Liquid-propellant rocket

Last updated
A simplified diagram of a liquid-propellant rocket.
Liquid rocket fuel.
Oxidizer.
Pumps carry the fuel and oxidizer.
The combustion chamber mixes and burns the two liquids.
Combustion product gasses enter the nozzle through a throat.
Exhaust exits the rocket. Liquid-Fuel Rocket Diagram.svg
A simplified diagram of a liquid-propellant rocket.
  1. Liquid rocket fuel.
  2. Oxidizer.
  3. Pumps carry the fuel and oxidizer.
  4. The combustion chamber mixes and burns the two liquids.
  5. Combustion product gasses enter the nozzle through a throat.
  6. Exhaust exits the rocket.

A liquid-propellant rocket or liquid rocket utilizes a rocket engine burning liquid propellants. (Alternate approaches use gaseous or solid propellants.) Liquids are desirable propellants because they have reasonably high density and their combustion products have high specific impulse (Isp). This allows the volume of the propellant tanks to be relatively low.

Contents

Types

Liquid rockets can be monopropellant rockets using a single type of propellant, or bipropellant rockets using two types of propellant. Tripropellant rockets using three types of propellant are rare. Liquid oxidizer propellants are also used in hybrid rockets, with some of the advantages of a solid rocket. Bipropellant liquid rockets use a liquid fuel such as liquid hydrogen or RP-1, and a liquid oxidizer such as liquid oxygen. The engine may be a cryogenic rocket engine, where the fuel and oxidizer, such as hydrogen and oxygen, are gases which have been liquefied at very low temperatures.

Most designs of liquid rocket engines are throttleable for variable thrust operation. Some allow control of the propellant mixture ratio (ratio at which oxidizer and fuel are mixed). Some can be shut down and, with a suitable ignition system or self-igniting propellant, restarted.

Hybrid rockets apply a liquid or gaseous oxidizer to a solid fuel. [1] :354–356

Advantages and disadvantages

The use of liquid propellants has a number of advantages:

Bipropellant liquid rockets are simple in concept but due to high temperatures and high speed moving parts, very complex in practice. Space Shuttle Main Engine Maintenance - GPN-2000-000548.jpg
Bipropellant liquid rockets are simple in concept but due to high temperatures and high speed moving parts, very complex in practice.

Use of liquid propellants can also be associated with a number of issues:

Principle of operation

Liquid rocket engines have tankage and pipes to store and transfer propellant, an injector system and one or more combustion chambers with associated nozzles.

Typical liquid propellants have densities roughly similar to water, approximately 0.7–1.4g/cm3. An exception is liquid hydrogen which has a much lower density, while requiring only relatively modest pressure to prevent vaporization. The density and low pressure of liquid propellants permit lightweight tankage: approximately 1% of the contents for dense propellants and around 10% for liquid hydrogen. The increased tank mass is due to liquid hydrogen's low density and the mass of the required insulation.

For injection into the combustion chamber, the propellant pressure at the injectors needs to be greater than the chamber pressure. This is often achieved with a pump. Suitable pumps usually use centrifugal turbopumps due to their high power and light weight, although reciprocating pumps have been employed in the past. Turbopumps are usually lightweight and can give excellent performance; with an on-Earth weight well under 1% of the thrust. Indeed, overall thrust to weight ratios including a turbopump have been as high as 155:1 with the SpaceX Merlin 1D rocket engine and up to 180:1 with the vacuum version. [5] Instead of a pump, some designs use a tank of a high-pressure inert gas such as helium to pressurize the propellants. These rockets often provide lower delta-v because the mass of the pressurant tankage reduces performance. In some designs for high altitude or vacuum use the tankage mass can be acceptable.

The major components of a rocket engine are therefore the combustion chamber (thrust chamber), pyrotechnic igniter, propellant feed system, valves, regulators, propellant tanks and the rocket engine nozzle. For feeding propellants to the combustion chamber, liquid-propellant engines are either pressure-fed or pump-fed, with pump-fed engines working in a variety of engine cycles.

Pressurization

Liquid propellants are often pumped into the combustion chamber with a lightweight centrifugal turbopump. Recently some aerospace companies have used electric pumps with batteries for this. In simpler small engines an inert gas stored in a tank at a high pressure is sometimes used instead of pumps to force propellants into the combustion chamber. These engines may have a higher mass ratio, but are usually more reliable, and are therefore used widely in satellites for orbit maintenance. [1]

Propellants

Thousands of combinations of fuels and oxidizers have been tried over the years. Some of the more common and practical ones are:

Cryogenic

One of the most efficient mixtures, oxygen and hydrogen, suffers from the extremely low temperatures required for storing liquid hydrogen (around 20 K or −253.2 °C or −423.7 °F) and very low fuel density (70 kg/m3 or 4.4 lb/cu ft, compared to RP-1 at 820 kg/m3 or 51 lb/cu ft), necessitating large tanks that must also be lightweight and insulating. Lightweight foam insulation on the Space Shuttle external tank led to the Space Shuttle Columbia's destruction, as a piece broke loose, damaged its wing and caused it to break up on atmospheric reentry.

Liquid methane/LNG has several advantages over LH2. Its performance (max. specific impulse) is lower than that of LH2 but higher than that of RP1 (kerosene) and solid propellants, and its higher density, similarly to other hydrocarbon fuels, provides higher thrust to volume ratios than LH2, although its density is not as high as that of RP1. [7] This makes it specially attractive for reusable launch systems because higher density allows for smaller motors, propellant tanks and associated systems. [6] LNG also burns with less or no soot (less or no coking) than RP1, which eases reusability when compared with it, and LNG and RP1 burn cooler than LH2 so LNG and RP1 do not deform the interior structures of the engine as much. This means that engines that burn LNG can be reused more than those that burn RP1 or LH2. Unlike engines that burn LH2, both RP1 and LNG engines can be designed with a shared shaft with a single turbine and two turbopumps, one each for LOX and LNG/RP1. [7] In space, LNG does not need heaters to keep it liquid, unlike RP1. [8] LNG is less expensive, being readily available in large quantities. It can be stored for more prolonged periods of time, and is less explosive than LH2. [6]

Semi-cryogenic

Non-cryogenic/storable/hypergolic

The NMUSAF's Me 163B Komet rocket plane Messerschmitt Me 163B USAF.jpg
The NMUSAF's Me 163B Komet rocket plane

Many non-cryogenic bipropellants are hypergolic (self igniting).

For storable ICBMs and most spacecraft, including crewed vehicles, planetary probes, and satellites, storing cryogenic propellants over extended periods is unfeasible. Because of this, mixtures of hydrazine or its derivatives in combination with nitrogen oxides are generally used for such applications, but are toxic and carcinogenic. Consequently, to improve handling, some crew vehicles such as Dream Chaser and Space Ship Two plan to use hybrid rockets with non-toxic fuel and oxidizer combinations.

Injectors

The injector implementation in liquid rockets determines the percentage of the theoretical performance of the nozzle that can be achieved. A poor injector performance causes unburnt propellant to leave the engine, giving poor efficiency.

Additionally, injectors are also usually key in reducing thermal loads on the nozzle; by increasing the proportion of fuel around the edge of the chamber, this gives much lower temperatures on the walls of the nozzle.

Types of injectors

Injectors can be as simple as a number of small diameter holes arranged in carefully constructed patterns through which the fuel and oxidizer travel. The speed of the flow is determined by the square root of the pressure drop across the injectors, the shape of the hole and other details such as the density of the propellant.

The first injectors used on the V-2 created parallel jets of fuel and oxidizer which then combusted in the chamber. This gave quite poor efficiency.

Injectors today classically consist of a number of small holes which aim jets of fuel and oxidizer so that they collide at a point in space a short distance away from the injector plate. This helps to break the flow up into small droplets that burn more easily.

The main types of injectors are

The pintle injector permits good mixture control of fuel and oxidizer over a wide range of flow rates. The pintle injector was used in the Apollo Lunar Module engines (Descent Propulsion System) and the Kestrel engine, it is currently used in the Merlin engine on Falcon 9 and Falcon Heavy rockets.

The RS-25 engine designed for the Space Shuttle uses a system of fluted posts, which use heated hydrogen from the preburner to vaporize the liquid oxygen flowing through the center of the posts [10] and this improves the rate and stability of the combustion process; previous engines such as the F-1 used for the Apollo program had significant issues with oscillations that led to destruction of the engines, but this was not a problem in the RS-25 due to this design detail.

Valentin Glushko invented the centripetal injector in the early 1930s, and it has been almost universally used in Russian engines. Rotational motion is applied to the liquid (and sometimes the two propellants are mixed), then it is expelled through a small hole, where it forms a cone-shaped sheet that rapidly atomizes. Goddard's first liquid engine used a single impinging injector. German scientists in WWII experimented with impinging injectors on flat plates, used successfully in the Wasserfall missile.

Combustion stability

To avoid instabilities such as chugging, which is a relatively low speed oscillation, the engine must be designed with enough pressure drop across the injectors to render the flow largely independent of the chamber pressure. This pressure drop is normally achieved by using at least 20% of the chamber pressure across the injectors.

Nevertheless, particularly in larger engines, a high speed combustion oscillation is easily triggered, and these are not well understood. These high speed oscillations tend to disrupt the gas side boundary layer of the engine, and this can cause the cooling system to rapidly fail, destroying the engine. These kinds of oscillations are much more common on large engines, and plagued the development of the Saturn V, but were finally overcome.

Some combustion chambers, such as those of the RS-25 engine, use Helmholtz resonators as damping mechanisms to stop particular resonant frequencies from growing.

To prevent these issues the RS-25 injector design instead went to a lot of effort to vaporize the propellant prior to injection into the combustion chamber. Although many other features were used to ensure that instabilities could not occur, later research showed that these other features were unnecessary, and the gas phase combustion worked reliably.

Testing for stability often involves the use of small explosives. These are detonated within the chamber during operation, and causes an impulsive excitation. By examining the pressure trace of the chamber to determine how quickly the effects of the disturbance die away, it is possible to estimate the stability and redesign features of the chamber if required.

Engine cycles

For liquid-propellant rockets, four different ways of powering the injection of the propellant into the chamber are in common use. [11]

Fuel and oxidizer must be pumped into the combustion chamber against the pressure of the hot gasses being burned, and engine power is limited by the rate at which propellant can be pumped into the combustion chamber. For atmospheric or launcher use, high pressure, and thus high power, engine cycles are desirable to minimize gravity drag. For orbital use, lower power cycles are usually fine.

Pressure-fed cycle
The propellants are forced in from pressurised (relatively heavy) tanks. The heavy tanks mean that a relatively low pressure is optimal, limiting engine power, but all the fuel is burned, allowing high efficiency. The pressurant used is frequently helium due to its lack of reactivity and low density. Examples: AJ-10, used in the Space Shuttle OMS, Apollo SPS, and the second stage of the Delta II.
Electric pump-fed
An electric motor, generally a brushless DC electric motor, drives the pumps. The electric motor is powered by a battery pack. It is relatively simple to implement and reduces the complexity of the turbomachinery design, but at the expense of the extra dry mass of the battery pack. Example engine is the Rutherford designed and used by Rocket Lab.
Gas-generator cycle
A small percentage of the propellants are burnt in a preburner to power a turbopump and then exhausted through a separate nozzle, or low down on the main one. This results in a reduction in efficiency since the exhaust contributes little or no thrust, but the pump turbines can be very large, allowing for high power engines. Examples: Saturn V's F-1 and J-2, Delta IV's RS-68, Ariane 5's HM7B, Falcon 9's Merlin.
Tap-off cycle
Takes hot gases from the main combustion chamber of the rocket engine and routes them through engine turbopump turbines to pump propellant, then is exhausted. Since not all propellant flows through the main combustion chamber, the tap-off cycle is considered an open-cycle engine. Examples include the J-2S and BE-3.
Expander cycle
Cryogenic fuel (hydrogen, or methane) is used to cool the walls of the combustion chamber and nozzle. Absorbed heat vaporizes and expands the fuel which is then used to drive the turbopumps before it enters the combustion chamber, allowing for high efficiency, or is bled overboard, allowing for higher power turbopumps. The limited heat available to vaporize the fuel constrains engine power. Examples: RL10 for Atlas V and Delta IV second stages (closed cycle), H-II's LE-5 (bleed cycle).
Staged combustion cycle
A fuel- or oxidizer-rich mixture is burned in a preburner and then drives turbopumps, and this high-pressure exhaust is fed directly into the main chamber where the remainder of the fuel or oxidizer undergoes combustion, permitting very high pressures and efficiency. Examples: SSME, RD-191, LE-7.
Full-flow staged combustion cycle
Fuel- and oxidizer-rich mixtures are burned in separate preburners and driving the turbopumps, then both high-pressure exhausts, one oxygen rich and the other fuel rich, are fed directly into the main chamber where they combine and combust, permitting very high pressures and high efficiency. Example: SpaceX Raptor.

Engine cycle tradeoffs

Selecting an engine cycle is one of the earlier steps to rocket engine design. A number of tradeoffs arise from this selection, some of which include:

Tradeoff comparison among popular engine cycles
Cycle type
Gas generatorExpander cycleStaged-combustionPressure-fed
AdvantagesSimple; low dry mass; allows for high power turbopumps for high thrustHigh specific impulse; fairly low complexityHigh specific impulse; high combustion chamber pressures allowing for high thrustSimple; no turbopumps; low dry mass; high specific impulse
DisadvantagesLower specific impulseMust use cryogenic fuel; heat transfer to the fuel limits available power to the turbine and thus engine thrustGreatly increased complexity &, therefore, mass (more-so for full-flow)Tank pressure limits combustion chamber pressure and thrust; heavy tanks and associated pressurization hardware

Cooling

Injectors are commonly laid out so that a fuel-rich layer is created at the combustion chamber wall. This reduces the temperature there, and downstream to the throat and even into the nozzle and permits the combustion chamber to be run at higher pressure, which permits a higher expansion ratio nozzle to be used which gives a higher ISP and better system performance. [12] A liquid rocket engine often employs regenerative cooling, which uses the fuel or less commonly the oxidizer to cool the chamber and nozzle.

Ignition

Ignition can be performed in many ways, but perhaps more so with liquid propellants than other rockets a consistent and significant ignitions source is required; a delay of ignition (in some cases as small as a few tens of milliseconds) can cause overpressure of the chamber due to excess propellant. A hard start can even cause an engine to explode.

Generally, ignition systems try to apply flames across the injector surface, with a mass flow of approximately 1% of the full mass flow of the chamber.

Safety interlocks are sometimes used to ensure the presence of an ignition source before the main valves open; however reliability of the interlocks can in some cases be lower than the ignition system. Thus it depends on whether the system must fail safe, or whether overall mission success is more important. Interlocks are rarely used for upper, uncrewed stages where failure of the interlock would cause loss of mission, but are present on the RS-25 engine, to shut the engines down prior to liftoff of the Space Shuttle. In addition, detection of successful ignition of the igniter is surprisingly difficult, some systems use thin wires that are cut by the flames, pressure sensors have also seen some use.

Methods of ignition include pyrotechnic, electrical (spark or hot wire), and chemical. Hypergolic propellants have the advantage of self igniting, reliably and with less chance of hard starts. In the 1940s, the Russians began to start engines with hypergols, to then switch over to the primary propellants after ignition. This was also used on the American F-1 rocket engine on the Apollo program.

Ignition with a pyrophoric agent: Triethylaluminium ignites on contact with air and will ignite and/or decompose on contact with water, and with any other oxidizer—it is one of the few substances sufficiently pyrophoric to ignite on contact with cryogenic liquid oxygen. The enthalpy of combustion, ΔcH°, is −5,105.70 ± 2.90 kJ/mol (−1,220.29 ± 0.69 kcal/mol). Its easy ignition makes it particularly desirable as a rocket engine ignitor. May be used in conjunction with triethylborane to create triethylaluminum-triethylborane, better known as TEA-TEB.

History

Russia–Soviet Union

Rocket 09 (left) and 10 (GIRD-09 and GIRD-X). Museum of Cosmonautics and Rocket Technology; St. Petersburg. Raketa 09 i 10.jpg
Rocket 09 (left) and 10 (GIRD-09 and GIRD-X). Museum of Cosmonautics and Rocket Technology; St. Petersburg.

The idea of a liquid rocket as understood in the modern context first appeared in 1903 in the book Exploration of the Universe with Rocket-Propelled Vehicles, [13] by the Russian rocket scientist Konstantin Tsiolkovsky.The magnitude of his contribution to astronautics is astounding, including the Tsiolkovsky rocket equation, multi staged rockets and using liquid oxygen and liquid hydrogen in liquid propellant rockets. [14] Tsiolkovsky influenced later rocket scientists throughout Europe, like Wernher von Braun. Soviet search teams at Peenemünde found a German translation of a book by Tsiolkovsky of which "almost every page...was embellished by von Braun's comments and notes." [15] Leading Soviet rocket-engine designer Valentin Glushko and rocket designer Sergey Korolev studied Tsiolkovsky's works as youths [16] and both sought to turn Tsiolkovsky's theories into reality. [17]

From 1929 to 1930 in Leningrad Glushko pursued rocket research at the Gas Dynamics Laboratory (GDL), where a new research section was set up for the study of liquid-propellant and electric rocket engines. This resulted in the creation of ORM (from "Experimental Rocket Motor" in Russian) engines ORM-1  [ ru ] to ORM-52  [ ru ]. [18] A total of 100 bench tests of liquid-propellant rockets were conducted using various types of fuel, both low and high-boiling and thrust up to 300 kg was achieved. [19] [18]

During this period in Moscow Fredrich Tsander, a scientist and inventor was designing and building liquid rocket engines which ran on compressed air and gasoline. Tsander used it to investigate high-energy fuels including powdered metals mixed with gasoline. In September 1931 Tsander formed the Moscow based 'Group for the Study of Reactive Motion', [20] better known by its Russian acronym "GIRD". [21] In May 1932, Sergey Korolev replaced Tsander as the head of GIRD. Mikhail Tikhonravov launched the first Soviet liquid propelled rocket, fueled by liquid oxygen and jellied gasoline, the GIRD-9, took place on 17 August 1933, which reached an altitude of 400 metres (1,300 ft). [22] In January 1933 Tsander began development of the GIRD-X rocket. This design burned liquid oxygen and gasoline and was one of the first engines to be regeneratively cooled by the liquid oxygen, which flowed around the inner wall of the combustion chamber before entering it. Problems with burn-through during testing prompted a switch from gasoline to less energetic alcohol. The final missile, 2.2 metres (7.2 ft) long by 140 millimetres (5.5 in) in diameter, had a mass of 30 kilograms (66 lb), and it was anticipated that it could carry a 2 kilograms (4.4 lb) payload to an altitude of 5.5 kilometres (3.4 mi). [23] The GIRD X rocket was launched on 25 November 1933 and flew to a height of 80 meters. [24]

In 1933 GDL and GIRD merged and became the Reactive Scientific Research Institute (RNII). At RNII Gushko continued the development of liquid propellant rocket engines ОРМ-53 to ОРМ-102, with ORM-65  [ ru ] powering the RP-318 rocket-powered aircraft. [18] In 1938 Leonid Dushkin replaced Glushko and continued development of the ORM engines, including the engine for the rocket powered interceptor, the Bereznyak-Isayev BI-1. [25] At RNII Tikhonravov worked on developing oxygen/alcohol liquid-propellant rocket engines. [26] Ultimately liquid propellant rocket engines were given a low priority during the late 1930s at RNII, however the research was productive and very important for later achievements of the Soviet rocket program. [27]

Peru

Pedro Paulet's Avion Torpedo of 1902, featuring a canopy fixed to a delta tiltwing for horizontal or vertical flight. Paulet Avion-Torpedo System.png
Pedro Paulet's Avion Torpedo of 1902, featuring a canopy fixed to a delta tiltwing for horizontal or vertical flight.

Peruvian Pedro Paulet, who had experimented with rockets throughout his life in Peru, wrote a letter to El Comercio in Lima in 1927, claiming he had experimented with a liquid rocket engine while he was a student in Paris three decades earlier. [28] [29] Historians of early rocketry experiments, among them Max Valier, Willy Ley, and John D. Clark, have given differing amounts of credence to Paulet's report. Valier applauded Paulet's liquid-propelled rocket design in the Verein für Raumschiffahrt publication Die Rakete, saying the engine had "amazing power" and that his plans were necessary for future rocket development. [30] Hermann Oberth would name Paulet as a pioneer in rocketry in 1965. [31] Wernher von Braun would also describe Paulet as "the pioneer of the liquid fuel propulsion motor" and stated that "Paulet helped man reach the Moon". [28] [32] [33] [34] [35] Paulet was later approached by Nazi Germany, being invited to join the Astronomische Gesellschaft to help develop rocket technology, though he refused to assist after discovering that the project was destined for weaponization and never shared the formula for his propellant. [36] [37] According to filmmaker and researcher Álvaro Mejía, Frederick I. Ordway III would later attempt to discredit Paulet's discoveries in the context of the Cold War and in an effort to shift the public image of von Braun away from his history with Nazi Germany. [38]

United States

Robert H. Goddard, bundled against the cold New England weather of March 16, 1926, holds the launching frame of his most notable invention -- the first liquid rocket. Goddard and Rocket.jpg
Robert H. Goddard, bundled against the cold New England weather of March 16, 1926, holds the launching frame of his most notable invention the first liquid rocket.

The first flight of a liquid-propellant rocket took place on March 16, 1926 at Auburn, Massachusetts, when American professor Dr. Robert H. Goddard launched a vehicle using liquid oxygen and gasoline as propellants. [39] The rocket, which was dubbed "Nell", rose just 41 feet during a 2.5-second flight that ended in a cabbage field, but it was an important demonstration that rockets utilizing liquid propulsion were possible. Goddard proposed liquid propellants about fifteen years earlier and began to seriously experiment with them in 1921. The German-Romanian Hermann Oberth published a book in 1922 suggesting the use of liquid propellants.

Germany

In Germany, engineers and scientists became enthralled with liquid propulsion, building and testing them in the late 1920s within Opel RAK, the world's first rocket program, in Rüsselsheim. According to Max Valier's account, [40] Opel RAK rocket designer, Friedrich Wilhelm Sander launched two liquid-fuel rockets at Opel Rennbahn in Rüsselsheim on April 10 and April 12, 1929. These Opel RAK rockets have been the first European, and after Goddard the world's second, liquid-fuel rockets in history. In his book "Raketenfahrt" Valier describes the size of the rockets as of 21 cm in diameter and with a length of 74 cm, weighing 7 kg empty and 16 kg with fuel. The maximum thrust was 45 to 50 kp, with a total burning time of 132 seconds. These properties indicate a gas pressure pumping. The main purpose of these tests was to develop the liquid rocket-propulsion system for a Gebrüder-Müller-Griessheim aircraft [41] under construction for a planned flight across the English channel. Also spaceflight historian Frank H. Winter, curator at National Air and Space Museum in Washington, DC, confirms the Opel group was working, in addition to their solid-fuel rockets used for land-speed records and the world's first crewed rocket-plane flights with the Opel RAK.1, on liquid-fuel rockets. [42] By May 1929, the engine produced a thrust of 200 kg (440 lb.) "for longer than fifteen minutes and in July 1929, the Opel RAK collaborators were able to attain powered phases of more than thirty minutes for thrusts of 300 kg (660-lb.) at Opel's works in Rüsselsheim," again according to Max Valier's account. The Great Depression brought an end to the Opel RAK activities. After working for the German military in the early 1930s, Sander was arrested by Gestapo in 1935, when private rocket-engineering became forbidden in Germany. He was convicted of treason to 5 years in prison and forced to sell his company, he died in 1938. [43] Max Valier's (via Arthur Rudolph and Heylandt), who died while experimenting in 1930, and Friedrich Sander's work on liquid-fuel rockets was confiscated by the German military, the Heereswaffenamt and integrated into the activities under General Walter Dornberger in the early and mid-1930s in a field near Berlin. [44] Max Valier was a co-founder of an amateur research group, the VfR, working on liquid rockets in the early 1930s, and many of whose members eventually became important rocket technology pioneers, including Wernher von Braun. Von Braun served as head of the army research station that designed the V-2 rocket weapon for the Nazis.

Drawing of the He 176 V1 prototype rocket aircraft Heinkel He 176 3-view line drawing.svg
Drawing of the He 176 V1 prototype rocket aircraft

By the late 1930s, use of rocket propulsion for crewed flight began to be seriously experimented with, as Germany's Heinkel He 176 made the first crewed rocket-powered flight using a liquid rocket engine, designed by German aeronautics engineer Hellmuth Walter on June 20, 1939. [45] The only production rocket-powered combat aircraft ever to see military service, the Me 163 Komet in 1944-45, also used a Walter-designed liquid rocket engine, the Walter HWK 109-509, which produced up to 1,700 kgf (16.7 kN) thrust at full power.

Post World War II

After World War II the American government and military finally seriously considered liquid-propellant rockets as weapons and began to fund work on them. The Soviet Union did likewise, and thus began the Space Race.

In 2010s 3D printed engines started being used for spaceflight. Examples of such engines include SuperDraco used in launch escape system of the SpaceX Dragon 2 and also engines used for first or second stages in launch vehicles from Astra, [46] Orbex, [47] [48] Relativity Space, [49] Skyrora, [50] or Launcher. [51] [52] [53]

See also

Related Research Articles

<span class="mw-page-title-main">Hybrid-propellant rocket</span> Rocket engine that uses both liquid / gaseous and solid fuel

A hybrid-propellant rocket is a rocket with a rocket motor that uses rocket propellants in two different phases: one solid and the other either gas or liquid. The hybrid rocket concept can be traced back to the early 1930s.

<span class="mw-page-title-main">Hypergolic propellant</span> Type of rocket engine fuel

A hypergolic propellant is a rocket propellant combination used in a rocket engine, whose components spontaneously ignite when they come into contact with each other.

<span class="mw-page-title-main">Rocket engine</span> Non-air breathing jet engine used to propel a missile or vehicle

A rocket engine uses stored rocket propellants as the reaction mass for forming a high-speed propulsive jet of fluid, usually high-temperature gas. Rocket engines are reaction engines, producing thrust by ejecting mass rearward, in accordance with Newton's third law. Most rocket engines use the combustion of reactive chemicals to supply the necessary energy, but non-combusting forms such as cold gas thrusters and nuclear thermal rockets also exist. Vehicles propelled by rocket engines are commonly used by ballistic missiles and rockets. Rocket vehicles carry their own oxidiser, unlike most combustion engines, so rocket engines can be used in a vacuum to propel spacecraft and ballistic missiles.

<span class="mw-page-title-main">Expander cycle</span> Rocket engine operation method

The expander cycle is a power cycle of a bipropellant rocket engine. In this cycle, the fuel is used to cool the engine's combustion chamber, picking up heat and changing phase. The now heated and gaseous fuel then powers the turbine that drives the engine's fuel and oxidizer pumps before being injected into the combustion chamber and burned.

<span class="mw-page-title-main">RS-25</span> Space Shuttle and SLS main engine

The Aerojet Rocketdyne RS-25, also known as the Space Shuttle Main Engine (SSME), is a liquid-fuel cryogenic rocket engine that was used on NASA's Space Shuttle and is used on the Space Launch System (SLS).

<span class="mw-page-title-main">Rocketdyne J-2</span> Rocket engine

The J-2, commonly known as Rocketdyne J-2, was a liquid-fuel cryogenic rocket engine used on NASA's Saturn IB and Saturn V launch vehicles. Built in the United States by Rocketdyne, the J-2 burned cryogenic liquid hydrogen (LH2) and liquid oxygen (LOX) propellants, with each engine producing 1,033.1 kN (232,250 lbf) of thrust in vacuum. The engine's preliminary design dates back to recommendations of the 1959 Silverstein Committee. Rocketdyne won approval to develop the J-2 in June 1960 and the first flight, AS-201, occurred on 26 February 1966. The J-2 underwent several minor upgrades over its operational history to improve the engine's performance, with two major upgrade programs, the de Laval nozzle-type J-2S and aerospike-type J-2T, which were cancelled after the conclusion of the Apollo program.

The highest specific impulse chemical rockets use liquid propellants. They can consist of a single chemical or a mix of two chemicals, called bipropellants. Bipropellants can further be divided into two categories; hypergolic propellants, which ignite when the fuel and oxidizer make contact, and non-hypergolic propellants which require an ignition source.

<span class="mw-page-title-main">Staged combustion cycle</span> Rocket engine operation method

The staged combustion cycle is a power cycle of a bipropellant rocket engine. In the staged combustion cycle, propellant flows through multiple combustion chambers, and is thus combusted in stages. The main advantage relative to other rocket engine power cycles is high fuel efficiency, measured through specific impulse, while its main disadvantage is engineering complexity.

<span class="mw-page-title-main">Gas-generator cycle</span> Rocket engine operation method

The gas-generator cycle, also called open cycle, is one of the most commonly used power cycles in bipropellant liquid rocket engines. Part of the unburned propellant is burned in a gas generator and the resulting hot gas is used to power the propellant pumps before being exhausted overboard, and lost. Because of this loss, this type of engine is termed open cycle.

<span class="mw-page-title-main">Aerojet M-1</span> One of the largest rocket engines to be designed

The Aerojet M-1 was one of the largest and most powerful liquid-hydrogen-fueled liquid-fuel rocket engine to be designed and component-tested. It was originally developed during the 1950s by the US Air Force. The M-1 offered a baseline thrust of 6.67 MN and an immediate growth target of 8 MN. If built, the M-1 would have been larger and more efficient than the famed F-1 that powered the first stage of the Saturn V rocket to the Moon.

The YF-75 is a liquid cryogenic rocket engine burning liquid hydrogen and liquid oxygen in a gas generator cycle. It is China's second generation of cryogenic propellant engine, after the YF-73, which it replaced. It is used in a dual engine mount in the H-18 third stage of the Long March 3A, Long March 3B and Long March 3C launch vehicles. Within the mount, each engine can gimbal individually to enable thrust vectoring control. The engine also heats hydrogen and helium to pressurize the stage tanks and can control the mixture ratio to optimize propellant consumption.

<span class="mw-page-title-main">YF-77</span> Chinese rocket engine

The YF-77 is China's first cryogenic rocket engine developed for booster applications. It burns liquid hydrogen fuel and liquid oxygen oxidizer using a gas generator cycle. A pair of these engines powers the LM-5 core stage. Each engine can independently gimbal in two planes. Although the YF-77 is ignited prior to liftoff, the LM-5's four strap-on boosters provide most of the initial thrust in an arrangement similar to the European Vulcain on the Ariane 5 or the Japanese LE-7 on the H-II.

<span class="mw-page-title-main">LE-7</span> Japanese hydrolox staged combustion rocket engine

The LE-7 and its succeeding upgrade model the LE-7A are staged combustion cycle LH2/LOX liquid rocket engines produced in Japan for the H-II series of launch vehicles. Design and production work was all done domestically in Japan, the first major (main/first-stage) liquid rocket engine with that claim, in a collaborative effort from the National Space Development Agency (NASDA), Aerospace Engineering Laboratory (NAL), Mitsubishi Heavy Industries, and Ishikawajima-Harima. NASDA and NAL have since been integrated into JAXA. However, a large part of the work was contracted to Mitsubishi, with Ishikawajima-Harima providing turbomachinery, and the engine is often referred to as the Mitsubishi LE-7(A).

<span class="mw-page-title-main">LE-5</span> Japanese hydrolox rocket engine used on the H3 upper stage

The LE-5 liquid rocket engine and its derivative models were developed in Japan to meet the need for an upper stage propulsion system for the H-I and H-II series of launch vehicles. It is a bipropellant design, using LH2 and LOX. Primary design and production work was carried out by Mitsubishi Heavy Industries. In terms of liquid rockets, it is a fairly small engine, both in size and thrust output, being in the 89 kN (20,000 lbf) and the more recent models the 130 kN (30,000 lbf) thrust class. The motor is capable of multiple restarts, due to a spark ignition system as opposed to the single use pyrotechnic or hypergolic igniters commonly used on some contemporary engines. Though rated for up to 16 starts and 40+ minutes of firing time, on the H-II the engine is considered expendable, being used for one flight and jettisoned. It is sometimes started only once for a nine-minute burn, but in missions to GTO the engine is often fired a second time to inject the payload into the higher orbit after a temporary low Earth orbit has been established.

The RL60 was a planned liquid-fuel cryogenic rocket engine designed in the United States by Pratt & Whitney, burning cryogenic liquid hydrogen and liquid oxygen propellants. The engine runs on an expander cycle, running the turbopumps with waste heat absorbed from the main combustion process. This high-efficiency, waste heat based combustion cycle combined with the high-performance liquid hydrogen fuel enables the engine to reach a very high specific impulse of up to 465 seconds in a vacuum. The engine was planned to be a more capable successor to the Aerojet Rocketdyne RL10, providing improved performance and efficiency while maintaining the installation envelope of the RL10.

<span class="mw-page-title-main">Cryogenic rocket engine</span> Type of rocket engine which uses liquid fuel stored at very low temperatures

A cryogenic rocket engine is a rocket engine that uses a cryogenic fuel and oxidizer; that is, both its fuel and oxidizer are gases which have been liquefied and are stored at very low temperatures. These highly efficient engines were first flown on the US Atlas-Centaur and were one of the main factors of NASA's success in reaching the Moon by the Saturn V rocket.

The RD-701 is a liquid-fuel rocket engine developed by Energomash, Russia. It was briefly proposed to propel the reusable MAKS space plane, but the project was cancelled shortly before the end of USSR. The RD-701 is a tripropellant engine that uses a staged combustion cycle with afterburning of oxidizer-rich hot turbine gas. The RD-701 has two modes. Mode 1 uses three components: LOX as an oxidizer and a fuel mixture of RP-1 / LH2 which is used in the lower atmosphere. Mode 2 also uses LOX, with LH2 as fuel in vacuum where atmospheric influence is negligible.

<span class="mw-page-title-main">Aerojet LR87</span> American rocket engine family used on Titan missile first stages

The LR87 was an American liquid-propellant rocket engine used on the first stages of Titan intercontinental ballistic missiles and launch vehicles. Composed of twin motors with separate combustion chambers and turbopump machinery, it is considered a single unit and was never flown as a single combustion chamber engine or designed for this. The LR87 first flew in 1959.

Fastrac was a turbo pump-fed, liquid rocket engine. The engine was designed by NASA as part of the low cost X-34 Reusable Launch Vehicle (RLV) and as part of the Low Cost Booster Technology project. This engine was later known as the MC-1 engine when it was merged into the X-34 project.

<span class="mw-page-title-main">Rocket propellant</span> Chemical or mixture used as fuel for a rocket engine

Rocket propellant is the reaction mass of a rocket. This reaction mass is ejected at the highest achievable velocity from a rocket engine to produce thrust. The energy required can either come from the propellants themselves, as with a chemical rocket, or from an external source, as with ion engines.

References

  1. 1 2 Sutton, George P. (1963). Rocket Propulsion Elements, 3rd edition. New York: John Wiley & Sons. pp. 25, 186, 187.
  2. NASA:Liquid rocket engines, 1998, Purdue University
  3. 1 2 Heister, Stephen D.; Anderson, William E.; Pourpoint, Timothée L.; Cassady, R. Joseph (2019-02-07). Rocket Propulsion. Cambridge University Press. doi:10.1017/9781108381376. ISBN   978-1-108-38137-6. S2CID   203039055.
  4. 1 2 History and principles of rocket propulsion, Springer Praxis Books, Springer Berlin Heidelberg, 2005, pp. 1–34, doi:10.1007/3-540-27041-8_1, ISBN   978-3-540-22190-6 , retrieved 2023-11-29
  5. "Thomas Mueller's answer to Is SpaceX's Merlin 1D's thrust-to-weight ratio of 150+ believable? - Quora". www.quora.com.
  6. 1 2 3 "About LNG Propulsion System". JAXA . Retrieved 2020-08-25.
  7. 1 2 Hagemann, Dr. Gerald (November 4, 2015). "LOX/Methane The Future is Green" (PDF). Retrieved November 29, 2022.
  8. "Methane Engine Just for Future Space Transportation" (PDF). IHI Corporation. Retrieved November 29, 2022.
  9. Landis (2001). "Mars Rocket Vehicle Using In Situ Propellants". Journal of Spacecraft and Rockets. 38 (5): 730–735. Bibcode:2001JSpRo..38..730L. doi:10.2514/2.3739.
  10. Sutton, George P. and Biblarz, Oscar, Rocket Propulsion Elements, 7th ed., John Wiley & Sons, Inc., New York, 2001.
  11. "Sometimes, Smaller is Better". Archived from the original on 2012-04-14. Retrieved 2010-06-01.
  12. Rocket Propulsion elements - Sutton Biblarz, section 8.1
  13. Russian title Issledovaniye mirovykh prostranstv reaktivnymi priborami (Исследование мировых пространств реактивными приборами)
  14. Siddiqi 2000, p. 1.
  15. Siddiqi 2000, p. 27.
  16. Siddiqi 2000, p. 6–7,333.
  17. Siddiqi 2000, p. 3,166,182,187,205–206,208.
  18. 1 2 3 Glushko, Valentin (1 January 1973). Developments of Rocketry and Space Technology in the USSR. Novosti Press Pub. House. pp. 12–14, 19. OCLC   699561269.
  19. Zak, Anatoly. "Gas Dynamics Laboratory". Russian Space Web. Retrieved 20 July 2022.
  20. Chertok 2005, p. 165 Vol 1.
  21. Siddiqi 2000, p. 4.
  22. Asif Siddiqi (November 2007). "The Man Behind the Curtain". Archived from the original on 2021-04-03.
  23. Albrecht, Ulrich (1993). The Soviet Armaments Industry. Routledge. pp. 74–75. ISBN   3-7186-5313-3.
  24. Tsander, F. A. (1964). Problems of Flight by Jet Propulsion-Interplanetary Flights (Translated from Russian) (PDF). Israel Program for Scientific Translations. pp. 32, 38–39, 58–59. Retrieved 13 June 2022.
  25. Gordon, E.; Sweetman, Bill (1992). Soviet X-planes. Bill Sweetman. Osceola, WI: Motorbooks International. p. 47. ISBN   978-0-87938-498-2. OCLC   22704082.
  26. Chertok 2005, p. 167 Vol 1.
  27. Siddiqi 2000, p. 8-9.
  28. 1 2 Paulet de Vásquez, Sara (2002). "Pedro Paulet: pionero peruano del espacio". Ciencia y tecnología. Lima: 5–12.
  29. Ordway, F. I. (September 1977). "The alleged contributions of Pedro E. Paulet to liquid-propellant rocketry". Nasa, Washington Essays on the History of Rocketry and Astronautics, Vol. 2. NASA.
  30. Mejía 2017, pp. 115–116.
  31. Fitzgerald, Michael (2018). Hitler's Secret Weapons of Mass Destruction: The Nazi Plan for Final Victory. pp. Chapter 3. Paulet was clearly a pioneer in the field of rocketry and it is unsurprising that the Nazis were keen to recruit him to assist their efforts. The German Astronautical Society invited him to Germany to become part of a team of researchers into rocket propulsion and he was initially interested, but when he discovered that the intention was to construct a weapon that would be used for military purposes he declined the invitation. As late as 1965, Oberth described him as one of the true pioneers of rocket science.
  32. "El peruano que se convirtió en el padre de la astronáutica inspirado por Julio Verne y que aparece en los nuevos billetes de 100 soles". BBC News (in Spanish). Retrieved 2022-03-11.
  33. Von Braun, Wernher; Ordway III, Frederick I. (1968). Histoire Mondiale de L'Astronautique. París: Larousse / Paris -Match. pp. 51–52.
  34. Fitzgerald, Michael (2018). Hitler's Secret Weapons of Mass Destruction: The Nazi Plan for Final Victory. pp. Chapter 3. Even Wernher von Braun described Paulet as 'one of the fathers of aeronautics' and 'the pioneer of the liquid fuel propulsion motor'. He declared that 'by his efforts, Paulet helped man reach the Moon'.
  35. Harding, Robert C. (2012). Space Policy in Developing Countries: The Search for Security and Development on the Final Frontier. Routledge. p. 156. ISBN   9781136257902. Peru holds a special place among Latin America's EMSAs because the country was home to Pedro Paulet, who invented the world's first liquid-propelled rocket engine in 1895 and the first modern rocket propulsion system in 1900. ... According to Wernher von Braun, 'Paulet should be considered the pioneer of the liquid fuel propulsion motor ... by his efforts, Paulet helped man reach the moon.' Paulet went on to found Peru's National Pro-Aviation League, a precursor of the Peruvian Air Force.
  36. "El peruano que se convirtió en el padre de la astronáutica inspirado por Julio Verne y que aparece en los nuevos billetes de 100 soles". BBC News (in Spanish). Retrieved 2022-03-11.
  37. Fitzgerald, Michael (2018). Hitler's Secret Weapons of Mass Destruction: The Nazi Plan for Final Victory. pp. Chapter 3. Paulet was clearly a pioneer in the field of rocketry and it is unsurprising that the Nazis were keen to recruit him to assist their efforts. The German Astronautical Society invited him to Germany to become part of a team of researchers into rocket propulsion and he was initially interested, but when he discovered that the intention was to construct a weapon that would be used for military purposes he declined the invitation. As late as 1965, Oberth described him as one of the true pioneers of rocket science.
  38. "Un documental reivindicará al peruano Paulet como pionero de la astronáutica". EFE (in Spanish). 2012-04-05. Retrieved 2022-03-11.
  39. "Re-Creating History". NASA. Archived from the original on 2007-12-01.
  40. Max, Valier, Raketenfahrt: Eine technische Möglichkeit Gebundene Ausgabe – Großdruck, 1. Januar 1930, De Gruyter Oldenbourg, Reprint 2019 ( ISBN   978-3-486-76182-5)
  41. "Fritz von Opel, Speech at Deutsches Museum, April 3, 1968, re-print in "Opel Post"" (PDF). May 1968. p. 4ff.
  42. Frank H. Winter, "1928-1929 Forerunners of the Shuttle: the 'Von Opel Flights'", SPACEFLIGHT, Vol. 21,2, Feb. 1979
  43. Boyne, Walter J. (September 2004). "Rocket Men" (PDF). Air Force Magazine.
  44. Magazines, Hearst (1 May 1931). Popular Mechanics. Hearst Magazines. p.  716 via Internet Archive. Popular Mechanics 1931 curtiss.
  45. Volker Koos, Heinkel He 176 – Dichtung und Wahrheit, Jet&Prop 1/94 p. 17–21
  46. "Astra Rocket Engine — Delphin 3.0". June 2020.
  47. "Orbex builds single-piece rocket engine 3D printed on SLM 800 - Aerospace Manufacturing". 13 February 2019.
  48. "Orbex unveiled largest 3D printed rocket engine in the world". 13 February 2019.
  49. "Relativity Space will 3D-print rockets at new autonomous factory in Long Beach, California". Space.com . 28 February 2020.
  50. "Launch startup Skyrora successfully tests 3D-printed rocket engines powered by plastic waste". 3 February 2020.
  51. "A tiny start-up based in Brooklyn has a 3D-printed rocket engine it says is the largest in the world". CNBC . 20 February 2019.
  52. "Air Force funding keeps Launcher development on track". 14 November 2019.
  53. "Meet Launcher, the rocket engine builder with just eight employees". 9 November 2020.

Sources cited