Category Archives: US

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 more 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 Stratofortresses 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.

Lockheed S-3 Viking

United States of America (1975)

Anti-Submarine Warfare Aircraft; 188 built, 160 upgraded to S-3B

An S-3 Viking comes in to land on the aircraft carrier USS Independence. [National Archives]
The Lockheed S-3 Viking was an anti-submarine warfare aircraft designed to replace the aging S-2 Tracker, later becoming one of the most important components of the US Navy’s anti-submarine strategy during the late Cold War. Designed in anticipation of modern Soviet Nuclear submarines, the Viking could boast of a host of cutting edge sensors and computerization that put it well above the curve, and all wrapped up in an airframe that was reliable and versatile. Its exceptional anti-submarine capabilities were augmented even further during its mid-life improvements which lead to the introduction of the improved S-3B. After the Cold War, the aircraft transitioned away from its traditional anti-submarine duties to surface surveillance, signals intelligence, and aerial tanker duties. A thoroughly reliable and advanced aircraft, the Viking easily ranked among the most important and versatile aircraft to ever serve aboard US carriers.

The Modern Submarine

The submarine of the Second World War was little more than a long range torpedo boat with the ability to submerge itself for short periods of time to avoid detection. Its offensive capabilities were rather modest, and apart from some outlying, but considerable, success against warships, it was typically seen as a tool for disrupting overseas shipping. Their comparatively low speed coupled with the need to transit on the surface for long periods, which snorkels could not entirely eliminate, would see them become a supporting vessel of most navies. However, advancements near the end of the war would transform the submarine from a raider and reconnaissance vessel, to one of naval warfare’s principal combatants.

Owing to the extreme desperation of the German U-boat force, a submarine built along new, revolutionary lines was designed. As the surface proved an exceptionally dangerous place to be, due to long range Allied patrol aircraft, the new boat would be designed to operate almost entirely submerged for the duration of its patrols. The new Type XXI was designed around the most modern features of any submarine thus built, featuring a much improved pressure hull construction, partially-automatic torpedo loading, a powerful sonar array, and a massive battery capacity which, combined with a hydrodynamically clean hull, allowed it to travel at double the speed of a conventional Type VIIC with over three times the range.

A Type XXI submarine ready to be assembled from prefabricated sections. Massive quality control problems prevented any hope of the submarine’s use in the Second World War, though this construction process was improved post-war world wide. [national archives]
The Type XXI only completed a single wartime patrol, but its effects on naval engineering and submarine design were dramatic. In effect, every submarine built before it was obsolete, effectively restarting a new naval arms race. In the context of the then brewing Cold War, this was the cause of no shortage of anxiety for Western Navies. While the Soviet Union’s shipbuilding capabilities were relatively meager, and greatly damaged during the war, their experience with the new German submarine could very well allow them to leap up to the position of the world’s most prominent navies, if only in the field of submarine design.

In addition to the new submarine’s capabilities, the Type XXI also demonstrated that submarines could also be built at an unheard of rate thanks to its modular construction. Submarine sections could be constructed at secondary factories before being shipped to main construction yards, where they would be assembled into completed boats. Initially, an intelligence survey estimated that the Soviet Union could have as many as 2000 modern diesel-electric submarines in 1960. However, a much more reasonable secondary survey noted that they were likely restricted to 400 boats, owing to available dockyard space, fuel, and bottlenecks in battery maintenance and production. Regardless, the US Navy began work on a modernized anti-submarine strategy to counter a potential flood of Soviet boats which could threaten intercontinental supply lines in a potential war.

The first Whiskey class submarines were only marginal improvements on their WWII era predecessors. Late models, pictured here, had snorkels and performance somewhat below the German Type XXI, but with hundreds made in a relatively short time, their numbers helped offset these deficiencies. [US Navy]
The first of the new Soviet boats was the Project 613 ‘Whiskey’, a somewhat shrunken derivative of the German Type XXI. It had more modest performance than the German boat in regards to speed, range, and endurance, but once it received a snorkel on later models, it had the same ability to remain underwater for long periods. The Whiskey was thus the most advanced submarine the Soviet Union had yet built. In countering these submarines, the US Navy would employ a modified version of the same strategy it had used in the Second World War. The primary anti-submarine weapon was to remain the airplane, in the form of long range patrol aircraft, like the P-2 Neptune, and carrier based planes, like the new models of TBM-3 Avenger. Their primary means of locating submarines were radar, which could detect snorkeling submarines, magnetic anomaly detectors, which were set off by a submarine’s magnetic signature, and sonobuoys, which determine the position of a transiting submarine if dropped close enough. Radar was the main means of detecting a submarine at range, with the other two systems being used to ‘fix’ its location before attacking with torpedoes and depth charges.

Unlike their land based counterparts, early carrier based ASW aircraft lacked the ability to carry both the sensors and weapons needed for the task and were thus placed in a pair of cooperating aircraft. The first such pair were the TBM-3W ‘warning’, for detection, and TBM-3S ‘strike’, for carrying out attacks on marked submarines. These hunter-killer teams operated aboard modified escort carriers and later switched to fleet carriers, when it became clear the small escort carriers could not reliably launch and recover the larger hunter-killers. In the early 50s, it was recognized that the entire system was extremely clumsy and would not provide adequate anti-submarine support.

The Hunter-Seeker ASW method proved far too unwieldy for further use. This ‘hunter’ Grumman Guardian has a search radar on one wing and a high powered searchlight under the other. Its torpedo was stored internally. These were the largest single engine piston aircraft in service at the time of their introduction. (US navy)

The CVS program was thus introduced, which brought several mothballed WWII era-fleet carriers back into service as dedicated anti-submarine warfare ships. The CVS’s, which were introduced in 1952, were soon joined by the S-2 Tracker two years after. The Tracker was large enough to carry both the sensors and the weapons, and the clumsy hunter-killers were finally dispensed with. The S-2 was an excellent ASW aircraft which would go on to serve in a number of roles, though by the mid 60s, the growing capabilities of Soviet submarines and operational troubles with operating a piston engined aircraft on increasingly jet dominated carriers began to highlight the need for a replacement.

The Soviet Nuclear Submarine

Through the 50s and early sixties, the existing strategies for sub hunting were predicated on the need for submarines to recharge their batteries, and that said batteries could be discharged during a drawn out search, thus rendering the submarine helpless. Advancements in Soviet nuclear engineering would end up negating most of these existing strategies. General Secretary Iosif Stalin would formally sign off on the program to build the first Soviet nuclear submarine in 1952. The boat was to be a delivery platform for a gigantic nuclear torpedo for use against harbors. It was completely impractical, and due to the extreme secrecy surrounding it, was rejected by Soviet Admiral Kuzntetsov upon learning of it. The Project 672 Kit (NATO reporting sign November) was then given a conventional torpedo armament and went out to sea in 1958. It was a fast boat, with a given maximum speed of 28 kts, but its turbines proved unreliable and its reactor developed leaks after 800 hours. Less concerning was its noisiness, a factor Soviet submarine designers felt was less important than top speed, and a design choice that would plague Soviet nuclear submarines into the 1970s.

The nuclear submarine was a far more capable and deadly opponent compared to its diesel electric counterparts. Without needing to rely on electric power for underwater propulsion, a nuclear submarine was not restricted to a small patrol area, nor did it need to expose itself to detection to recharge. Furthermore, it was fast. As loud as the Novembers were, they were nearly twice as fast as contemporary diesel electric submarines. Lastly, their larger size enabled them to carry larger, more sensitive sonar systems and greater complements of weapons. In short, it was a faster, more alert, and better armed threat than anything the US Navy ever had to contend with.

The Novembers proved to be a wake up call to the US Navy, but their operational restrictions kept them from being perceived as a massive threat. For instance, they were not deployed to the Caribbean during the Cuban missile crisis, as the distance was deemed a hazard. The turbines aboard these boats were unreliable, and there was no wish to have their most advanced submarine being seen under tow. Subsequent developments would however be a more considerable concern to the US surface fleet. General Secretary Nikita Kruschev’s plan for the Soviet Navy was to be one that was capable of defending its own coasts using light warships armed with anti-ship missiles, and submarines which could stalk shipping lanes for enemy vessels. As opposed to Stalin’s views, Kruschev’s plan heavily favored the development of cruise missiles and submarines over a balanced fleet, and largely handicapped the development of larger warships.

The torpedo shaped November class was a massive, if clumsy, step forward for the Soviet Navy. While unreliability and loud acoustic emissions plagued these boats, they showed the promise of nuclear submarines to future Soviet naval planners. [US Navy]
The immediate products of this philosophy were the Echo class nuclear submarines, and to a lesser extent, the conventional Juliet class. These new boats carried heavy, anti-ship cruise missiles and were initially considered a serious threat to US carriers. They were not, however, without serious limitations. They required cooperating patrol planes to share radar data for over the horizon targets, and needed to stay on the surface for up to thirty minutes before carrying out the attack with their long range missiles. They were accordingly extremely vulnerable when operating in areas without a substantial Soviet air presence. The more advanced Charlie class materialized after Kruschev’s fall, and was capable of submerged launches, but of slower and short ranged missiles. With Kruschev gone, the Soviet Navy largely abandoned any plans of Atlantic convoy raiding to pursue building better defenses against American Polaris missile subs, and later to focus on denying potential enemies access to bastions where their own SSBNs patrolled. Largely under Admiral of the Fleet Sergei Gorshkov’s direction, the Soviet fleet would try to right itself to become a more balanced force, so that it might better assist Soviet foreign policy, and to build up a defense against wartime incursions from enemy aircraft carriers and modern nuclear submarines.

In spite of the limitations of the Soviet nuclear submarine fleet of the sixties, their growing capabilities would prompt the US into developing their anti-submarine forces even further. Throughout the sixties, new aircraft ASW tactics were employed to replace the old snorkel-chasing methods. A greater focus was placed on the use of sonobuoys, which could be used to survey larger patrol areas, and the newer versions of which were growing ever more sensitive and sophisticated. Greater coordination with surface vessels was also employed, with newer destroyers and frigates mounting considerably more powerful sonar systems. Overall, US nuclear subs would take up an ever more important role in anti-submarine warfare, massive new hydrophone lines were laid in strategically important areas, and the aircraft carrier was soon to take a primary position in anti-submarine strategy.

Viking

In the world of the nuclear submarine and the jet carrier air group, the S-2 Tracker was becoming an ever more inconvenient asset. As carriers began to carry an ever greater number of jet aircraft, there was some frustration with having to still carry stores of aviation gasoline for the S-2s. The situation was not improved by the retirement of the WWII era converted CVS, which would be entirely out of service by the early 1970s. As a result, the entire surface ASW framework was to be restructured. Among the earliest moves was to announce a competition for the S-2 replacement in 1964, under the designation VSX. The new plane was required to have at least twice the speed, twice the range, and twice the ceiling of the aging Tracker. Lockheed was among the most promising entrants due to their previous history in designing maritime patrol aircraft, though their lack of experience with carrier based aircraft saw them partner with LTV Aviation, and the new ASW gear was to be designed by Univac Federal Systems.

A wooden mockup of Lockheed’s entry into the VSX competition. [US Navy]
Lockheed’s Viking was a robust, high wing aircraft which featured a pair of turbofan engines for their power and fuel economy. The plane also carried nearly every modern airborne submarine detection system of the time. Its four crewmen operated the aircraft’s systems in coordination with a central, general purpose digital computer, which greatly aided the crew in processing the information gathered by the aircraft’s sensors. Further crew integration was accomplished through the use of multi-purpose displays that could show information from any of the aircraft’s crew positions. In addition to the MAD, radar, and sonobuoy systems, the plane was equipped with a FLIR system mounted to an extendable turret which was capable of detecting snorkeling, or near surface submarines and sea mines. To complement its sensors, the aircraft had a maximum speed of approximately 429 kts, a ceiling of 40,000 ft, and a maximum endurance of over six hours. Of the entries from Grumman, General Dynamics, and Convair, Lockheed’s design won out.

They were formally awarded the contract in 1969. The first of eight YS-3A prototype and pre production aircraft flew only three years after the contract was finalized in 1972, with the aircraft entering service two years later. This program was also the first to have a formalized set of milestones to ensure costs were kept low and technical risks were reduced. All program milestones were met ahead of schedule, and the plane was prototyped, built, and delivered in quantity in only five years. Their carriers too were modified to better suit ASW operations. In 1971, the USS Saratoga was the first to receive an ASW analysis center and support shops for ASW gear and weapons. All carriers but the older, smaller Midway class were able to receive the improvements. Prior to the introduction of the Viking, carriers operated S-2’s, with the introduction of the new aircraft vastly improving the anti-submarine capabilities of US carrier battle groups. The plane could perform an ASW search quickly at 35,000 feet at a speed of over 300 kts, a massive improvement over the S-2’s 135 kts at 10,000 ft. Even before considering the massive improvements in sensors and the centralized computer integration, the Viking could patrol truly massive stretches of ocean for a plane of its size. With a payload of four lightweight torpedoes and 60 sonobuoys, the Viking could fly out 826 nmi from its carrier, and conduct a two hour search before having to return. The use of external stores and airborne tankers could push this already phenomenal range out even further.

The unified CV concept brought together the anti-submarine and surface distinctions, as the old sub-hunting legacy carriers began to be decommissioned. The carrier’s air wing was tailored to its deployment goals. [US Navy]
VS-21, the first S-3A squadron, was deployed aboard the USS John Fitzgerald Kennedy in the summer of 1975. During its Mediterranean deployment, the Kennedy was able to truly demonstrate the universal carrier concept. Previously, carriers were divided between the CVS, sub hunting carriers, and the CVA’s, which hunted everything else. The introduction of the Viking enabled the consolidation of all US carriers into CV’s, the new concept seeing carriers equipped for every conceivable mission. However, the S-3A was not the only newcomer to the ASW mission. The year prior to its first deployment saw the introduction of the Kaman Sh-2F Seasprite. This light anti-submarine helicopter would soon be found aboard most US warships, extending both their maximum search and offensive ranges. In short, the US surface fleet’s ASW capabilities had been thoroughly improved through the adoption of these two aircraft, well in advance of the predicted improvements in Soviet nuclear submarines.

An A-6 Intruder and S-3A Viking overfly a surfaced Project 641 ‘Foxtrot’ class submarine. These boats had improved fire control and sensors over the older Whiskey and Zulu class boats, but were otherwise built along the same post-WWII lines. Significant improvements in regards to quieting and hull form would not be achieved until the later Project 641B ‘Tango’. [US Navy]
In service, the S-3A was primarily a screening element for the carrier group and any surface groups it might be supporting. A US carrier group is typically deployed alongside independent surface action groups and nuclear submarines, these often being the outermost defenses for the carrier group. The carrier’s offensive range and ability to survey thousands of miles of ocean make it the center of naval operations, and the most well defended asset. It was the job of the outer forces to screen the path and potential approaches to the carrier from enemy submarines, and to a lesser extent surface ships, though those more often fell under the purview of other aircraft and vessels.

Given the distance between these forces, gaps inevitably form, and these areas are typically patrolled by aircraft. In wartime, the Viking could quickly fly out to these locations and deploy a grid of sonobuoys, which it could maintain for several hours before being relieved by other aircraft. In addition to screening the path of the carrier, the S-3A could also be tasked to patrol the open ocean to search for older cruise missile submarines, which had to surface for long periods to fire their weapons. The S-3 would eventually receive Harpoons for this role, but initially, it would carry Hydra 2.75 inch rockets or unguided bombs. By the late 70s, these submarine ‘Shaddock’ missiles were easily defeated by the new EW systems and defensive weapons added to destroyers, cruisers, and carriers, but they still posed a threat to lighter warships and shipping. In addition to open ocean patrols and barrier searches, the Vikings could be quickly dispatched to support patrolling frigates and destroyers which were tracking submarines.

While the Victor class boats were primarily designed around the anti-submarine mission, they could fire salvos of two heavy weight, long range Type 65 torpedoes for use against large surface groups. The boat first entered service in 1977 and represented a major success in achieving low acoustic emissions in Soviet submarine design. [US Navy]
Though the Soviet fleet consisted of a large number of these older submarine classes, new models of Soviet nuclear submarines would pose a greater challenge. A change in design philosophy would see a shift in focus away from achieving the best possible speed, to a balanced approach which placed greater importance on lower acoustic emissions. When commissioned in 1974, K-387, a Project 617RT ‘Victor II’, was the first Soviet nuclear submarine to incorporate rafted equipment. With its turbines suspended on vibration dampening mounts and its hull clad in anechoic rubber tiles, it was remarkably quieter than its forebearers. Further improvements to this class resulted in the Project 617RTM ‘Victor III’, with the first boat being commissioned in 1978. However, sound reduction was only marginally improved, with much of the focus being placed on new sensors, with the main mission for the sub being ASW. With 48 total Victors of all classes being produced, it represented the modern workhorse of the Soviet submarine force. More concerning to the carrier, however, were the successors to the Echo and Charlie class SSGN. The Project 949 ‘Oscar’ was a massive vessel which carried 24 P-700 ‘Shipwreck’ missiles, three times as many missiles as the Echo. Capable of submerged launches and engaging surface targets at long range, the Oscar lacked the handicaps of the earlier boats, and its state-of-the-art missiles boasted high speed and countermeasure resistance. A single Oscar could put the air defenses of a carrier battle group to the test, and thus long range anti-submarine screening became key for naval planners. The character of the Soviet submarine force of the eighties was rather peculiar, being composed mostly of obsolete to somewhat up to date vessels, but with a small and growing pool of cutting edge submarines.

Vikings among A-7 Corsair II and A-6 Intruder strike aircraft aboard the nuclear aircraft carrier Dwight D. Eisenhower, 1980. [National Archives]
These ever-advanced models of Soviet submarines were anticipated and largely matched by the US Navy’s efforts to build a defense against them. All new warships possessed powerful new sonar systems and light ASW helicopters, and the carrier based S-3A sat at the center of fleet-wide anti-submarine strategy.

Into the 80s

While the S-3A proved an incredible new addition to the fleet, it soon encountered an unexpected challenge. As a result of the post-Vietnam defense cuts, the spare parts program for the Viking was among the worst affected. Stocks of replacement parts began to grow tight by 1977, though they would not pose a serious issue until the turn of the decade. As a result of stricter rationing of components, the mission readiness level of the Viking squadrons often fell to below 40% in 1981. However, the problem was soon identified and the procurement of more replacement parts began the following year, along with a new series of maintenance programs to increase readiness. Thanks to these efforts, the mission readiness of these squadrons climbed to 60% in 1983 and rose to 80% in the coming years, the highest in the fleet.

While the Navy was procuring additional parts, they also initiated a program to drastically improve the offensive and sensor capabilities of the aircraft. The Weapon Systems Improvement Program would seek to prepare the S-3 Viking for its service into the new millennium. Most of these improvements were focused around the aircraft’s sensor systems, most notably its new inverse synthetic aperture radar, which boasted a much higher capability in regards to periscope and snorkel detection, and its acoustic sensor suite. The acoustic data processor was improved through the use of a standardized naval signal processor which ran on a software shared among new naval maritime patrol aircraft, a new sonobuoy receiver boosted the available channels from 31 to 99, and it received a new, more reliable tape recorder for storing gathered acoustic data.

An S-3 passes a Kilo class submarine. While much of the Soviet diesel-boat fleet consisted mostly of obsolete classes like the Foxtrot and Romeo into the 1980s, the Kilo was thoroughly modern. [The Drive]
In addition to its sensor improvements, the Viking received the new ALE-39 countermeasure system, and its electronic support measures were improved to allow better classification of contacts by their radar and radio emissions. Lastly, it finally received the capability to utilize the AGM-84 Harpoon missile, with the pair of missiles being mounted on the outer hardpoints. With a range of approximately 75 nmi’s, the sea skimming Harpoon could prove very difficult to detect and shoot down. As more effective air defenses against sea skimming missiles would not become widespread for almost a decade, the inclusion of this weapon would make the Viking a considerable anti-surface asset, along with its already impressive anti-submarine capabilities.

The sum of these upgrades would end up seeing the modified aircraft identified as S-3B’s, as squadrons began to receive the improvements in 1984. In addition to these upgrades and after the parts shortage, the scope of duties for the aircraft began to grow over the years. Among the first new tasks assigned to the Viking was to act as an airborne tanker. The long endurance of the aircraft, coupled with its incredibly fuel efficient turbofan engines, made it extremely capable in the new role. Carrying ‘buddy stores’, the S-3 could increase the range and endurance of cooperating carrier-borne aircraft in a much more efficient manner than the Ka-6d tanker, or a fighter or strike aircraft carrying the fuel tank and drogue system.

As the 1980s drew on, the Navy began to push the operational limits of the aircraft out ever further, and to great success. The S-3 took on the aerial mining mission, and during the Northern Wedding and United Effort exercises of 1982 and 1983, the operational search range of the Viking was pushed out to 1000 nmi with the use of airborne tankers. Even more noteworthy, they were able to detect and track submarines at that range during the exercise. While the S-3 Viking was initially introduced to serve a single, and very specialized purpose, the aircraft would end up proving extremely versatile and provided a number of new services to the carrier fleet, far beyond the expectations of its designers.

Operation Desert Storm and Late Career

An S-3B tanker launches from the deck of the USS Nimitz. [National Archives]
As the Cold War came to a close, events in the Middle East soon culminated in the largest armed conflict since the end of the Vietnam war. As Iraq invaded the neighboring country of Kuwait over oil disputes, a coalition was built among Arab and Western militaries to oust Iraqi forces from Kuwait and deal a serious blow to Saddam’s forces. Along with a massive USAF contingent, the US Navy would deploy six aircraft carriers in order to dislodge the Iraqi army from Kuwait. Of the force, USS Kennedy, Saratoga, America, Ranger, and Roosevelt carried embarked squadrons of S-3B’s. USS Midway lacked a squadron of Vikings, as it did not possess an ASW analysis center.

A total of 43 Vikings would be active across these carriers by February 1991, where they would serve in a number of roles. Ironically, due to Iraq’s lack of a submarine force, ASW was not a role they performed during this conflict. These aircraft flew a total of 1,674 sorties between January 17 and February 28, 1991. The majority, with 1043 flights, were aerial refueling missions supporting other coalition aircraft. However, they also flew a number of reconnaissance, electronic warfare, and several surface air combat patrol sorties, these numbering 263, 101, and 20, respectively. The rest of their flights were categorized as unspecified support missions, or ‘other’.

Apart from aerial tanker duties, these Vikings flew most of their patrols to survey the Persian gulf, in order to track what few warships Iraq had, and to mark the location of mines. Some Vikings were also involved in the search for Iraq’s short range Scud ballistic missiles, a great fear at the time being that some of them may have carried chemical weapons payloads. They also performed a number of unorthodox tasks. For instance, the US carrier air groups could not electronically receive their daily air tasking orders from the coalition headquarters in Riyadh, Saudi Arabia. Their solution was to dispatch an S-3B to pick them up on a near daily basis. Among the most imaginative uses of the aircraft was in delivering photos from carrier based reconnaissance services to units fighting on the ground. This was done by placing the photos in an empty sonobuoy tube and parachuting them to units. One Viking also sank an unidentified class of Iraqi patrol ship on February 20, 1991, after dropping three unguided mk82 bombs on it. Offensive patrols were comparatively restricted and were conducted in areas with limited anti-aircraft threats.

A long exposure shot of deck crew preparing an S-3B on the USS Truman during Operation Desert Shield. [National Archives]
After the end of the war in the Gulf, the S-3B was used for continued surveillance of the area and to uphold the sanctions on Iraq during Operation Desert Shield. It likewise performed similar support roles in the numerous NATO air operations over the former Yugoslavia. Their roles during those conflicts were almost entirely restricted to airborne tanker duties, though a number of Vikings, including a specialized ELINT model, performed signals intelligence missions.

As a result of the collapse of the USSR, the global submarine threat to the US Navy declined to almost nothing, and thus the Viking squadrons transitioned from anti-submarine, to surface control units to better represent their more multipurpose role. They would eventually discard their ASW equipment, with the anti-submarine mission being made the purview of the US submarine fleet and long range maritime patrol squadrons. Several new upgrades were initiated during the turn of the millennium, mostly in regards to new avionics and improvements to carrier landing aids. They would also include the Maverick Plus upgrade, which would enable the S-3B to use IR guided models of the AGM 65 missile, and the AGM-84H family of ground attack missiles. However, after the KA-6D left the service in the late 90s, the Viking would become the fleet’s primary aerial tanker.

The last major operation where the Vikings saw use was during the later invasion of Iraq, during which they primarily acted as tankers. There were, however, some strikes carried out by S-3Bs using the new Maverick Plus system, but these were comparatively rare. As the 2000’s came to a close and the US carrier force wished to divest itself of all but the most essential airframes, the Viking had fully left the service by 2010. The fleet was thus without a dedicated aerial tanker, and instead employed F/A-18s carrying ‘buddy stores’ to refuel other fighters.

Perhaps its later most publicized use was in delivering President George H.W. Bush aboard the carrier USS Abraham Lincoln after the invasion of Iraq. There, he delivered an address to the nation regarding the end of Operation Iraqi Freedom, in front of the long derided banner which simply read ‘Mission Accomplished’.

NASA

While the Viking’s military career came to a close, a number of aircraft were transferred to NASA as test aircraft in 2004. One of these planes was further developed into a dedicated testing platform in 2006 and was subsequently demilitarized. Most of the existing avionics were replaced with more contemporary civilian equipment and provisions for adding experimental equipment were installed. The Viking was given the civil air registration code N601NA and would see heavy use by the administration for the next 15 years, with the remaining Vikings being used for ground testing.

The NASA Viking proved to be an ideal platform to run a variety of experiments that required steady, precise flying at low speeds. [aionline]
The plane was used for a variety of missions regarding aeronautic safety, aerodynamic studies, and Earth studies. Operating out of the Glenn research center, the plane tested engine icing under harsh conditions, flew communication equipment tests over much of the US to determine FAA guidelines for unmanned aircraft, and flew over the Great Lakes to study algal blooms. This Viking was the last airworthy example of the entire line, and was finally retired in July of 2021. NASA has since donated the plane to the San Diego Air and Space Museum.

Operating Characteristics

The high and broad wings of the Viking presented good low speed flight characteristics and high maneuverability. This was also aided by the lateral control system of the aircraft, which consisted of a set of small outboard ailerons, a pair of spoilers above the wing and one on the underside, and a leading edge flap. Pilots in both the Navy and NASA test programs praised the responsiveness and stability these systems provided. This ability was well valued during low altitude MAD searches and during low level communications testing for NASA and the FAA. However, at higher speeds, pilots needed to be aware of a degree of oversensitivity, as the aircraft did not possess a fly by wire system.

A view of a carrier flight deck from the cockpit of a Viking. [The Drive]
The Viking had an extremely high carrier boarding rating thanks to its dynamic lift system, which in combination with the spoilers, gave the pilot a high degree of control during their final approach. The slow descent of the aircraft also permitted both the pilot and the LSO considerable time to make alterations. This is not to say this was a simple affair, as the aircraft was fairly sensitive to the air disturbance that forms immediately behind the moving carrier, and thus the pilot is still required to make the approach with caution. The engines had to be practically idled on the glide slope, and still often felt overpowered. The DLC system was essential, though the flaps were not, with many recoveries being flap up. Getting off the carrier was a far easier affair, as the aircraft only had a speed requirement of 120 kts under a normal load. Off wind catapult launches were made easily and some pilot and ground crew would remark that the Viking seemed like it could just fly off on its own. In both launching and recovery, the aircraft was remarked to handle well under poor conditions.

The addition of an APU in this aircraft greatly simplified and accelerated the start up procedure, as it did not require the pilot to request external power from the deck crew. A relatively simple start up enabled the aircraft to be ready some 15 to 20 minutes before its scheduled launch, and helped in speeding up the turn-over in deck operations. The only inconvenience the aircraft presented was that its low mounted engines were considerably quieter than most other embarked aircraft, meaning ground crew needed to pay particular attention to these aircraft as they moved across the deck. In short, the S-3 was very well suited to carrier operations.

A technician checks over the TACCO’s multipurpose display on an S-3B. The displays at each station were of slightly different dimensions. The TACCO station’s monitor was enlarged on the B model of the aircraft. [National Archives]
A high level of crew cooperation was possible on the Viking thanks to the centralized nature of its avionics, sensors, and weapon systems. In managing all of these functions through its central computer, most crew functions were visible across all stations and, in some cases, could actually be managed between them. This was primarily achieved through the multipurpose displays at each station, which allowed crewmembers to share information. This made the SENSO and TACCO stations far more capable than they were on other aircraft, allowing for some division and management of the workload. The TACCO position likewise possessed a high degree of integration with the pilot and copilot, and in certain autopilot modes, could guide the aircraft to the weapon release point. All stations effectively had a high degree of situational awareness outside the aircraft, as the multipurpose displays could be set to show various airborne, surface, and subsurface contacts relative to their positions from the aircraft. The computer system itself proved fairly easy to manage, and designed with self-maintenance in mind. In the event of a system error, the computer could run a diagnostic and be reset in flight. Thanks to this level of digital integration, the Viking was viewed as being as capable as a number of patrol aircraft with significantly larger crews.

In the submarine hunting role, the Viking was in no shortage of equipment. The primary means of conducting an anti-submarine search were its sonobuoys. The aircraft carried a variety of these devices, which allowed for passive listening, or actively sending out an echolocating ping which revealed the positions of nearby submarines. These were often arrayed out in a grid like pattern in an aircraft’s patrol area to allow for the surveillance of a much larger area. They were typically dropped in line-like, or triangular patterns when used to try and get an accurate fix on the submarine’s location. Through acoustic analysis, the Viking was able classify submarine contacts by comparing them to existing sound profiles and was capable of gathering new profiles on vessels which had not yet had one compiled. Sonobuoys were usually dropped from the aircraft’s cruising altitude of 35,000 ft, though often from lower altitudes when a contact had been found and a finer pattern of the devices needed to be sown. The sonobuoy system was the first of its kind capable of accurately pinpointing the position of each device.

Sonobuoys provided a screen through which a transiting submarine could be detected, though they were not used for basic open ocean searches. The limited effective range of the individual devices meant that they were used for screening areas ahead of surface groups, filling gaps between other patrol areas, and investigating contacts that were beyond the range of other warships. The aircraft could hand off its sonobuoys to other aircraft from a shared channel, and could receive information from other, off-aircraft sensor sources through their datalink. Thus, in the submarine hunting role, the aircraft could either be a proactive tool, in performing its own searches, or reactive, in responding to suspicious or identified subsurface contacts from other aircraft and vessels. In concert with more modern anti-submarine assets, like the Spruance class destroyer or underwater hydrophone lines, the Viking could prove an incredible asset well beyond the limitations of its own sensors. The Viking was one of, if not the, best equipped ASW aircraft of the entire Cold War. Designed primarily around countering the threat of nuclear submarines, it would of course prove even more capable against diesel-electric submarines which presented more opportunities for detection.

Carrier deck crew load a sonobuoy into the aircraft’s external chutes. Viking’s could carry passive, active, dual purpose, oceanographic, and search and rescue beacon buoys. [National Archives]
In conjunction with sonobuoys, the aircraft possessed its radar, FLIR optics, and a magnetic anomaly detector. The radar of the aircraft was designed to detect periscopes and snorkels deployed by near surface submarines. The key was to look for contacts that either appeared or disappeared from the scope without explanation, and was otherwise a very straight forward system. The FLIR sensor was used to detect heat sources, and in the submarine hunting mission, was used to spot submarines at a shallow depth, and the extended snorkel and exhaust of diesel-electric submarines recharging their batteries. Last was the MAD, which detected the magnetic field of a submarine, which caused slight disruptions in measurements of the earth’s magnetic field, hence the anomaly. This was the only sensor which required the aircraft to fly low, and the limited range of the sensor also meant a contact was typically only revealed if it was overflown. The radar and infrared systems were also important tools in conducting long range surface reconnaissance for the entire fleet. These systems were also necessary in providing targeting data for the Harpoon anti-ship missile.

In employing weapons, the majority of the work was done through the TACCO position. This crewmember assigned weapons to targets, and in coordination with the pilot and copilot/COTAC, delivered them. Originally, this meant he would deploy the Mk.46 lightweight torpedoes and depth charges, with the plane being capable of deploying nuclear models as well. Unguided munitions, typically Zuni rockets, Mk 82 iron bombs, or Mk 20 Rockeye cluster bombs, were the responsibility of the pilot and would have been used against surfaced guided missile submarines, or damaged warships. Later in the aircraft’s career, the TACCO would deploy mines, launch AGM 84 Harpoons, and later operate a variety of air to ground missiles with the introduction of the Maverick Plus upgrade.

The aircraft later excelled as an airborne tanker, where its ability to operate for long periods and at range from the carrier were crucial. The task was relatively simple enough, fly straight ahead while offloading fuel onto another aircraft through a hose and basket fuel transfer line. The asymmetric load of the fuel tank and drogue mount required constant trimming, which grew worse as the tank was drained, but this was a largely simple job the plane was well suited for.

Construction

A general diagram of the S-3B. [S-3B manual]
The S-3 was a high wing, twin engine, carrier based anti-submarine aircraft. It possessed a very durable semi-monocoque airframe with three folding flight surfaces, being the wings and the vertical stabilizer. The fuselage was wide enough to permit the fitting of a considerable set of ASW gear, and an internal weapons bay. The general construction of the aircraft was fairly conventional in comparison to other carrier based aircraft. Lockheed was the primary contractor for the aircraft,  with LTV building the wings, engine pods, tail assembly, and landing gear, while Univac provided the digital computer and some of the avionics.

The nose of the aircraft contained the radar, followed by the cockpit which seated the pilot and copilot, behind them were the weapons and sensor operators. Aft of the crew sat the forward avionics bay, which itself was over the internal weapons bay, and to the rear of it all was the MAD boom and rear avionics bay. On the underside of the aircraft were the sonobuoy shoots, which in addition to the 48 outer slots, held additional internal stores for 12 more devices. All critical systems had redundancy built in.

The landing gear, and catapult launch bar, were derived from those of the LTV F-8 Crusader and A-7 Crusader II. These consisted of a forward, upward retracting gear and two rear landing gear which retracted inward toward the fuselage. These are hydraulically actuated, though in an emergency, they could be extended by bypassing the hydraulics and letting gravity, and a leaf spring to force the gears into the extended position.

The wings of the aircraft were designed to permit a high degree of control and stability at both low and high speeds at low engine power, up to the maximum permitted speed of 429 kts. These were mounted high on the fuselage and possessed a high aspect ratio of 7.73 and a rearward sweep of 15 degrees. The wings consisted of an outer panel, which could fold inward, and an inner panel, roughly a third the length of the outer panel, which contained a fuel tank, and supported an engine nacelle and a pylon which could fit external fuel or weapons. The tall vertical stabilizer also folded down and to the port side to permit the aircraft to fit the carrier’s hangar doors. The extendable airborne refuel probe was stored just ahead of the wings.

Spoiler, aileron, and flap deployment diagram. [S-3B Manual]
The Viking possessed an unusual flight control system which combined six large spoilers with a set of small ailerons and a leading edge flap. Lateral control was greatly aided by the inclusion of the spoilers in combination with the leading edge flap, which permitted effective control at very low speeds with low engine power settings. All control surfaces on the aircraft were deflected using hydraulically actuated servos, with an artificial feel system designed to give the pilot an idea of the extent of control surface deflection. These controls did however prove to be somewhat oversensitive at high speed. Overall, the control surfaces were very effective on patrols at low speed, though they could prove rather clumsy in a carrier landing pattern. This was largely due to the overpowered engines, which gave the aircraft a somewhat unorthodox glide slope and its large wings increased its sensitivity to the ‘burble’ air disruption behind the carrier. To compensate for this, the Viking was equipped with a dynamic lift control system which provided 12 degrees of speed brake extension and retraction through the upper spoilers.

The S-3 was powered by a pair of either General Electric T34-GE-2 or T34-GE-400A high bypass turbofan engines. These both produced 9,275 pounds of thrust at sea level, and the former was used only on pre-production aircraft. These engines used a dual-rotor, single stage, front-fan configuration with a bypass ratio of 6.23 to 1. These were divided into four major sections, being the fan, compressor, combustor, and turbine. The fan was driven by the low pressure turbine and produced 85 percent of the engine’s total thrust. The compressor was composed of 14 stages which compressed air prior to the combustion section, and provided the air for the pneumatic systems aboard the aircraft. The combustor section was where the compressed air was mixed with a fuel air mixture and ignited. The resultant flow drove the high and low pressure turbines within their own section, the high pressure turbine being responsible for driving the compressor, and the low, the fan. The air flow continues out the back of the low-pressure turbine to comprise the remainder of the engine’s thrust.

Standard and exploded views of the General Electric T34-GE-400A turbo fan engine. [S-3B Manual]
The aircraft was fitted with a number of surface and subsurface sensors. The Viking originally possessed an AN/APS-116 search radar, which was primarily designed to detect the masts of submarines near the surface, but doubled as a general purpose surface search radar. This was replaced on the S-3B with the APS-137 inverse synthetic aperture radar which was more than twice as effective at detecting low RCS masts and had improved surface search capability. Specifically, it gained the ability to identify surface vessels at long range by comparing radar returns to existing 2D profiles of vessels. The aircraft also carried an AN/ASQ-81 magnetic anomaly detector fitted to an extendable boom at the rear of the aircraft often called the ‘Stinger’. The boom allowed the sensor to be placed farther away from ferrous objects on the aircraft, which might interfere with any measurements taken. They also carried the OR-89 FLIR sensor, it being mounted in an extendable turret on the forward, starboard side of the aircraft. The sensor would display surface and near surface contacts, making it extremely useful in detecting mines, submarines at a shallow depth, and the exhaust of diesel-electric submarines charging their batteries.

The Viking’s FLIR turret in its deployed position. [replane]
What could be considered the primary anti-submarine sensor were the aircraft’s sonobuoys. Up to 60 could be carried in the chutes that sat flush with the outside of the aircraft and internal stowage. The aircraft communicated with minimal signal emissions and was capable of displaying their exact positions. Data from the devices was processed using an OL-82/AYS data processor and, coupled with its original receiver, was initially capable of monitoring 31 buoys. When upgraded, the acoustic data processor incorporated a new standardized UYS-1 signal process which had improved reliability and had parts and software commonality with other ASW platforms. A more advanced sonobuoy reference system, AN/ARS-2, would also boost the number of usable sonobuoy channels from 31 to 99 and an automatic channel scanning capability to search for available RF channels. The last upgrade to this system saw the analogue tape recorder switched from AN/ASH-27 to the AQH-4(V)2, which was both smaller, more reliable, and was compatible with the new UYS-1 signal processor.

The rear two stations of the S-3A Viking. The SENSO’s dual screens allowed him to monitor multiple sonobuoys simultaneously, this information being only partially available to the single screen displays at all of the other positions. [S-3B Manual]
The aircraft’s countermeasures initially came in the form of the AN/ALR-47, a passive sensor which displayed radar emissions from search and track radars. This was later supplemented with the ALE-39, which included countermeasure dispensers. It also received electronic support measures, which allowed it to detect a wider variety of radar emissions to allow it to classify their emitters. In the event of being shot down, the aircraft was equipped with ejection seats. These could be used on the ground while the aircraft was still, and had a preset ejection sequence to prevent any collisions in air.

All of these systems were managed through a single Univac AN/AYK-10 digital computer. The system allowed for a much higher ability to process information compared to the isolated systems in use on virtually all other maritime patrol craft. Additionally, and perhaps much more importantly, it allowed the crewmembers to display information from their own stations to one another through a set of multipurpose display screens at every station. This allowed for the sharing of most sensor data across all four positions, though it was more limited in the case of sonobuoy readouts, as they were half displayed on a secondary screen at the SENSO station. These displays would give crews the ability to coordinate during surface and subsurface searches, and improve planning when preparing to attack. This was particularly valuable to the copilot/COTAC, whose job was to essentially direct the aircraft in achieving its mission. The addition of this system essentially gave them access to every senor and allowed them to work closer with the TACCO when it came time to deploy weapons.

Initially, the Viking could be armed with up to four Mk 46 torpedoes, being either the Mod 1 or Mod 5 NEARTIP during the 1980s. Both types measured 8.5 ft long with a diameter of 12.75 inches, and both carried a 95 lb warhead. The Mod 1 possessed a maximum speed of 45 kts,with the NEARTIP being considerably faster. The NEARTIP provided better tracking of faster targets and better countermeasure rejection, having incorporated a new sonar transducer, control and guidance group, and a new engine which switched from solid propellant to liquid monopropellant. The Viking would also receive the new electric Mk 50 torpedo in the early nineties, but it would shortly after transition away from the ASW mission. There were provisions for mounting up to four torpedoes internally from hardpoints rated up to 600 lbs each. The bomb bay could also carry up to four mines and depth charges, or two B57 nuclear depth charges.

Crewmen prepare to load a Mk 46 torpedo aboard an S-3A. [National Archives]
The pair of external hard points could carry a combination of weapons, external fuel tanks, and airborne refueling systems. Initially, this was restricted only to unguided weapons and fuel tanks. Each hardpoint had a mounting capacity of 2,500 lbs and could carry up to three bombs through the use of the TER-7 bomb mount. The S-3B upgrade would allow the aircraft to use the AGM 84 Harpoon and was able to carry two of these sea skimming missiles. The last major upgrade package, which was installed around 2002, included a variety of avionics improvements, and the Maverick Plus system. This allowed the Viking to mount the AGM 65 Maverick, one per hardpoint, and the AGM-84E SLAM. The Maverick was to be used mostly against light shipping, with the SLAM providing stand off capability against ground targets. The SLAM could be guided manually after launch if a guidance pod was installed on one of the outer hardpoints. Both missiles were otherwise supported and targeted through a common display.

The S-3B could use any of the AGM-84’s in the Navy’s arsenal by the time of its introduction. The first of these was the Block 1B introduced in 1982, which had improved radar guidance allowing it to fly at lower altitudes. The subsequent 1C entered service in 1984 and incorporated a denser fuel, which increased its range by five nautical miles out to 80 nmi when launched from sea level, and added an alternate pop-up attack mode. The 1D entered service in 1992, with the lengthened missile possessing a range of 150 nmi and re-engagement capability, which allowed the missile to circle back to its target if it was deceived by chaff or electronic countermeasures on its first pass.

These sea skimming, turbojet powered missiles were exceptionally difficult to detect and intercept during the Cold War and flew at a constant Mach .85. These utilized active radar terminal homing, carried a 510 lb high explosive warhead, and had a flight reliability of over 93 percent.

Conclusion

A Viking prepares to launch after an F-14B Tomcat aboard the USS Nimitz during Operation Southern Watch, 1999. [National Archives]
With the exception of the parts shortage, the Viking can be said to be among the most reliable and versatile tools the US Navy has ever possessed. The aircraft proved a phenomenally capable anti-submarine aircraft, which entered service long before high capability threats entered service in the Soviet Navy. When that particular threat had gone, the plane continued to serve ably, as a tanker, a reconnaissance aircraft, and limited strike aircraft. Finally, the venerable aircraft ended its career as a research aircraft.

S-3A Viking

Specification

Engine T34-GE-400A
Maximum Continuous Engine Output (Maximum) 6,690 lbs (7,365 lbs for 5 minutes)
Combat weight at catapult 44,947 lbs
Gross Weight 36,574 lbs
Empty weight 26,581 lbs
Range [4x Mk.46 60xSonobuoys] 2,506 nmi
Combat radius [4x Mk.46 60xSonobuoys] 826 nmi for 6.9 hours at 346 kts
Maximum speed 429 kts at sea level
Crew Pilot, Copilot/COTAC, TACCO, SENSO
Length (Folded) 53.33 ft (49.42 ft)
Height (Folded) 22.75 ft (15.25 ft)
Wingspan (Folded) 68.67 ft (29.50 ft)
Wing Area 598 sq.ft

S-3 variant

General Description

Number built/converted

YS-3A Prototype/Preproduction 8 built
S-3A ASW Aircraft 180 built
S-3B ASW/ASuW Aircraft 160 converted from S-3A
US-3A Cargo Aircraft 4 converted from YS-3A
KS-3A Airborne Tanker 1 converted from YS-3A
ES-3A ELINT Aircraft 16 converted from S-3A

Viking Squadrons

VS-21 ‘Fighting Redtails’ VS-31 ‘Topcats’
VS-22 ‘Checkmates’ VS-32 ‘Maulers’
VS-24 ‘Scouts’ VS-33 ‘Screwbirds’  
VS-27 ‘Grim Watchdogs’ VS-35 ‘Blue Wolves’
VS-28 ‘Gamblers’ VS-37 ‘Sawbucks’
VS-29 ‘Dragonfires’ VS-38 ‘Red Griffins’
VS-30 ‘Diamondcutters’ VS-41 ‘Shamrocks’

(wikimedia, popular patch)

Credits

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

Illustrations

Gallery

The Viking flying alongside the older S-2 Tracker maritime patrol aircraft. The S-3A rapidly replaced the Tracker from 1974 to 78, when the last Viking left the production line. [jrdavis]
An S-3 is brought up to the flight deck in its stowed condition. The vertical stabilizer folds just below the rudder. [National Archives]
A member of the USS Enterprise’s deck crew warms their hands in a turbine. Taken during the Fleet EX’88 Exercise off the coast of Alaska. [National Archives]
A Viking prepares to launch from USS America. [National Archives]
The evaluation S-3B aircraft passed its final trials in 1985. A rapid upgrade program would begin in 1987. [flight manuals online].
S-3Bs on the crowded deck of the USS John C. Stennis in 2007. [National Archives]
An SH-60 Seahawk comes in to land on the USS Kitty Hawk. [National Archives]
A Sikorsky Sea King comes in to land on the USS Theodore Roosevelt. These helicopters and the Sh-60’s represented the inner circle of fleet anti-submarine defense. [National Archive]
 

A Viking, among other aircraft, aboard the USS John F. Kennedy. [National Archives]

An aircraft prepares to take on fuel from an S-3B tanker. Note the missing MAD boom and the covered sonobuoy chutes. [National Archives]
The most publicized use of the Viking. ‘Navy One’ lands on the USS Abraham Lincoln with President George W. Bush aboard to deliver a less than well received speech after the end of Operation Iraqi Freedom. [US Navy]
The ES-3 Shadow was an electronic surveillance aircraft which replaced the aging Skywarrior. It saw considerable use during the NATO intervention in the former Yugoslavia, where it monitored communications and radar emissions. It is easily distinguished by its dorsal equipment fairing [FAS]
A Viking with its MAD ‘stinger’ deployed. [The Drive]

Sources

Primary

Standard Aircraft Characteristics Navy Model S-3A Aircraft. Commander of the Naval Air Systems Command. NAVAIR 00-110AS3-1. January 1973.

NATOPS Flight Manual Navy Model S-3B Aircraft. Commander of the Naval Air Systems Command. NAVAIR 01-S3AAB-1. September 2000.

NATOPS Weapon System Manual Navy Model S-3B Aircraft. Commander of the Naval Air Systems Command. NAVAIR 01-S3AAB-1.1. December 2002.

Fiscal year 1976 and July-September 1976 transition period authorization for military procurement, research and development, and active duty, selected reserve, and civilian personnel strengths : hearing before the Committee on Armed Services, United States Senate, Ninety-fourth Congress, first session, on S. 920

NASA fiscal year 2010 budget request : hearing before the Subcommittee on Science and Space of the Committee on Commerce, Science, and Transportation, United States Senate, One Hundred Eleventh Congress, first session, May 21, 2009.

Department of Defense authorization for appropriations for fiscal year 1982 : hearings before the Committee on Armed Services, United States Senate, Ninety-seventh Congress, first session, on S. 815.

Department of Defense appropriations for 1984 hearings before a subcommittee of the Committee on Appropriations, House of Representatives, Ninety-eighth Congress, first session / Subcommittee on the Department of Defense.

NASA’s aeronautics R & D program : status and issues : hearing before the Subcommittee on Space and Aeronautics, Committee on Science and Technology, House of Representatives, One Hundred Tenth Congress, second session, May 1, 2008.

Department of Defense authorization for appropriations for fiscal years 1988 and 1989 : hearings before the Committee on Armed Services, United States Senate, One hundredth Congress, first session on S. 1174.

Department of Defense appropriations for 1985 hearings before a subcommittee of the Committee on Appropriations, House of Representatives, Ninety-eighth Congress, second session / Subcommittee on the Department of Defense.

Department of Defense authorization for appropriations for fiscal year 1983 : hearings before the Committee on Armed Services, United States Senate, Ninety-seventh Congress, second session, on S. 2248.

Secondary

Chambers, Joseph R.. Partners in freedom: contributions of the Langley Research Center to U.S. military aircraft of the 1990’s.

Brown, Ronald J. Humanitarian operations in northern Iraq, 1991: with marines in Operation Provide Comfort.

Knaak, Jerry. A Hunting We Will Go. Naval Aviation News. March-April 1997.

Vikings Sweep the Seas & Viking. Naval Aviation News February 1983.

LSO School and the Paddles’s Profession. Naval Aviation News V70, November-December.

Benjamin, Dick. A Sea Rover for ASW. Naval Aviation News January 1972.

Richman, John P. The Viking at Home in the Fleet. Approach, July 1975.

Francillon, Rene J. Lockheed Aircraft Since 1913. Naval Institute Press. 1987.

Polmar, Norman & Moore, Kenneth J. Cold war Submarines The Design and Construction of U.S. and Soviet Submarines. Potomac Books. 2004.

Polmar, Norman. Aircraft Carriers a History of Carrier Aviation and its Influence on World Events Volume II 1946-2005. Potomac Books. 2007.

Kaman SH-2F Seasprite

United States of America (1974)

Anti-Submarine & Utility Helicopter

190 total airframes built: 85 converted to SH-2F w/ 48 new airframes.

A SeaSprite takes on fuel aboard the Destroyer USS Briscoe. (National Archives)

Introduction

Kaman’s SH-2 proved an exceptional asset for the US Navy through the mid to late Cold War, serving a variety of roles across nearly the entirety of the surface fleet. Beginning its service as a multipurpose naval helicopter designed to ferry equipment and rescue downed fliers, the light helicopter soon played an even greater role as an anti-submarine aircraft. Replacing the outdated and clumsy DASH drone, the Seasprite incorporated cutting edge sensors to become a sub chaser that could fit on even the lightest modern frigates in the US Navy. Spanning the early sixties to the new millenium, the Seasprite served as an able light transport, search and rescue, and anti-submarine helicopter before finally being phased out by the UH-60 Seahawk.

Whirlybirds

Of all the world’s navies, that of the United States was the first to employ helicopters enmasse. While helicopters had undergone considerable development since the first usable designs had been conceived in the 1920s, they remained a clumsy novelty into the 1940s. This was until the Sikorski R-4 was developed. Igor Sikorski, born in the Kiev Governorate in the reign of Alexander II, was already an aviation legend before the Russian Civil War saw him emigrate to the United States in 1919. Having previously designed four engine biplane airliners in the Russian Empire, and several of the flying boats that saw Pan Am span half the globe, Sikorski was a name known for breaking new ground. His R-4 helicopter would build this reputation further. The greatest advantage the R-4 had over its foreign contemporaries, most notably the Focke-Anchleis 223, was its simplicity and ruggedness. The use of a main lifting rotor and anti-torque tail rotor would prove a far lighter, and more robust method of control than the transverse and intermeshing rotors that drove a number of contemporary types.


Igor Sikorskiy (right) aboard a test flight of his R-4 helicopter (wikimedia).

The R-4 reached the notice of the US armed forces through Commander William J. Kossler of the Coast Guard, after the officer had seen the XR-4 undergo a test flight in April 1942. Impressed, he invited fellow officer CDR W.A. Burton to see the helicopter. The report on the aircraft took note of its ability to conduct patrols at low speeds, and unlike US Navy airships, did not require a large hangar for storage. Initially skeptical, the Navy was later convinced of the aircraft’s anti-submarine and convoy surveillance properties. Limited production began in 1942 and testing was conducted through 1943 and ‘44, though its sub chasing capabilities were not pursued. Instead, the helicopter proved itself as an air rescue vehicle. Its first trial came on January 3, 1944, when it delivered vital blood plasma from New York City to Sandy Hook, New Jersey, through a violent storm, in order to treat sailors after a fire had sunk the destroyer USS Turner. In all, several dozen R-4s would be delivered to the Coast Guard and Navy, where they took part in a number of rescue missions across North America and the Pacific.

While the R-4 was still limited in its carrying capacity and presented pilots with challenging flight characteristics, it demonstrated the utility of helicopters to every branch of the US armed forces. Sikorski would capitalize on this over the coming decade with their heavy H-19 and H-34 helicopters. Entering service in the early fifties, these helicopters were all metal and equipped with heavy radial engines. In civilian and military service, they would prove exceptional, capable of airlifting cargo to otherwise unreachable areas. However, a new, revolutionary advancement would soon render them obsolete. In 1955, the French Allouette II became the first production helicopter to feature a geared gas turbine. The turbine provided a far better power to weight ratio than the radial engines, and it was compact, allowing it to be placed at the center of the helicopter and thus avoided the forward engine placement that made some earlier helicopters nose heavy. This engine also allowed the nimble Alloutte to possess a speed and range far beyond comparable piston engined models. From then on, it was clear that turbine power would be the future of helicopter design.

 

A Sikorsky ‘Choctaw’ helicopter hovers to recover astronaut Alan Shephard and a Mercury reentry capsule after the first manned US space flight. The addition of a powerful radial engine made these among the first successful heavy lift helicopters. (wikimedia)

In the US, the first experiments for this type of helicopter propulsion were pioneered by Charles Kaman’s aircraft company. The first successful experiment was achieved through combining the Boeing 502 turbine with his company’s K-225. Kaman, a former employee of Sikorsky, would develop this new helicopter along with his head designer, Anton Flettner, a German engineer who pioneered the use of intermeshing rotors. The experimental K-225 proved promising enough to warrant further development, and soon, the Kaman Aircraft company would produce a new utility helicopter along its lines. The firm’s HH-43 Huskie fire fighting and rescue helicopter fit the bill, and its later models were equipped with turboshaft engines in the late 50s.

 

However, the firm’s greatest success was soon to arrive, when the navy sent out a request for a new carrier-borne, lightweight helicopter.

Seasprite

The US Navy’s request for a light multipurpose and rescue helicopter was soon met with Kaman’s newest design, the Kaman Seasprite. The helicopter would settle the requirements, being capable of carrying up to 12 people, remaining compact and fuel efficient, and taking up little space aboard aircraft carriers. In the 1956 competition, Kaman’s design won handily and the next year saw a contract issued for procurement. The helicopter was the first Kaman design to feature a single main rotor, and in conjunction with the servo-flap rotor system, it was cutting edge, reliable, and possessed smooth flight characteristics.

The design, then named HU2K, first flew on July 2, 1959, and was introduced fully in December 1962. It proved to be robust with good handling, however, the single General Electric T58GE turbine left it fairly underpowered. This prevented it from taking on any new missions, but it was sufficient for the basic role it was designed for. These helicopters, later designated UH-2A and UH–2B, though largely identical, were produced until 1965, with a total of 142 airframes built.

A Kaman UH-2A/B flies alongside the USS Enterprise as a plane guard as it launches a Grumman E-2a Hawkeye. (wikimedia)

The Seasprites, supplied to utility helicopter squadrons, were distributed amongst US aircraft carriers and saw widespread use during the Vietnam War. There, they served largely as plane guards, where they took up a position alongside aircraft carriers when large scale air operations were underway. In case of an accident during take off or landing, the Seasprites would move in quickly to recover downed pilots. Search and rescue also fell under their purview, and alongside a number of other models, they pulled hundreds of airmen from the sea. As a fleet utility helicopter, they also flew ashore and between various vessels in order to transfer personnel and equipment. Medical evacuations were also among tasks these helicopters performed, moving injured personnel to ships with more substantial medical facilities. The small size and smooth controls of the Seasprite made landing on the basic helicopter facilities of most ships an easier affair compared to the bulkier Sikorsky Sea King. Its only drawback was the relatively little power offered by its small turbine engine. It could make for tricky takeoffs as the small helicopter was slow to climb.

In spite of it being underpowered, it proved to be a valuable asset to the fleet and was respected by its pilots. Naturally, the Navy wished for improved models. Kaman’s first move was to add a second turbine engine to the helicopter, the improved model being the UH-2C. As the production run had already been completed, the Navy sent Kaman the older A and B models back to the company in order to receive the upgrade. The C model was introduced in 1966, though now with its much higher speed and carrying capacity, it was soon deemed that the Seasprite was to take on a much wider scope of duties.

Sub Chaser

During the late sixties, the increased threat posed by ever more advanced models of submarines was of great concern to the US surface fleet. Even more concerning was a lack of long range anti-submarine weapons. While many ASW vessels did carry the ASROC missile, tipped with either a nuclear depth charge or a Mk 46 torpedo, there was some concern of submarines attacking from beyond the 6 to 8 mile range of this weapon. The existing long range anti-submarine weapon was the Gyrodyne DASH drone, a small drone helicopter capable of carrying depth charges and torpedoes. While it was compact, it was inflexible, and with no means of collecting additional data in the area of the suspected submarine, accuracy was very poor.

The UH-2D was an interim ASW model to test the helicopters ability to carry the equipment needed for the role. These are differentiated from the later 2F’s by their tail wheel being further out. This aircraft lacks the sonobuoy rack. (wikimedia)

This left most of the US Navy’s light surface forces, which often operated too far from the carrier to be covered by its airborne ASW umbrella, under threat from more modern submarines. The solution was found in the re-engined Seasprite. The new SH-2D represented the greatest change thus far, with the new aircraft sporting a chin mounted surface search radar, a rack to carry a Mk 46 lightweight torpedo, and a 15 chute sonobuoy rack. The small size of the helicopter would allow it to operate aboard some of the lightest frigates in the fleet, these being the Garcia-class.

The performance of the helicopter, and its ability to operate on nearly every major surface combatant, would see this mission expanded even further. Thus came the Light Airborne Multi-Purpose System, a fleet-wide program to equip most warships with helicopters in order to boost their anti-submarine and anti-surface capabilities. LAMPS I would place a now standardized SH-2F aboard nearly every frigate, destroyer, and cruiser in the fleet. In addition to the long standing utility missions, the helicopters were datalinked to their host ship to allow them to prosecute possible submarine contacts, provide long range surface surveillance, and allow for more effective over the horizon targeting of enemy surface threats.

The new SH-2F was largely the same as the proceeding UH-2D model, though it standardized the use of composite rotor blades which existed on some previous models, and its tail wheel was moved forward to enable it to better operate off of smaller ships. Some 85 Seasprites were converted to this type, and a further 48 were produced in the early 80s in order to cover a shortfall before the introduction of the SH-60B Seahawk. The new, standard LAMPS helicopter entered service in 1973.

LAMPS I

The LAMPS I program vastly increased the offensive and surveillance capabilities of participating vessels. This encompassed some half dozen ship classes ranging from the workhorse frigates of the fleet, such as the Knox and Oliver Hazard Perry, to the nuclear guided missile cruiser, Truxton. In the ASW mission, on detecting a suspected submarine, whether attacking or transiting, the ship would launch its SH-2F. Capable of using sensor data from the ship, the helicopter would move in and begin to deploy its sonobuoys, being either passive AN/SSQ-41’s or active AN/SSQ-47’s. The helicopter then relayed the sonobuoy data back to the ship for processing, and if the contact was found and classified, the helicopter would move in to attack with its Mk 46 torpedo. The onboard magnetic anomaly detector could also mark the position of a submarine if over flown by the helicopter. A ship equipped with ASROC could also join the helicopter in the attack, provided the target was in range. In the ASW role, the helicopter was a largely reactive measure, as it was unable to process its own sonobuoy data and lacked a dipping sonar, and thus required other platforms to detect the submarine first. This is not to say it lacked considerable offensive potential, as the powerful hull mounted sonar arrays aboard the Knox class frigates and Spruance class destroyers, and the OHP’s short range but highly sensitive sonar, were among the most advanced systems of their kind and could give early warning to submerged threats. The presence of the helicopter thus allowed ships to prosecute, classify, and engage submerged contacts that would otherwise be beyond the effective range of their sensors and weapons.

The Spruance class Destroyers were among the most capable anti-submarine warships used during the Cold War. With their advanced sonar systems and two helicopters, they could pose a serious threat to even the most modern nuclear submarines. (National Archives)

The Spruance class in particular could prove very dangerous to submarines at range thanks to its convergence zone sonar. The AN/SQS-53 could make use of the aforementioned phenomenon, and under ideal conditions, detect submarines at extreme ranges. These zones are where sounds are bounced off the seafloor or thermal layers into a concentrated area and are thus made dramatically louder. Convergence zones are exploited by all ASW vessels, though the specialized sonar aboard these ships allowed them to exploit sound propagated at distances far in excess of the norm. A Spruance class ship making use of a convergence zone could dispatch helicopters against submarines potentially dozens of miles away, making them among the most capable ASW vessels of the Cold War. In the absence of a convergence zone, it switched to a short to medium range mode. It shared this system with the Ticonderoga class guided missile cruiser, and the Kidd class destroyer, both of which used the same hull, however their role was air defense. These ships all transitioned to LAMPS III once it became available in the mid 1980s.

The LAMPS system featured most prominently in escort and screening vessels, namely the Knox and Oliver Hazard Perry (OHP) class frigates. The Knox class was an anti-submarine frigate with limited anti-surface capability that entered service in 1969, with 46 vessels being commissioned in all. These ships carried a single Seasprite and were armed with an ASROC launcher, which later received the capability to launch Harpoon anti-surface missiles. The OHP class carried no ASROC launcher, though they instead carried two helicopters. The last 26 of the class were LAMPS III ships and carried the heavier and more capable Sikorski Seahawk. In place of the ASROC launcher was a Mk 13 mod 4 launcher for Standard missiles and Harpoons. Both frigates carried hull sonar and towed arrays, the Knox possessing a larger hull array, and the OHP carrying a short range, high resolution hull sonar system, with a towed array being used for longer range surveillance. The difference in systems was due to the OHP being designed as a fast escort, and needed the capability to conduct passive sonar searches at speeds faster than a typical surface group. The resulting hull sonar system was thus highly sensitive, but had a decreased maximum effective range.

The Knox class was initially classified as a destroyer escort and later designated as a frigate. For mid to late Cold War vessels, they were very capable anti-submarine patrol vessels for their size with good anti-surface capabilities, featuring both a dual purpose ASROC-Harpoon launcher and a LAMPS I helicopter. (wikimedia)

In addition to the added anti-submarine mission, the Seasprite performed anti-surface support and anti-ship missile defense roles. In performing these missions, the Seasprite used its search radar to track and identify potentially hostile surface vessels. This allowed the host vessel to build a picture of enemy forces while putting itself in comparatively little direct danger. With this information, any LAMPS I vessel had early warning against potentially hostile surface vessels, and could also use the relayed information to more accurately fire Harpoon and Standard missiles over the horizon, without using its own radar and revealing itself. The extended surveillance range of a LAMPS vessel was pushed beyond 170 miles with the use of the Seasprite.

LAMPS I thoroughly improved the anti-submarine and anti-surface capabilities of much of the US fleet, with the Seasprite itself being an almost perfect off the shelf solution. While there were limitations, like the inability to perform an independent ASW search, the overall benefit of the ship not needing to prosecute sub surface contacts alone or having to reveal itself to perform a radar search in its patrol area was well worth the resources devoted to the Seasprite.

Late Career

Beyond ASW duties, Seasprites also allowed their host vessels to conduct surface surveillance over a much wider area. Here, an SH-2F identifies a natural gas carrier during Operation Desert Shield. (National archives)

By the end of the Cold War, the Seasprite had incorporated a number of improvements. These comprised a number of on board and weapon systems, perhaps most notably the introduction of the Mk 46 Mod 5, or NEARTIP, lightweight torpedo. The new model was designed to counter the latest advancements in Soviet nuclear submarine design, with the torpedo possessing an improved engine to make for a higher speed, an improved sonar transducer to increase the effective detection range of the weapon and add better countermeasure resistance, and had a new guidance and control group. The new weapon entered service in 1979, with kits being produced to convert old stocks to the new standard.

An improved model of the helicopter equipped with T700-GE-401 engines was also developed in 1985, though few were procured, as the Navy sought to increase supplies of the SH-60 Sea Hawk. Some of the improvements from the scaled back Super Seasprite did however make their way into the SH-2F. A number of LAMPS I helicopters during the mid 80s were equipped with FLIR pods for IR searches, IR jammers, chaff and flare dispensers, and an infrared sea mine detection system. Their service during the Gulf War saw them mostly perform ship to ship material and personnel transfers, mine detection, and medical evacuation roles, as Iraq possessed no submarines. Their primary mission in the theater was mine hunting duties, for which they used IR sensors in their search. They were only carried aboard lighter surface combatants during Operation Desert Storm, and weren’t present among the air wings of any of the aircraft carriers during the conflict.

After almost thirty years of service, the SH-2F was withdrawn along with most of the vessels that carried them. Its end was hastened by the withdrawal of the Knox class frigates from service and the sale of most of the short hull OHP frigates to foreign navies. The Navy would fully transition over to the Sikorsky Seahawk, a much larger and more powerful helicopter which carried two torpedoes, a dipping sonar, and incorporated sonobuoy processing capabilities.

Construction and Flight Characteristics

The Kaman SH-2F Seasprite was compact, and while conventional for a modern helicopter, was very advanced for its day. Its fuselage was watertight, possessed forward retractable landing gear, and was equipped with a variety of onboard sensors. While it could not perform waterlandings, its sealed canopy allowed it to float until the helicopter’s crew could be recovered. The pilot sat on the port side of the cockpit and the copilot/tactical coordinator, who operated the weapon systems, was seated starboard. The systems operator sat behind the pilot and operated the sonobuoy dispenser, the magnetic anomaly detector, and radar system. The systems operator lacked the equipment to process the sonobuoy data, which was instead processed aboard the LAMPS I host vessel and sent back via a data link.

An SH-2F instrument panel (wikimedia).

At the nose of the helicopter was the LN-66 surface search radar, designed for detecting both surface vessels and submarine snorkels. On the starboard pylon was the MAD streamer which worked in conjunction with an extendable antenna on the underside of the helicopter. This system worked by measuring the local strength of Earth’s magnetic field, and would spike if it encountered a large magnetic object, or in other words, a submerged submarine. Triggering a readable detection required the aircraft to over fly the contact and was thus typically used to pin the exact position of the submarine while preparing to attack after closing in during the sonobuoy search. The Seasprite carried a mix of AN/SSQ-41A passive and AN/SSQ-47B active sonar sonobuoys. The AN/SSQ-41A omni-directional passive sonobuoys operate at a depth of 60 ft for shallow searches and 300 ft for deep, and have a frequency range of 10 Hz to 20 kHz. Depending on their settings, they lasted between one to eight hours. The SSQ-47B active sonobuoy provided ranging information and operated at either 60 or 800 ft and possessed a maximum endurance of thirty minutes. Sonobuoy data was processed aboard the supporting ship and was used to localize submarine contacts that were otherwise too distant or quiet to be effectively tracked by the ship’s sensors alone. The information provided from the data link allowed the helicopter to detect, classify, and engage subsurface contacts in cooperation with the host vessel.

Re-detecting a submarine at longer ranges from the ship was difficult, as passive sonobuoys laid out in a large search pattern gave little chance of success. The best chances of re-detection on a lost contact was when it was near the surface, transiting, or maneuvering to avoid attacks from other vessels and aircraft. The standard procedure for sub chasing was to head down the azimuth of the ship’s sonar contact and to begin to lay a sonobuoy field to uncover its exact location.

The Systems operator station. To the left is the MAD readout, in the center is a scope for the surface search radar, and on the right is the (shuttered) sonobuoy display. (National archives)

The Seasprite was initially powered by a single General electric T58-GE-8F turboshaft before a second was installed on the UH-2C. These each produced up to 1,350 shp and allowed the SH-2F to travel at a top speed of 152 mph at sea level and allowed the small helicopter to carry up to 2000 lbs worth of equipment in the vertical replenishment role, with a maximum cargo hook capacity of 4000 lbs. To save fuel during emergencies, the helicopter could run on one engine on the way back to the ship. These engines were well regarded and considered very reliable.

The helicopter’s lift was provided by a 44 ft main rotor which used composite blades which were directed with servo operated flaps. These flaps are easily visible on the rotors, each having a wider chord than the rest of the blade. The flap is used to change the angle of attack of the rotor in flight and allows for smooth altitude adjustment. The anti-torque rotor at the rear of the helicopter had its blades increased from three to four going from the C to D model. The Seasprite handled well and was easy to perform a hover in, an important capability when it comes to search and rescue, and transfers to vessels without any landing areas. This was particularly important when landing on Knox class frigates, which both had significant air disturbance aft of the ship, and a very claustrophobic landing area.

In the air rescue role, the copilot would coordinate with divers and rescue crew. The cargo space of the helicopter could fit two stretchers or three seats. For water recovery of personnel, divers were carried aboard and recovered downed airmen through the use of a rescue hoist mounted on the starboard side of the helicopter. Mechanically driven, it had a capacity of 600 lbs.

Throughout the 1980’s, Seasprites were often equipped with a variety of new devices. This aircraft features two ALQ 144 IR jammers for missile defense, chaff and flare dispensers, and a FLIR imager. Crews also often removed the doors from these helicopters for faster entry and exit. (National Archives)

The Seasprite could carry a variety of unguided weapons, but rarely carried anything other than the Mk 46 torpedo, being either the Mod 0, or Mod 5 NEARTIP during the 1980s. On paper, the Seasprite could carry two torpedoes, but in practice, the second equipment position was taken up by an external fuel tank on ASW patrols. Both torpedo types measured 8.5 ft long with a diameter of 12.75 inches. The Mod 0 weighed 568 lbs, and both carried a 95 lb warhead. The Mod 0 possessed a maximum speed of 45 kts, with the NEARTIP being considerably faster. The NEARTIP provided better tracking of faster targets and better countermeasure rejection, having incorporated a new sonar transducer, control and guidance group, and a new engine which switched from solid propellant to liquid monopropellant. Prior to the introduction of the Mod 5, there was little hope for successful attacks against the fastest nuclear submarines of the 1970s. However, in confirming the location of a submarine, its position also became revealed to long range ASW aircraft which could make follow up attacks.

Other weapons included unguided 2.75 inch unguided rockets, and some rare, late examples possessed FLIR optics and could carry AGM-65 Maverick missiles. These weapons, however, were rarely ever carried. Later Seasprites carried a variety of countermeasures including an ALQ-144 tail mounted IR jammer and an ALE-39 flare and chaff dispenser. A considerable number of these helicopters were equipped with infrared jammers and flares during the 1980s.

Conclusion

An SH-2F is being used to evacuate a sailor who received severe burns, necessitating treatment off-vessel. (National Archive)

The Kaman Seasprite can be said to be among the most versatile aircraft ever operated by the US Navy. Entering service as a plane guard, the number of roles it served grew considerably over the years to encompass everything from medical evacuation, to anti-submarine duties. As the core of the LAMPS program for nearly 10 years, it gave US warships a boost in their offensive and defensive qualities against both surface and subsurface opponents.

Specification

SH-2F Seasprite Specification
Engine 2x General Electric T58-GE-8F
Output (maximum) 2300 SHP (2700 SHP)
Maximum Weight 12800 lbs
Empty Weight 8652 lbs
Range for Utility 234 N.MI
Radius of Action for Utility 111 N.MI
Endurance for Utility (ASW) [Ferry] 2 hours (1.9 hours) [2.8 hours]
Standard Armament 1 Mk 46 Mod 0/5 Lightweight torpedo
Crew Pilot, copilot/tactical coordinator, systems operator
Length of fuselage 40.5 ft
Width of fuselage 10 ft
Designation Sub type
HU2K/UH-2A Basic single engine utility helicopter
UH-2B Minor differences in avionics, later made identical to A model
UH-2C First two engine model
H-2 Army project, single engine
HH-2C Combat rescue model, 7.62 side door gun emplacements, M134 rotary gun turret. Two engines.
HH-2D Same as HH-2C but without armament. Used to test ASW equipment and loading. Two engines.
NUH-2C/D Test helicopter, two engines.
YSH-2E Testing helicopter for radar and ASW gear for canceled LAMPS II program
SH-2D Early ASW model
SH-2F Standard LAMPS I helicopter
SH-2G SH-2F with T700 turboshaft engines, improved avionics. Small production run.
Avionics Type
Surface Search Radar LN-66HP
IFF AN/APX-72
Transponder Computer KIT-1A/TSEC
UHF Radio Set AN/ARC-159
Secure Speech KY-28
ICS AN/AIC-14
TACAN AN/ARN-52
Doppler Radar AN/APN-182
Attitude Heading AN/ASN-50
NAV Computer AN/AYK-2
Plotting Board PT-492
UHF Direction Finder AN/ARA-25
OTPI R1047A/A
Radar Altimeter AN/AP-171
RAWS AN/APQ-107
Sonobuoy receiver AN/ARR-52
Acoustic Data Processor AN/ASA-26B
Data Link AN/ASK-22
Magnetic Anomaly Detector AN/ASQ-81
Radar Warning Receiver AN/ALR-54

Profile:

The SH-2F Seasprite was a simple, but excellent conversion of a proven airframe. Installed aboard much of the US surface fleet, it was a potent force multiplier.
During the mid 80’s, the Seasprite fleet received a number of improvements. These included the ALE-39 countermeasure dispenser, the AN/ALQ-144 IR jammer for use against heat seeking missiles, and later FLIR optics.

Gallery:

 

The Knox class’s helicopter facilities were quite claustrophobic, and precluded the use of a larger helicopter. (National Archive)
A forward view of a Seasprite aboard a Spruance class Destroyer. (National Archives)
Despite its small size, the Seasprite could carry a considerable sling load between vessels. (wikimedia).

A Knox class frigate during a visit to La Roche, France with its LAMPS helicopter on deck. Curiously, this ship’s Sea Sparrow launcher has been removed. (Wikimedia)
The colorful MAD streamer. (Wikimedia)
A Seasprite responds to a medical emergency aboard a freighter near a naval exercise. (National Archives)

A Seasprite flies as a plane guard alongside the USS America. An Essex class refit carrier sails in the background. (National Archives)
An SH-2F undergoes checks aboard the USS Iowa during the Northern Wedding naval exercise, 1986. (National Archives)

A small number of combat rescue helicopters were converted to recover airmen from potentially dangerous coastal areas. In practice, the nose mounted gun was typically not retained. (wikimedia)
With its rotors folded, the crew of the USS John Hancock prepare to stow their Seasprite. (National Archives)
A snapshot taken by a Seasprite: Soviet Submarine K-324 and frigate USS McCloy (Knox class) were engaged in mutual surveillance when the submarine’s screw became entangled in the frigate’s towed sonar array. The emergency was responded to by the Soviet oceanic survey ship SSW 506 and the American destroyer USS Peterson. The K-324 was a Victor III class nuclear submarine, this type being the most numerous modern Soviet nuclear submarine of the late Cold War.

Credits: 

  • Article written by Henry H.
  • Edited by  Stan L. and Henry H.
  • Ported by Henry H.
  • Illustrations by Godzilla

Sources

Primary:

Standard Aircraft Characteristics Navy Model SH-2F aircraft. NAVAIR 00-110AH2-8. Commander of the Naval Air systems Command. July 1974.

Andrews, Harold. Sea Sprite. Naval Aviation New 1983 (Feb).

Naval Aviation News 1985 (May-June)

Naval Aviation News 1983 (Jan-Feb & May-Aug)

Department of Defense authorization for appropriations for fiscal year 1982 : hearings before the Committee on Armed Services, United States Senate, Ninety-seventh Congress, first session, on S. 815.

Department of Defense appropriations for 1984 hearings before a subcommittee of the Committee on Appropriations, House of Representatives, Ninety-eighth Congress, first session / Subcommittee on the Department of Defense.

Department of Defense authorization for appropriations for fiscal year 1986 : hearings before the Committee on Armed Services, United States Senate, Ninety-ninth Congress, first session, on S. 674.

Department of Defense authorization for appropriations for fiscal year 1979 : hearings before the Committee on Armed Services, United States Senate, Ninety-fifth Congress, second session, on S. 2571

Department of Defense authorization for appropriations for fiscal year 1980 : hearings before the Committee on Armed Services, United States Senate, Ninety-sixth Congress, first session, on S. 428.

CDR Rausa Rosario. LAMPS MK III. Naval Aviation News 1980 (June).

Defense Department authorization and oversight hearings on H.R. 5167, Department of Defense authorization of appropriations for fiscal year 1985, and oversight of previously authorized programs before the Committee on Armed Services, House of Representatives, Ninety-eighth Congress, second session.

Secondary:

Polmar, Norman. Ships and Aircraft of the U.S. Fleet. Fifteenth Edition. US Naval Institute Press. 1993.

Sikorsky HNS-1 “Hoverfly”. United States Coast Guard.

Stuyvenberg, Luke. Helicopter Turboshafts. University of Colorado at Boulder, Department of Aerospace Engineering. 2015.

Garcia Class Frigate. NAVsource online.

Aero Spacelines PG-2 Princess Guppy

sweden flag USA (1964)
Oversized Cargo Aircraft – None Built

The slightly smaller PG-3 used eight jet engines and kept the wings off the Princess without changing them too drastically. Here it is seen carrying the S-II stage. [allaboutguppys.com]
The Aero Spacelines PG-2 was an oversize cargo aircraft with an extremely large cargo hold, designed to airlift the first and second stages of the Saturn V rocket from their factories to Cape Canaveral, Florida, for final assembly. To save on costs, the aircraft would use components from existing aircraft, and most interestingly would use several key components from the British Saunders Roe Princess flying boat, hence the nickname “Princess Guppy”. Unfortunately, due to opposition from Congress, and the deterioration of the Princess aircraft, none of the type would be built.

NASA and the Transportation Problem

The S-II stage of the Saturn V. [James Vaughan- Flickr]
The race to the moon in the 1960s between the United States and Soviet Union introduced a number of challenges upon the growing aerospace industry. The task at hand was one of the biggest endeavors in human history, requiring manpower, materials, logistics, training, and calculations never used before to achieve such a tremendous goal. In America, the Apollo program was well underway, composed of the Apollo spacecraft and the massive Saturn V rocket it would be launched from. The Saturn V (at this point called the Saturn C-5), was a multistage launch platform with 3 different stages. Logistically, there was a problem with its design. The Saturn V was meant to be assembled and launched from Cape Canaveral in Florida, but the first and second stages of the rocket were assembled in completely different states. The first stage, S-IC, was assembled in New Orleans by Boeing, while the second stage, S-II, was produced on the opposite end of the country in California by North American Aviation. This created a massive challenge regarding transportation. The two stages were massive in size, each having a diameter of 33 ft (10 m). The first stage had a length of 138 ft (42 m) while the second stage had a length of 83 ft (24.9 m). Transporting these two components was a major issue, as almost nothing could quickly move these to Cape Canaveral. This led to NASA deciding to use an aircraft to transport the 1st and 2nd stages. However, this brought on yet another problem. At the time, no aircraft then in service could carry such a large and ungainly cargo, leading several aircraft companies to propose concept aircrafts to complete such a task. Due to the nature of the challenge, the proposals often were unorthodox in their design to accommodate the large load. One company, however, was formed deliberately to fill NASA’s airlifting needs.

An example of Aero Spacelines’ other work, the Super Guppy. These aircraft would transport the 3rd stage of the Saturn V.

Aero Spacelines was formed in 1960 by Jack M. Conroy with NASA as their main customer in mind. Jack, being a former Air Force and commercial pilot, knew of their transportation issue regarding rocket components even before the Saturn V rocket, beginning with their previous multistage rocket designs. He proposed using modified Boeing 377 Stratocruiser airliners with large overhead cargo holds to carry these rocket components from their manufacturers to their assembly points. The first of his “Guppy” designs as they were called, the Pregnant Guppy, first flew in 1962 and was awarded contract work for NASA as an airlifter. The Pregnant Guppy was still too small, however, to carry the large 1st and 2nd stages of the Saturn V, and so a larger design began to be drawn up. An early study was done with an entirely new fuselage using B-36 wings and control surfaces to save on parts. This design would have a cargo hold with a diameter of 40 ft (12.2 m), allowing it to carry both boosters. Little is known of this design outside of this but it would be quickly changed on January 30th, 1964 when John M. Conroy announced Aero Spacelines would design a new oversized load airlifter based on the Saunders Roe Princess.

The Saunders Roe Princess: A Dead Dream Revived.

The Saunders-Roe Princess in flight. The size of the aircraft is evident in this photo. [Tom Wigley – Flickr]
The Saunders Roe Princess was the biggest flying boat design built in Britain, and the biggest all metal flying boat ever built. Originally designed as an innovative transatlantic passenger liner, it would first fly in August of 1952. However the Princess encountered two major issues. The ten Proteus engines used were underpowered, causing performance to suffer. On a more pressing matter, the Princess found itself being quickly outdated as it was developed. With the arrival of the De Havilland Comet, the world’s first jet powered airliner, in the same year as the Princess, it was quickly shown that piston-engine airliners, let alone floatplanes, was a dying breed of travel. Jet aircraft could fly faster, smoother, and further than piston engine airliners, and the Princess couldn’t find buyers because of this change in the market. In addition, the amount of airfields left in Britain after the Second World War nullified the benefits of flying boats and their lack of need for airfields. A single Princess would be built and tested, with two more being completed, but not flown, when the program was ended. The three Princesses were put into storage, cocooned away in hopes that a buyer would eventually come and save them.

Over the years several interested parties would look at the Princesses but no deal ever came to fruition. The three airframes would sit in storage for a decade (1954-1964) when they came to the attention of Jack M. Conroy. Interestingly, this wouldn’t be the first time the US considered acquiring the three aircraft, as the Navy had once proposed to convert the three into flying nuclear-powered test beds, but this plan never progressed past a few models and drawings. At the time, Conroy was still working on his booster carrier concept using B-36 components, but the large design of the Princess gave him an idea. Instead of the B-36, Conroy had the idea of reusing the same parts from the Princess. The plans were quickly reworked and came to be known as the PG-2 Princess Guppy. The PG-2 would reuse the wings from the Princess but had several enhancements. Instead of using the ten Proteus engines, these would be swapped out for six Rolls-Royce Tyne turboprop engines. The Tyne engines were originally planned for the Princess during its development, but the engines weren’t ready and couldn’t be used by the time the Princess was built. Now a decade later, the engine was fully operational and ready. The wing length would also be stretched to 40 feet (16.2 m), and the cargo hold would have a 38 foot (11.6 m) minimum diameter and a length of 100 feet (30.5 m). This reduction in length would no longer allow the aircraft to carry the S-IC booster. Maximum cargo capacity would be up to 200,000Ib (104,600 Kg). The aircraft would be reworked once again later in 1964 as the PG-3. The PG-3 would be reduced in size to some degree. The Princess wings would no longer be lengthened to save on costs, and the Tyne engines would no longer be used on the PG-3, instead they’d be replaced by jet engines. A total of eight jet engines would be used on this design, with 4 pairs of engines being used on B-52H engine pods. Other than wing design, the rear of the aircraft was also changed, with the fuselage not angling upward and instead being more of a straight point.

Aero Spacelines had full intentions of seeing this project through, and eventually a representative of the company was sent to inspect the three Princesses at their storage facility. However, a terrible revelation was discovered upon inspection. At some point, maintenance on the three Princesses in storage was stopped, and so they were left to rot for nearly a decade. Being near the sea and exposed to the elements, the three aircraft had deteriorated to such an extent they would no longer be usable. With this discovery, Aero Spacelines had to unfortunately cancel the project and the three Princesses were scrapped. Work on a large carrier was halted for Aero Spacelines and they focused on their smaller Super Guppy aircraft instead, which carried the 3rd stage of the Saturn V.

Design

The Aero Spacelines PG-2 concept. Note the modified wings and tail empennage off the Saunders-Roe Princess [allaboutguppys.com]
The Aero Spacelines PG-2 was a large oversized cargo aircraft designed to carry the first and second stages of the Saturn V rocket. To do so, it would have a very large fuselage to accommodate the rocket stages. The aircraft would have an all metal fuselage that was 200ft (61 m) in length. The lower section of the fuselage contained the huge cargo bay for the rocket stages. On the original plan this section had a diameter of 40 feet (16.2 m) but was shortened to 38 feet (11.6 m) on the PG-2. Cargo was loaded into the aircraft by means of a ramp. The cargo bay had a large clamshell door in front of the aircraft. Landing gear was divided into six pairs of wheels on the underside. Two pairs of wheels were closer to the front of the aircraft while the remaining four were towards the rear. The cockpit and crew section was located above the cargo bay in the aircraft. The cockpit itself bears a striking resemblance to the cockpit section of the Douglas C-133. A crew of 3 to 4 was expected for operations. Initially, the wings would be reused from the Convair B-36 bomber. The engines for this version were never specified. On the PG-2, it was decided at this point that the wings of the Saunders-Roe Princess would be used over for the B-36’s. The wings would be lengthened an additional 40 feet in total for stabilization. The Princess’ original ten Proteus engines would be replaced with six Rolls Royce Tyne engines to improve performance. The tail section of the PG-2 would also be reused from the Princess.

The PG-2 would be reworked into the PG-3 design. The overall proportions were diminished to save on labor. The specifications of this version are relatively unknown aside from one or two estimates based on promotional images. The cargo bay was to remain the same in length. The cockpit section and most of the fuselage remain unchanged aside from the rear. The rear of the fuselage no longer tapered upward and instead transitioned straight back into a cone shape. The tail section of the aircraft remained in the same location but was now supported by a large support to accommodate the height difference of the rear of the aircraft. The wings of the PG-3 remove the 40ft (16.2 m) extension off of the Princess wings and keep the original length. The six Tyne engines were removed in favor of eight jet engines. These engines would be paired together in four B-52 engine pods on the wings. The jet engine intended for use isn’t stated but it’s likely they were Pratt and Whitney TF33 engines. Promotional art also depicts the PG-3 having wingtip mounted fuel tanks.

Conclusion

The final result of the transportation issue. The S-IC and S-II boosters would be transported via barge from their factories to assebmly. Here the S-II booster is carried by the barge Poseidon. [Wiki]
With the cancellation of the Princess Guppy, Aero Spacelines moved on to other means to assist NASA regarding transportation, however they weren’t the only company to offer an aircraft design to carry the larger rocket stages. Several other companies had offered proposals to NASA for the same function, such as Convair and Fairchild. Many of these designs reused existing aircraft as their base or for parts to save on costs. None of these would come to fruition either. Despite reusing components from existing aircraft, many members of Congress found building a new aircraft for this role unnecessary for the amount of funding it needed. Instead it was decided that the 1st and 2nd stages would continue to be transported to Cape Canaveral via barge. For the 2nd stage, this was a very long journey that involved going through the Panama Canal to reach Florida. Despite being time consuming, this method was one Congress found cost effective. No oversized aircraft proposals would be built aside from Aero Spacelines’ own Super Guppy design, which was used to transport the 3rd stage of the rocket, and one of which is still in service to this day by NASA.

The Princess Guppy was a well researched design using prior knowledge of Aero Spacelines’ Pregnant Guppy. The design would have brought back to life a decade-old dream but unfortunately it was crushed due to negligence. Had it been built, it would be questionable if the aircraft would even be airworthy. The immense size of the fuselage and the small amount of engines in comparison to said size could have prevented the aircraft from even lifting off. Regardless, none of the types were built.

Variants

  • Early Design (PG-1?) – The first design of the booster carrying aircraft reused components of the Convair B-36 bomber. It would have a large cargo hold to carry the oversized load.
  • PG-2 – Second design of the booster carrier. The PG-2 Princess Guppy would use the modified wings and tail components of the Saunders-Roe Princess and would be powered by six Rolls Royce Tyne engines.
  • PG-3 – Reduced size version to lessen the work needed to build the aircraft. It was powered by 8 jet engines in B-52H engine pods. The Princess’ wing returns to its normal size for this version.

Operators

  • United States of America – The Princess Guppy was designed specifically to be used by NASA for the transport of the first and second stages of the Saturn V rocket. None were be built.

Aero-Spacelines PG-2 specifications

Wingspan 259.8 ft / 79.2 m
Length 200 ft / 61 m
Height 86 ft / 26.2 m
Wing Area 6328 ft² / 587.8 m²
Engine 6 x 4,616 hp (3,442 kW) Rolls Royce Tyne RTy.12 turboprop engines
Propeller 6 x De Havilland 4-blade propellers
Powerplant Ratings
Horsepower output Altitude
Take Off 5730 hp Sea Level
Weights
Useful 250,000 lb / 113398 kg
Minimum Flying Weight 180,000 lb / 81646.6 kg
Maximum Take Off 430,000 lb / 195044.7 kg
Maximum Landing 400,000 lb / 181436.9 kg
Crew 3 to 4

Gallery

Artist Concept of the PG-2 by Godzilla

Credits

  • Written by Medicman
  • Edited by Henry H. & Ed J.
  • Illustrations by Godzilla

Sources

  • COX, G. (2019). AMERICAN SECRET PROJECTS 3 : u.s. airlifters since 1962. Place of publication not identified: CRECY PUB.
  • Keeshen, J. & Hess, A. (2013). Secret US proposals of the Cold War : radical concepts in military aircraft. Manchester North Branch, MN: Crécy Publishing Limited,Distributed in the USA by Specialty Press.
  • https://wightaviationmuseum.org.uk/princess-flying-boat/

Northrop’s Early LRI Contenders

USA flag old United States of America (1953)
Long Range Interceptor Proposals [None Built]

Detailed drawing of the N-144, with cutaway section

Born from the Long Range Interceptor program, the first of Northrop’s contenders were three aircraft that had large delta wings and overall similar shapes and designs. The first, the N-126, started as a modified version of Northrop’s F-89D Scorpion fighter but would become its own unique aircraft by 1954. The second, the N-144, was a large four-engine interceptor design that dwarfed current bombers of the time and could carry an impressive arsenal. The third, the N-149, differed the most from its two siblings. It was much smaller and used General Electric engines over Wright engines. The N-144 was the most successful out of the entire program, but would prove to be too costly and a maintenance nightmare if produced. The N-126 and N-149 would also not meet expectations, as did none of the other competitors in the doomed program.

The LRI Competition

At the start of the Cold War, it was realized that if a Third World War would ever happen, defending the mainland United States from airborne threats would be a top priority. ICBMs and nuclear missiles are the go-to threat everyone imagines when they think of the Cold War, but these wouldn’t be operational until the late 1950’s. In the early years, nuclear weapons would be deployed by strategic bombers and these would be the major threat. Intercepting these long range aircraft would be of the utmost importance if the war went hot in the 1950’s. Developing an aircraft able to reach these bombers and destroy them led to the creation of the modern interceptor. Most countries had begun developing an interceptor of their own. At the forefront was the United States Long Range Interceptor program (LRI). This program originated in early 1952, with Major General L.P. Whitten of the Northeast Air Command noticing that a capable aircraft would be able to takeoff and intercept enemy bombers using the warning time of the Semi-Automatic Ground Environment (SAGE) system, which was an integrated defense network of SAM, radar and fighters across the US and Canada, able to intercept enemy bombers well before they were able to reach the United States. Although the idea was put out, no official requirements for the idea came about until December of 1953, when the Air Council put out extremely demanding needs. The aircraft would need to be airborne in two minutes from getting the scramble alert. Maximum speed would be Mach 1.7 with a range of 1,000 nm (1,850 km). Combat ceiling would be 60,000 ft (18,000 m) with a climb rate of 500 ft/min (150 m/min). The aircraft would be minimally armed with forty-eight 2.75 inch rockets, eight GAR-1A Falcon AAMs or three unguided nuclear rockets. This requirement became known as Weapon System WS-202A. Most companies developed submissions, but McDonnell and Northrop had an early start with a long range interceptor design being conceived very early on, well before an official requirement had been requested. Northrop had three aircraft designs that would fit the requirement for WS-202A; the N-126, N-144 and N-149. All three were visually similar to each other and shared concepts and equipment with one another.

Northrop N-126: The Delta Scorpion

Bottom view of the N-126 Delta Scorpion model [US Secret Fighter Projects]
The first of the designs Northrop submitted was the N-126 Delta Scorpion. This aircraft actually began development months before an official requirement was put out. The design was submitted in February of 1953 and was essentially a Northrop F-89D Scorpion modified with a new delta wing design and Wright YJ67 engines. The aircraft received a performance review sometime in 1953 along with McDonnell’s two-seat version of the F-101 Voodoo. Neither design was chosen for production. The N-126 did show promise, as it came close to meeting the very first requirements and it was supported by the Air Defense Command. However, the predicted first flight in twenty-one months was a bit too optimistic and the design was disliked by the United States Air Force Headquarters, as it didn’t exactly meet requirements compared to the F-101 variant. Northrop pushed this early design and adamantly tried to acquire production.

Front quarter view of the N-126 Delta Scorpion model [US Secret Fighter Projects]
They were quick to begin working on an improved design that would be longer and yield better results. It took over fifty concept designs before they found a suitable improvement. The aircraft itself no longer resembled the F-89D Scorpion it got its name from, but the name would stick until the end of the project. This new design was submitted in August of 1954. The N-126 was now much sleeker, with a forty-five degree delta wing and two underwing Wright J67-W-1 engines (Allison J71-A-11 engines were a weaker alternative choice). The delta wings all three projects used provided lower weight than generic straight wings and minimized drag. The trailing edge of the wing would have a split speed brake on the outer surface, an aileron located in the middle and a feature on the inboard section only referred to as an “altitude flap”. For the landing gear, a bicycle configuration with two wheels on each gear would be mounted directly under the aircraft, with a smaller landing gear being placed under the wings.

For armament, the aircraft would use the required eight Falcon AAMs and forty-eight rockets being mounted in a 20 ft weapon bay. Four external hardpoints would allow extra ordnance to be carried, such as bombs or extra missiles. Alternative loadouts included any combination of four AIR-2A unguided nuclear rockets, six Sidewinders, or two Sparrow guided missiles. The N-126 would use the Hughes E-9A fire control system, one of the few remnants carried over from the F-89. The E-9A would be linked to a long-range search radar that would have a range of 100 nm (185 km). For fuel, one large internal tank and two smaller tanks in the wings would hold 4,844 gal (22,025 l). Extra drop tanks could be mounted under the wings and offer an additional 1,600 gal (7,275 lit). For its predicted mission, the N-126 would be able to launch and engage enemy bombers twenty-seven minutes after scramble. Northrop expected a prototype would be ready for a first flight by June of 1957.

Northrop N-144: The Monstrous Interceptor

Color photo of the N-144 model [US Secret Fighter Projects]
The N-144 was the second design Northrop submitted. It was made to offer the best results in regard to the WS-202A requirements. It resembled the N-126 but was much larger and had four J67 engines. The N-144 dwarfed its siblings, competitors, and even several current bombers of the time. With a wingspan of 78 ft and a length of 103 ft, this was no small aircraft. In comparison, the Convair B-58 supersonic bomber had a wingspan of 56 ft and a length of 96 ft (interesting to note, a plan to convert the B-58 into a long range interceptor was proposed).

Its appearance wasn’t the only thing carried over from the N-126. The E-9A fire control system, its accompanying scanner, and its landing gear design (now with four wheels on the main gear) were all reused in the N-144. The N-144 also had a forty-five degree delta wing like the N-126. The N-126 and N-144 would both have their engines on pylons on the wings. This configuration allowed much more powerful engines to be used and a simpler intake system compared to having the engines be built into the body, not to mention the layout being much safer in the event of a fire.

Top down view of the N-144 model. Note the 45 degree delta wing [US Secret Fighter Projects]
The N-144 utilized many features that would directly improve the aerodynamics of the aircraft. The aircraft would have low wing loading which would increase its cruise altitude and improve takeoff and landings. The addition of a horizontal tail, which isn’t often seen in delta wing designs, gave the N-144 improved handling and stability over designs that lacked the horizontal tail (see the Convair F-102 Delta Dagger for example). When the aircraft would be supersonic, the wing would have a chord flap that would retract into the wing to reduce drag. Area ruling was a feature involving tapering the center of the fuselage which would reduce drag while the aircraft was flying at supersonic speeds. Most current delta wing designs utilized area ruling, but none of Northrop’s interceptors surprisingly did. Northrop ruled that the advantages would only affect supersonic flight, and not provide anything useful during subsonic flight. Having no area rule also made the aircraft simpler in design and easier to produce. Northrop’s studies into the delta wing expected to see performance increase as time went on, with more modifications and better engines being used on the N-144 if it went into production. With these expected improvements, Northrop theorized a 14% improvement in top speed and service ceiling.

Frontal view of the massive N-144 model. The size of its engine pods are evident. [US Secret Fighter Projects]
For armament, the N-144 would still utilize the standard eight Falcon AAMs and forty-eight rockets, but could also carry twelve Falcon AAMs, six AIR-2A Genie (Ding Dong) rockets, 452 2.75 in FFAR rockets or 782 2 in (5.1 cm) rockets internally in any order. External hardpoints could also be fixed for carrying bombs or more ordnance. For fuel, a large fuel tank would be in the wings and fuselage and could carry 6,910 gal (31,419 l) of fuel. Given the size of the aircraft, Northrop advertised that it could be used in alternative roles.

Northrop N-149: The Opposite End

Model of the N-149. The additional fuel tanks can be seen. [US Secret Fighter Projects]
The N-149 was the third and final design submitted by Northrop for WS-202A. Submitted in July of 1954, the N-149 was almost the polar opposite of the N-144. Instead of opting for raw power and utilizing four engines, the N-149 was meant to be the smallest option available while still performing just as well as its competitors. In comparison, the N-126 would be 85 ft (25.9 m)long with a wingspan of 62 ft (19 m), while the N-149 would be 70 ft (21.5 m) long with a wingspan of 50 ft (15.5 m). This size decrease would save cost, space and fuel consumption. The N-149 used the same wing layout as the previous entries and would also retain the E-9A fire control system and accompanying radar. Given the advancements of the N-144’s wings, it is likely the N-149 would also benefit from them as well. The N-149 did not use Wright J67 jet engines like the N-126 and N-144, but would instead use General Electric J79 engines. These engines were longer than the J67 but would benefit the aircraft, given its small size, to achieve the required speed and rate of climb. The bicycle landing gear with outer wing gear was once again used, but now with two wheels on each gear like the N-126. The armament for the N-149 was less than its predecessors, but it would make up for weapons in the amount able to be built. Once again, eight Falcon AAMs and forty-eight 2.75in rockets were standard, but alternative armaments would be a single Sparrow AAM, four Sidewinder AAMs, another 105 2.75 in rockets or 270 2 in rockets. Additional armament could be mounted on four external hardpoints like the N-126 and N-144, however, two of these would be taken up by external fuel tanks. These tanks would be 600 gal (2,730 l). The majority of the fuel would be in a large tank that spanned the fuselage and into the wings and would carry 2,050 gal (9,320 l) of fuel. Northrop expected a first flight of the aircraft by the summer of 1957.

The Program Concludes

Detailed drawing of the N-149 with cutaway

Although Northrop is the center of this article, Boeing, Douglas, Lockheed, Martin, McDonnell, North American, Chance-Vought, Grumman and Convair all submitted designs. When the assessment of all the designs was completed, it was concluded that none of the proposals exactly met up the set requirements. The N-144, however, came the closest to meeting the specification. After assessment, the N-144 had a predicted speed of Mach 1.76, a combat ceiling of 58,500 ft (17,800 m) and a combat range of 1,015 nm (1,880 km).

McDonnell’s design came close, as it could go faster and reach the same altitude, but its range was much less compared to the N-144. Materials Command was not too keen of the N-144 and it is obvious why. The cost, production and maintenance of it would be tremendous. Given its four engines, the aircraft would require much more maintenance compared to its two-engine competitors. Producing such a large aircraft would be extremely costly given its size and engine count. The best option for performance would also be the worst option considering its cost.

Its siblings didn’t meet the specifications as well. No reason was put out as to why the N-126 failed the competition, but given the state of the program, it can easily be assumed it didn’t meet either the range, speed, or altitude requirements. The N-149 did have a specified reason for its rejection, though. After taking off at full power and reaching its maximum height, it would only offer 20 minutes of flight, with 5 minutes at full power for combat. Having your aircraft destroy as many bombers before reaching their target is necessary and only 5 minutes wouldn’t be sufficient to fulfill its duty. Ultimately, WS-202A wouldn’t produce any aircraft. The requirements had gone too high, and the companies wouldn’t be able to produce a cost effective aircraft in time that would meet the expected specifications. The program would go on to become the new LRI-X program in October of 1954, and Northrop would be one of three companies tasked with creating a new interceptor, which their Delta-Wing trio would surely influence in a number of ways.

Variants

  • Northrop N-126 (February 1953) – The 1953 N-126 Delta Scorpion was an improvement upon the F-89D Scorpion by having a delta wing and YJ67 engines.
  • Northrop N-126 (1954) – The 1954 version of the N-126 no longer resembled the F-89 but was now longer and more streamlined.
  • Northrop N-144 – The N-144 would be the second design submitted to the LRI competition. It was much larger than the other two submissions and would utilize four engines.
  • Northrop N-149 – The N-149 was the smallest of the three designs and was meant to be the best performing for its size. It looked visually similar to the N-126 but would carry slightly less ordnance and utilize Gen Elec XJ79-GE-1 jet engines over the Wright J67-W-1s.

Operators

  • United States of America – All three designs would have been operated by the United States Air Force had they been constructed.

Northrop N-126 Delta Scorpion (1954) Specifications

Wingspan 62 ft 3 in / 19 m
Length 85 ft / 25.9 m
Wing Area 1,050 ft² / 97.7 m²
Engine 2x 13,200 Ibs ( 58.7 kN ) Wright J67-W-1 Jet engines
Weights 75,830 lbs / 34,400 kg (Gross)
Fuel Storage 4,844 gal / 22,025 l
Maximum Speed 1,183 mph / 1,903 km/h at 35,000 ft / 10,700 m
Cruising Speed 793 mph / 1,276 kmh
Range 800 nm / 1,500 km
Climb Rate 2.45 minutes to 40,000 ft / 12,000 m
Maximum Service Ceiling 59,600 ft / 18,000 m (Point Interception Role)

56,200 ft / 17,000 m (Area Interception Role)

Crew 1 Pilot

1 Radar Operator

Main Proposed Armament
  • 8x GAR-1 Falcon AAM
  • 48 2.75in (7 cm) FFAR
Alternative Armament Loadouts
  • 4x Ding Dong Unguided Nuclear Rockets
  • 6x Sidewinder AAMs
  • 2x Sparrow AAMs
  • 1x 1,640 lbs (744 kg) bomb

Northrop N-144 Specifications

Wingspan 78 ft 10 in / 24 m
Length 103 ft 6 in / 31.5 m
Wing Area 1,700 ft² / 158.1 m²
Engine 4x 13,200 Ibs ( 58.7kN ) Wright J67-W-1 Jet engines
Weights 113,700 lbs / 51,500 kg (Gross)

91,600 Ibs / 41,550 kg (Combat)

Fuel Storage 6,910 gal / 31,420 l

44,940 Ib / 20,390 kg

Maximum Speed (Mach 2.04) 1560 mph / 2520 km/h at 34,000 ft / 10,000 m
Cruising Speed (Mach 1.06) 810 mph / 1300 km/h
Range 1,015 nm / 1,880 km
Climb Rate 1.9 minutes to 40,000 ft / 12,000 m
Maximum Service Ceiling 63,000 ft / 19,202 m (Point Interception Role)

60,000 ft / 18,288 m (Area Interception Role)

Crew 1 Pilot

1 Radar Operator

Main Proposed Armament
  • 8x GAR-1 Falcon AAM
  • 48 2.75in (7 cm) FFAR
Alternative Armament Loadouts Internal Storage

  • 12x Falcon AAM
  • 6x AIR-2 Genie (Ding Dong) Missiles
  • 452 2.75 in FFAR
  • 782 2in (5.1cm) Rockets

External Hardpoints

  • Unknown type of bombs mounted on 4 hardpoints.

Northrop N-149 Specifications

Wingspan 50 ft 10 in / 15.5 m
Length 70ft 6 in /21.5 m
Wing Area 700 ft² / 65.1 m²
Engine 2x 9,300 Ibs ( 41.3 kN ) Gen Elec XJ79-GE-1 Jet engines
Weight 43,400 Ibs / 19,700 kg
Fuel Storage 2,050 gal / 9,320 lit

13,310 Ibs / 19,690kg

Maximum Speed (Mach 1.51) 1160 mph / 1860 km/h at 35,000 ft / 10,700 m
Cruising Speed (Mach 1) 770 mph / 1230 km/h
Range 770 nm / 1,430 km
Climb Rate 3.1 minutes to 40,000 ft / 12,000 m
Maximum Service Ceiling 55,700 ft / 17,000 m (Point Interception Role)

52,800 ft / 16,000 m (Area Interception Role)

Crew 1 Pilot

1 Radar Operator

Main Proposed Armament
  • 8x GAR-1 Falcon AAM
  • 48 2.75in (7 cm) FFAR
Alternative Armament Loadouts Internal Storage

  • 1x Sparrow II AAM
  • 4x Sidewinder AAMs
  • 105x 2.75in (7 cm) rockets (original 48 on top of this)
  • 270x 2in (5.1 cm) rockets

External Hardpoints

  • 4x Hardpoints for additional weapons (2 are used for fuel tanks)

Gallery

Northrop N-126 – Artist Impression of the Delta Scorpion in USAF Prototype Stage
Northrop N-144 – Artist Impression of the N-144 the in Late Prototype Stage
Northrop N-149 – Artist Impression of the N-149 in service with the 171 Fighter Interceptor Squadron, Michigan, circa 1960s

 

3-Way drawing of the N-126 Delta Scorpion [US Secret Fighter Projects]
3-Way drawing of the N-149 [US Secret Fighter Projects]
Underside quarter view of the N-126 model [US Secret Fighter Projects]

3 view drawing of the N-126 Delta Scorpion
A photo of the N-126 Delta Scorpion in wind tunnel testing

3-Way drawing of the N-149 [US Secret Fighter Projects]
Colored photo of the N-149 model. Note the tail has been slightly damaged. [US Secret Fighter Projects]
Rear view of the N-149 model. Damage to the tail is evident here. [American Secret Projects: Fighters & Interceptors, 1945-1978]
Credits

Operation Plumbbob – Pascal B Cap

USA flag old United States of America (1957)
Underground Nuclear Test Shaft Cap – 1 Built

This photo depicts the smoke after the detonation of Ranier, an underground nuclear test very similar to Pascal B

The brainchild of one ambitious American astrophysicist during the course of U.S. nuclear tests yielded the first manmade object in Earth’s orbit. The four foot round steel cap was launched into orbit in late August 1957 by the United States, beating the USSR’s Sputnik 1 to orbit by one month and nine days, scoring a major victory in the space race for the Americans. This feat has gone largely unrecognized by most historians.

History

Dr. Robert R. Brownlee

During Operation Plumbbob, which was a series of nuclear tests performed by the United States in 1957, Dr. Robert Brownlee was tasked with determining methods for containing nuclear blasts underground. Initially working from a detonation performed at the bottom of an open shaft, and progressively adding additional ‘plugs’ of concrete to ‘tamp’ the explosion.

The first empty shaft test was called Pascal A, and performed on July 26, 1957. It’s significance was characterized by the fact that it was the first contained underground nuclear test ever performed. The bomb was placed at the bottom of a shaft of about 500 feet in depth, around 3 feet in diameter. The blast yield was much greater than anticipated, estimated at around 55 tons which caused quite a stir at the test site when it was detonated. A concrete collimator with a thickness of five feet was lowered about halfway down the shaft with a detector installed on top. The concrete and detector were presumably vaporized in the explosion, which occured at night and caused a “big blue glow in the sky,” according to Test Director Robert Campbell.

Pascal B

Nevada Test Site Entrance

The next test, Pascal B, attempted to measure the effect of installing a concrete plug just above the bomb, still deep at the bottom of a 500 foot shaft, with a steel cap installed at the end, where the shaft met the surface. The concrete plug, also serving as a collimator for test instruments as in Pascal A, was placed above the bomb. The plug was estimated to have weighed 2 tons.

The shaft diameter for Pascal B was 4 feet in diameter, with a round solid steel cap, 4 inches thick welded to the top. The weight of the cap was estimated to be 2,000 lb (900 kg). Dr. Brownlee designed his calculations to estimate the time and measurements of the nuclear blast’s shockwave in meeting the cap. The estimated time for the shockwave’s arrival was 31 milliseconds. It was anticipated that the pressure and temperature would launch the cap away from the shaft at an extremely high velocity, although this would not necessarily be directly a result of the explosion, since the cap was located too far from the bomb at the bottom of the shaft. Rather, the vaporization and resulting superheated gas of the 2 ton concrete collimator plug placed above the bomb would actually turn the shaft into a ‘giant gun.’ The cap was estimated to achieve a velocity six times the escape velocity of the Earth. A high speed camera was installed nearby with the hopes of capturing the cap’s departure, to hopefully obtain a calculation of the cap’s speed as it left the shaft.

At the Nevada Test Site on August 27, 1957 at 3:35 PM local time, Pascal B was detonated with a yield of 300 tons. The fireball reached into the blue Nevada sky, launching the cap as expected. The high speed camera recorded the cap above the hole in only one frame of the resulting film. The anticipated velocity values combined with the framerate of the camera did not yield any specifically useful measurements, leading Dr. Brownlee to sum up the speed of the cap as “going like a bat!” The original calculation of six times the escape velocity of the Earth of 41.75 miles per second (67.2 km/sec) seemed to have been approximately correct. Other calculations by Carey Sublette that attempt to estimate the expanding gas of the vaporized concrete collimator indicate a similar figure of around five times the escape velocity.

First Manmade Object in Earth Orbit

Contemporary satellite photo of the test site, NTS U3d

Whether or not the cap actually made it to space is still a topic of debate. No trace of the cap was ever found anywhere near the test site. Some say it would have been vaporized in the same manner as a meteorite burning up upon entry into Earth’s atmosphere. Still others theorize that the object may have made it into Earth’s orbit. For the purposes of this article, it is assumed that the cap made it into Earth’s orbit.

The cap would not be the first manmade object in space. That honor belongs to a V-2 rocket launch in Nazi Germany on October 3, 1942, which crossed the Kármán line which is considered to be the boundary of space at an altitude of 100 km (62 miles).

Aside from the Pascal B cap, the most generally agreed upon first manmade object in Earth’s orbit is Sputnik 1, launched on October 4, 1957. If the cap in fact achieved orbit, it would have beaten Sputnik by 1 month and 9 days. This fact has yet to be widely recognized, with most people and historians believing that the USSR achieved the first object in orbit.

Design

3 View Drawing of the Cap by Ed Jackson

The steel cap round, 4 inches thick, was welded to the end of the round metal test shaft, 4 feet in diameter. The cap was presumably not painted or covered with any sort of coating. More than likely the cap was machined locally along with the other significant large scale industrial milling, machining, and fabrication to facilitate the testing operations in support of Operation Plumbbob.

The cap has yet to be found in orbit by NASA, however its exact position still may yet be discovered. At only 4 feet in diameter, dark in color, and at an unknown orbital position, it is difficult to estimate its potential location.

Operators

  • United States – Originally launched from the Nevada Test Site in 1957, the Pascal B Cap remains in service in Earth’s orbit despite its unknown location.

Pascal B Cap Specifications

Diameter 4 ft / 1.22 m
Thickness 4 in / 10.16 cm
Initial Propellant 64.6 lb Plutonium Pit Nuclear Bomb with PBX 9401 and 9404 explosives
Weight 2,000 lb / 907 kg [estimated]
Climb Rate 41.75 miles per second (67.2 km/sec) [estimated]
Maximum Speed 150,300 mph / 241,884 kmh
Range ∞ mi / ∞ km
Maximum Orbital Altitude 574,147 ft / 924,000 km [estimated]
Crew Unmanned
Armament
  • Ramming Impact Capability
  • Sharp Edges
  • High Velocity Steel Fragment & Debris

Gallery

Artist conception of the current state of the cap in orbit by Ed Jackson

Sources

North American F-86A Sabre

USA flag old United States of America (1947)
Jet Fighter – 554 Built

F-86A-1-NA Sabre 47-630 in flight (North American Aviation)

The iconic F-86A got its first official production underway with the A series in 1947, with the initial examples fulfilling many testing duties, followed by a larger second production batch for active service. The development of these first Sabres would address many teething problems with the aircraft’s engines, speed brakes, and weaponry.  The A models, alongside many other first generation American jet aircraft would go on to see a few short years of service in the Korean theatre as well as defense of the United States before being eclipsed by the relatively rapid development of more advanced jet designs.

History

The P-86A was the first production version of the Sabre. North American had received an order for 33 production P-86As on November 20, 1946, even before the first XF-86 prototype had flown. The P-86A was outwardly quite similar to the XP-86, with external changes being very slight. About the only noticeable external difference was that the pitot tube was moved from the upper vertical fin to a position inside the air intact duct.

Major Richard L. Johnson, USAF with F-86A-1-NA Sabre 47-611 and others at Muroc AFB, 15 September 1948. (F-86 Sabre, by Maurice Allward)

The first production block consisted of 33 P-86A-1-NAs, ordered on October 16, 1947. These were known as NA-151 on North American company records. Serials were 47-605 through 47-637. Since there were officially no YP-86 service test aircraft, this initial production block effectively served as such.

The first production P-86A-1-NA (serial number 47-605) flew for the first time on May 20, 1948. The first and second production machines were accepted by the USAF on May 28, 1948, although they both remained at Inglewood on bailment to North American for production development work. Aircraft no. 47-605 was not actually sent to an Air Force base until April 29, 1950. It remained at WPAFB until May of 1952, when it was retired to storage at the Griffiss Air Depot.

In June of 1948, the P-86 was redesignated F-86 when the P-for-pursuit category was replaced by F-for-fighter

By March of 1949 the last F-86A-1-NA (47-637) had been delivered. Most of the 33 F-86A-1-NAs built were used for various tests and evaluations, and none actually entered squadron service.

The first production block to enter squadron service was actually the second production batch, 188 of which were ordered on February 23, 1949. They were assigned the designation of F-86A-5-NA by the USAF, but continued to be carried as NA-151 on company records. Serials were 48-129 to 48-316. These were powered by the J47-GE-7 jet engine. Deliveries began in March of 1949 and were completed in September of 1949.

A contract for 333 additional F-86As was received on May 29, 1948, and the final contract was approved on February 23, 1949. These aircraft were assigned a new designation of NA-161 on North American company records, but continued to be designated F-86A-5-NA in USAF records. Their serials were 49-1007 to 49-1229. These were powered by the General Electric J47-GE-13 engine which offered 5200 pounds of static thrust. The cockpit wiring was simplified. New 120-gallon drop tanks, developed specifically for the F-86, were introduced during this production run. Deliveries commenced in October of 1949 and were completed by December of 1950. The 282nd F-86A aircraft had a redesigned wing trailing edge with shorter chord aileron and greater elevator boost. Deliveries commenced October 1949 and ended in December 1950.

First Deployment

The first USAF combat organization to receive the F-86A was the First Fighter Group based at March AFB in California, with the famous “Hat in the Ring” 94th Squadron being the first to take delivery when they traded in their F-80s for the F-86A-5-NA during February of 1949. The 27th and 71st Squadrons were equipped with F-86A-5-NAs next, and by the end of May of 1949 the group had 83 F-86As on strength. This group was charged with the aerial defense of the Los Angeles area, which, coincidentally, is where the North American Aviation factory was located. Next to get the F-86 the the 4th Fighter Group based at Langley AFB, charged with the defense of Washington, D.C, and then the 81st Fighter Group, based at Kirtland AFT and charged with the defense of the nuclear bomb facilities at Alamogordo, New Mexico. Next came the 33rd Fighter Group based at Otis AFB in Massachusetts, charged with defending the northeastern approaches into the USA. In January of 1950, all air defense units were redesignated as Fighter Interceptor Groups (FIGs) or Fighter Interceptor Wings (FIWs) as a part of the Air Defense Command.

Origin of the “Sabre” Name

In February of 1949, there was a contest held by the First Fighter Group to choose a name for their new fighter. The name *Sabre* was selected, and was made official on March 4, 1949.

Reserves

The first Sabres that went to Reserve units were assigned to the 116th Fighter Interceptor Squadron of the Air National Guard, which received its first F-86As on December 22, 1950.

The following Wings were issued with the F-86A:

  • 1st Fighter Interceptor Wing (27th, 75st and 94th Squadrons)
  • 4th Fighter Interceptor Wing (334th, 335th, 336th Squadrons)
  • 33rd Fighter Interceptor Wing (58th, 59th and 60th Squadrons)
  • 56th Fighter Interceptor Wing (61st, 62nd, 63rd Squadrons)
  • 81st Fighter Interceptor Wing (78th, 89st, 92nd Squadron)

The F-86A was replaced in active USAF service by the F-86E beginning in the autumn of 1951. As F-86As left active USAF service, they were refurbished, reconditioned and transferred to Air National Guard units in the United States. The first ANG units to get the F-86A were the 198th Squadron in Puerto Rico, the 115th and 195th Squadrons at Van Nuys, California, the 196th at Ontario, and the 197th at Phoenix, Arizona.

Record Breaker

In the summer of 1948, the world’s air speed record was 650.796 mph, set by the Navy’s Douglas D-558-1 Skystreak research aircraft on August 25, 1947. Like the record-setting Lockheed P-80R before it, the Skystreak was a “one-off” souped-up aircraft specialized for high speed flight. The USAF thought that now would be a good time to show off its new fighter by using a stock, fully-equipped production model of the F-86A to break the world’s air speed record.

Major Richard L. Johnson on the day of his record-breaking flight, September 15th, 1948 (

To get the maximum impact, the Air Force decided to make the attempt on the speed record in the full glare of publicity, before a crowd of 80,000 spectators at the 1948 National Air Races in Cleveland, Ohio. The fourth production F-86A-1-NA (serial number 47-608, the cold weather test aircraft) was selected to make the record attempt, and Major Robert L. Johnson was to be the pilot. According to Federation Aeronautique Internationale (FAI) rules, a 3km (1.86 mile) course had to be covered twice in each direction (to compensate for wind) in one continuous flight. At that time, the record runs had to be made at extremely low altitudes (below 165 feet) to enable precise timing with cameras to be made.

On September 5, 1948, Major Johnson was ready to go and flew his F-86A-1-NA serial number 47-708 on six low-level passes over the course in front of the crowd at Cleveland. Unfortunately, timing difficulties prevented three of these runs from being clocked accurately. In addition, interference caused by other aircraft wandering into the F-86A’s flight pattern at the wrong time prevented some of the other runs from being made at maximum speed. Even though the average of the three runs that were timed was 669.480 mph, the record was not recognized as being official by the FAI.

Further attempts to set an official record at Cleveland were frustrated by bad weather and by excessively turbulent air. Major Johnson then decided to move his record-setting effort out to Muroc Dry Lake (later renamed Edwards AFB), where the weather was more predictable and the air less turbulent. On September 15, 1948, Major Johnson finally succeeded in setting an official record of 670.981 mph by flying a different F-86A-1-NA (serial number 47-611, the armaments test aircraft) four times over a 1.86-mile course at altitudes between 75 and 125 feet.

Design

F-86A-1 47-611 Conducting a Static 5-inch HVAR Rocket Firing Test (U.S. Air Force Photo)

The P-86A incorporated as standard some of the changes first tested on the third XP-86 prototype. The front-opening speed brakes on the sides of the rear fuselage were replaced by rear-opening brakes, and the underside speed brake was deleted. However, the most important difference between the P-68A and the three XP-86 prototypes was the introduction of the 4850 lb.s.t. General Electric J47-GE-1 (TG-190) in place of the 4000 lb.s.t. J35. The two engines had a similar size, the J47 differing from the J35 primarily in having a twelfth compressor stage.

The F-86A-1-NA fighters could be recognized by their curved windshields and the flush-fitting electrically-operated gun muzzle doors that maintained the smooth surface of the nose. These muzzle doors opened automatically when the trigger was pressed to fire the guns, and closed automatically after each burst.

The cockpit of the F-86A remained almost the same as that of the XP-86, although certain military equipment was provided, such as an AN/ARC-3 VHF radio, an AN/ARN-6 radio compass, and an AN/APX-6 IFF radar identification set. The IFF set was equipped with a destructor which was automatically activated by impact during a crash or which could be manually activated by the pilot in an emergency. This was intended to prevent the codes stored in the device from being compromised by capture by the enemy. The F-86A was provided with a type T-4E-1 ejection seat, with a manually-jettisoned canopy.

The F-86A-1-NA’s empty weight was up to 10,077 pounds as compared to the prototype’s 9730 pounds, but the higher thrust of the J-47 engine increased the speed to 673 mph at sea level, which made the F-86A-1-NA almost 75 mph faster than the XP-86. Service ceiling rose from 41,200 feet to 46,000 feet. The initial climb rate was almost twice that of the XP-86.

In the autumn of 1948, problems with the J-47-GE-1 engine of the early F-86As forced a momentary halt to F-86 production. It was followed by a few J47-GE-3s, and in December the J47-GE-7 became available, which offered 5340 lb.s.t. and full production resumed.

A close up of the early A models’ retractable gunport covers. (Julien of Britmodeller)

The F-86A-5-NA had a V-shaped armored windscreen which replaced the curved windscreen of the F-86A-1-NA. The A-5 would dispense with the gun doors at some point in its production in the interest of maintenance simplicity, although many A-5 examples can be seen with gun doors, many of them with the doors permanently open. A jettisonable cockpit canopy was introduced. The A-5 introduced underwing pylons capable of carrying a variety of bombs (500 and 1000-pounders) or underwing fuel tanks of up to 206 gallons in capacity. A heating system was provided for the gun compartments, and stainless steel oil tanks and lines were provided for better fire resistance.

In May of 1949, beginning with the 100th F-86A aircraft, an improved canopy defrosting system was installed and a special coating was applied to the nose intake duct to prevent rain erosion. Earlier airframes were retrofitted to include these changes. The 116th F-86A was provided with a new wing slat mechanism which eliminated the lock and provided a fully automatic operation.

Gun Sight & Radar

The P-86A was equipped with the armament first tested on the third XP-86 six 0.50-inch machine guns in the nose, three on each side of the pilot’s cockpit. The guns had a rate of fire of 1100 rounds per minute. Each gun was fed by an ammunition canister in the lower fuselage holding up to 300 rounds of ammunition. The ammunition bay door could be opened up to double as the first step for pilot entry into the cockpit. The P-86A had two underwing hardpoints for weapons carriage. They could carry either a pair of 206.5 US-gallon drop tanks or a pair of 1000-lb bombs. Four zero-length stub rocket launchers could be installed underneath each wing to fire the 5-inch HVAR rocket, which could be carried in pairs on each launcher.

An innovation introduced with the NA-161 production batch was a new type of gun aiming system. All earlier F-86As had been equipped at the factory with Sperry Mark 18 optical lead computing gunsight, which was quite similar to the type of gunsight used on American fighter aircraft in the latter parts of World War 2. When the pilot identified his target, he set the span scale selector lever to correspond to the wingspan of the enemy aircraft he was chasing. He then aimed his fighter so that the target appeared within a circle of six diamond images on the reflector. Next, he rotated the range control unit until the diameter of the circle was the same as the size of the target. When the target was properly framed, the sight automatically computed the required lead and the guns could be fired.

Beginning with the first NA-161 aircraft (49-1007), the A-1B GBR sight and AN/APG-5C ranging radar were provided as factory-installed equipment. This new equipment was designed to automatically measure the range and automatically calculate the appropriate lead before the guns were fired, relieving the pilot of the cumbersome task of having to manually adjust an optical sight in order to determine the range to the target. When activated, the system automatically locked onto and tracked the target. The sight image determined by the A-1B was projected onto the armored glass of the windscreen, and the illumination of a radar target indicator light on the sight indicated time to track target continuously for one second before firing. This system could be used for rocket or bomb aiming as well as for guns.

In the last 24 F-86A-5-NAs that were built, the A-1B GPR sight and AN/APG-5C ranging radar were replaced by the A-1CM sight that was coupled with an AN/APG-30 radar scanner installed in the upper lip of the nose intake underneath a dark-colored dielectric covering. The APG-30 radar was a better unit than the AN/APG-5C, with a sweep range from 150 to 3000 yards. The A-1CM sight and the APG-30 ranging radar were both retrofitted to earlier A-5s during in-field modifications. These planes were redesignated F-86A-7-NA. However, some F-86A-5-NAs had the new A-1CM GBR sight combined with the older AN/APG-5C radar. These were redesignated F-86A-6-NA.

Engines

Some consideration given to replacing the J47 engine with the improved J35-A-17 that was used in the F-84E. This engine was tested in the first XP-86. Flight tests between November 28, 1949 and March 1951 indicated that the performance remained much the same as that of the F-86A-1-NA but with a slightly better range. However, the improvement was not considered significant enough to warrant changing production models.

Some F-86As were re-engined with the J47-GE-13 engine, rated at 5450 lb.s.t., but their designation did not change.

All F-86As were initially delivered with the pitot head located inside the air intake duct. It was found in practice that false airspeed readings could be obtained due to the increased airflow within the intake duct, so North American decided to move the pitot head to the tip of a short boom that extended from the leading edge of the starboard wingtip. All F-86As were later retrofitted with the wingtip boom when went through IRAN (Inspect and Repair as Necessary). However, the pitot tube in the intake was never designed to provide airspeed input to the pilot, and the pitot tube in the intake was still there and was used to provide input for the engine.

Fuel

Internal fuel capacity of the F-86A was 435 gallons, carried in four self-sealing tanks. Two of the tanks were in the lower part of the fuselage, one of them being wrapped around the intake duct just ahead of the engine and the other being wrapped around the engine itself. The other two fuel tanks were in the wing roots. Usually the F-86A carried two 120-gallon drop tanks, although 206.5 gallon tanks could be fitted for ferry purposes.

Weapons

Ground attack weapons could be installed in place of the jettisonable underwing fuel tanks. Choices include a pair of 100, 500 or 1000-pound bombs, 750-pound napalm tanks, or 500 pound fragmentation clusters. Alternatively, eight removable zero-rail rocket launchers could be installed. These mounted sixteen 5-inch rockets. When external armament was fitted in place of the drop tanks, combat radius was reduced from 330 to 50 miles, which was not a very useful distance.

F-86A in Korea

Even though the initial skirmishes with MiGs in Korea had demonstrated that their pilots lacked experience and an aggressive approach, the MiG threat was very real and threw the USAF into a near panic. The USAF had nothing in Korea that could provide an effective counter if the MiG-15s were to intervene in large numbers.

In order to counter the MiG threat, on November 8 the 4th Fighter Interceptor Wing, which consisted of the 334th 335th, and 336th Squadrons, based at Wilmington, Delaware and equipped with the F-86A Sabre was ordered to Korea. Most of their pilots were seasoned veterans of World War 2 and they had shot down over 1000 Germans during that conflict. Prior to flying to the West Coast, the 4th FIG exchanged their older ’48 model F-86As for some of the best “low-time” F-86As taken from other Sabre units. The 334th and 335th FIS flew to San Diego and their planes were loaded aboard a Navy escort carrier. The 336th FIS went to San Francisco and was loaded aboard a tanker. Their F-86A aircraft arrived in Japan in mid-December. The aircraft were then unloaded and flown to Kimpo airfield in Korea.

However, before any of these Sabres could reach the front, on November 26, 1950, Chinese armies intervened with devastating force in Korea, breaking through the UN lines and throwing them back in utter confusion. The MiGs did not provide any effective support for this invasion, being unable to establish any effective intervention below a narrow strip up near the Yalu. The MiG pilots were relatively inexperienced and were poor marksmen. They would seldom risk more than one pass at their targets before they would dart back across the Yalu. Had the MiGs been able to establish and hold air superiority over the battle area, the UN forces may well have been thrown entirely out of Korea.

The first advanced detachment of 336th FIS F-86As arrived at Kimpo airfield south of Seoul on December 15. The first Sabre mission took place on December 17. It was an armed reconnaissance of the region just south of the Yalu. Lt. Col. Bruce H. Hinton, commander of the 336th Squadron, succeeding in shooting down one MiG-15 out of a flight of four, to score first blood for the Sabre. The rest of the MiGs fled back across the Yalu. On December 19, Col. Hinton led another four-plane flight up to the Yalu, where his flight met six MiGs who flew through his formation without firing a shot before dashing back across the Yalu. On December 22, the MiGs managed to shoot down a single Sabre out of a flight of eight without loss to themselves, but later that day the Sabres got their revenge by destroying six MiGs out a flight of 15. This loss spooked the MiG pilots, and they avoided combat for the rest of the month.

During December, the 4th Wing had flown 234 sorties, clashed with the enemy 76 times, scored eight victories, and lost one aircraft.

By the end of 1950, Chinese armies had driven UN forces out of North Korea and had begun to invade the South. The Sabres were forced to leave Kimpo and return to Japan which put them out of range of the action up at the Yalu.

Even though the Yalu was now out of range, on January 14, an F-86A detachment appeared at Taegu to participate as fighter bombers to try to halt the Chinese advance. The F-86A was not very successful in the fighter-bomber role, being judged much less effective than slower types such as the F-80 and the F-84. When carrying underwing ordinance, the F-86A’s range and endurance were much too low, and it could not carry a sufficiently large offensive load to make it a really effective fighter bomber. In these attacks, the underwing armament was usually limited to only a pair of 5-inch rockets.

Eventually, the Chinese advance ground to a halt due to extended supply lines and the relentless UN air attacks. The Chinese advance was halted by the end of January, and the UN forces began pushing them back. Kimpo airfield was recovered on February 10. The halting of the Chinese advance can be blamed largely on the inability of the MiGs to provide any effective support for the Chinese attack. Not only had no Chinese bombers appeared to attack UN troops, but no MiGs had flown south of the Yalu region to provide any air support.

The Chinese apparently did have plans for a major spring offensive to complete the task of driving the UN out of Korea. This plan was to be based on the construction of a series of North Korean air bases and for Chinese MiGs to use these bases as forward landing strips to provide air superiority over the North, preventing UN aircraft from interfering with the advance.

In early March, the MiGs began to become more active in support of this offensive, On March 1, MiGs jumped a formation of nine B-29s and severely damaged three of them. Fortunately, by this time the UN base at Suwon was now ready, and the Sabres were now able to return to Korea and reenter the fray over the Yalu. The Sabres of the 334th Squadron began their first Yalu patrols on March 6th, and the rest of the squadron moved in four days later. At the same time, the 336th Squadron moved to Taegu from Japan, so that they could stage Sabres through Suwon. The 4th Wing’s other squadron, the 335th, stayed in Japan until May 1.

MiG Alley

The strip of airspace in western Korea just south of the Yalu soon became known as “MiG Alley” to the Sabre pilots. The Sabres would arrive for their 25-minute patrols in five minute intervals. The MiGs would usually cruise back and forth at high altitude on the other side of the Yalu, looking for an opportune time to intervene. Very often they would remain on the north side of the river, tantalizingly out of reach. When the MiGs did choose to enter battle, the Sabres would usually have only a fleeting chance to fire at the enemy before the MiGs broke off and escaped back across the Yalu. The MiGs had the advantage of being able to choose the time and place of the battle. The MiG-15 had a better high-altitude performance than the F-86A. The MiG had a higher combat ceiling, a higher climb rate, and was faster at higher altitudes than the F-86A. Its superior high-altitude performance enabled the MiG to break off combat at will. Despite these handicaps, the F-86A pilots were far more experienced than their Chinese opponents and they were better marksmen. The Sabre was a more stable gun platform and had fewer high-speed instabilities than did the MiG-15. In addition, the F-86A was faster than the MiG-15 at lower altitudes, and an effective strategy was for the Sabre to force the battle down to lower altitudes where it had the advantage.

In April of 1951, the MiGs got a little bolder, and they would often make attempts to intercept B-29 formations that were attacking targets in the Sinuiju area up near the Yalu. The biggest air battle of that spring took place on April 12, when a formation of 39 B-29s escorted by F-84Es and F-86As were attacked by over 70 MiGs. Three B-29s were lost, whereas 14 MiGs were claimed destroyed, four by the escorting Sabres and ten by B-29 gunners.

On May 20, 1951, F-86A pilot Captain James Jabara became the world’s first jet ace when he shot down a pair of MiGs to bring his total to six.

No F-86As were lost in action during the first five months of 1951, and they flew 3550 sorties and scored 22 victories. Most of the attrition was caused by accidents rather than by losses in actual combat.

In June of 1951, the MiGs began to show more aggressive behavior, and their pilots began to get somewhat better. In air battles on June 17th, 18th, and 19th, six MiGs were destroyed but two Sabres were lost. Another Sabre was lost on June 11 when the 4th Wing covering an F-80 attack on the Sinuiju airfield shot down two more MiGs.

As the first year of the Korean War came to an end, it was apparent that the Sabre had been instrumental in frustrating the MiG-15’s bid for air superiority. Without control of the air, the Red Chinese were unable to establish their series of air bases and they were not able to carry out effective air support of their spring offensive, and the Korean War settled down to a stalemate on the ground.

The more-advanced F-86E began to enter action in Korea with the 4th Wing in July of 1951, replacing that unit’s F-86As on a one-by-one basis. The conversion to the F-86E was rather slow, and the last F-86A was not replaced until July of 1952.

Operators

  • U.S. Air Force – The U.S. utilized the F-86A extensively for the air defense of the Continental United States, while also seeing action in Korea in MiG Alley.

North American F-86A-5-NA Specifications

Wingspan 37 ft 1.5 in / 11.32 m
Length 37 ft 6.5 in / 11.44 m
Height 14 ft 9 in / 4.5 m
Wing Area 287.9 ft² / 26.8 m²
Engine 1x General Electric J47-GE-13 Turbojet Engine

5200 lbst

Weights
Empty 10,093 lb / 4,578 kg
Maximum Take Off 14,108 lb / 6,399 kg
Combat 13,791 lb / 6,255 kg
Climb Rate
Rate of Climb at Sea Level 7,470 ft / 2,277 m per minute
Time to 40,000 ft / 12,192 m 10.4 minutes
Maximum Speed
Sea Level 679 mph / 1,092 kmh
35,000 ft / 10,668 m 601 mph / 967 kmh
Takeoff Run 2,430 ft / 741 m
Range (with Drop Tanks) 660 mi / 1,062 km
Maximum Service Ceiling 48,000 ft / 14,630 m
Crew 1 pilot
Armament
  • 6x Browning M3 machine guns, 300 rounds per gun
  • A-1B GBR Gun Sight
  • AN/APG-5C Ranging Radar
  • 8x 5-inch HVAR Rockets
  • 2x 1000 lb bombs
  • 2x Drop Tanks – 206.5 U.S. Gal / 781.7 Liters

Gallery

Illustrations by Ed Jackson

F-86A-1 Sabre 47-611 – September 1948
F-86A-1 Sabre 47-630 – 1948
F-86A-5 48-0158 – 1949
F-86A-5 48-1257 – Korea 1951 – Flown by Capt. James Jabara
F-86A-5 Sabre 49-1080 February 1952 – Note the 5 inch HVAR Rocket Mounted inboard of the fuel tank

Sources:

  1. F-86 Sabre in Action, Larry Davis, Squadron/Signal Publications, 1992.
  2. The North American Sabre, Ray Wagner, MacDonald, 1963.
  3. The American Fighter, Enzo Angelucci and Peter Bowers, Orion, 1987.
  4. The World Guide to Combat Planes, William Green, MacDonald, 1966.
  5. Flash of the Sabre, Jack Dean, Wings Vol 22, No 5, 1992.
  6. North American F-86 Sabre, Larry Davis, Wings of Fame, Volume 10, 1998

 

North American XP-86 Sabre

USA flag old United States of America (1945)
Prototype Fighter – 3 Built

The first XP-86 Prototype 45-59598, flown by George Welch

The North American F-86 Sabre is one of the most well-known fighter aircraft of all time, marking the transition from the propeller to the jet turbine. It first entered service with the newly formed U.S. Air Force in 1949, and was instrumental in denying air superiority to Communist forces during the Korean War. After the war ended, many Sabres entered service with dozens of foreign air arms, becoming the primary fighter equipment of many Allied nations. It was built under license in Canada, Japan, Italy, and Australia. Its service was so long-lived that the last operational F-86 was not withdrawn from service until 1993.

History

The F-86 Sabre began its life as North American Aviation’s company project NA-134, which was originally intended for the US Navy. As the war in the Pacific edged toward its climax, the Navy was making plans to acquire jet-powered carrier-based aircraft, which it was could be pressed into service in time for Operation Olympic-Coronet, the invasion of Japan planned for May 1946. The Navy had planned to acquire four jet fighters, the Vought XF6U-1 Pirate, the McDonnell XFD-1 Phantom, the McDonnell XF2D-1 Banshee, and the North American XFJ-1 Fury.

Work on the NA-134 project began in the late autumn of 1944. The NA-134 had a straight, thin-section wing set low on a round fuselage. It featured a straight through flow of air from the nose intake to the jet exhaust that exited the aircraft under a straight tailplane. The wing was borrowed directly from the P-51D, and had a laminar-flow airfoil. It was to be powered by a single General Electric TG-180 gas turbine which was a license-built version of the de Havilland Goblin. The TG-180 was designated J35 by the military and was an 11-stage axial-flow turbojet which offered 4000 lb.s.t. at sea level. The Navy ordered three prototypes of the NA-134 under the designation XFJ-1 on January 1, 1945. On May 28, 1945, the Navy approved a contract for 100 production FJ-1s (NA-141).

At the same time that North American was beginning to design the Navy’s XFJ-1, the U.S. Army Air Force (USAAF) issued a requirement for a medium-range day fighter which could also be used as an escort fighter and a dive bomber. Specifications called for a speed of at least 600 mph, since the Republic XP-84 Thunderjet already under construction promised 587 mph. On Nov 22, 1944, the company’s RD-1265 design study proposed a version of the XFJ-1 for the Air Force to meet this requirement. This design was known in company records as NA-140. The USAAF was sufficiently impressed that they issued a letter contract on May 18, 1945 which authorized the acquisition of three NA-140 aircraft under the designation XP-86.

The Navy’s XFJ-1 design had to incorporate some performance compromises in order to support low-speed carrier operations, but the land-based USAAF XP-86 was not so constrained and had a somewhat thinner wing and a slimmer fuselage with a high fineness ratio. However, the XP-86 retained the tail surfaces of the XFJ-1.

The XP-86 incorporated several features not previously used on fighter aircraft, including a fully-pressurized cockpit and hydraulically-boosted ailerons and elevators. Armament was the standard USAAF equipment of the era–six 0.50-inch Browning M3 machine guns that fired at 1100 rounds per minute, with 267 rounds per gun. The aircraft was to use the Sperry type A-1B gun/bomb/rocket sight, working in conjunction with an AN/APG-5 ranging radar. Rocket launchers could be added underneath the wings to carry up to 8 5-inch HVARs. Self-sealing fuel tanks were to be fitted, and the pilot was to be provided with some armor plating around the cockpit area.

In the XP-86, a ten percent ratio of wing thickness to chord was used to extend the critical Mach number to 0.9. Wingspan was to be 38 feet 2.5 inches, length was 35 feet 6 inches, and height was 13 feet 2.5 inches. Four speed brakes were to be attached above and below the wings. At a gross weight of 11,500 pounds, the XP-86 was estimated to be capable of achieving a top speed of 574 mph at sea level and 582 mph at 10,000 feet, still below the USAAF requirement. Initial climb rate was to be 5,850 feet per minute and service ceiling was to be 46,000 feet. Combat radius was 297 miles with 410 gallons of internal fuel, but could be increased to 750 miles by adding a 170 gallon drop tank to each wingtip. As it would turn out, these performance figures were greatly exaggerated.

A mock-up of the XP-86 was built and approved on June 20, 1945. However, early wind tunnel tests indicated that the airframe of the XP-86 would not be able to reach the desired speed of 600 mph. It is highly likely that the XP-86 project would have been cancelled at this time were it not for some unusual developments.

Saved by the Germans

After the surrender of Germany in May of 1945, the USAAF, along with a lot of other air forces, was keenly interested in obtaining information about the latest German jet fighters and in learning as much as they could about secret German wartime research on jet propulsion, rocket power, and ballistic missiles. American teams were selected from industry and research institutions and sent into occupied Germany to investigate captured weapons research data, microfilm it, and ship it back to the US.

The First XP-86 Prototype in Flight Testing [San Diego Air & Space Museum]
By the summer of 1945, a great deal of German data was pouring in, much of it as yet untranslated into English. As it turned out, German aeronautical engineers had wind-tunnel tested just about every aerodynamic shape that the human mind could conceive of, even some ideas even only remotely promising. A particular German paper dated 1940 reported that wind tunnel tests showed that there were some significant advantages offered by swept wings at speeds of about Mach 0.9. A straight-winged aircraft was severely affected by compressibility effects as sonic speed was approached, but the use of a swept wing delayed the effects of shock waves and permitted better control at these higher speeds. Unfortunately, German research also indicated that the use of wing sweep introduced some undesirable wing tip stall and low-speed stability effects. American researchers had also encountered a similar problem with the swept-wing Curtiss XP-55 Ascender, which was so unstable that it flipped over on its back and stalled on one of its test flights.

In 1940, these German studies were of only theoretical interest, since no powerplants were available even remotely capable of reaching such speeds. However, such studies caught the attention of North American engineers trying to develop ways to improve the performance of their XP-86.

Going Supersonic

The first XP-86 prototype in what would be a temporary white paint scheme

The optimal design for an aircraft capable of high speeds produces a design that stalls easily at low speeds. The cure for the low-speed stability problem that was worked out by North American engineers was to attach automatic slats to the wing leading edges. The wing slats were entirely automatic, and opened and closed in response to aerodynamic forces. When the slats opened, the changed airflow over the upper wing surface increased the lift and produced lower stalling speeds. At high speeds, the slats automatically closed to minimize drag.

In August of 1945, project aerodynamicist L. P. Greene proposed to Raymond Rice that a swept-wing configuration for the P-86 be adopted. Wind tunnel tests carried out in September of 1945 confirmed the reduction in drag at high subsonic speeds as well as the beneficial effect of the slats on low speed stability. The limiting Mach number was raised to 0.875.

Based on these wind-tunnel studies, a new design for a swept-wing P-86 was submitted in the fall of 1945. The USAAF was impressed, and on November 1, 1945 it readily approved the proposal. This was one of the most important decisions ever made by the USAAF. Had they not agreed to this change, the history of the next forty years would undoubtedly have been quite different.

North American’s next step was to choose the aspect ratio of the swept wing. A larger aspect ratio would give better range, a narrower one better stability, and the correct choice would have to be a tradeoff between the two. Further tests carried out between late October and mid November indicated that a wing aspect ratio of 6 would be satisfactory, and such an aspect ratio had been planned for in the proposal accepted on November 1. However, early in 1946 additional wind tunnel tests indicated that stability with such a narrow wing would be too great a problem, and in March the design reverted to a shorter wingform. An aspect ratio of 4.79, a sweep-back of 35 degrees, and a thickness/chord ratio of 11% at the root and 10% at the tip was finally chosen.

All of these changes lengthened the time scale of the P-86 development in comparison to that of the Navy’s XFJ-1. The XFJ-1 took to the air for the first time on November 27, 1946, but the XP-86 still had almost another year of work ahead before it was ready for its first flight.

The first XP-86 prototype in flight during testing [North American Aviation]
On February 28, 1946, the mockup of the swept-winged XP-86 was inspected and approved. In August of 1946, the basic engineering drawings were made available to the manufacturing shop of North American, and the first metal was cut. The USAAF was so confident of the future performance of the XP-86, that on December 20, 1946 another letter contract for 33 production P-86As was approved. No service test aircraft were ordered. Although the 4000 lb.s.t. J35 would power the three XP-86 prototypes, production P-86As would be powered by the General Electric TG-190 (J47) turbojet offering 5000 lb.s.t.

The first of three prototypes, 45-59597, was rolled out of the Inglewood factory on August 8, 1947. It was powered by a Chevrolet-built J35-C-3 turbojet rated at 4000 pounds of static thrust. The aircraft was unarmed. After a few ground taxiing and braking tests, it was disassembled and trucked out to Muroc Dry Lake Army Air Base, where it was reassembled.

Test pilot George “Wheaties” Welch took the XP-86 up into the air for the first time on October 1, 1947. The flight went well until it came time to lower the landing gear and come in for a landing. Welch found that the nosewheel wouldn’t come down all the way. After spending forty minutes in fruitless attempts to shake the nosewheel down into place, Welch finally brought the plane in for a nose-high landing. Fortunately, the impact of the main wheels jolted the nosewheel into place, and the aircraft rolled safely to a stop. The swept-wing XP-86 had made its first flight.

On October 16, 1947, the USAF gave final approval to the fixed price contract for 33 P-86As, with the additional authorization for 190 P-86Bs. The P-86B was to be a strengthened P-86A for rough-field operations.

XP-86 number 45-59597 was officially delivered to the USAF on November 30, 1948. By that time, its designation had been changed to XF-86. Phase II flight tests, those flown by USAF pilots, began in early December of 1947. An Allison-built J35-A-5 rated at 4000 lbs of static thrust was installed for USAF tests. The second and third XP-86 prototypes, 45-59598 and 45-59599 respectively, joined the test program in early 1948. These were different from the first prototype as well as being different from each other in several respects. Numbers 1 and 2 had different fuel gauges, a stall warning system built into the control stick, a bypass for emergency operation of the hydraulic boost system, and hydraulically-actuated leading-edge slat locks. The number 3 prototype was the only one of the three to have fully-automatic leading-edge slats that opened at 135 mph. Numbers 2 and 3 had SCR-695-B IFF beacons and carried the AN/ARN-6 radio compass set.

The original XP-86 prototype was used for evaluating the effects of nuclear blasts on military hardware at Frenchman Flats. It was later scrapped. [This Day in Aviation]
In June of 1948, the new US Air Force redesignated all Pursuit aircraft as Fighter aircraft, changing the prefix from P to F. Thus the XP-86 became the XF-86. XP-86 number one was officially delivered to the USAF on November 30, 1948. The three prototypes remained in various test and evaluation roles well into the 1950s, and were unofficially referred to as YP-86s. All three prototypes were sold for scrap after being used in nuclear tests at Frenchman Flats in Nevada

Design

The three XP-86 prototypes flying in formation together in 1948 [National Archives]
Evolving from the NA-134 project with wings borrowed from a P-51, the XP-86 would eventually end up with a low swept wing mounted to a tubular fuselage, with a large jet intake opening at the nose. The plexiglass bubble canopy gave the pilot great visibility, and afforded the pilot a pressurized cockpit. The tail featured a swept back rudder with tailplanes angled upwards, marking a departure from the largely perpendicular angles seen on most of the Sabre’s propeller driven predecessors. The landing gear was a tricycle configuration, which helped balance the weight of the jet engine at the rear.

The wing of the XP-86 was to be constructed of a double-skin structure with hat sections between layers extending from the center section to the outboard edges of the outer panel fuel tanks. This structure replaced the conventional rib and stringer construction in that area. This new construction method provided additional strength and allowed enough space in the wing for fuel tanks.

The wing-mounted speed brakes originally contemplated for the XP-86 were considered unsuitable for the wing design, so they were replaced by a hydraulic door-type brake mounted on each side of the rear fuselage and one brake mounted on the bottom of the fuselage in a dorsal position. The speed brakes opened frontwards, and had the advantage that they could be opened at any attitude and speed, including speeds above Mach One.

The maximum speed of the XP-86 was over 650 mph, 75 mph faster than anything else in service at the time. The noise and vibration levels were considerably lower than other jet-powered aircraft. However, the J35 engine did not produce enough thrust, and the XP-86 could only climb at 4,000 feet per minute. However, this was not considered an issue, since the production P-86As were to be powered by the 5000 lb.s.t. General Electric J47.

The XP-86 could go supersonic in a dive with only a moderate and manageable tendency to nose-up, although below 25,000 feet there was a tendency to roll which made it unwise to stay supersonic for very long. Production Sabres were limited to Mach 0.95 below 25,000 feet for safety reasons because of this roll tendency.

For the second and third prototypes, the ventral brake was eliminated, and the two rear-opening side fuselage brakes were replaced by brakes which had hinges at the front and opened out and down. These air brakes were adopted for production aircraft.

Prototype number 3 was the only one to be fitted with armament. The armament of six 0.50-inch M3 machine guns were mounted in blocks of three on either side of the cockpit. Ammunition bays were installed in the bottom of the fuselage underneath the gun bay, with as many as 300 rounds per gun. The guns were aimed by a Mk 18 gyroscopic gunsight with manual ranging.

Possibly the First Supersonic Aircraft

George Welch Circa 1947 – [San Diego Air & Space Museum]
There is actually a possibility that the XP-86 rather than the Bell XS-1 might have been the first aircraft to achieve supersonic flight. During some of his early flight tests, George Welch reported that he had encountered some rather unusual fluctuations in his airspeed and altitude indicators during high speed dives, which might have meant that he had exceeded the speed of sound. However, at that time, North American had no way of calibrating airspeed indicators into the transonic range above Mach 1, so it is uncertain just how fast Welch had gone. On October 14, 1947, Chuck Yeager exceeded Mach 1 in the XS-1. Although the event was kept secret from the general public, North American test crews heard about this feat through rumors and persuaded NACA to use its equipment to track the XP-86 in a high-speed dive to see if there was a possibility that the XP-86 could also go supersonic. This test was done on October 19, five days after Yeager’s flight, in which George Welch was tracked at Mach 1.02. The tests were flown again on October 21 with the same results. Since Welch had been performing the very same flight patterns in tests before October 14, there is the possibility that he, not Chuck Yeager, might have been first to exceed the speed of sound.

In any case, the fact that the XP-86 had exceeded the speed of sound was immediately classified, and remained so for several months afterward. In May of 1948, the world was informed that George Welch had exceeded Mach 1.0 in the XP-86, becoming the first “aircraft” to do so, with an aircraft being defined as a vehicle that takes off and lands under its own power. The date was set as April 26, 1948. This flight did actually take place, but George Welch was not the pilot. In fact, it was a British pilot who was evaluating the XP-86 who inadvertently broadcasted that he had exceeded Mach 1 over an open radio channel. However, the facts soon became common knowledge throughout the aviation community. The June 14, 1948 issue of Aviation Week published an article revealing that the XP-86 had gone supersonic.

Variants

  • XP-86 45-59597 – The first prototype Sabre produced, was reconfigured many times with various test configurations. May have been the first aircraft to have gone supersonic in October 1947 with George Welch at the controls.
  • XP-86 45-59598 – The second prototype, had different production model speedbrake and flap configuration, various sensors and equipment installed for testing purposes.
  • XP-86 45-59599 – The third prototype, and the only Sabre prototype to have been armed, fitted with the standard six M3 Browning guns

Operators

  • United States – The prototypes were extensively tested by North American Aviation before being handed over to the U.S. Air Force in 1948.

North American XP-86 Specifications

Wingspan 37 ft 1.5 in / 11.32 m
Length 37 ft 6.5 in / 11.44 m
Height 14 ft 9 in / 4.5 m
Wing Area 299 ft² / 27.8 m²
Engine 1x Chevrolet J35-C-3 Turbojet Engine

4000 lbst

Fuel Capacity 410 US Gal / 1,552 L

750 US Gal / 2,839 L with wingtip drop tanks

Weights
Empty 9,730 lb / 4,413 kg
Gross 13,395 lb / 6,076 kg
Maximum Take Off 16,438 lb / 7,456 kg
Climb Rate
Rate of Climb at Sea Level 4000 ft / 1219 m per minute
Time to 20,000 ft / 6,096 m 6.4 minutes
Time to 30,000 ft / 9,144 m 12.1 minutes
Maximum Speed
Sea Level 599 mph / 964 kmh
14,000 ft / 4267 m 618 mph / 995 kmh
35,000 ft / 10,668 m 575 mph / 925 kmh
Takeoff Run 3,030 ft / 924 m
Range 297 mi / 478 km
Maximum Service Ceiling 41,300 ft / 12,588 m
Crew 1 pilot
Armament
  • 6x Browning M3 machine guns, 267 rounds per gun
  • Sperry type A-1B gun/bomb/rocket sight
  • AN/APG-5C ranging radar
  • Underwing Rocket Launchers, up to 8x 5-inch HVAR

Gallery

Illustrations by Ed Jackson

XP-86 – 1st Prototype 45-59597 circa 1947 note it bears the P for “Pursuit”
XP-86 – 1st Prototype 45-59597 circa June 1948 in white paint scheme, note the wingtip pitot probes
XP-86 – 1st Prototype 45-59597 circa 1948, note the additional test equipment behind the pilot’s seat
XP-86 – 2nd Prototype 45-59598 circa 1948
XP-86 – 3rd Prototype 45-59599 circa 1948

Credits

Rockwell B-1A Lancer

USA flag United States of America (1974)
Prototype Supersonic Heavy Bomber – 4 Built

B-1A 74-0159

The B-1A program arose out of a need for a long-range, supersonic, low-flying heavy bomber. The program’s initial development was pushed forward through an ever-shifting geopolitical landscape, as well as opposition and contention among the the top levels of the U.S. government. Even with advanced features such as variable sweep wings, and variable air intake and exhaust capability, it was derided as a ‘dinosaur’ in the age of ICBMs. The opposition and political infighting nearly ended the Lancer, before it was given a miraculous second chance.

History

B-1A 74-158 taxiing on ground. (U.S. Air Force photo)

The origin of the Rockwell B-1 can be traced back to 1961, when the Air Force began to consider alternatives to the North American B-70 Valkyrie, which had just been downgraded from production to test aircraft status. At that time, the long range strategic missile was assumed to be the weapon of the future, with manned long-range bombers being relegated to a secondary role. The B-70 had been designed to fly at extremely high altitudes and at Mach 3 speeds, and increasingly effective Soviet anti aircraft defenses had made such an aircraft rather vulnerable.

Nevertheless, the Air Force commissioned several studies to explore possible roles for manned bombers in future planning. If successful, these would replace the B-52. At this time, the ability to fly through enemy airspace at extremely low altitudes was was thought to be the key for survival in the face of sophisticated air defenses.

The first such study was known as the Subsonic Low Altitude Bomber (SLAB), which was completed in 1961. It envisaged a 500,000 pound fixed-wing aircraft with a total range of 11,000 nautical miles, with 4300 nm of these miles being flown at low altitudes. This was followed soon after by the Extended Range Strike Aircraft (ERSA), which had a weight of 600,000 pounds and featured a variable sweep wing. The ERSA was supposed to be able to carry a payload of 10,000 pounds and achieve a range of 8750 nautical miles, with 2500 of these miles being flown at altitudes as low as 500 feet. In August of 1963, a third study known as Low-Altitude Manned Penetrator(LAMP) was completed. It called for a 20,000 payload and a 6200 nautical mile range, 2000 miles being flown at low altitude. None of these projects ever got beyond the basic concept stage.

In October of 1963, the Air Force looked over these proposals and used the results as the foundation of a new bomber proposal, termed Advanced Manned Precision Strike System (AMPSS). In November of that year, 3 contractors were issued Requests for Proposals for the AMPSS. The companies were Boeing, General Dynamics, and North American. However, Secretary of Defense Robert McNamara kept a tight rein on funds, and expressed doubts about the assumptions behind AMPSS, so the RFPs only involved basic concept studies and did not focus on a specific aircraft. In addition, the contractors all agreed that some of the suggested USAF requirements either did not make much sense or else were prohibitively costly.

In mid-1964, the USAF had revised its requirements and retitled the project as Advanced Manned Strategic Aircraft (AMSA). The AMSA still envisaged an aircraft with the takeoff and low-altitude performance characteristics of the AMPSS, but in addition asked for a high-altitude supersonic performance capability. The projected gross weight for the aircraft was 375,000 pounds, and the range was to be 6300 nautical miles, 2000 of which would be flown at low altitude.

Secretary McNamara was never very excited about the AMSA, since he thought that strategic missiles could do a better job of “assured destruction” than manned bombers, and thought that the cost of the AMSA would probably be excessive. Nevertheless, there was a potential gain in avionics and propulsion technology that could be achieved if the project were to proceed, and McNamara released a small amount of funding for preliminary AMSA studies. The airframe for the AMSA would be worked on by Boeing, General Dynamics, and North American, whereas Curtiss-Wright, General Electric, and Pratt & Whitney would work on the engines. Both IBM and Hughes aircraft looked at potential avionics systems. These contractors issued their reports in late 1964. General Electric and Pratt & Whitney were given a contract to produce two demonstrator engines, but no airframe and avionics contracts were issued at that time.

74-0160 on display at Edwards AFB in 1980. (U.S. Air Force photo)

A bit of confusion entered the picture when the Defense Department selected the FB-111A as the replacement for the B-52C, B-52F, and B-58. The Air Force had not requested a bomber version of the controversial F-111, and was not all that enthusiastic about the choice. Nevertheless, a low-cost interim bomber did have some attractive features, and the Air Force went along with the choice of the FB-111A provided it did not interfere with AMSA development.

By 1968, an advanced development contract was issued to IBM and the Autonetics Division of North American Rockwell. On September 22, 1967, North American Aviation had merged with Rockwell Standard Corporation to create North American Rockwell. Earlier in that year, the Joint Chiefs of Staff had recommended the immediate development of the AMSA, but Secretary McNamara was still opposed, preferring instead to upgrade the existing FB-111 and B-52 fleet. McNamara vetoed the proposal.

When Richard Nixon became President in January of 1969, his Secretary of Defense Melvin Laird reviewed Defense Department needs and announced in March of 1969 that the planned acquisition of 253 FB-111s would be reduced to only 76, since the FB-111 lacked the range and payload required for strategic operations, and recommended that the AMSA design studies be accelerated.

The AMSA was officially assigned the designation B-1A in April of 1969. This was the first entry in the new bomber designation series, first created in 1962.

New Requests For Proposals were issued in November of 1969. IBM and Autonetics were selected for the avionics work on December 19. The selection of airframe and engine contractors was delayed by budget cuts in FY 1970 and 1971. On December 8, 1969 North American Rockwell and General Electric were announced as the winners of the respective airframe and engine contracts for the B-1A.

The original program called for 2 test airframes, 5 flyable aircraft, and 40 engines. This was cut in 1971 to one ground test aircraft and 3 flight test articles (74-0158/0160). First flight was set for April of 1974. A fourth prototype (76-1074) was ordered in the FY 1976 budget. This fourth plane was to be built to production standards. At one time, some 240 B-1As were to be built, with initial operational capability set for 1979.

Design

B-1A Orthogonal Projection. Note the difference between the wings at maximum and minimum sweep. (U.S. Air Force photo)

The fuselage of the B-1A was fairly slim, and seated a crew of four in the nose. There was a large swept vertical tail, with a set of all-flying slab tailplanes mounted fairly high on the vertical tail. The aircraft’s fuselage blended smoothly into the wing to enhance lift and reduce drag. In addition, the fuselage was designed to reduce the aircraft’s radar cross section in order to minimize the probability of detection by enemy defenses.

In order to achieve the required high-speed performance and still be able to have a good low-speed takeoff and landing capability, a variable-sweep wing was used. This made it possible for the aircraft to use short runways that would be inaccessible to the B-52. The outer wing panels were attached to a wing carry-through attachment box which faired smoothly into a slim, narrow fuselage. Each outer wing had full-span slats and slotted flaps, but used no ailerons. Lateral control was provided by a set of spoilers on the wing upper surface, acting in conjunction with differential operation of the slab tailplanes.

The engines were four afterburning General Electric F101-100 turbofans. The engines were installed in pairs inside large nacelles underneath the wing roots,, and close to the aircraft’s center of gravity to improve stability while flying at high speed through highly-turbulent low-altitude air. The nacelles were far enough apart so that the main landing gear members could be installed in the wing roots between them with enough clearance to retract inwards. In order to achieve the required Mach 2 performance at high altitudes, the air intake inlets were variable. In addition, the exhaust nozzles were fully variable.

Initially, it had been expected that a Mach 1.2 performance could be achieved at low altitude, which required that titanium rather than aluminum be used in critical areas in the fuselage and wing structure. However, this low altitude performance requirement was lowered to only Mach 0.85, enabling a greater percentage of aluminum to be used, lowering the overall cost. Titanium was used primarily for the wing carry-through box, the inner ends of the outer wings incorporating the pivots, and for some areas around the engines and rear fuselage.

Eight integral fuel tanks were planned, one in each outer wing panel, and the rest in the fuselage. About 150,000 pounds of fuel could be carried. There were three 15-foot weapons bays in the lower fuselage, two ahead and one behind the wing carry-through box. Each bay could carry up to 25,000 pounds of conventional or nuclear weapons. The total weapons load was almost twice what a B-52 could carry. All of the offensive weapons were to be carried internally, with no provision for externally-mounted pylons. A key weapon was to be the AGM-69A SRAM (Short-Range Attack Missile), 8 of which could be carried on a rotary launcher in each of the weapons bays.

No defensive armament was planned, the B-1A relying on its low-altitude performance and its suite of electronic countermeasures gear to avoid interception.

An extensive suite of electronics was planned, including a Litton LN-15 inertial navigation system, a Doppler radar altimeter, a Hughes forward-looking infrared, and a General Electric APQ-114 forward-looking radar and a Texas Instruments APQ-146 terrain-following radar.

The B-1A carried a crew of four–a pilot, copilot, offensive systems officer, and defensive systems officer. The crew escape system resembled that of the F-111 crew escape module. In an emergency, a capsule containing all four crewmembers would separate from the aircraft and be steered and stabilized by various fins and spoilers. A rocket motor would fire and lift the capsule up and away from the aircraft. Three parachutes would then open and would lower the capsule along with the crew safely to the surface. Once down, the capsule would serve as a survival shelter for the crew members.

Development

The B-1A mockup review occurred in late October of 1971. There were 297 requests for alterations.

The first B-1 flight aircraft (74-0158) rolled out from USAF Plant 42 at Palmdale, CA on October 26, 1974. It made its first flight on December 23, 1974, a short hop to Edwards AFB where the flight testing was to be carried out. The crew was Rockwell test pilot Charlie C. Bock,; Jr, Col. Emil Sturmthal, and Richard Abrams. The third aircraft (74-0160) was to be the avionics testbed and flew for the first time on March 26, 1976. The second aircraft (74-0159) was initially used for some static ground testing and did not make its first flight until June 14, 1976.

The B-1A test program went fairly smoothly. However, there were numerous modifications introduced throughout the program and some items of additional equipment were added. The avionics suite of the B-1A was perhaps the most complex yet used on an aircraft. The Initial Operational Test and Evaluation tests were successfully passed in September of 1976. The Phase 1 flight test program was completed on September 30, 1976. In December of 1976, the Air Force concluded that the B-1A was to go into production, with contracts placed for the first three aircraft and plans were made for an initial Block 2 production batch of 8 aircraft.

It seemed that the B-1A was well on its way to a full production run of 240 aircraft. However, the cost of the B-1A program began to escalate, and there were still some unresolved issues concerning the avionics suite. In 1970, the estimated per-unit price was $40 million, and by 1972, the cost had risen to $45.6 million. Although this sounds like small-change by today’s standards, this was considerably greater than the figure for any previous production aircraft. Moreover, by 1975, this number had climbed to $70 million.

Alarmed at these rising costs, the new presidential administration of Jimmy Carter (which had taken office on January 20, 1977) began to take a second look at the whole B-1A program. On June 30, 1977, President Carter announced that plans to produce the B-1A would be cancelled, and that the defense needs of the USA would be met by ICBMs, SLBMs, and a fleet of modernized B-52s armed with ALCMs. President Carter genuinely wanted to reduce the arms race, but he was unaware at the time of the secret projects that would ultimately lead to the F-117A stealth attack aircraft and the B-2 Spirit stealth bomber.

B-1A during the B-1B flight test program. (U.S. Air Force photo)

Despite the cancellation of the production program, the Carter administration allowed the flight testing of the B-1A to continue. Most of the effort involved the avionics, in particular the defensive systems. In addition, General Electric continued to work on improvements for the F101 engine, and most of the contractors kept their engineering teams intact. Perhaps most important, work continued in reducing the radar cross section of the aircraft. Less than a month after the cancellation, 74-0160 launched a SRAM on July 28, 1977 at an altitude of 6,000 feet over the White Sands missile range. This aircraft was later modified with an advanced electronic countermeasures system mounted in a dorsal spine, and Doppler beam sharpening was added to the forward-looking radar. 74-0158 had achieved Mach 2.0 in April of 1976, and after completing its stability and control tests was placed in storage in 1978. On October 5, 1978, 74-0159 achieved a speed of Mach 2.22, the highest speed achieved during the B-1A program.

74-0158 was retired from flying in April of 1981 after having flown 138 sorties, the largest number of flights of any of the prototypes. By this time, it had acquired a three-tone desert camouflage scheme. It was eventually dismantled and used as a weapons trainer at Lowry AFB.

74-0159 was later used as a flight test article in the B-1B program. It was modified by having B-1B flight control system features installed. It began flying on March 23, 1983. Unfortunately, it crashed on August 29, 1984 when the aircraft’s center of gravity got unbalanced during fuel transfer management procedures, causing it to lose control. The escape capsule deployed successfully, but the parachute risers did not deploy properly. The capsule hit the ground at a steep angle, so steep that the inflatable cushions could not shield the impact. Chief test pilot Doug Benefield was killed, and two other crew members were seriously injured.

74-0160 was later converted to a ground trainer under the designation GB-1A and is now on display at the Wings Over The Rockies Air and Space Museum (formerly Lowry AFB), near Denver, Colorado.

76-0174 had been ordered to serve as a pre-production B-1A aircraft and was configured with full avionics systems. When the B-1A program was cancelled, work on this aircraft was well under way. Unlike the first three B-1s, 76-0174 was equipped with four conventional ejector seats in place of the escape capsule. This change was made after tests had determined that the crew escape module was unstable if ejected at speeds above 347 knots. It flew on February 14, 1979 and carried out 70 sorties. This plane was later used as a test article in support of the B-1B program. It resumed flying on July 30, 1984. Externally, the main change was the removal of the long dorsal spine but many of the B-1B avionics systems were installed internally. It is now on display at the USAF Museum at Wright Patterson AFB in Ohio.

Variants

  • B-1A – The initial prototype run of four aircraft

Operators

  • U.S. Air Force – The sole operator of the B-1A was the USAF

 

B-1A Lancer

Wingspan
(at max sweep)
78 ft 2.5 in / 23.84 m
Wingspan
(at min sweep)
136 ft 8.5 in / 41.67 m
Length 143 ft 3.5 in / 43.8 m
Height 34 ft 0 in / 10.36 m
Wing Area 1,950 ft² / 181.2 m²
Engine 4x General Electric F101-GE-100 turbofans, 17,390 lbf dry, 30,000 lbf with afterburner
Fuel Capacity 29,755 US Gal / 11,2634 L
Loaded Weight 389,000 lb / 176,450 kg
Maximum Take Off Weight 395,000 lb / 179,170 kg
Maximum Speed Mach 2.2 / 1,688 mph / 2716.5 kmh at 50,000 ft / 15,240 m
Maximum Service Ceiling 62,000 ft / 18,900 m
Crew 1 pilot, 1 copilot, 1 offensive systems officer, 1 defensive systems officer

Gallery

Illustrations by Basilisk https://basilisk2.deviantart.com

B-1A 74-0158 seen in Anti-Flash White
B-1A 74-0160 seen in a SAC Low Level Livery
B-1A 76-0174 seen in camouflage paint scheme
B-1A 76-174 seen in camouflage during testing. (U.S. Air Force photo)
A right side ground view of a B-1A aircraft wearing dark green camo. (U.S. Air Force Photo)
B-1A 76-174 in flight with wings extended in the 25-degree sweep position. (U.S. Air Force photo)

Sources

1.-F-14A-VF-1-BuNo-162597_03 Tomcat

Grumman F-14 Tomcat

usa flag USA (1974)
Tactical Fighter Plane – 712 Built

The F-14 Tomcat is the most iconic Cold War US Naval fighter, next to the McDonnel Douglas F-4 Phantom. It is also a replacement for the F-4 Phantom and the failed F-111B, incorporating the lessons and experiences acquired during Vietnam as well, like the F-15 Eagle. It has a similar origin to that of the F-15, but it is also the result of two additional factors. First, the Navy’s quest to find a Fleet Air Defence asset, with long-range and high-endurance interceptor characteristics to defend the aircraft carrier battle groups, mainly against long-range anti-ship missiles launched from Soviet bombers and submarines, in addition to intercepting those same Soviet bombers. It also needed a more capable radar and provision for longer range missiles. The role of then Secretary of Defence Robert McNamara was also crucial in this case, as he directed the Navy to take part in the Tactical Fighter Experimental program. But the Navy stepped out in fears that the USAF’s need for a low-attack aircraft would hamper the fighter abilities of the new airplane. Second, the ongoing TFX F-111B project was facing a large number of issues in the late 60s that made both the Navy and Grumman, which happened to be the builder of the F-111B alongside General Dynamics, to consider a new option with better capabilities and less operational and development issues. The F-111B proved unsuitable for the conditions of the Vietnam War and had no long-range missile capability. The Naval Air Systems Command (NAVAIR) also had a role, as it issued requirements for a tandem two-seat, twin-engine fighter with mainly air-to-air capacities capable of reaching speed of up to 2.2 match and able to operate with a new generation missiles. It was also directed to have a secondary Close Air Support (CAS) role and incorporate an internal M61A1 20mm Vulcan cannon, correcting the mistake made with the previous Phantom F-4, as it had no internal gun for close-range combat. A feat achieved by the Tomcat was that it had its first flight 23 months after the contract was awarded, making the of the Tomcat a milestone in the development of new air assets. NASA also had an important role during the development stage as it did with the F-15 through the Langley Research Centre, mainly related to the F-14’s most advanced feature: the geometrically variable wings. But it also played a role in the overall design of the fighter, working very closely with Grumman providing the company with technical assistance and data.

Characteristics

F-14B Tomcat aircraft from VF-101 circa 2004

The F-14 Tomcat is a double-seat tandem, twin-engine, double-tail, all-weather carrier-based fighter and interceptor and later gaining multi-role capability, with numerous remarkable features. The glove-mounted swept wings have variable geometry capability, in the same manner as the General Dynamics F-111, the Mig-23 Flogger, and the Panavia Tornado. When the wings were positioned rearwards, it was fitted for high-speed intercept missions. When swept outwards, the wings naturally increased drag, allowing lower speed flight and a lower stall speed. The control of the wing movement was automatic with manual control if needed. The flat area between the engines nacelles, at the rear of the fighter, purposed to contain fuel and avionics components, such as the controllers for the wing-sweep mechanism, flares and chaff and other flight assist functions. This results in a wide space between the two nacelles giving the Tomcat it’s characteristic shape. Its design is based in the aforementioned requirements, which required the new fighter to carry a combination of AIM-9 Sidewinder short-range missiles, AIM-7 Sparrow medium-range missiles, and long-range AIM-54 Phoenix missiles, alongside the 20mm M61A1 Vulcan cannon. As Grumman was awarded with the contract in 1968, it incorporated two features of the unsuccessful F-111B project: the two Pratt & Whitney TF-30-P3 engines and the required AWG-9 radar for the AIM-54 Phoenix. If one observes carefully, it can be concluded that there are many similarities between the F-111 and the F-14, not only the geometrically variable wings.

The F-14 Tomcat became the Naval equivalent of the F-15, as it was equally as capable as the Eagle, with the addition role of an embarked fighter, performing maritime air superiority, fleet defense, long-range interception, and tactical aerial reconnaissance missions. Despite the quite similar structure of the Eagle, the two fighters are very different, and not only because of their purposed missions. The F-14 reportedly relied more on airborne surveillance and identification systems for beyond visual range firing.
The F-14 structure is made of 25% titanium, such as the wing structure, pivots, and both upper and lower wing flight surfaces, with electron beam welding used in their construction. The same fuselage, in combination with the wing, provide the F-14 with exceptional performance in combination with the capability provided by the variable sweep wings, provided the between 40-60% of the airframe’s lift. In fact, it allowed a Tomcat to land safely after suffering a mid-air collision that removed more than the 50% of its right wing. The wings, with their variable geometry, allowed the aircraft to reach an optimum lift-to-drag ratio according to the variation in speed, which in turn permitted the aircraft to perform various missions at different speeds. The aircraft’s twin tail configuration helped it in maneuvers at high angles of attack and contributed in reducing the height of the aircraft, making it more conducive to storage in the limited height of an aircraft carrier’s lower decks. The powerplant also allowed the Tomcat to have a good performance, but it suffered from teething problems in its early years, later requiring modifications. The Tomcat had its first flight in December 1970. The first versions were powered by two Pratt & Whitney TF-30-P412A turbofan engines, yielding speeds of up to 1,563 mph (2,517 km/h) at high altitude. But this initial engine was deemed as unreliable as it caused 28% of the Tomcat’s accidents, mainly due to compressor stalls. As a result, the powerplant had to be improved, and later versions had the were replaced with the General Electric F-100-GE-400 turbofan engine. The Tomcat also had advanced avionics that gave it superior air-to-air and later on, enhanced air-to-ground capability.

An F-14D Tomcat attached to the “Bounty Hunters” of VF-2 makes a sharp pull-up in full afterburner circa 2003

The F-14, was subject to numerous improvement programs in avionics and engines, as well as weaponry. For instance, in 1994, the Low Altitude Navigation and Targeting Infrared System for Night (LANTIRN) was incorporated on the right wing glove pylon, which enhanced the Tomcat’s CAS and air-ground attack capabilities. In addition to the pod, other upgrades in avionics and cockpit displays allowed the usage of precision-guided weaponry, enhanced defensive systems, displays and control devices and even structural improvements. A Global Positioning System and Inertial Navigation System (GPS-INS) was integrated in the LANTIRN pod. Between the late 80s and early 90s, the Tomcat was able to operate with free-fall iron bombs, thus having limited ground-attack capabilities that were enhanced by the aforementioned improvements in avionics. Many proposed improved versions were drafted, but they were ultimately rejected given technical assessments and political reluctance to develop and introduce them, considering that new and more advanced and/or comparatively lower costs alternatives were already introduced or were at their late stage of development.

Tomcats in Combat

F-14D Tomcat on the deck of the USS John C. Stennis

The F-14 saw its good share of action after being introduced in September 1974, with the first missions being implemented in the last days of the Vietnam War, providing top cover for the evacuation air route through combat air patrols. During the Cold War and in the North Atlantic, it was a routine for the F-14 to execute long-range interceptions of Soviet bombers and maritime reconnaissance aircraft that were flying too close to the aircraft carrier groups, such as the Tupolev Tu-95, Tupolev Tu-16 Badger, Tupolev M-4 Bison, Antonov An-12 Cub and Illyushin Il-38 May. In addition, NATO exercises in the Northern region of the Atlantic usually garnered the attention of the Soviets, while their routine flights from the Kola Peninsula to Cuba prompted these interceptions on a weekly, if not daily basis. The F-14 also saw some action in the Lebanese Civil War, with combat air patrols while American nationals were evacuated in 1976, and again between 1982 and 1986, with further combat air patrols and Tactical Air Reconnaissance Pod System (TARPS) missions to spot artillery positions firing against the international peacekeeping force and to provide naval gunfire support with intelligence on targets. During these operations, many F-14s were attacked by Syrian anti-aircraft fire that never managed to strike any targets, prompting retaliatory strikes where the F-14 provided cover to attacking airplanes, and also prompting the battleship USS New Jersey to open fire against Syrian AA batteries. Syrian Migs engaged but did not attack the Tomcat. Tomcats also took part in the failed operation to free the American hostages in Iran.
It was in Libya where the F-14 became very famous, during a series of incidents between the USA and Libya throughout the 80’s, where the F-14 managed to shoot down 4 Libyan aircraft, 2 Sukhoi Su-22 Fitters, and 2 MiG-23 Floggers, while also sinking a corvette and a patrol boat, and damaging many more, including surface-to-air missile (SAM) sites. During these incidents, the F-14 provided combat air patrols and interceptions, supporting various missions, such as Operation Arid Farmer, Prairie Fire and El Dorado Canyon, even outmanoeuvring 2 MiG-25 Foxbats that were intercepted. During these interventions, Tomcats were also attacked by SAMs and air-to-air missiles fired by Libyan air assets, suffering no casualties. Similar incidents took place in Somalia in 1983, where two F-14s were attacked by SAMs while performing photo-reconnaissance over the port of Berbera, being confused with Ethiopian MiG-23s. Photo-reconnaissance, damage assessment, and combat air patrols were also executed by Tomcats during the Invasion of Grenada. During the hijacking of the Italian cruise ship Achille Lauro, the F-14s monitored activities around the vessel alongside combat air patrols, managing also to force the airliner carrying the terrorists that hijacked the ship to land in a NATO air base in Italy. During the “Tanker War”, an episode of the Iran-Iraq, the F-14 provided Navy vessels with combat air patrols and escort missions, alongside fighter cover during Operation Nimble Archer and Operation Praying Mantis.

F-14D on the deck of the USS Harry S. Truman in the Persian Gulf circa 2005

The last scenarios where the F-14 saw action was in Iraq during Desert Shield and Desert Storm, where it provided combat air patrols in protection of naval and land forces deployed at sea and in Saudi Arabia, deterring Iraqi advances. Escort for attack aircraft, long range defence of naval assets, combat air patrols, and TARPS patrols were among the additional missions carried out by the Tomcats during the campaign, pinpointing SCUD launchers, and performing battle damage assessments. A single F-14 was lost due to a SAM missile, while an Mi-8 helicopter was the only air kill achieved by the Tomcat, as Iraqi air assets tended to flee when engaged by the Tomcat, being shot down by other fighters instead. After the 1991 Gulf War, Tomcats enforced no-fly zones and executed bombings with advanced ordnance, such as the GBU-24 Paveway III and GBU-10/16/24 laser-guided bombs, making use of the LANTIRN pod and of night vision systems for the first time. During the Second Gulf War and its aftermath, and during Operation Enduring Freedom in Afghanistan, Tomcats executed strike and CAS missions, deploying the JDAM bombs for the first time in combat and against high profile targets. They also acted as Forward Air Controllers for other air assets. Another scenario was in the Balkans, where the F-14 was also deployed, using laser-guided bombs and performing combat air patrol, escort, strike missions, Forward Air Controllers and TARPS tasks.

As Iran was a key US ally up until the 1979 Revolution, it received F-14s to ward-off Soviet MiG-25 reconnaissance flights over Iran. After the Revolution and the following Iran-Iraq War, the Iranian Tomcats saw extensive combat, scoring several air kills, reportedly 160, and managing to intimidate its adversaries, against the loss of 16 Tomcats due to combat and accidents. This was an impressive feat as the Tomcats were not operational and crews lacked training and experience. Reportedly, Iranian Tomcats were escorting Russian bombers performing air strikes against ISIS in 2015, the last to remain in active service.
US Navy Tomcats were retired from service in September 2006, marking the end of an era to a plane that has reached an almost mythical fame in service. They were replaced by the Boeing F/A-18E/F Super Hornet. 712 units were produced between 1969 and 1991, of which 79 were delivered to Iran in the second half of the 70’s.

Design

Tomcat of VF-101 circa at an airshow 2004

The F-14 is composite-construction fighter, with aluminium around 25% of the structure and boron among its structural components, with glove-mounted wings, powered by 2 Pratt & Whitney TF-30-PA412A on the earlier F-14A, and 2 General Electric F-110-GE-400 on the F-14B and F-14D, located within two engines nacelles on either side of the aircraft. These engines are fed by two rectangular air intakes placed at each side of the fuselage, located right just aft of the second crewman’s position. These intakes are equipped with movable air ramps and bleed doors to regulate airflows and to prevent disruptive shockwaves. A bleed system was also installed to reduce engine power during missile launches. The nacelles and engine exhausts are widely separated by a flat area containing avionics systems. A small flat and rectangular radome, fuel tanks and the air brakes are also located midship. A fuel dump is located at the very rear. It has machined frames, titanium main longerons and light alloy stressed skin, with the center fuselage possessing fuel-carrying capacity. The radome at front hinges upwards to allow access to radar.
Although the shape of the Tomcat’s airframe significantly contributed to its lift and light maneuverability, it was still one of the largest and heaviest fighter in service with the US Navy. Another outstanding characteristic of the F-14 is the geometrically variable wings, which are swept and can variate from 20° to 68°, and up onto 75° to overlap the horizontal stabilizers and facilitate storage in the aircraft carrier hangars. The wings can be automatically or manually varied inflight and by the Central Air Data Computer, that gives the variation according to the speed. The wings on asymmetric configuration manage to keep the plane flying and to land; even landings with an angle of 68°in case of emergencies. At high-speed interception, they are entirely swept back, while in low-speeds they are swept forwards. The wing pivot points in the wing gloves are spaced enough to allow instalment of weaponry by a pylon on each side, and the centre of lift moved less, reducing trim drag, at the point of allowing the required high-speed of 2.0 Mach. There are no ailerons and wing-mounted spoilers provide control during roll. There are full-span slats and flaps. The superior and inferior surfaces of the wings are of titanium, with the wing carry-through is a one-piece electron beam-welded aluminium alloy structure with a 6.71m span. Fins and rudders are of light alloy honeycomb sandwich. The aft part of the Tomcat is also where the two twin tails are placed, right at the top of the engine nacelles, in the middle, and with the horizontal stabilizers placed side to side of the aft area of the nacelles. The tails have multiple spars, honeycomb trailing-edges and boron/epoxy composite skins. The landing gear is of the characteristic tricycle type, with the forward gear being beneath the nose, and the rear gears which are retractable, located at the “shadow” of the wings. This area was reinforced in order to withstand with the force that landing and taking-offs from aircraft carriers usually require. An arresting hook is placed beneath the rear fuselage area, in a small ventral fairing.

F-14D Super Tomcat maneuvers in the Persian Gulf circa 2005

The cockpit is placed at the forward fuselage of the fighter, having two seats in tandem where the crew consisting of a pilot and radar intercept officer are seated. The seats are Martin-Baker GRU-7A ejection seats. Flight controls are hybrid analog-digital type with the pilot being the one only in charge of controls. The avionics within the cockpit comprise of a Kaiser AN/AVG-12 HUD along a AN/AVA-12 vertical and horizontal situation display, communications and direction-finders embedded in the AWG-9 radar display, the Central Air Data Computer (CADC) made by GarretAiResearch with a MOSFET-based Large-Scale Integration chipset MP944. This is reportedly one of the first chip microprocessors in history. In addition, a Northrop AN/AXX-1 Television Camera Set (TCS) for long-range target identification, mounted in the undernose pod and having two cockpit selectable Fields of View (FOV), which replaced the original AN/ALR-23 IRST with idium antimonide detectors. This device allows pilots to visually identify and track objectives within distances of 97 km (60 mi). Information gathered from the pod can be recorded by the Cockpit Television System (CTS). An AN/ALR-45 radar warning and control system, a Magnavox AN/ALR-25 radar warning receiver, a Tracor AN/ALE-29/39 chaff and flare dispenser device, which is installed at the very rear, and a Sanders AN/ALQ-100 deception jamming pod. The canopy is a bubble-shape that provides 360° view, being beneficial in air-to-air combat, which is complemented and enhanced by a set of four mirrors for each crew member.

The wings do not carry any weapon stations, but the wing pivot point beneath the wing glove and the fuselage itself are the areas where the payload is carried. The normal configuration of weaponry was 4 AIM-54 Phoenixes, 2 AIM-7 Sparrows and 2 AIM-9 Sidewinders, but this configuration varied depending of operational needs. In addition, bombs such as Mk-80 free-fall iron bombs, Mk-20 Rockeye II cluster bombs, JDAM precision bombs and Paveway laser-guided bombs were also part of the payload, mainly in case of CAS and strike/attack missions. AGM-88 HARM and AGM-84 HARPOON were tested and deemed possible for use in the Tomcat. For close-quarter-combat, the F-14 is fitted with an internal multi-barrel M61A1 Vulcan Gatling gun of 675 rounds, located at the left area of the nose. TARPS pods for reconnaissance, LANTIRN targeting pod and 2 external fuel tanks are also among the payload that the Tomcat could carry in missions.

The F-14 Tomcat owes its exceptional performance to the combination of powerplant, avionics, the swept variable wings and the fuselage. For instance, the relatively wide airframe provided the Tomcat with 40-60% of its aerodynamic lifting position in conjunction with the wings, thanks to the structure’s components that reduced weight while increasing resistance to G forces. In addition, the range, payload, acceleration and climb were enhanced by these factors. The engine gave the Tomcat remarkable acceleration, speed and climbing characteristics, with a maximum speed of 1,584 mph (2548 km/h). The wings also provided good capability, such as variable speeds, enabling the Tomcat to accomplish a wide array of missions, and better capacity to hold at a designated area for a prolonged period of time. Agility is also a strong suit for the Tomcat, being able to perform high-performance maneuvers, thanks to the pitch authority resulting from the design of the airframe. The deadly and spectacular characteristics of the F-14 are complemented by the very capable and advanced avionics systems that enabled it to carry out its missions, enhanced by the aforementioned improvements in this area. The Hughes AN/AWG-9X radar with integrated Identification Friend-Foe (IFF) can track up to 24 targets thanks to the Track-While-Scan (TWS), Range-While-Search (RWS), Pulse-Doppler Single-Target Track (PDSTT), and JAT (JAT). 6 targets located within distances of up to 97km (60 mi) can be engaged through the TWS while devising and executing fire control solutions for these targets. While the Pulse-Doppler mode allows firing of cruise missiles thanks to the same radar detecting, locking and tracking small objects at very low altitude. For self-defence and situational awareness, the F-14 is fitted with electronic countermeasure (ECM), Radar Warning Receivers (RWR) which could calculate direction and distance of enemy radars and even to differentiate between the varied types of radars, chaff/flare dispensers, a precise inertial navigation system, and fighter-to-fighter data link. These were complemented later by the installation of a GPS device to enhance navigation. Upgrades in avionics allowed the F-14 to depend less on USAF AWACS or other air assets with target designators, as during Desert Storm and the interventions in the Balkans the Tomcat depended of other air assets to identify its targets.
The Tomcat’s capacity to receive upgrades along its flight and combat capacities were made evident during its service time, as new avionics were fitted in the early 90’s, and as the Tomcat in American and Iranian hands was capable of scoring and outperforming adversarial air assets, let alone their capacity to damage and sink naval assets and AA assets of the adversary. It even managed to avoid missile fire and to retaliate under US Navy service, with the exception of the one unit that was shot down during Desert Storm.
A legendary and fearsome cat beyond the screens: naval power in the air
Grumman has had a tradition of designing and building some of the most legendary and almost unmatched naval fighters in history, like the Grumman F6F Hellcat. The F-14 Tomcat was a continuation of such traditions, being considered the best naval interceptor built ever made. It also honored its predecessor, the venerable Phantom F-4 II, as it maximized US naval power by taking it into the air. Like an enraged cat protecting its territory and even fighting back, it was able to defend the aircraft carrier groups and the airspace it was ordered to defend, and even to strike back against its aggressor when needed. Its sole presence was so imposing that after Iraqi air assets suffered heavily at the hands of the Tomcat with both the U.S. and Iran, they usually elected to flee when Tomcats were detected. But like a cat ambushing its prey, the enemy air assets fled from the Tomcat only to be destroyed by other fighters. The Libyans and Syrians who opened fire with their SAM missiles against the Tomcat had to watch in shock how the Tomcats paid them back either by attacking the AA themselves or by directing fire against such positions. What is more astonishing is that losses from SAMs were almost zero, with only one F-14 lost during Desert Storm. In other incidents, the missiles never scored a hit. The Tomcat also let its might to be felt during the series of crises between the US and Libya in the 80’s, destroying 4 fighters and delivering a heavy blow to Libyan naval assets and AA artillery. Even downgraded versions of the Tomcat, facing limited supplies and logistics, managed to yield very impressive performance. During the Iran-Iraq War it scored a large number of air kills with few losses of its own, evidencing that even with trimmed claws, it was able to terrify and eliminate its prey.

But the F-14 was also able to impose itself without firing a single shot. When not hunting, it was able to guard the skies and waters it was tasked to protect. It managed to monitor the surroundings of a hijacked cruise line ships, and to force an airliner carrying the terrorists who hijacked the vessel to land in a base where they were apprehended. It also enforced the no-fly zone over Iraqi skies after the First Gulf War and punished the Serbians hard along with other air assets during the Kosovo intervention. It also intercepted aircraft that were a serious threat for its aircraft carriers. The Tomcat was also an avid sentinel, as it executed very effective and successful surveillance of enemy territory and assets.
The Tomcat was further immortalized in the movie Top Gun, where it was the main star of the film. Despite this well-deserved fame and exceptional performance, the Tomcat saw service only until the early days of the 21st century, as it was deemed “outdated” given its age, and was admittedly very expensive to maintain, operate, and upgrade. Like the F-15, it was a product of the experiences the US faced during the Vietnam War. Considering the performance the Tomcat had and its very active service throughout its career, it fulfilled its purpose. If the Tomcat were further modernized with the proposed versions by Grumman, it could have been an overhauled Cold War-era air asset still able to deliver a powerful punch in the modern era. Yet financial restrictions and the emergence of new technologies doomed this fighter to be retired from service sooner than its half-brother the F-15. The mark it left in aviation and history will be hardly matched in the future: many remain as monuments or museum pieces, as a memory from a bygone era. The remaining Tomcats still in service are those of Iran as of this writing.

Variants

  • F-14 Prototypes (YF-14A) – The first 12 F-14A were used initially as prototypes. Two were lost during trials.
  • F-14A – It is the first basic version of the Tomcat, powered by two Pratt & Whitney TF-30-P412A turbofan engines, and equipped with the AWG-9 radar for the AIM-54 Phoenix missiles originally intended for the F-111B. This version received upgrades in electronics, such as AN/ALR-67 Countermeasure Warning and Control System (CWCS), a LANTIRN pod and Programmable Tactical Information Display, improved engines, and a Digital Flight Control System which enhanced flight safety and control in the 90’s, and new precision strike munitions. 478 F-14A models were delivered to the US Navy, with 79 delivered to Iran. The 80th F-14A intended for Iran was delivered to the US Navy instead. There were plans for replacing the TARPS pod with a TARPS Digital Imaging System.
  • F-14B (or F-14+ / F-14B Upgrade or “Bombcat”) – Both an upgraded version of the F-14A and also a very limited new-built version of the same airframe, initially denominated as F-14A+. The previous engine was replaced with new General Electric F-110-GE-400 engines, enhancing capability and maneuverability while eliminating throttle restrictions or engine trimming, and even the need for afterburner launches. The avionics were similar to that of the F-14A except in the newly acquired advanced ALR-67 Radar Homing and Warning (RHAW). Further avionics were fitted during a life extension and upgrade program, including: Fatigue Engine Monitoring System, AN/ALR-67 Countermeasure Warning and Control System, Gun Gas Purge redesign, Direct Lift Control/Approach Power Compensator, AN/AWG-15F Fire Control System, Engine Door Tension Fittings and an Embedded GPS Inertial (EGI) navigation system. Other upgrades comprised a MIL-STD-1553B Digital Multiplex Data Bus, programmable multi-Display indicator group, another AN/AWG-15H fire control system, a AN/ALR-67D(V)2 Radar Warning Receiver, and Mission Data Loader, among others. It took part in the 1991 Gulf War. Further upgrades packages made the airplane to be denominated also a F-14B Upgrade “Bombcat”. 48 F-14A airframes were upgraded to the F-14B standard, while 38 new F-14B examples were manufactured. The upgraded airframes were denominated as F-14B after a proposed enhanced F-14B interceptor was rejected.
  • F-14D Super Tomcat – This was the final version of the legendary Tomcat, after the F-14B version was restricted by the Navy, prompting further modifications and upgrades to existing airframes and building some new ones under this standard. It was powered by 2 General Electric F-110-GE-400 engines, which provides the fighter with a higher top speed, improved thrust and quicker response. It also provided more endurance and striking range, increased climb rate and no need to use afterburner, although safety concerns were the main reason for this. New avionics were installed in this version, including a more powerful AN/APG-71 radar, better controls and digital displays that facilitates better control and navigation by automation and simplicity, decreased Weapon Replaceable Assemblies (WRA), new signal processors, data processors, receivers and antenna. IRSTS and the Air Force’s Joint Tactical Information Distribution System (JTIDS) were installed, enhancing security of digital data and voice communication and providing accurate navigation capabilities. A proposed new computer software to allow operation with AIM-120 AMRAAM missiles was considered but not implemented. In the mid 2000’s, a Remotely Operated Video Enhanced Receiver (ROVER III) upgrade was fitted in some F-14D airframes. 37 new units were built and delivered, while 18 F-14A were modified to the new standard. This was the most capable and powerful version of the Tomcat.
  • F-14B interceptor versions and F-14C – The F-14B was intended to be an enhanced version of the previous F-14A with better Pratt & Whitney F-401 turbofan engines that was rejected. The F-14C was a proposed enhanced version of the F-14B (or F-14A+ for clarity) with better avionics and weapons, better radar and fire control systems. Although rejected, many of the intended improvements were later on incorporated in other operational versions. A proposed enhanced interception version based on the F-14B to replace the Convair F-106 Delta Dart was also cancelled.
    F-14D Super Tomcat (proposed) improved versions
    These were proposed versions of the F-14D by Grumman to the US Navy and Congress, which were ultimately rejected.
  • F-14D Quickstrike – A proposed enhanced version of the F-14D Super Tomcat fitted with navigational and targeting PODS, additional hardpoints and a radar with ground-attack capacities, intended to replace the then retiring Grumman A-6 Intruder.
  • F-14D Super Tomcat 21 – As the Quickstrike was rejected by the US Congress, Grumman proposed the Super Tomcat 21 version as a cheaper version to the Navy Advanced Tactical Fighter programme. Among the proposed improvements were a better AN/APG-71 radar, new and more powerful General Electric F-100-129 engines capable of providing supercuise speeds of up to 1.3 Mach and having thrust vectoring nozzles, along enhanced control surfaces and fuel capacity. They would have improved takeoff and landing approaches at lower speeds.
  • F-14 Attack Super Tomcat – It was reportedly the last of the Super Cat proposed enhanced versions, with even more improvements in control surfaces, fuel capacity and an Active Electronically Scanned Array (AESA) radar from the also cancelled McDonnell-Douglas A-12 Avenger II attacker.
  • F-14 Advanced Strike Fighter (ASF) – Another rejected proposed version proposed under the Navy Advanced Tactical Fighter programme, as it was deemed too costly. The Navy then decided to pursue the F/A-18E/F Super Hornet.

Operators

  • United States of America
    The US Navy was the main operator of the Tomcat, which began operating it in 1974 in squadrons VF-1 “Wolfpack” and VF-2 “Bounty Hunters” embarked in the aircraft carrier USS Enterprise. It began operations during the American evacuation of Saigon, being also very active in performing fleet defence interceptions especially in the North Atlantic, escorting many Soviet bombers and maritime reconnaissance airplanes. During the Lebanese Civil War it executed combat air patrols and TARPS missions to detect targets for naval gun fire. Noteworthy to point out that it began its career also as a photo-reconnaissance platform, as it replaced the RA-5C Vigilante and RF-8G Crusaders in such missions. Tomcats were attacked by Syrian air assets and AA without any losses and often fleeing once engaged by the F-14s. It also had a very limited role during the failed operation to free the American hostages in Iran.
    Libya and the Mediterranean Sea was one of the areas where US Navy-operated Tomcats saw intensive action, as incidents and tensions between the US and Libya were common during the 80’s. The F-14s contained and pushed back Libyan air assets, as they managed to shoot down 2 Sukhoi Su-22 Fitters and 2 MiG-23 Floggers, and even to outmaneuver 2 incoming MiG-25 Foxbats. They also managed to destroy two Libyan naval units and damage another two, whilst additionally taking out several SAM sites. It was during these incidents that the F-14 proved its value and capacities, by successfully defending the aircraft carrier group, avoiding enemy fire and even returning fire. The F-14s were also active in Somalia, where they were attacked by mistake, and in Grenada, where they supported intervention on the island. The F-14 also had a remarkable anti-terrorist action, as it monitored activity near the hijacked Italian cruise Achille Lauro, and then managed to intercept the Egyptian airliner carrying the terrorists that hijacked the cruise ship, forcing it to land at a NATO air base in Italy, where the terrorists were apprehended by Italian and American security forces.
    The Persian Gulf was another area where the US Navy Tomcats saw a good share of action, with the combat air patrols and escort missions it provided to US air and naval assets, as well as with fighter cover during two retaliatory operations after Iran attacked and threatened commercial and US Navy vessels. With the First Gulf War, Tomcats executed combat air patrols protecting allied forces in the area and preventing a potential Iraqi incursion into Saudi Arabia, along with escorting attack aircraft, long range defence of naval assets, combat air patrols and TARPS patrols. Tomcats also identified individual SCUD missile-launchers. During this conflict, a single F-14 was shot down by a SAM missile, with one of the crew falling prisoner to the Iraqis. The F-14 managed to score a single air kill, a Mi-8 helicopter, as its sole presence usually prompted Iraqi air assets to flee, only to be shot by other American air assets in the area, such as the F-15. In the period between the 1990 and 2003 wars, it enforced the no-fly zone and took part in punitive air strikes against Iraqi assets as well, using advanced ordnance like GBU-24 Paveway III and GBU-10/16/24 laser-guided bombs, and making use of the LANTIRN pod and night vision technology for the first time. Further CAS and strike missions were executed during the Second Gulf War in 2003 and afterwards, using JDAMS bombs for the first time against important military and governmental targets, acting also as Forward Air Controllers for other warplanes. In Afghanistan they had similar missions, spearheading Operation Enduring Freedom and taking off from the Indian Ocean in some of the longest range missions for Tomcats.
    And a final area where the Tomcats saw considerable action was in the Balkans, where they used laser-guided bombs, conducted combat air patrols, escorts, strike missions, Forward Air Controllers and TARPS missions. As they were not fitted with LANTIRN pods, F/A-18s had to assist in pinpointing the designated targets.
    The first US Navy female pilot had her first flight in an F-14 Tomcat.
    The US Navy retired the F-14 from service in 2006, with its role being taken now by the F/A-18E/F Super Hornet.
  • Iran
    Iran is the only foreign operator of the F-14 Tomcat, as it received 79 units in the late 70’s thanks to its strategic alliance with the US in the region during the Cold War and up until the Iranian Revolution of 1979. They saw extensive action in the 1980-1988 Iran-Iraq war, engaging Iraqi air assets on numerous occasions. It is reported that the Iranian Tomcats scored 160 air kills, which included: 58 MiG-23, 33 Dassault Mirage F-1, 23 MiG-21, 23 Su-20 and Su-22, 9 Mig-25, 5 Tu-22, 2 MiG-27, one MiL Mi-24 helicopter, 1 Dassault Mirage 5, 1 B-6D (Xian H-6), 1 Aerospatiale Super Frelon helicopter, and two unspecified aircraft. The only losses in combat were 3 Tomcats downed by Iraqi air assets and 4 losses from SAMs, 2 that disappeared and 7 that were lost to non-combat incidents. During this conflict, the F-14 Tomcat demonstrated its capabilities, at the point of intimidating and deterring the Iraqi Air Force, and despite being a downgraded version of the Tomcat in terms of avionics. By 2015, an estimated of 20-30 airframes remained on active duty with the Islamic Republic Iran Air Force (IRIAF), and were reported to escort Russian Tu-95 Bear bombers carrying out bombing against ISIS terrorists’ positions.

 

F-14D Specifications

Wingspan  64 ft / 19.55 m (wings extended)

38 ft / 11.65 (wings swept)

Length  62 ft / 19.1 m
Height  16 ft / 4.88 m
Wing Area  565 ft² / 52.49 m²
Engine  2 x General Electric F-100-GE-400 afterburning turbofans
Maximum Take-Off Weight  74,350 lb / 33,720 kg
Empty Weight  43,735 lb / 19,838 kg
Loaded Weight  61,000 lb / 27,700 kg
Climb Rate  over 45,000 ft/min (230 m/s)
Maximum Speed  At high altitude: Mach 2.34 ( 1,544 mph / 2,485 kmh )
Range  575 mi / 926 km for combat radius; 1,840 / 2,960 for ferry
Maximum Service Ceiling  50,000 ft / 15,200 m
Crew  2 (pilot and radar intercept officer)
Armament
  • 1 X 20mm M61A1 Vulcan 6-barrel rotary cannon
  • 10 hardpoints – six under the fuselage, two under the nacelles, and two on the wing gloves, all allowing up to 6600 kg (14,500 lb) of ordnance and fuel tanks. The payload was varied in deployment and type, usually being 6 AIM-7 Sparrow, 4 AIM-9 Sidewinder and/or 6 AIM-54 Phoenix (and MIM-23 Hawk in the case of the IRIAF). Up to 6622 kg (14,599 lb) of air-to-ground were also carried, including Mk 80 free-fall iron bombs, Mk 20 Rockeye II cluster bombs, Paveway laser-guided bombs, and JDAM precision-guided munition bombs. 2x 267 1010 l fuel tanks were carried as well.
  • The fighter/naval interceptor had avionics both part of its structure and carried in the hardpoints. Among those at the hardpoints were the TARPS and the LANTIRN targeting pods. Among its onboard avionics were a Hughes AN/APG-71 radar, an AN/ASN-130 inertial navigation system (INS), Infra-Red Search and Track (IRST) and Track Control System (TCS). It also had a AN/ALR-45 and AL/ALR-67 (F-14D) RWR, a AN/ALQ-167 ECM pod and a AN/ALQ-50 towed decoy (the two last ones in the F-14D).

Gallery

1.-F-14A-VF-1-BuNo-162597_03 Tomcat
F-14A VF-1 “Wolfpack” BuNo 162597 circa 1987 loaded with AIM-9s, AIM-7s, and AIM-54s
F-14B Tomcat aircraft from VF-101 circa 2004
An F-14D Tomcat attached to the “Bounty Hunters” of VF-2 makes a sharp pull-up in full afterburner circa 2003
Tomcat of VF-101 circa at an airshow 2004
F-14D Tomcat aircraft of VF-124 flying over part of California circa 1991
F-14D Tomcat makes a near supersonic fly-by above the flight deck of the USS Theodore Roosevelt on July 28, 2006 just prior to its retirement in September 2006
Test fire of AIM-54C Phoenix from F-14D “Jolly Rogers” circa 2002
F-14D on the deck of the USS Harry S. Truman in the Persian Gulf circa 2005
F-15D Tomcat on the deck of the USS John C. Stennis
F-14D Super Tomcat maneuvers in the Persian Gulf circa 2005
Tomcat in formation with Croatian Air Force MiG 21s circa 2002
Tomcat on display, note it’s low height profile to facilitate carrier operations and storage
A Tomcat with its wings fully extended for low speed maneuvers
F-14A as viewed from rear, note the space between the engines affording the Tomcat significant lift from its airframe
The 6 barrel, 20mm vulcan cannon with its panels removed

Sources

Berger, R (Ed.). Aviones [Flugzeuge, Vicenç Prat, trans.]. Colonia, Alemania: Naumann & Göbel Verlagsgessellschaft mbH., Chambers, J. R. (2000). Partners in Freedom. Contributions of the Langley Research Center to US Military Aircraft of the 1990’s (NASA monograph NASA SP-2000-4519). NASA History Division: Washington DC, USA.,  Cooper, T. (2006). Persian Cats. Air&Space., Donald. D. (2009). Aviones Militares, Guia Visual [Military Aircraft. Visual Guide, Seconsat, trans.]. Madrid, Spain: Editorial Libsa (Original work published in 2008)., Dudney, R. S, &. Boyne, W. J. (January 2015). Airpower Classics. F-14 Tomcat. Air Force Magazine, 98 (1), 76., GlobalSecurity.org (2016). F-14 Tomcat. GlobalSecurity.org., Goebel, G. (2016). [1.0] Creating the Tomcat. AirVectors.net., Goebel, G. (2016). [2.0] Iranian Tomcats / Tomcat Improvements. AirVectors.net., Lemoin, J. (2002). Fighter Planes. 1960-2002., N.R.P. (2015). Origins – The Story of the Legendary F-14 Tomcat., National Naval Aviation Museum (2016). F-14A Tomcat. National Naval Aviation Museum., Sharpe, M (2001). Jets de Ataque y Defensa [Attack and Interceptor Jets, Macarena Rojo, trans.]. Madrid, Spain: Editorial LIBSA (Original work published in 2001)., Sponsler, G. C., Gignoux, D., Dare, E., & Rubin N. N. (1973). The F-4 and the F-14., U.S. General Accounting Office. (1972). The F-14 Aircraft., F-14 Tomcat operational history. (2017, June 14). In Wikipedia, The Free Encyclopedia.Grumman F-14 Tomcat. (2017, June 19). In Wikipedia, The Free EncyclopediaImages: Tomcat Gun by Jeff Kubina / CC BY-SA 2.0, Tomcat Rear Engines by kevinofsydney / CC BY 2.0Tomcat Wings Extended by D. Miller / CC BY 2.0Tomcat Display by Eric Kilby / CC BY-SA 2.0Side Profile Views by Ed Jackson – Artbyedo.com,  Note: Images not credited are in the Public Domain