Dobbins ARB, stopped in Howard AFB, and then proceeded to Mendoza International Airport in Argentina. A day later, the three C-130s took off, setting course to and later landing at the 1st Air Brigade base in El Palomar, Buenos Aires province.

Later that month, they received their official designation numbers, TC-61, TC-62, which in 1977 was modernized to the H variant, and TC-63, which was also later modified to the H variant .

On April 11, 1970, TC-61 was the first Hercules to land at the Marambio Joint forces military base in Argentine Antarctica. TC-61 is still in operational service as of today and her only remarkable duties were transport missions and airdrop missions over the Malvinas during the Conflict.

After some time in service, the Argentine crews began to call their C-130s “chanchas” (Spanish for female pig, remarking their big size and relatively slow speed). As of today, the crews still give this nickname to these planes.

TC-62

Assigned to meteorological duties on the flight route of Rio Gallegos, Santa Cruz province – Petrel Naval Base, Argentine Antarctica, this plane performed its tasks up until around mid-1975. It also performed mail delivery and cargo delivery duties to said base and to Marambio base.

By that time, It was transferred to perform logistical duties to fight off the Peronist and Communist guerrillas that were waging an insurrection in the northern part of the country. On the 28th of August 1975, it was destroyed by a terrorist attack while taking off from the international airport of Tucuman.

An IED was detonated on the runway, causing the aircraft to explode, and engulfed in flames as it skidded off the runway; 6 of the 144 Gendarmerie officers onboard (the Gendarmerie being the Argentine paramilitary force that performs both police and COIN/border patrol duties) were killed, and 28 others were injured, among them the 6 crewmembers.

TC-62 destroyed after the terrorist attack. (La Razon)

During the Malvinas/Falklands war, on the 1st of June 1982, TC-63, war codename “Tiza” (chalk), was given the order to scramble towards the last known position of the enemy fleet. The inclement weather, thick clouds and low visibility gave this plane the perfect opportunity to fly undercover and deliver supplies to the area of operations. After landing in Puerto Argentino (Port Stanley) and delivering its payload, TC-63 took off from the landing strip at 8:53 AM to perform a recon mission. It was tasked with looking for enemy vessels over the San Carlos Bay and designating them as targets for the Canberras and A-4s of the FAA.

At 10:25 AM, TC-63 reached her waypoint, and her radar signal was detected by HMS Minerva, the British vessel designated to control the airspace over that region. 20 miles north of the San Carlos Bay, a pair of Sea Harriers from the Royal Navy’s 801st Naval Air Squadron which were on combat air patrol duties. Piloted by Lieutenant Commander Nigel Wards and Lieutenant Steve Thomas, they scrambled towards the last known location of the Argentine Hercules.

Once intercepted, Ward launched an AIM-9L Sidewinder that plunged into the sea, as it was fired at an exaggerated range from its target. Ward pushed the throttle on his Sea Harrier, closed the gap with TC-63, and fired his second Sidewinder, which successfully impacted the right wing of the plane between engine 3 and 4.

Despite this severe damage, the Hercules, being loyal to its name, kept on flying firm and strong. Ward got even closer to the aircraft and fired his cannon upon it. The right wing finally broke off and the C-130 began to spiral down into the sea, disintegrating on impact. All 7 crew members died.

Illustration of the shoot down of TC-63. (Keith Woodcock)

TC-64

In early 1971, a fourth and fifth unit were purchased, this time both being H variants that differed from earlier E models by having updated T56-A-T5 turboprops, a redesigned outer wing, updated avionics, and other minor improvements. Similarly, both the C-130E and H carried 6,700 gallons of fuel in six integral wing tanks. Under each wing of the C-130E/H was an external pylon fuel tank with a capacity of 1,300 gallons. A pressure refueling point was in the aft side wheel well fairing for ground refueling. These new units were then designated with the identification numbers TC-64 and TC-65. One of these aircraft (TC-64) was the third to land in Puerto Argentino (also called Port Stanley by the British) under the war callsign Litro 3. On May 20, 1982, Litro 3 was shot at by friendly anti-air fire and by ground troops’ small arms fire over Fox bay, while it was on an airdrop supply mission to aid the Argentine 8th Marine division. Luckily, TC-64 managed to drop all her 10 cargo containers to aid the Navy marine division and fled the AO untouched.

The author has a personal connection with TC-64 or simply “64”, as the crew call it. It is common for crews to refer to individual Hercules by their designation numbers alone. That specific aircraft flew the author to see his father in 2019, who was stationed as an Officer in Command of the Resistencia AFB in Chaco province, Northern Argentina.

Hercules TC-64 taxiing in the Posadas International Airport, Misiones province, unknown date. This civilian airport also has the role of military airport for the Air Force. (FAA)

TC-65

TC-65 did not have a future as bright or lengthy as TC-64. It was assigned to the 1st Squadron, Transport Group 1 in late December 1971 and performed numerous duties over the Antarctic Argentine base of Marambio, from transport of supplies, MedEvacs, and VIP/personnel transport.

With the start of the Malvinas/Falklands war, 65 was assigned to perform transport duties on the Islands’ theater of operations. It was the last Hercules to flee the Islands. It was later abandoned in 2006 at El Palomar Airbase, with her nose severely damaged, and the plane lacking its engines. It was then cannibalized and salvaged for parts.

TC-65 in late December 2021. The poor state of the aircraft as of today can be appreciated. It was deprived of all critical and useful elements; its fuselage was sold as scrap in the final week of that month. (aerospotter)

TC-66

TC-66 was purchased by the FAA in mid-1972 and delivered to Argentina in September of the same year. It was also assigned to the 1st Squadron, Transport Group 1 and became the first plane to perform a trans-antarctic intercontinental flight in the world, connecting the International Airport “Jorge Newbery” in Buenos Aires, to McMurdo base in Antarctica, Christchurch in New Zealand, and Canberra in Australia. The trip was made in December 1973 and took 17 hours and 55 minutes.

During landing on Rio Grande base in Tierra del Fuego on the 11th of February 1998, it suffered an estimated 15% structural damage after hydroplaning and skidding away from the landing strip, halting abruptly on the surrounding soil. No casualties or injuries were reported, and it was then repaired by Lockheed Martin Argentina, in Córdoba Province Argentina.

TC-66 refueling in Marambio base, Argentine Antarctica. (FAA)

TC-67

This aircraft was acquired in late 1974, and delivered on the 3rd of March 1975. TC-67 was assigned to the 1st Squadron, Transport Group 1, where it performed Airdrop duties over the Orkney Islands Navy base.

TC-67 taxiing in Puerto Argentino. (Historias Individuales)

It suffered an accident landing at Marambio base, where it severely damaged her nose landing gear on the 23rd of November 1981, and because of this, TC-67 was the only Hercules not to be reported to have been deployed over the Malvinas War Theater. However, pictures of it parked in Puerto Argentino have been found suggesting its use during the conflict.

After attempting to land at the Tandil Airbase on the 16th of May 1996, it skidded off the runway and the subsequent impact caused around 9% of structural and landing gear damage.

Repairing this aircraft proved not to be worth it considering the budgetary restrictions of the Air Force. The wings were sold off to the Pakistani Air Force, and the fuselage was donated to the Malvinas War Museum in Pilar, Buenos Aires Province, where her cargo bay is being used as a cinema. Her engines were salvaged.

TC-68

Perhaps the most interesting specimen of them all, TC-68, was acquired alongside TC-67 and delivered a week after it. Acquired in early March 1975 and delivered on the 10th of March same year, it performed regular transport duties during peace time for the 1st Squadron, Transport Group 1 located in El Palomar AFB.

It was the first Hercules to land at Puerto Argentino and performed an airdrop supply mission over Darwin on the 19th of May 1982.

Using what’s known in Argentina as “Viveza Criolla” or Criollo quick thinking, some field officers decided to modify TC-68 to perform Frontline Bomber duties after performing recon missions over the exclusion zone.

It was first used against the tanker HMS British Wye with relative success, landing two of her bombs on target. However, this is not the mission it is most known for.

Not many were the ships whose captains were brave enough to venture or deliver payload inside the war’s exclusion zone. The ironically named tanker “Hercules” was one of these.

Being a tanker operated by the United States under a Liberian flag, the Hercules was heading towards Ascension Island when it was spotted by an “unidentified military aircraft”. That aircraft was another FAA C-130 which was flying over the area in search of enemy vessels at around 13:00 PM. Captain Bataliari ignored the event, as he identified the plane as being a military cargo plane, and he thought it was likely an RAF or Royal Navy aircraft.

Exactly six minutes later, a pair of planes, consisting of a FAA Canberra MK-62 and a C-130H, were spotted flying in a straight intercept course with the vessel. No one onboard the tanker could believe what they were about to encounter: TC-68 was fitted with a SFON system ,the ground-attack oriented sights available in the IA-58 Pucará; some fielded officers say that this sights may have been extracted not from a damaged Pucará, but from the Canberra, as the jet bomber used the same kind of sights. It also carried a pair of wing mounts, extracted from a damaged Canberra, and loaded with a total of fourteen 250 kg dumb bombs. These were mounted where , regularly, drop tanks were installed.

TC-68 restored with mock-up bombs installed to show how they were installed during the war, as the aircraft today serves as a museum piece. (Aztec Models)

The plane had taken off from Comodoro Rivadavia on the 8th of June 1982 at 10:00 AM, tasked to attack this tanker. The mission was conducted by the pilot of the Hercules, Vice commodore Alberto Vianna.

Vianna recalled the mission:

“The day before, a Boeing (The Air Force’s pair of Boeing 707s were regularly used for recon missions over the islands as well as for troops transport) detected the ship traveling towards where the British fleet was stationed. They (HQ) sent us to intercept it.

When we arrived, the first surprise was that the oil ship was 320 meters long, and the second one was that it was called Hercules!

If I had to say that a Hercules was going to face another Hercules in the middle of the Atlantic, no one would believe me.”

Once encountered, many radio warnings were sent in both Spanish and English over different frequencies to order the ship to change its course, however, the tanker opted to ramp up her speed and escape towards the British fleet.

The order to attack was sent to Vianna. He dived with his Hercules from 1000 ft of altitude, expecting to release his bombs at 150 ft, 550 kph, and at a 45° angle.

The Vice commodore released eight of the fourteen loaded bombs. Four of them impacted the sea without exploding, three then exploded near the port (left side) of the ship, and the last one ricocheted off the ship’s deck and exploded over the sea. Despite the poor accuracy, structural damage was achieved on the ship by the shockwaves.

Illustration of the attack of the Hercules. (Pakistan Defense)

The coup de grâce was given to it by the Canberra flying alongside TC-68, when it dropped her three 500 kg MK-17 bombs over the ship. One of them got stuck in one of the ship’s orlops unexploded. Hercules fled the area towards Rio de Janeiro, while the two planes flew back to mainland Argentina. There, the damage and the impossibility of removing the bomb rendered the ship permanently disabled. Most of the oil was then salvaged. The ship was dragged to Brazilian controlled high seas and the bomb was detonated remotely, sinking it.

TC-68 performed 5 more bombing runs in 5 different missions with relative effectiveness.

Sadly, 27 years later, it was decommissioned and salvaged for parts. Abandoned and engineless, it was left behind as scrap inside the base of El Palomar in March 2009.

TC-68 at her final resting place. (Foro Militar)

However, unlike some other examples, TC-68 had a well-deserved restoration ordered and completed in 2018. It now serves as a museum in said base.

KC-130 TC-69 and KC-130 TC-70

The addition of these tankers to the air fleet proved to offer an invaluable tactical advantage, as the Air Force could now deploy their planes and perform missions at much farther ranges.

Nicknamed today as “Puerto Argentino(TC-69) and BAM Malvinas(TC-70), they were acquired at the same time and delivered with 2 weeks of difference between each other, TC-69 on the 24th of April 1979 and TC-70 on the 10th of May 1979. Both “chanchas” operated under the 1st squadron, Transport Group 1 in El Palomar AFB, and TC-69 was the first aircraft to perform an air refueling mission for the FAA on the 12th of June 1979, refueling a squadron of A-4Cs.

During the Malvinas war, it also performed air refueling missions to different CAPs and Strike groups.

TC-69 after being modernized, having her batteries charged, late December 2021. (Via Pais)

It was sent on the 15th of October 2014 to Waco, Texas, for a full modernization of her avionics and cockpit elements, installing a modernized and all-digital navigation system, new communications equipment, and a new controls system and digital instrumentation. Due to a lack of payment from the Argentine Air Force, the aircraft was retained for some months but later released, as the Argentine Ministry of Defense paid out the debt and modernization duties continued back in her home country. After the main modifications and modernizations ended on the 27th of April 2019, it was presented and delivered to El Palomar AFB once again. Smaller updates and modifications are still ongoing as of the time of writing this article, in her home country and it still is in active service today.

Screenshot of a video captured of TC-70 performing air refueling over San Luis province to a pair of A4AR Fightinghawk, circa 2019. (Foro militar)

TC-70 performed similar duties during peacetime and war time. It is most well known for being the Hercules that refueled the squadrons that performed the attack and sinking of HMS Sheffield, HMS Atlantic Conveyor, and HMS Invincible.

It was the third Hercules to be modernized and the first one to do so fully in FAdeA (Fábrica Argentina de Aviones, Argentine Airplane Factory, the state-owned aircraft manufacturing factory) on 19th of September 2018.

TC-100

Lastly, there is the newest Hercules in the air fleet, TC-100.

Accepted into service in December 1982, it was the technology demonstrator aircraft of the L-100-30 variant for Lockheed. It was then leased by Karu Kinka Airways, a regional southern Argentinian airline that operated solely between Rio Grande in Tierra del Fuego, and Buenos Aires. For legal issues, it received the plate LQ-FAA, then LV-APW (while operated by Karu Kinka), and finally in 1998, TC-100, when it was returned to the FAA. It is the biggest cargo plane at the disposal of the Air Force thanks to her elongated fuselage.

As an interesting side note, it is also the only C-130 to receive a different camouflage compared to her sisters.

TC-100 was initially operated with a white livery with red lines. (Pedro Borsero)

 

It is still being used today, and more recently, performed flights to repatriate Argentine citizens from abroad when international lockdowns were enacted at the beginning of the COVID-19 pandemic. It also took Argentine medics to Cuba and returned home with Cuban medics to help relieve those infected with the virus.

Conclusion

Replacing the venerable, and very dated Junkers Ju 52, the C-130 provided the FAA with many capabilities. Able to operate from some of the harshest conditions in South America, the C-130 has proven to be an invaluable and versatile asset for the Argentine Air Force, acting as a transport plane, refueling aircraft, and even as a long-range frontline bomber.

Specifications (C-130H)

  • CREW: Five (two pilots, one navigator, one flight engineer, and one cargo bay master)
  • CAPACITY:
  1. Troop transport: 92 soldiers, 72 soldiers with full combat load or 64 paratroopers
  2. Medevac: 72 stretchers and 2 Medic crews with full equipment.
  3. Cargo: 6 fully loaded pallets.
  4. Vehicle transport: 2-3 Humvees or 1 M113 APC.
  • MAXIMUM LOAD CAPACITY: 20,000 kg or 44,080 lb.
  • LENGTH: 29.8 m or 97.8 ft.
  • WINGSPAN: 40.4 m or 132.5 ft.
  • HEIGHT: 11.6 m or 38.1 ft.
  • WING AREA SURFACE: 162.1 m2 or 1744.9 ft2.
  • EMPTY WEIGHT: 34,400 kg or 75,817.6 lb.
  • MAX CARGO WEIGHT: 33,000 kg or 72,732 lb.
  • MAX TAKEOFF WEIGHT: 70,300 kg or 154,941.2 lb.
  • ENGINES: 4x Turboprop Allison T56-A-15
  1. Power: 3430 kW (4729 hp; 4664 CV) each.
  • PROPELLER WIDTH: 4.17 m, 13.68 ft
  • MAX SPEED/CRUISE SPEED: 592 km/h at 6060 m. / 540 km/h. 367.85 mph at 19882 ft.
  • RANGE: 2052 nmi
  • MAXIMUM OPERATION CEILING: 10,058 m empty / 7010 m with 19,000 kg of load. 33,000ft  empty, 23,000 with 42,000lbs of load.
  • TAKE OFF DISTANCE: 1093 m with full load, 427 m empty. 3586ft loaded, 1400ft empty.
  • AVIONICS: Meteorological and navigation radar Westinghouse Electric AN/APN-241 (cheesy side note, my father worked with these ones! He’s a radar engineer and operator in the Air Force)

Illustrations

Credits

  • Article written by Wilhelm
  • Edited by  Henry H. & Stan L.
  • Ported by Henry H.
  • Illustrated by Godzilla

Sources:

https://loudandclearisnotenought.blogspot.com/2011/06/tc-100-lockheed-l-100-30-hercules-cn.html

https://www.jetphotos.com/airline/Karu%20Kinka%20Lineas%20Aereas

https://www.aviacionline.com/2020/08/fotogaleria-los-hercules-de-la-fuerza-aerea-argentina/

https://defensayarmas.blogspot.com/2018/09/probaron-el-kc-130-en-reabastecimiento.html

https://en.wikipedia.org/wiki/Lockheed_L-100_Hercules

https://airlinehistory.co.uk/airline/karu-kinka-lineas-aereas/

https://www.zona-militar.com/2021/12/24/el-ministerio-de-defensa-y-fadea-avanzan-en-la-recuperacion-del-hercules-tc-69/

https://deyseg.com/malvinas/627

https://www.gacetaeronautica.com/gaceta/wp-101/?p=29058

https://loudandclearisnotenought.blogspot.com/2011/06/tc-70-lockheed-kc-130h-hercules-cn-4816.html

https://loudandclearisnotenought.blogspot.com/2011/06/tc-69-lockheed-kc-130h-hercules-cn-4814.html

https://es.wikipedia.org/wiki/Lockheed_C-130_Hercules#:~:text=El%20KC%2D130H%20TC%2D69,de%20las%20Malvinas%20en%201982.

https://www.lanacion.com.ar/politica/los-locos-de-la-guerra-pilotearon-el-hercules-que-se-volvio-en-una-pesadilla-para-las-tropas-nid01042022/

https://malvinasguerraaerea.blogspot.com/2019/04/el-hercules-bombardero-tc-68-fue-recuperado-del-olvido.html

Sopwith T.F.2 Salamander

United Kingdom (1918)

Ground Attack Aircraft [300-500+ Built]

A frontal view of a production Sopwith Salamander. The entire front section of this aircraft was armored. (Wikipedia)

The Sopwith Salamander was a dedicated ground attack aircraft, at this point known as a trench fighter, designed for use by the Royal Air Force in the First World War. The Salamander was based off of the Sopwith Snipe fighter and reused many components, but was much more armed and armored. Only a few Salamanders would be assigned to squadrons for testing during the war and none would see frontline combat. After the war, the Salamander was in service with squadrons in British territory until at least 1922. The aircraft was interesting as, in addition to its other modifications, it would be one of the first aircraft to be officially painted by the RAF in camouflage, most likely being the first in RAF aircraft to do so.

The Trench Fighter: Birth of the Ground Attacker

Rear view of the T.F.1 Camel. This was Sopwith’s first attempt at a dedicated Trench Fighter before the Salamander. (Sopwith Aircraft from 1912-1920)

Late into the First World War, the British Royal Air Force began using single-engine fighters to deliberately attack enemy trenches. This was seen at the Battle of Ypres and Cambrai in 1917. Oftentimes, the types used for this role could not perform well enough to dogfight or had some other glaring issue that prevented them from seeing widespread service. Although not their original purpose, these “Trench Fighters” were the first evolutionary step to creating what is now known as dedicated ground attack and close air support aircraft. The Sopwith Aviation Company began experimenting with dedicated, purpose-built trench fighters in 1918. The first of these was a derivative design based on their famous Sopwith Camel fighter. The T.F.1 Camel, TF standing for Trench Fighter, was a modified Sopwith F.1 Camel that had additional armor and was to be used to strafe trenches with a machine gun or bombs. Despite work being done on the T.F.1, it was only considered as a test for a trench fighting aircraft and was never meant to enter service nor production.

Instead, the Royal Air Force was looking for an aircraft with a more powerful engine, which the Camel airframe could not accommodate. Sopwith looked instead to their recently developed Snipe fighter. The Sopwith Snipe aircraft had been designed in late 1917 as a successor to the esteemed Sopwith Camel. It would not enter widespread service until September of 1918 and would only see combat for three months before the end of the war. Despite its short combat service, the Snipe proved itself as one of the most advanced fighters of the time, thanks to its powerful engine and excellent maneuverability. All of this had yet to be proven, however, when the trench fighter derivative design was being drawn up, as the Snipe had only just started testing in late 1917.

Official work began on the trench fighter Snipe in January of 1918. This machine was seen to have several advantages over the TF1. The newer design of the Snipe proved to be much more agile and it was able to carry the powerful 230 hp Bentley BR2 rotary engine. There were three factors that sought to specialize the design of this new aircraft; engine, armor and armament. A rotary engine was favored over an inline on the aircraft because an armored cowling could easily fit over the engine and was thus less likely to be hit from ground fire. For armament, it was planned to have a single forward facing Vickers machine gun with two more in a downward firing position, akin to the armament of the TF1. This idea was ultimately scrapped and two forward facing Vickers were chosen instead, like the armament on the Snipe. Relating to the armor, the front section of the fuselage was made to be a heavily armored box that would protect the pilot and engine from enemy fire. It was optimistically thought only three things would be able to shoot this new aircraft down; a direct hit from anti-air artillery, damage to the flying wires or heavily damaging the main spar. Three prototypes of the new trench fighter aircraft began construction in late January 1918. The first of these would be airworthy and ready in April. By now, the aircraft had received an official name; the Sopwith T.F.2 Salamander.

An example of a production Sopwith Snipe. This would be one of the best aircraft the RAF would field in the later stages of the First World War, and is the aircraft the Salamander would be based on. (Pilots and Planes)

Design

A cockpit view of the aircraft. (Imperial War Museum)

The Sopwith T.F.2 Salamander was an early ground attack aircraft based on the Sopwith Snipe fighter. The two aircraft shared many components, but the Salamander would have a number of features that would make its design unique. It had a wingspan of 19ft 6in (9.5 m). The wings were of two bay construction and consisted of a frame covered in canvas. The fuselage was of all wooden construction and covered in fabric, like the Snipe. It had a length of 19ft 6in (5.9 m). In total, the aircraft had a height of 9ft 4in (2.8 m). The sides of the fuselage were flat, being a change from the rounder fuselage of the Snipe.

In the front of the aircraft would sit the 230-hp Bentley B.R.2 air-cooled radial engine. The eleven-cylinder Clerget 11E engine was an alternative to the Bentley, but no Salamander would be equipped with this engine. The engine and cockpit section of the aircraft would sit in an armored box that would protect its most vital assets. The armored box was 8 mm thick in the front (the armor over the engine and the engine itself also factored in as frontal protection), 6 mm for the sides, 11 mm for the floor, and 10-gauge sheet metal with an additional 6-gauge sheet at the rear. In addition to the armored box, the engine would have an armored cowling over it. The aircraft had around 650 Ibs of armor in total. The sheer amount of armor was meant to protect the aircraft from German anti-armor rounds fired from short range, something it would no doubt deal with at the frontlines.

The controls and cockpit were likely carried over from the Snipe. Behind the cockpit was an armored head fairing that was not present on the Snipe. This detail is a distinct visual difference that one can use to identify the Salamander over the Snipe. Beneath the cockpit was the undercarriage and landing gear. During testing, it was found the armor made the aircraft quite hard to land, and the landing gear was further reinforced during development to assist in this area. The fuselage would taper towards the rear and tailplane. Beneath the tail was a simple landing skid. The tail and rudder were small at first on the prototype Salamanders, like on earlier Snipes, but this would be replaced by a larger rudder and tailfin on the production versions. At first, the tailplane was rigged via wires but this was replaced by four steel tubes connecting at the top and bottom.

A view of the armored front section of the aircraft. (Weapons and Warfare)

For fuel, the Salamander would carry less than the Snipe to accommodate the extra weight of the armor. The fuel delivery system was composed of a Badin vacuum-feed system with a Weyman hand pump connected to the main petrol tank for standby use. The fuel delivery system was protected with armor and rubber along the piping to prevent leaks or fire. In addition to the main petrol tank, there was an oil and gravity tank connected via piping.

The armament of the Salamander went through a number of iterations before its final layout. Originally, the aircraft was going to have a single forward facing Lewis machine gun, with two more facing downwards into the hull, but this was replaced by two synchronized Vickers guns that were staggered to house more ammunition (1000 rounds each). There exist other known layouts pf the Salamander but it is unknown if any of these were tested at any point. These included eight downward firing guns in one layout and two downward facing Lewis guns with two more over the center (in addition to the standard two Vickers). No photos of these two layouts exist. For special missions, the Salamander could carry up to four 20 Ib (9 kg) bombs or a single 112 Ib (51 kg) bomb.

A direct frontal view of the Sopwith Salamander. (Wikipedia)

The Sopwith Salamander: World War Woes

Rear view of the 3rd prototype Salamander. This example has the early rudder. Unfortunately this particular aircraft would be lost in a crash. (Pilots and Planes)

The Salamander would have its first flight on April 27th at Brooklands. The prototype Salamander, E5429, shared the wing mainplane, ailerons and tail control surfaces with the early model Snipe, but these would be improved later on the production models. The improvements were the same as done on the Snipe, which included increasing the size of the rudder. On May 9th, the first Salamander prototype was sent to France for service testing. There is a strange overlap in information with the prototype. Some sources claim that it returned to England on June 30th for further testing at Martlesham Heath, but others claim the prototype was lost to a crash in France on May 19th. Perhaps this was confused with the 3rd prototype, which did crash at a later unknown date. By this point, the other two prototypes were completed (E5430 and E5431). Testing found that the aircraft performed well, but problems appeared with the controls, which were found to be sluggish due to the extra armor.

The Salamander did have its fair share of critics, with several pilots being harsh towards the slower controls of the aircraft and some even finding the concept of an armored aircraft a waste of resources. Many of those who were strong critics of the aircraft criticized it as they did conventional fighters of the time, glossing over its specialized role of ground attack and arguing its armor would make it sluggish in a dogfight, when the aircraft was never intended to operate as a dogfighter. Originally, a plan for 6 prototypes was made but the last 3 were canceled. The 3rd prototype would stagger its machine guns to accommodate the increased amount of ammunition the Salamander had over the Snipe. This change would be present on all Salamanders going forward. With the aircraft performing well in testing, an initial order of 500 aircraft was requested in the early summer months of 1918. Sopwith would begin building production Salamanders at their factories, being constructed alongside the Snipe. In addition to Sopwith, several other aircraft manufacturers would begin constructing Salamanders as well; Air Navigation Co Ltd, National Aircraft Factory No.1, Palladium Autocars Ltd, Glendower Aircraft Co Ltd and Wolseley Motors Ltd. The production versions differed from the prototype Salamanders, having the larger tail fin and rudder as well as the ailerons from the production Snipes being fitted, as well as the staggered machine guns from the 3rd prototype.

A production line at a Sopwith factory where both Salamanders and Snipes are under construction. The first row are incomplete Salamanders. (Armament of British Aircraft)

As the year went on, production for the Salamander increased, as the order jumped from from 500, to 600 to 1400 by the war’s end. Producing the Salamander was found to be more difficult than the Snipe, thanks to its complicated wiring due to the extra steps of creating the armored cockpit area. Problems also began to be found with the armor, as the box was found to warp after some time and distort the frame. This was not a known problem at first, but it plagued many of the early production versions after the war. In October, production Salamanders began being painted in unique disruptive camouflage patterns. This practice started on the 3rd prototype. This would be one of the first times the RAF would officially camouflage paint aircraft, something that would eventually become a mainstay in the next World War. By early November, two Salamanders were sent over and stationed in France, with one being assigned to No 86 Squadron at Phalempin. No 86 Squadron had just been assigned as a dedicated ground attack unit when it arrived. Back in Britain, squadrons No 95 at Weyton, and No 157 at Upper Heyford were also reworked to be dedicated trench fighting squadrons and equipped with five Salamanders each. No 157 Squadron was scheduled to leave for the front on November 21st. With production rapidly increasing and the aircraft soon to be used at the front, all of this was suddenly brought to a halt when the Armistice was signed on November 11th.

 

Postwar Mediocrity

A Sopwith Salamander showcasing its unique camouflage livery (RAF Museum)

With the signing of the armistice, all plans to ship the Salamander-equipped squadrons to the front were canceled. Production was soon to be cut short as well, as the need for such a specialized aircraft disappeared. Gradually, the order of 1400 was decreased to a much smaller number. Sopwith and Glendower continued producing the Salamander until mid 1919, when total production was completely halted. The other companies mentioned before either stopped production entirely or produced only a few more Salamanders after the Armistice. The Salamander was prepared to be used in full force had the war continued into 1919, with an expected thirteen full Salamander squadrons stationed in France by May. There were expected delays with the production of the Bentley engine, so five of these squadrons were to be equipped with the aforementioned Clerget engines. The exact number of Salamanders produced varies from source to source. The most common number found is that 210 were produced in total, but other sources claim that the actual number is closer to 300. Others claim that almost 500 were built. None of these numbers can truly be confirmed but it is likely much more than the commonly thrown around 210.

Rear view of a Sopwith Salamander (Imperial War Musuem)

Postwar, the Salamander did not find itself too popular, as many issues rose up with the design. The warping of the armor began to become a serious problem on early production Salamanders and it was also found the first 70 Salamanders built by Sopwith had upper wings from Snipes, which were not capable of supporting the heavier Salamander. All of these 70 aircraft were found to be extremely dangerous to fly and it took until December of 1918 for the problem to be realized and fixed. From what can be gathered, most of the production Salamanders were put into storage after the Armistice, with many being finished and immediately sent into storage. Flight testing of the type continued until 1920 despite all interest in the Salamander seemingly being lost in mid 1919.

In addition to the disruptive camo, there is mention of a Salamander being painted in a type of lozenge camo, similar to German aircraft schemes in the war, but no photos are known to exist. It was to be tested at Farnborough alongside the regular camo in July of 1919 but it was unlikely anything became of the tests. Despite the lack of interest, the Salamander did occupy a number of squadrons post war, however the details of where and when are sparse. The latest Salamanders mentioned in RAF service were a squadron stationed out of Egypt in 1922. This would have coincided with the Chanak Crisis against Turkey. A few Salamanders were sent to foreign nations for testing. An unknown Salamander was sent to France to be tested by the Section Technique de l’Aéronautique (Aeronautical Technical Section) in Villacoublay, France. Salamander F6533 was sent overseas to America for trials and testing by their Army Air Service. No further orders or Salamanders were made by America after this and the sole example was known to have been still at McCook Airfield as late as 1926. It is likely the warping issue happened with this particular aircraft, as beneath the cockpit “This machine is not to be flown.” was printed and was seen in photographs of the aircraft.

Salamander F6533 at Mccook Airfield (Pilots and Planes)

Many combat aircraft of the First World War found new life in the following years in the hands of private collectors or attending airshows for spectacular performances. The Salamander was unfortunately not one of these aircraft due to its specialized nature and slower performance compared to the fast aircraft that were featured in such displays. With the purpose of the aircraft now gone and with no future in sight, the Salamander was left to be forgotten as newer aircraft replaced it in squadrons and eventually all would be scrapped. None survive to this day.

Conclusion

The Salamander was one of the first British attempts to create a dedicated ground attack aircraft. In addition, it first tested camouflage patterns on RAF aircraft. Unfortunately, it came too late, if only by a few weeks, to be tested in combat. With the war over and the need for such an aircraft gone, the dream of the Salamander strafing enemy positions died and it fell into obscurity as the type was eventually completely scrapped. Had it entered combat, it would have encountered the same problems it did postwar, which would have left the aircraft prone to accidents of its own design and would have taken time to repair in the field. A strange, and perhaps sad, note is the Salamander was the last Sopwith aircraft to enter service with the RAF before the company became defunct in 1920.

Variants

 

  • Sopwith T.F.2 Salamander Prototypes – The first prototypes for the Salamander had many of the same features as the Snipe, including sharing the mainplane, unstaggered guns and the tailplane was supported by wires.
  • Sopwith T.F.2 Salamander Production – The production version of the Salamander had staggered guns, provisions for carrying bombs, and the tailplane was supported by four steel rods. The first 70 production aircraft accidentally were equipped with the upper wings of the Sopwith Snipe.

 

Operators

 

  • United Kingdom – The Sopwith Salamander was built as a dedicated Trench Fighter for the Royal Air Force, but hostilities would stop before it could be sent to the frontlines. After the war, most Salamanders would be put in storage, but a few would be sent abroad, such as to Egypt.
  • United States of America – A single T.F.2 Salamander (F6533) was sent to McCook Field for testing.
  • France – A single T.F.2 Salamander was sent to France for testing with the Section Technique de l’Aéronautique in Villacoublay, France.

Sopwith T.F.2 Salamander Specifications

Wingspan 31 ft 2 in / 9.5 m
Length 19 ft 6 in / 5.9 m
Height 9 ft 4 in / 2.8 m
Wing Area 272 ft² / 25.3 m²
Engine 1x 230 hp ( 171.5 kW ) Bentley B.R.2 Radial Engine
Propeller 1x 2-blade wooden propeller
Weights
Empty 1844 lb / 836 kg
Maximum 2512 lb / 1139 kg
Climb Rate
Time to 5,000 ft / 1,525 m 6 minutes 5 sec
Time to 6,500 ft / 1,980 m 9 minutes 6 sec
Time to 10,000 ft / 3,050 m 17 minutes 5 sec
Maximum Speed 117 mph / 188 km/h at 10,000 ft / 3,050 m

123 mph / 198 km/h at 6,500 ft / 1,980 m

125 mph / 201 km/h at 3,000 ft / 915 m

Cruising Speed 125 mph / 201 kmh
Endurance 1 ½ hours
Maximum Service Ceiling 13,000 ft / 3,690 m
Crew 1 pilot
Armament
  • 2x synchronized Vickers .303 machine guns (1000 rounds per gun)
  • 4x 20 Ib (9 kg) bombs or 1x 112 Ib (51 kg) bomb

Illustrations

The Salamander in standard RAF livery

 

Several Salamanders would receive a standardized camouflage pattern, they were among the earliest RAF planes to use an official camouflage livery.

Credits

  • Article written by Medicman11
  • Edited by  Henry H. & Stan L.
  • Ported by Henry H.
  • Illustrated by Carpaticus

 

Sources

https://www.baesystems.com/en/heritage/sopwith-salamander

https://www.rafmuseum.org.uk/blog/salamandrine-fire/

King, H. F. Sopwith Aircraft, 1912-1920. Putnam, 1981.

Mason, Francis K. The British Fighter since 1912. Naval Institute Press, 1992.

Green, W. and Swanborough, G., n.d. The complete book of fighters.

 

 

Focke-Wulf Ta 152H

Nazi flag Nazi Germany (1945)

Fighter: Approximately 60 built

A Ta 152H undergoing compass calibration. (flugrevue)

Introduction:

Throughout the Second World War, the job of the interceptor would become ever more challenging. Their targets, mostly bombers and photo reconnaissance aircraft, would fly ever higher and faster thanks to new advancements in turbo and supercharging. With Germany under a state of permanent siege and surveillance by aircraft like the Boeing B-17 and De Havilland Mosquito, it was clear the Luftwaffe needed a specialized interceptor to effectively reach these high flying threats and the multitude of new fighters that were appearing in growing numbers. After several failed attempts to develop the Fw 190 into such an interceptor, Kurt Tank designed the Ta 152H. The short lived design incorporated all of the available developments in high altitude flight available to German aviation in an attempt to create the ultimate high altitude fighter.

High altitude threats and Interceptors

In the summer of 1941, the Mosquito was making its first reconnaissance sorties and becoming one of the gravest threats to German aerial defenses. Operating above 7km and capable of reaching speeds upwards of 560 km/h, the aircraft was almost untouchable after it had reached its destination. Once they had taken their photos, they turned for home and entered a shallow dive that allowed them to accelerate to speeds beyond those of pursuing fighters who were not already chasing them from a higher altitude. With such a small interception window, they were a chief concern to the Luftwaffe. Doubly so were the bomber variants of the aircraft, which raided targets all over North Western Europe.

The following year saw the entrance of the United States into the Second World War, their air force possessing some of the most capable high altitude aircraft at the time. Investments in engine turbocharging allowed them to field a number of bombers and fighters with exceptional high altitude performance. B-17’s were conducting regular operations above 7 km. At first, they undertook operations at significantly lower altitudes, never straying too far from their air bases in southern England, but it was becoming clear that they would soon pose a threat that the Luftwaffe was ill equipped to combat.

The De Havilland Mosquito was quickly recognized as a serious threat not long after its introduction. Photorecon versions, like this later Mk XVI here, could surveil Germany with little fear of interception. (wikimedia)

The only two fighters of consequence employed by the Luftwaffe, the Bf 109 and Fw 190, were effective low to medium altitude fighters. However, through 1942, both were operating with engine power restrictions, and supercharger related performance bottlenecks. While inferior alloys and lubricants were causing a variety of issues, that was less of a concern than the engines themselves not being designed for use at high altitudes. The Bf 109G’s DB 605A, with its variable single stage blower, provided a full throttle height of roughly 6.5 km, depending on the variant. The Fw 190’s BMW 801, with its significantly simpler, single stage, double speed supercharger, was even worse off. Its critical altitude was only roughly 6 km, leaving it, and the 109, distinctly lacking in power at the over 7.5km B-17’s often flew at. Above these altitudes, neither engine could maintain the manifold pressure needed for combat power, putting them at a distinct disadvantage in trying to catch the Mosquito, or fighting American high altitude fighters which were soon making forays into German airspace. As the USAAF began its strategic bombing campaign against Germany proper, there were deep concerns within the Luftwaffe about the battle they were soon to fight, and for which they were clearly technically unprepared for. Even more concerning was the fear that the RAF would soon be operating the Vickers Wellington V bomber, which was reportedly capable of operating at an almost untouchable altitude of 12 km. They never entered service, but were the impetus for the creation of a specialized high altitude fighter with the Höhenjäger program.

With these anxieties building, the RLM convened a conference on the development of high altitude fighters on May 20, 1942 at Messerschmitt’s plant in Augsburg. In addition to the high altitude British bomber, further concerns were spelled out over the recent study of the new Merlin 61 engine, which, with its two stage, two speed supercharger, promised to make the Spitfire and Mosquito even more challenging opponents at high altitude. Of particularly grave concern was that the German aviation industry could not simply follow the same development path as the Allies. The poor qualities of their available alloys and the inadequate supplies of high octane fuels meant that even, if they had a factory furnished with all the tools to manufacture an engine like the Merlin 61, they simply could not build or operate it with the materials at hand.

As such, they had to pursue less conventional means of improving performance. Messerchmitt proposed a redesign of a former naval fighter proposal for high altitude use. The Me 155 carrier based fighter design, with its very long wingspan, was proposed to be converted for high altitude use, the work being done mostly at the S.N.C.A.N plant in Paris. The design would later be taken up and heavily altered by Blohm & Voss, who went on to design the Bv 155, with turbochargers and GM-1 nitrous boosting. Neither design came to fruition. A secondary design, the Bf 109H, would involve stretching the wingspan of a Bf 109F, and later G, and installing the high altitude GM-1 engine boost system. Likewise, this design was not pursued. In the end, Messerschmitt would go on to design a mass production, high altitude variant of their standard Bf 109G with a pressurized cockpit and nitrous boosting. While it would prove fairly adequate for the time, it was held back by the need for GM-1, which was difficult to transport in large quantities without a pipeline.

Focke-Wulf would face an even greater challenge with their program. While their Fw 190 was proving to be among the best medium altitude fighters of the war, its short wingspan and outdated supercharger meant it would take a considerable effort to make a high altitude fighter out of it.

The Höhenflieger Fw 190

Focke-Wulf first pursued turbocharging to get their fighter to reach the adequate level of performance for the Höhenjäger project. Almost immediately, they ran into the issue that it was almost impossible to fit a suitable turbocharger into a Fw 190A, though an externally mounted, and almost completely unwieldy unit was suggested. The first serious effort came with the proposal for the Fw 190B fighter, or Höhenjäger 1, in August of 1942. The design would take the then in production Fw 190A-3, increase its wingspan from 10.5 to 12.4 meters (increasing its area from 18.3 to 20.3 m^2), and install a pressurized canopy. The engine was initially unmodified and nitrous boosting was not pursued, in the hope a suitable turbocharger would be developed. The prototype, Fw 190V-12, began testing, but was abandoned in favor of using older, pre-production Fw 190A-0 prototypes before moving on to pre-production. The Fw 190B-0 received the new BMW 801 D-2 and several other modifications going into the new A-5 fighter. It began testing in December of 1942, and despite some faults with the pressurized canopy, which were later corrected, the aircraft had considerably better high altitude handling than the original A model. All four of the A-0’s were converted, but the program showed little promise. Despite the effort, the improvements were not enough and the aircraft was still too slow at high altitude. It was clear that the aircraft needed a heavily modified, or entirely different engine, in order to attain the level of performance needed.

Perhaps the most promising development for the next generation of the Fw 190, the C series, hoped to install a much more powerful DB 603 engine. Harsh teething issues and limited supplies for the engine doomed the project. (grafiq)

In parallel with the B-project, the decision was made to re-engine the aircraft with either the Junkers Jumo 213, or Daimler Benz’s DB 603. Both promised better high altitude performance over the BMW 801 along with a considerable overall increase in engine output. The DB 603 project would proceed with the designation Fw 190C, and the Jumo 213, Fw 190D. The first Fw 190C prototype, V13, had a DB 603 installed, with an annular radiator at the nose of the aircraft and its supercharger intake mounted between its two oil coolers, these modifications presenting a longer, but more streamlined profile. Little drag was added to the airframe with the modifications initially, but they would be forced to mount the supercharger scoop externally. The aircraft first flew in March of 1942, and overheating, along with general teething issues would be noted. Two more prototypes were converted, V15 and 16, receiving the longer wing from the B-project and GM-1 equipment. Turbocharging was also proposed, but not pursued until much later on. The program continued through May at a decent pace and they were achieving high speeds, one aircraft reaching 696 km/h at 6,950 m, but overheating and engine failure remained serious issues. Similar problems were likewise being experienced with the Jumo 213. The results, however, prompted Focke-Wulf to expand the program with six more prototypes, V13,15,16, 19, 20, 21, 25, 26, and 27 carrying the DB 603, and V22 and 23 using the Jumo 213. Despite the focus on the DB 603, the company was prepared to switch to the Jumo 213, which they could obtain a much larger supply of.

The large, drag inducing turbocharger scoop earned this aircraft the moniker ‘kangaroo’. Aerodynamically unsound, it proved unsatisfactory for service. (wwiiforum)

The final design for the Fw 190C featured the DB 603A with its supercharger intake mounted on the port engine cowling, with various provisions for an armament of MG 131 machineguns, MG 151/20, and MK 108 autocannons. Its highest tested speed was an impressive 722 km/h at 9 km, without armament or armor plates. Production was strongly considered, and then canceled. The DB 603, in its fighter configuration, was still proving troublesome, and V13 was written off after an engine failure forced the pilot to crash land. The engine itself had a comparatively small production run compared to the Jumo 213, and was being shared with a number of twin engine bombers and night fighters. As the older, and massive Jumo 211 production lines were transitioning to the more powerful Jumo 213, it was by far the better choice for a new mass production fighter.

The Fw 190D or ‘Dora’ project continued, though its development path did not lead to a mass produced, high altitude fighter. Rather, it became a project to facilitate getting the Jumo 213 into a fighter as fast as possible, as it was one of the few German engines capable of competing with Western Allied models in most areas. The only mass produced variant, the D-9, is often mislabeled as a high altitude fighter, though its engine was designed for low to medium altitude use. A small number of high altitude models, with the appropriate engines, were produced, but were nothing compared to the D-9’s production run of well over a thousand aircraft by the end of the war.

Shifting programs aside, Focke-Wulf would continue with the new Höhenjäger II project, now seeking to build a truly superb high altitude fighter by taking several of the Fw 190C prototypes and equipping them with Hirth TK 11 turbo-superchargers. With the Fw 190B improvements, the 2000 hp DB 603 S, and a pressurized cockpit, it was hoped that a number of exceptional high altitude fighters could be produced, even if they could never reach the production figures of the Dora. They attempted to solve the earlier issue with the unmanageable size of the turbosupercharger by installing it partially outside of the fuselage, with an air scoop at its front. V18 received the necessary modifications and flew in December of 1942, with serious cooling problems being noted. Further modifications were made after the first several flights, most notable being that a larger oil cooler was mounted, the tail was enlarged to improve high altitude control, and the next prototype, V30, was re-equipped with a four bladed Schwarz propellor. Their extreme high altitude performance was superior to the C, with the aircraft reaching a speed of 670 km/h at 11 km, though they were proving far more temperamental. Turbine and engine issues continued to cut test flights short, though more prototypes were constructed through early 1943, V29 to V33. However, turbine issues persisted, and the entire scoop set up was found to be aerodynamically poor, and the design was proving very disappointing in comparison to the fully recessed models in service with the USAAF. Its performance too was deemed inadequate, and the project was canceled.

Falling behind

Apart from expedient designs, like the GM-1 boosted Bf 109’s, German efforts to produce a high altitude fighter had largely stagnated during 1943, and by the beginning of 1944, they were at a distinct disadvantage. For the past two years, most of the aero engine industry was working hard to modify their existing models to run at their full power using the inferior materials and fuel that were available to them. Among the clearest problems this caused was with the Messerschmitt Bf 109 G, or ‘Gustav’ model, which was only finally cleared to run at its full combat power in the summer of 1943, almost two years after its introduction. Under such conditions, developing new engines was a mostly hopeless effort, and to make matters worse, Allied developments in this field were unfolding brilliantly. While Focke-Wulf and Messerschmitt had failed to deliver on their high altitude fighters, the RAF began to fully transition to the use of the two-stage Merlin in their Spitfires, while the even more powerful Griffin was in development. By the end of 1943, USAAF finally introduced the P-51B, using a licensed Packard Merlin engine, and the P-47 had seen significant performance improvements which gave it unparalleled performance above 9 km. The P-51 proved perhaps the most concerning, as it not only had the benefit of a significantly more advanced engine, but it had been designed with aerodynamic concepts that were not available to aircraft designers before the war. It was an altogether modern aircraft, whereas the German air force would remain dependent on modified versions of planes which had been flying before the war had begun. The Bf 109G had fallen behind its Western contemporaries in most areas of performance, while the Fw 190 still clung to a competitive edge in low to medium altitude engagements. At high altitudes, especially above 7.5 km, there were only a comparative handful of GM-1 boosted Bf 109G’s that could really challenge the Allies, and even then, not on equal terms.

Germany did not possess the materials needed for robust and reliable exhaust valves, bearings, or more efficient, high pressure, high temperature radiators like those on Western Allied planes. However, there were areas of hopeful improvement. Foremost was that, by the autumn of 1943, German engine manufacturers had developed nickel coatings for engine pistons to overcome corrosion problems, and had modified the DB 605’s oil scavenge system to allow it to run at its originally planned combat power. While they would not be able to produce engines as reliable as those in the service of the RAF and USAAF, it was clear that the performance disparity could be reduced. Just as crucially, improvements were being made in regards to radiator design, particularly the annular units which were being tested on the high altitude Focke-Wulf projects. The new AJA 180 on the Fw 190 series was both approaching the pressure and temperature tolerance of Allied models, and was very compact, allowing the Fw 190 to retain its aerodynamic sleekness even when it switched engines.

The Fw 190D project transitioned away from high altitude fighters when the demand for better general purpose fighters grew. Rushed into production, it had more than its fair share of blemishes but, nonetheless, was an effective successor to the earlier ‘Anton’. (worldwarphotos)

While Messerschmitt had already succeeded in producing an acceptable high altitude fighter in the GM-1 boosted Bf 109G, Focke-Wulf’s projects took a different turn. The high altitude Fw 190D project shifted focus to produce a medium altitude fighter, the Fw 190D-9, and another project would seek to build a successor to the Fw 190, the new plane being named Ta 153. The designation changed to reflect Kurt Tank’s role as the head designer at Focke-Wulf. With this new design, hopes for significant high altitude improvements were again stoked, but as had become clear by their earlier failures, such improvements could not come from any unfamiliar solutions or technically complex methods, like turbocharging.

The Successor

The Ta 153 was so designated as it was not a variant, but a successor to the original aircraft. It featured a new fuselage and wings and the occasionally troublesome electrically driven landing gear actuators were changed for hydraulically driven ones. Being almost entirely divorced from the Fw 190’s supply chain, it was thus denied for production in March of 1943, given the amount of labor and time it would take to set up tooling. A compromise model between the design and the Fw 190D was selected, designated the Ta 152.

The Ta 152C program attempted to replace the Fw 190 A&D and correct the faults of the Dora. In the end, it proved too similar to justify shifting production before the end of the war. (ta-152.de)

There were several types planned, namely Ta 152 A,B,C and H. These were standard fighters, heavy fighters for use against bombers, fighter bombers, and a high altitude interceptor. The A and B were designed to use the Jumo 213A & E, respectively, the C the DB 603, and the H, the Jumo 213E. To avoid impacting the production of the Fw 190D, the high altitude model was the first to be developed. These planes featured a hydraulic landing gear system as opposed to the electric actuators on the Fw 190, an improved vertical stabilizer from the Fw 190C program, larger wings, and a half meter fuselage extension in the rear fuselage, with the ensuing redistribution of weight helping to correct for an issue with the aircraft’s center of gravity.

While it may seem odd that they were essentially pursuing two fighter designs to succeed the Fw 190A, the Luftwaffe was desperately looking for higher performance fighters. Hopes were placed on the new Jumo 213 in the Fw 190D, and the new DB 605D in the Bf 109K, to keep pace with the Allies. The Dora was an expedient solution which could use the same supply chain as the original fighter, and the Ta 152 would be a more thoroughly improved model which would be transitioned to once the Dora’s supply chain was well established. In any case, only the high altitude Ta 152 variant was pursued with any substantial amount of resources, given it would be assigned a mission the Bf 109K and Fw 190D models were not suitable for. Jets were, of course, also quite promising, but they were still an immature technology, and it was clear that the leap from pistons to turbines could not be made in 1944.

The new fighter would be designed with both high altitude and low altitude performance in mind. To meet this challenging requirement, both the GM-1 high altitude, and MW 50 low altitude engine boost systems were to be installed aboard the aircraft. Kurt Tank selected several of the old Fw 190C prototypes to be converted for the new program, these being V18, 29, 30, 32, and 33. V33 was the first to undergo modification and was redesignated V33/U1, now featuring a three bladed VS 9 propeller, a forward fuselage lengthening of .5 meters, a rear fuselage lengthening of 0.772 m, a new high aspect wing with an area of 23.5m^2, a hydraulically actuated undercarriage, and two 20 mm MG 151/20’s mounted in the wing roots.

It first flew on July 13, 1944, and was lost after it crashed during its 36 minute test flight at Vechta. The second prototype, V30/U1, flew on August 6, and like the first, was again lost, though this time resulting in the death of its pilot, Alfted Thomas. More success was had with the third prototype, V29/U1, which flew on September 29, 1944, and the fourth, V18/U2, which flew shortly after. With pre-production beginning in November, this left them about a month to perform flight tests on their surviving prototypes.  Serious trouble with the program was encountered as late as November, when test pilot Hans Sander had to crash land his aircraft after his engine seized due to fuel starvation. It was found a hydraulic valve had been installed in the fuel line, an accident most likely a result of the aircraft’s rushed development.

The losses and damages experienced at this point in testing were threatening to seriously interrupt the pace of the project, but in the end, they rushed through development with some of the stability issues unresolved. This effectively led to the aircraft entering production with only slight adjustments from the prototypes. However, the plane was achieving good high altitude performance, both in terms of speed and ceiling. Test pilot Friedrich Schnier would fly V29/U1 to an incredible height of 13.6 km on January 20th, 1945. Beyond this, the fourth and final converted aircraft was V32/U1, which was fitted with a four bladed Schwarz propeller and the new MG 213 revolver cannon. It first flew in January of 1945, though none of the equipment would be worked into any production aircraft.

The eccentric profile of the aircraft, with its high aspect ratio wings. (destination’s journey)

The H was unique among the Ta 152 series, with its long, high aspect wings designed for high altitude use, a pressurized cockpit, and the installation of both the GM-1 high altitude, and MW 50 low altitude boost systems. While together, they promised incredible performance at any height, GM-1 was never carried aboard any of the operational fighters due to its container’s adverse effects on stability. Eager to have this aircraft as soon as possible, Focke-Wulf sprinted through its development, and the Ta 152H entered pre-production in November of 1944. The extremely rapid pace of development was emblematic of the very desperate situation the German air force was in at the time. This resulted in the delivery of an aircraft that was effectively unfinished.

The Ta 152H-0 entered service without several of the key features that the plane was set to carry, lacking the outer wing fuel tanks, and the engine boost systems. As such, it was considerably lighter, and better handling than the planned production model, but without the boost systems, it was much slower.  For the time being it judged necessary, as there were serious weight distribution issues with wing fuel tanks and boost systems aboard. While it was designed with wing tanks, GM-1, and MW 50, the production model of the aircraft would not be permitted to fly with all three. In the end only the MW 50 and the wing tanks were permitted to be used together, but the GM-1 system would prove more troublesome. A stop gap solution late in the war would allow for the use of GM-1, but only GM-1. By the time the war ended, there was still no solution on how all three pieces of equipment would be added to the plane without jeopardizing its flying characteristics.

It was in this rough state when it was delivered to the Luftwaffe for testing in December. Due to supply chain issues, production was slow and the aircraft were finally delivered to the Luftwaffe until January 27, 1945.

Operational History

Given their very late introduction during the war, the Ta 152H saw very little action and its combat record is extremely limited. The aircraft was only supplied to the Stab, the squadron staff group, and Gruppe III of JG 301, a dual night and day fighter squadron which transitioned to them from Fw 190A-8’s on January 27. The squadron had a good pool of experienced pilots already familiar with Focke-Wulf aircraft, though their mechanics would have a far more difficult task, as the Ta 152H-0 had been pushed into service without maintenance manuals. At the airfield at Alteno, they received 11 aircraft, with 16 others having been destroyed or damaged on the ground before they could reach the unit. Familiarization and training proceeded until the end of February and was not without incident. One aircraft (150037) was lost in a training incident, a second damaged but repaired, and serviceability fell from 75% to 30% after an incident with water contaminating fuel supplies. The squadron would go on to receive several more aircraft before rebasing to Sachau when Alteno was overrun. They would attempt to engage Allied bombers on March 2, but the 12 Ta 152H’s would fail to reach them, as they were attacked by the Bf 109s of another squadron which mistook their unfamiliar planes for the enemy. No aircraft were lost in the engagement. A second high altitude interception against a DeHavilland Mosquito was also attempted, though engine trouble forced the pilot to return to base before contact was made.

One of the only photographs of Ta 152H’s in operation with JG 301. (destination’s journey)

The unit rebased again to Stendal near Erfurt, where they joined JG 301’s Gruppe II, during which one aircraft was lost, and the pilot, Jonny Wiegeshoff, was killed on the landing approach. This was believed to be the result of the propeller reduction gear failing and becoming stuck in an almost feathered position. By March 14th, the understrength unit was supplied with several Fw 190A-9s. Outnumbered and with little security, the Ta 152H’s often flew top cover for the rest of the unit during what few operations were undertaken. On April 10, Erfurt was contested, and during the fighting, the eight serviceable Ta 152’s engaged a flight of fifteen P-47’s near Brunswick, resulting in one victory claim.

Gruppe III’s last actions were conducted from Neustadt-Glewe. On April 15th, the unit suffered its first combat loss. During operations that day, four Ta 152s sortied to attack a pair of RAF Hawker Tempests engaging in a low level sweep. According to Obfw. Willi Reschke, the Ta 152H in the number two position, flown by Obfw. Sepp Sattler, suddenly lost control and crashed before contact was made, seemingly suffering a fatal malfunction, while other accounts claim he was brought down by one of the RAF Tempests. The remaining two Ta 152’s engaged the Tempests of No. 486 Squadron. In the ensuing battle, Obfw. Willi Reschke entered an intense, low level dogfight with one of the Tempests. Near the beginning of the engagement, he fired on and struck the tail of a Hawker Tempest flown by Lt. Mitchell, his gun’s electrical circuit seemingly failing shortly after. However, when Mitchel attempted to turn away from his opponent, he lost control of his damaged aircraft and crashed. Reschke swore by the low speed maneuverability of the Ta 152, which he felt was critical in this engagement, and his survival through the last days of the war. The Ta 152H flown by the Schwarm leader, Oberstleuteneant Fritz Auffhammer, suffered an engine failure, though the pilot successfully restored power and returned to base with his supercharger broken. Sattler and Mitchel were both buried at a cemetery in Neustadt-Glewe.

The last actions of the squadron were in the last stages of the Battle for Berlin, and on April 24th, the Ta 152s and Fw 190As of the IInd and IIIrd Gruppe attacked Soviet positions and engaged Yak 9’s. The final mission was flown over Berlin in poor conditions, and during an engagement with a flight of four Yak 9’s, Hauptman Hermann Stahl was killed during the engagement, with the four Yak-9’s being claimed by the unit. After the surrender, the unit rebased to Leck in Schleswig-Holstein, where they were disbanded and one of the serviceable Ta 152H’s was transferred to England by the RAF so that it could be evaluated. A second Ta 152H was also claimed by the USAAF for evaluation purposes, the plane being another H-0 which likely belonged to a testing unit at Rechlin.

One of the only serviceable planes was taken to the UK for testing. The other was taken in hand by the US. (Alessandro Orseniga)

In all, the Ta 152H was never actually used for any high altitude combat operations and its service was restricted to a single under strength unit. With at most ten victories and four operational losses, it is difficult to give any appraisal for its performance from its brief career with JG 301. Obfw. Josel Keil, was the only pilot to qualify as an ace on the Ta 152H, and together with Willi Reschke, who had two credits in the Ta 152H, and 24 in other aircraft, held nearly all of the aircraft’s combat credits between them.

Handling and Flying Characteristics

While the Ta 152H’s combat record leaves a lot of questions left unanswered, most pilots who had the chance to get behind the controls of the aircraft can at least agree that the aircraft flew very well. Among its most famous advocates was Royal Navy Test pilot Eric Brown. He would praise its excellent climb performance, maneuverability at high altitude, stability, and good landing characteristics. His only negative remarks were that its roll rate was reduced over the older Fw 190A, that its stick forces were notably heavier, and that its wheel brakes were still awful and prone to fade after a few moments of use. He otherwise considered it an excellent aircraft and the best high altitude piston engined fighter he had flown, comparing it favorably to a Spitfire Mk IXX. It must be noted that he misidentifies the aircraft as an H-1 in his book, and not the substantially lighter H-0, which is visually identical.

Captain Brown’s remarks are matched by those of the pilots who assessed the aircraft in the Stab and III/JG 301. The Ta 152H-0 had the best evaluation received by a front line operator of a Focke-Wulf aircraft. The aircraft possessed most of the best qualities of the earlier Fw 1,0D-9 without having its poor accelerated stall characteristics. While still described as uncomfortable like the Fw 190D it was so similar to, it was much improved and less prone to the aggressive snap rolling. So, while the aircraft was less maneuverable, generally speaking, most pilots were more comfortable pulling harder turns. In tests at the unit, some new pilots in Ta 152H’s were able to turn with seasoned pilots in Fw 190A’s. Take off runs were short, and the landing approach could be conducted at low speeds. Generally speaking, it was a fairly forgiving aircraft. The only negative notes on the aircraft were from the findings of the Rechlin Test Center, which found the aircraft became seriously unstable in dives exceeding 600 km/h and that level flight required excessive trimming of the horizontal stabilizer.

The enlarged tail and redesigned wings helped give these aircraft better handling characteristics than the Fw 190D. (Grafiq)

The stick forces were notably fairly high, but they were harmonized well, and the push rod control system ensured inputs were very responsive. Stability about the vertical axis was poor, and there was a tendency to skid. This tendency grew worse at higher altitudes and motivated them to install a level flight autopilot. The aircraft possessed good visibility to the back, sides, and rear, with the view over the nose being mediocre to poor. The controls were placed conveniently, with the instrument panel layout being clean and easy to read.

Most of its good qualities were not found in the fully equipped H-1 production model of the aircraft. Numerous problems were encountered when the full set of engine boosting equipment and fuel tanks were installed and filled. The added weight of the boost systems and wing tanks was substantial, and asymmetric. The GM-1 system and the wing tanks were particularly problematic, and the aircraft was unstable if the GM-1 container and fuel tanks were filled. Stability with the GM-1 system was only possible with a ballast kit, empty wing tanks, the removal of the MW 50 system, and a set fuel limit for the rear fuselage fuel tank. These issues were not resolved by the time the war ended, and there was no way the aircraft could use any combination of these systems without seriously jeopardizing its flying characteristics. MW 50 was usable aboard only the H-1 production model, but it may not have been available to JG 301 in the field. The squadron was still mostly composed of BMW 801 equipped Fw 190A’s which did not use the system.

Mechanics generally found the aircraft easier to maintain than the Fw 190, however there were some issues. The new hydraulic system for the landing gear was experiencing teething and quality control issues. The position of the landing gear wheel well was also found to be at issue, as when launching from damp conditions, the propeller cast mud and water into the well, which made its way inside the wing. This caused issues with the hydraulic systems and the autocannons fitted in the wing root.

Comparisons with contemporary fighters

The Ta 152H represented a leap in Germany high altitude fighter design, though not necessarily one that took them beyond the competition. (flugrevue)
Aircraft (manifold pressure) Speed at Sea Level (km/h) Speed 3050 m (10,000 ft) (km/h) Speed 6096 m (20,000 ft) (km/h) Speed 9144 m (30,000 ft) (km/h) Speed 9.5 km (31,168 ft) (km/h)
Ta 152H-1 (1.92 ata) 580 640 690 725 732
Fw 190D-9 (1.82 ata) 611 645 689 653 645
P-51B-15 (75″ Hg) 616 675 709 688 685
P-47N-5-RE (72″ Hg) 587 643 708 740 759
P-47M (72” Hg) 587 646 701 753 762
P-38L (60” Hg) 550 608 646 663 659
Spitfire Mk 21 (+21 lbs) 592 658 700 704 703
Me 262 A-1a 800 x 870 845 x

*The Ta 152H-1 could reach a maximum speed of 760 km/h at 12.5 km using the GM-1 boost system. While it was never cleared for operational use, on paper, it made the Ta 152H the fastest fighter at that altitude. The Fw 190D-9 represents a late model, having received an MW 50 boost system, as was available near the end of 1944.

The Ta 152 entered service on a battlefield where the Western Allies already had high altitude supremacy, and had a number of improved designs that had yet to make their debuts by the time the war in Europe was ending. By January of 1945, the German air force was no longer dealing just with long range escort fighters over its own soil, but virtually every fighter the Allies could throw at it, such as P-47’s, Spitfires of several marks, La-7’s, and Tempests, just to name a few.

Fw 190D-9 (Graphiq)

Against its contemporary Fw 190D-9 counterpart, it is clear that the Ta 152H did not represent a comprehensive upgrade. The Dora shared much of the same fuselage, though it retained the wings and tail sections of the older Anton series fighter, and it carried the Jumo 213A engine designed for use at lower altitudes. In regards to linear speed and acceleration below 6 km, the Dora roughly matched or exceeded the Ta 152H. This, however, was not the case at higher altitudes, where the high altitude specializations of the fighter showed their worth. The Ta 152H was known to be more maneuverable in flat turns and much more forgiving in most aggressive maneuvers, a result of its high aspect ratio wings which lacked the less than ideal tendency for snap rolling without much warning that the older Fw 190’s were known for. In a dive, the Dora was notably superior, as the aforementioned wings of the Ta 152H made it notably unstable at high speed. The H-1 carried, but was not cleared to use GM-1, nor does it seem they would have ever been supplied with the mixture. This is a discrepancy of several hundred kilograms, leaving the true climb performance of the aircraft somewhat ambiguous, with a claimed 20 m/s at sea level without MW 50.

The P-51B’s and D’s had marginal differences in performance They were among the most aerodynamically clean fighters of the war, boasting an extremely streamlined fuselage, laminar flow wings, and a radiator scoop which produced thrust that offset upwards of 90% of its own drag. To increase maneuverability in high speeds and in power dives, the control surfaces were internally sealed and used a diaphragm to reduce stick forces. The engine was a Packard Merlin V-1650-7 with an intercooled, two stage, two speed supercharger. Even though the engine was actually geared for lower altitude use than its predecessor, the combination of these features made the aircraft a very fast, maneuverable fighter which could boast of high performance at most altitude ranges.

Against the Ta 152H-1, the Mustang held to a higher top speed at low to medium altitude, better maneuverability at high speed, and far better dive performance. At extreme altitudes, the H-1 outstripped the Mustang in top speed, and across most altitudes would have had better low speed maneuverability. The high aspect ratio wings of the Ta 152 both gave it better handling at high altitude, and much improved stall characteristics over its predecessors down low. Curiously enough, both the Ta 152H and the Mustang were far more maneuverable than their wing loading would suggest, a result of high aspect ratio and laminar flow wing designs, respectively. However, in the Ta 152’s case, this came at the cost of a slower roll rate, and unstable high speed dive characteristics. While the Ta 152H could prove an exceptionally challenging high altitude opponent to all of the contemporary Allied fighters, it was a competitive, but not particularly impressive aircraft at lower altitudes. Performance wise, it could be said to fly like a more maneuverable, if slower, Fw 190D when at lower altitudes.

There is of course the story of Kurt Tank himself escaping a pair of P-51’s at low altitude in a Ta 152 prototype. Near the end of 1944, the designer himself was flying one of the prototypes to a conference in Cottbus, Germany, where he was happened upon by two P-51’s. Using the MW 50 boost system in the aircraft, Tank slipped away from his pursuers and arrived in Cottbus unscathed. Some laud this encounter a sign of the aircraft’s superiority, however, it is not a useful measure of the performance of any of the combat models of the aircraft. At Kurt Tank’s instruction, the prototype in question was unarmed and, more than likely, carrying no armor plate, which would have made the aircraft substantially lighter than any operational Ta 152H fighter.

The Spitfire Mk 21 represented the final evolution of the wartime Spitfire, by then nearing its tenth year in the air. A far echo from the Mk I, the 21 featured a vastly more powerful Griffin 61 engine. Much like its late Merlin powered predecessors, it possessed an intercooled, two stage, two speed supercharger. Unlike them, it was massive and much more powerful. After incorporating structural improvements and modifying controls for high speed, the Spitfire aged perhaps the best of any fighter of the war. Compared to the Ta 152H, it lacked the sheer distance in top speed performance of the P-51, but more than challenged the Focke-Wulf in linear speed and climb rate across most altitudes. However, at and above 7 km, the 152H had a confident advantage in speed and maneuverability.

The P-47N represented the final evolution of the Thunderbolt. Though not destined for Europe, it performed similarly to the P-47M which debuted in combat roughly the same time as the Ta 152H. (wikimedia)

Compared to the most modern Allied high altitude fighters, the Ta 152H lost most of its edge. The P-47N and M represented the final evolution of the American high altitude fighter, featuring a new 2800 hp, R-2800 turbocharged engine, and a variety of aerodynamic improvements to increase control at high speed. By the late Summer of 1944, the Western Allies had already gained air superiority over Europe, and so the new aircraft was stockpiled in the US for use in the Pacific, with the first deliveries being made in September of 1944. There was a similar performing model in Europe, the P-47M, though it was a limited production aircraft designed for chasing V-1 flying bombs and other high speed targets. Teething issues would keep it from entering service roughly until the Ta-125H did, in March of 1945. In the end though, the Luftwaffe had become so degraded that clearly no new updated models would be required and the performance increases would not justify the effort to refamiliarize pilots and maintenance personnel.

In terms of top speed, the P-47M&N handily outperformed the Ta 152H at all altitudes, the only exception being at extremely high altitudes when the Ta 152H employed GM-1. In contrast, the Focke-Wulf enjoyed a better climb rate and was likely the more maneuverable of the two, although it was certainly less capable in a dive. The late war Thunderbolts were certainly the fastest high altitude fighter which saw combat, the Ta 152H’s of JG 301 never having carried GM-1.

The P-38L was the last fighter variant of the Lightning fighter, the first model having been in service prior to the US entry to the war. With its turbo supercharged Allison engines, it was among the first fighters of the war that was designed for high altitude use. However, by the end of the war, it left something to be desired in terms of both its top speed, and like the Ta-152H, its high speed dive performance. Its low critical Mach number meant that the plane encountered compressibility at lower speeds than all of the fighters presented here. At high speeds and altitudes, the plane locked up and would remain uncontrollable until its high speed breaks were deployed, or it had descended into lower, denser air. Of all the Allied high altitude fighters, the Lightning compared fairly unfavorably with the Focke Wulf.

Most easily glossed over is the performance compared to jet fighters, which by the time the Ta 152H was introduced, could not exactly be called new. The Messerschmitt 262 had re-entered service in November of 1944 after earlier operational problems, and once training and maintenance programs were revised, the plane quickly proved itself. While it was slow to accelerate and climb, it was unapproachable in terms of top speed. Extreme high altitude use of the temperamental Jumo 004 turbojet engine was limited, though as a means of attacking high altitude formations of Allied bombers, it was by far the best equipped aircraft Germany possessed. Its slow acceleration meant that any energy-demanding maneuvers were largely off the table, but when flown by a pilot that understood its strengths, the plane was untouchable save for when it was taking off or landing. Though largely an issue post war, the Me 262 demonstrated the difficulty in justifying further piston engine fighter development at this point in aircraft development.

The Ta 152H began production after the Me 262 jet fighter had already entered service, so it, and many other programs, had to compete with it for resources. As the turbojet was already showing to be the future of fighter design, the Ta 152 was difficult to sell to the Luftwaffe. (I.PINIMG)

Overall, the Ta 152H certainly was not a Wunderwaffe by any means. At all but the highest altitudes, the aircraft was not a particularly better performer than its preceding, and much more numerous, Fw 190D counterpart. Even at extreme altitudes, it more than had competition in the form of the Thunderbolt N and M, which not only outstripped it in performance in a number of areas, but beat it into production by several months. It’s only truly exceptional performance was achieved using a high altitude engine boost system that was never made available to the unit carrying the aircraft, and in any case, it would have required a redesign of the aircraft to be used properly. Nevertheless, it represented a stark improvement in high altitude performance over previous German fighters. It too, could boast of extreme maneuverability at high altitudes, even if it didn’t lead the pack in pure speed. Top speed aside, its wings lent it a great degree of maneuverability at high altitude, and its overall performance at and above the altitudes Allied bombers flew at was considerable. This is also to say nothing of its trio of cannons; two 20mm MG151/20’s and its single 30mm MK108, which leant it incredible striking power. While the incorporation of the Jumo 213E, MW 50, and on paper, GM-1, did not produce the pinnacle of fighter design, the result was still a capable high altitude interceptor capable of engaging the highest flying targets of its day.

Construction

The construction of the Ta 152H’s fuselage was essentially that of a modified Fw 190A-8. The fuselage was largely the same with the following modifications: the forward fuselage was lengthened by 0.772 m in order to fit in a Mk 108 autocannon, the wing connecting section was moved forward 0.420 m to correct for the center of gravity, and the rear fuselage was lengthened by 0.5 m. The leading edge of the tail was exchanged for that on the Fw 190C, being considerably larger. Given the deteriorating situation near the end of the war, the new tail surfaces were wood, rather than metal skinned. The fin and rudder were enlarged for better control, with the new surface area of the tail stabilizers measuring 1.77 m2 for the vertical and 2.82 m2 for the horizontal. The changes to the fuselage necessitated strengthening, which saw some duralumin framing elements replaced with steel. In order to reduce the number of assembly jigs they needed to produce, the forward fuselage extension was bolted through the former engine attachment points.

The Ta 152H-1 featured all the fuel tanks pictured here, the preproduction H-0 had only those in the fuselage. (Deutchesluftwaffe.de)

The Ta 152H-1 featured all the tanks pictured here, the preproduction H-0 had only those in the fuselage. (Deutchesluftwaffe.de)

The wings were entirely redesigned from the Anton and changed to a high aspect model which increased the wingspan to 14.4 m, and to an area of 23.3 m2. Structurally, it remained a monocoque structure, but its rear spar and leading edge were used to absorb transverse forces and it was structurally reinforced with additional stiffening ribs. The landing gear were the same as the Fw 190A-8’s, but they were hydraulically and not electrically operated. They mounted 740 mm by 210 mm wheels to accommodate the increased weight of the aircraft. The inboard section of the wing mounted an MG 151/20 autocannon with provisions for 175 rounds of ammunition each.

Clean instrumentation, a high level of engine automation, and good visibility made the Ta 152H a fairly straightforward aircraft to fly. (destination’s journey)

The aircraft possessed a pressurized canopy to reduce the physiological stresses of high altitude flight. It was a very rudimentary system, with the cockpit rivets being sealed with DHK 8800 paste, and the sliding hood being sealed by means of a cylindrical rubber tube liner. Pressurization was regulated by means of a 1 liter air bottle supplied by a Knorr 300/10 air compressor which was geared to the engine with no intermediate gearing. The system was engaged at 8 km and maintained a constant .36 atmospheres. To prevent windscreen fogging, it was double-paned, with silica packets installed in the gap. Quality control issues saw varying effectiveness at altitude. On the record setting flight, Friedrich Schnier reported the system leaked badly above 12 km and shortly after he suffered joint pain, impaired vision, and numbness in his extremities due to low air pressure.

The Ta 152H carried an armament of two MG 151/20 20 mm cannons in each wing root and a centerline MK 108 30 mm cannon which fired through the propeller hub. The 20 mm guns were supplied with 175 rounds per gun, and the 30 mm with 90. The gunsight was the standard Revi 16b sight, which was eventually supposed to be replaced by the new EZ 42 gyroscopic sight which, when properly used, gave the pilot an accurate gunsight lead against his target. The aircraft was well armored with two engine plates, and six to protect the pilot, with a combined weight of 150 kg. The 8 mm plate behind the pilot was judged inadequate, though plans to increase its thickness to 15 mm were not carried out. A single hardpoint could be attached to the underside of the aircraft to install a 300 liter drop tank, but there were no provisions for carrying bombs.

The engine was a 35 liter Jumo 213E inverted V-12. Originally developed from the Jumo 211, which saw heavy use in bombers much earlier in the war, the new Jumo 213 was what most of the Luftwaffe’s hopes were placed on to compete with newer, more powerful Allied engines. It featured a new AJA 180 streamlined annular radiator that supported the oil and engine coolant. Critically, it was able to operate at significantly higher temperatures and pressures than older models, though not quite at the standards of the Western Allies. However, unlike Allied models, the Jumo was heavily automated. The Bediengerat, or control device, was a hydro-mechanical computer that managed the propeller RPM, mixture, supercharger speed, and radiator based on the pilot’s throttle inputs. This helped to relieve the pilot’s workload, as the Kommandogerat did on the BMW 801 powered models.

The Jumo 213E was the high altitude model which featured an intercooled, two stage, three speed supercharger. To further improve on high altitude performance, the aircraft would use a GM-1 nitrous boosting system. The system consisted of an 85 liter tank behind the pilot, and a crescent shaped liquid nitrous tank that sat at the right front side of the cockpit. The mixture was fed into the supercharger by a pump when the system was activated. As an oxygen carrier, the job of the nitrous is to provide an oxygen rich mixture to the engine when the supercharger is operating at altitudes where it is unable to provide the compression, and thus enough oxygen, needed to maintain a high manifold pressure. For the Jumo 213E, this was above 11 km. The drawbacks of the system were its uselessness below 11 km, and the bleed off of the evaporating liquid nitrous, which prevented it from being efficiently stored aboard the aircraft beyond several hours. Unlike its use on other aircraft, like Bf 109’s and Ju 88’s, the position of the nitrous tank aboard the Ta 152H proved dangerous, as it severely impacted the plane’s stability. It is unlikely the system would have been very effective without a major redesign of the fuel and mixture tanks, as even with a ballast kit that stabilized a GM-1 carrying plane, the aircraft could not carry anywhere near its full fuel load or its MW 50 boost system. While, on paper, the system promised unparalleled performance at extreme altitudes, it was almost unusable given its unstable configuration.

The MW 50 system was the low altitude boost system. It consisted of a 70 liter tank in the port wing containing MW 50, being roughly 49% methanol and 49% water, with the remainder being an anti corrosion measure. When active, the solution was pumped into the supercharger. The system was designed to boost engine power and overcome the less than ideal quality of German aviation fuels. Poor detonation characteristics, especially of the lower octane B4 fuels, forced the Germans to run at lower manifold pressures and thus lower power to avoid damaging their engines. Methanol boosted the octane rating of the fuel-air mixture entering the manifold, and the water cooled the mixture, with both factoring to bring major improvements in engine power via their combined anti-detonation, or knock, effects. The system made its debut in the summer of 1944, and was essential in allowing the later Bf 109G and Fw 190D series aircraft to stay competitive with their Allied counterparts. However, it was not without its drawbacks. It could not be used effectively above around 6 kilometers, and it was highly corrosive, severely limiting the lifespans of corrosion prone German engines. Aboard the Ta 152, it was to be installed in either a 70 liter wing tank or a standard 115 liter tank behind the pilot.

The Jumo 213E Kraftei. The entire assembly was bolted to the front of the fuselage and streamlined engine swaps. (ta152.de)

The engine had a bore and stroke of 150 mm and 165 mm, a compression ratio of 6.5:1, and a dry weight of 1040 kg. It differed from the standard model in that it had a slightly smaller bore, and the larger supercharger assembly and the associated intercooler added some 300 kg. It used B4 fuels which had a minimum octane rating of 87. The engine drove a constant speed 3.6 m VS 9 wooden propeller with a reduction gear of 1:2.40, and produced a maximum of 1753 PS (1729 hp) at sea level and 1260 PS (1242hp) at an altitude of 10.7 km. The oil header tank sat atop the front of the engine, and the coolant tank sat at the rear. On the Jumo 213A, these had a capacity of 55 and 115 liters respectively. The entire engine assembly was a Kraftei, or power-egg, consolidated unit, allowing the engine and its associated coolant systems to be easily removed or added to the aircraft.

Its radio and navigation systems included the FuG 16ZY ground control transceiver to allow it to be tracked and directed from ground based stations, a FuG 25A erstling IFF, and a FuG 125 radio direction finder for beacon homing. Some aircraft were also fitted with a K 23 level autopilot to reduce fatigue when flying the aircraft at high altitudes and in poor weather. The autopilot was accompanied with a heated windscreen and a FuG 125 Hermine radio navigation system as part of the R11 Rüstzustand equipment package.

Production of the Ta 152H

The Ta 152H was introduced in an environment where all quality control measures had already been cut down for every aspect of production. The lack of skilled labor and poor materials meant that building a reliable aircraft engine in Germany had become almost impossible by the spring of 1944. Slave labor and foreign, drafted workers had become the base of the labor pool, as most of Germany’s factory workers had been drafted to fight, resulting in a sharp decrease in quality. This was not only a result of poor working conditions and the inexperience of the workers, but sabotage became widespread, especially among those pulled to work from concentration camps. Even more desperate measures began to be instituted in the summer of 1944, as the re-use of parts from salvaged aircraft became more commonplace, and engine test runs were ever more limited to conserve dwindling fuel supplies.

The first Ta 152H-0 was completed in November of 1944 after considerable delays due to several sets of blueprints being found to be inaccurate, and sets of jigs had been lost in France the previous summer. The first planes were sent to the Rechlin test center in December of 1944, while Focke Wulf considered how to accelerate production. While doing so, they were hobbled when the Jagerstab, which managed strategic fighter production, shifted more and more resources to jet fighters and older, established piston engined fighters. Ta 152H production standards continued to decline in the midst of the widespread economic collapse of Germany. Near the end of January 1945, it became almost impossible to build any more Ta 152H’s, as the decentralized production system began to collapse, the rail system became unusable, and the wing and fuselage production center at Pozen was overrun by the Allies.

By the war’s end, approximately 60 Ta 152H fighters had been completed at the Focke Wulf facility at Cottbus. The series suffered extreme quality control issues in service with JG 301, which included supercharger surging and the failure of a propeller reduction unit, which resulted in the death of a pilot. In April of 1945, the plans were sold and shipped to Japan, where unsurprisingly, there was no new production of the aircraft.

Conclusion

The sole remaining Ta 152H is in storage at the Smithsonian Air and Space Museum, where it awaits restoration. (Smithsonian)

The Ta 152H is often seen as one of the great ‘what if’s’ of the Luftwaffe, but in reality, the aircraft was a good, rather than truly exceptional fighter. While on paper, the Ta 152H was to be an incredible aircraft at high altitude, it’s rushed development, and hasty introduction into service saw it fly without the GM-1 boost system that it needed to achieve these feats, and in a rather regrettable state in terms of build quality. It stacked up well against many of the older aircraft in the theaters it fought in, like the Yak-9, Spitfire Mk IX, or the P-38L, and against its contemporary Allied rivals, it was a competitive fighter at high altitudes.

Specification:

Specification Ta 152H-0 H-1
Engine Junkers Jumo 213E Junkers Jumo 213E
Engine Output 1753 PS, 2050 PS w/ MW50 1753 PS, 2050 PS w/ MW50
Empty Weight 4031 kg
Loaded Weight 4730 kg 5220 kg
Maximum Range 2000 km
Maximum Endurance 3.3 hrs
Maximum Speed [At altitude] approximately 720 km/h [10.9 km] 760 km/h w/GM-1 [12.5 km]
Service Ceiling 15 km w/ GM-1 (estimated)
Armament 1×30 mm MK 108, 2×20 mm MG 151/20 same
Crew 1x pilot same
Length 10.82 m 10.82 m
Wingspan 14.44 m 14.44 m
Wing Area 23.3 m^2 23.3 m^2
Height 3.38 m 3.38 m

 

Variants:

Ta 152H-0: Pre-production model, no wing fuel tanks, no MW 50 provisions, GM-1 capability but never cleared for operational use.

Ta 152H-0/R11: Poor weather pre-production series with level autopilot. Most pre-production aircraft were built in this configuration.

Ta 152H-1: Production model, wing fuel tanks, 85 liter GM-1 provisions but not supplied due to operational concerns. 70 liter MW 50 low pressure system installed. Fuel tankage increased from 595 liters to 995 liters with unprotected bag tanks in wings.

Ta 152H-1/R11: Poor weather model, autopilot. Most production aircraft were built in this configuration.

Ta 152H-1/R21: Equipped with Jumo 213EB intercooled engine, high pressure MW 50 system installed. Not operational.

Ta 152H-1/R31: Jumo 213EB, ballast kit to allow GM-1 use. No MW 50 and fuel capacity restricted. Not operational.

Ta 152H-2: FuG 15 radio set instead of FuG 16. Canceled in December 1944.

Ta 152H-2/R11: Bad Weather model.

Ta 152H-10: Photoreconnaissance model based on H-0.

Ta 152H-11: Photoreconnaissance model based on H-1.

Ta 152H-12: Photoreconnaissance model based on H-2.

Illustrations

The unique paint scheme of this aircraft was an identification measure, as the plane was largely unknown to German Flak and fighter crews. It was flown in this state to a conference.

 

 

Credits

  • Article written by Henry H.
  • Edited by  Henry H. & Stan L.
  • Ported by Henry H.
  • Illustrated by Hansclaw

Sources:

Primary:

Aeroplane and Armament Experimental Establishment Boscombe Down Spitfire F. Mk. 21 LA.187 (Griffon 61) Climb and Level Speed Trials. 10 October 1945.

Einmotorige Jäger: Leistungsdaten, 1.10.44

Ersatzteil-Liste TA 152. Konstruktionsgruppe 7 Triebwerksanlage. Focke-Wulf Flugzeugbau G.M.B.H. Bremen.

Fighter Offensive Performance at Altitude Model P-47N-5RE Engine P&W R-2800-73 GP=45:1 Propeller-4 Blades- 13’0” DIA. (Curtis 836) War Emergency- 2800/2800 S.L. to Critical Altitude G.W.=13962 LBS. Republic Aviation Corporation. Farmingdale L.I., New York.

Horizontalgeschwindigkeit über der Flughöhe mit Sonderleistung. Leistungsvergleich Fw 190 – Ta 152. Focke-Wulfe Flugzeugbau G.M.B.H. 3.1.45

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.

Smith F., M.A. and Brotherton J. Note on the performance in flight of the German jet-propelled aircraft Messerschmitt 262, Heinkel 162, and Arado 234. Royal Aircraft Establishment, Farnborough. October 1945.

Secondary:

Brown, Eric Melrose. Wings of the Luftwaffe. Hikoki, 2010.

Douglas, Calum E. Secret Horsepower Race: Second World War Fighter Aircraft Engine Development on the Western Front. TEMPEST, 2020.

Green, William. The Warplanes of the Third Reich. Doubleday & Company. 1970.

Harmann, Dietmar. Focke-Wulf Ta 152 the Story of the Luftwaffe’s Late-war, High-Altitude Fighter. Schiffer Military History. 1999.

Smith, J. & Creek, Eddie. Focke-Wulf Fw 190, Vol. 3: 1944-1945. Specialty Pr Pub & Wholesalers. 2015.

Smith, J. & Creek, Eddie. Me 262 Volume Two. Crecy Publishing. 2007.

Weal, John. Focke-Wulf Fw 190 Aces of the Western Front. Osprey Publishing. 1996.

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.