The Value of Short-Course Technical Training

I could not imagine living in a house that does not have a basement.  While most people would simply discard anything that doesn’t fit in the main level of their house, I simply move it to the basement.  So, my basement is quite full of junk.  Sometimes, I need to find something in my basement, […]

I could not imagine living in a house that does not have a basement.  While most people would simply discard anything that doesn’t fit in the main level of their house, I simply move it to the basement.  So, my basement is quite full of junk.  Sometimes, I need to find something in my basement, and I search endlessly, and eventually give up and declare the item “missing”.  In the process of searching, however, I often find things that had previously been declared “missing”.  So, I have come to learn that nothing is ever “missing”, it simply is waiting for a later time to be found.

I thought about my basement when I read about a recent discovery by scientists studying data received from the James Webb Space Telescope.  This telescope is the largest optical telescope in space.  It’s high-resolution and high-sensitivity instruments make it capable of viewing objects too distant or too faint for the Hubble Telescope.  This telescope has been in space for only a little more than a year, and it is already sending back data and images that are simply amazing.

Similar to my basement adventures, scientists were recently looking for one thing using the James Webb Space Telescope, and they discovered something entirely different.  NASA scientist were looking for a previously-discovered asteroid, but were unable to find it due its brightness and an offset in the telescope’s direction.  While they could not see the asteroid they were looking for, they did discover another asteroid which had never been seen before.  The new asteroid was very small, demonstrating that the James Webb Telescope was capable of finding asteroids smaller than anything which was previously discoverable with Hubble.  The mission had been declared a failure, but was now declared a great success.

I found this story particularly interesting because in addition to reminding me of my cluttered basement, it also made me think about how many different scientists, and how many different disciplines, and how many different engineering achievements were necessary to ultimately find this asteroid.  It was not a single person, or even a single team of people that got us here.  Designing and building the Telescope was the first task at hand, and that required massive amounts of Systems Engineering and manufacturing expertise.  Launching the Space Telescope into space with a Ariane 5 Rocket was also a huge feat, which required the skills of another team.  Daily operations of the telescope and managing the data from the telescope require even more attention from a completely different set of scientists.  There is huge number of people that had their hands on this discovery, and the future discoveries of the James Webb Space Telescope.

To be a well-rounded scientist or engineer, one should have a basic level of understanding of each of the disciplines that contribute to his or her area of expertise.  Short-Course Technical Training like what is offered by Applied Technology Institute is a great way to acquire that basic level of understanding.  ATI can not replace the intensive training a scientist acquired in his or her field of expertise; there is no way a 4-day short course can substitute for a long and rigorous college education.  ATI short-courses can, however, offer a way for a scientist or engineer to become more aware of the many disciplines which work in unison with their field of expertise.  And, even within a scientist’s field of expertise, short-courses can help refresh certain areas of their training.

A complete list of upcoming ATI short-courses, as well a complete list of available short-courses can be found at the ATI homepage ( www.aticourses.com ).  We hope to see you in an upcoming ATI short-course, or an upcoming ATI Free-Session soon.

Psyched for Mission PSYCHE

NASA’s Psyche Mission is similar to other NASA missions in some ways, but different in other ways.  Psyche is similar in that bold and innovative technologies are being used to push the boundaries of deep-space exploration.  Psyche is different however, in that the launch has been pushed forward for one year due to a delay […]

NASA’s Psyche Mission is similar to other NASA missions in some ways, but different in other ways.  Psyche is similar in that bold and innovative technologies are being used to push the boundaries of deep-space exploration.  Psyche is different however, in that the launch has been pushed forward for one year due to a delay in critical testing.  Launch of Psyche is now expected in October 2023.

Psyche will be launched from Earth using a SpaceX Heavy Falcon Rocket.  This launch system has been used before, and should be effective for its purpose.  Once in deep space, however, an alternate method will be required for propelling Psyche to its ultimate destination, the Comet Psyche. As explained by NASA, “The unique, metal-rich Psyche asteroid may be part of the core of a planetesimal, a building block of rocky planets in our solar system. Learning more about the asteroid could tell us more about how our own planet formed and help answer fundamental questions about Earth’s own metal core and the formation of our solar system.”

Once beyond the orbit of the moon, Psyche will use solar electric propulsion for its 1.5 billion ( with a B ) mile trip to the asteroid Psyche which will conclude in 2026.  This will be the first spacecraft to use “Hall-Effect Thrusters” for propulsion.  As explained by NASA, this thruster technology “traps electrons in a magnetic field and uses them to ionize onboard propellant, expending much less propellant than equivalent chemical rockets.”

As a secondary mission for this spacecraft, Psyche will be used to demonstrate and test Deep Space Optical Communications.  This capability will become increasingly important as future missions are planned for areas so deep in space that current communication methods may become infeasible.

As spacecraft and space missions become more complex, the rockets that propel them will also need to become more complex.  Rocket advances must keep up with Spacecraft advances, and the Psyche Mission is one indication that Rocket scientists are up to the challenge.

If you want to learn more about Rocket Science, consider taking ATI’s upcoming course on the subject.  You can learn more about the course, and register for it, at Rockets & Launch Vehicles – Selection & Design

This four-day course provides an overview of rockets and missiles, including a fourth day covering advanced selection and design processes. The course provides a wide practical knowledge in rocket and missile issues and technologies. 

The course is right around the corner in May, so if you are interested, register today.

And, as always, if want to see the full list of courses offered by ATI, you can find that, and other interesting information at www.aticourses.com

Yep, We Really Train Rocket Scientists!

It is so exciting that we are going back to the moon.  NASA is planning a bold set of missions.  Although one of the missions will visit the moon again, the ultimate goals are much more far-reaching.  The intent is to learn from the moon visit and apply knowledge to future manned missions which will […]

It is so exciting that we are going back to the moon.  NASA is planning a bold set of missions.  Although one of the missions will visit the moon again, the ultimate goals are much more far-reaching.  The intent is to learn from the moon visit and apply knowledge to future manned missions which will visit places far beyond the moon.

We are only one month until the Artemis I mission.  For this first mission, the uncrewed Orion Spacecraft will spend four to six weeks in Space, and go far beyond the moon. To do this, a very powerful rocket is needed to accelerate an Orion spacecraft fast enough to overcome the pull of Earth’s gravity.  This will be accomplished by NASA’s Space Launch System Rocket.  This is the most powerful rocket ever used by NASA, generating 8.8 million pounds of thrust.

As spacecraft and space missions become more complex, the rockets that propel them will also need to become more complex.  Rocket advances must keep up with Spacecraft advances, and the Space Launch System is one indication that Rocket scientists are up to the challenge.

If you want to learn more about Rocket Science, consider taking ATI’s upcoming course on the subject.  You can learn more about the course, and register for it, at Rockets & Launch Vehicles – Selection & Design

This four-day course provides an overview of rockets and missiles, including a fourth day covering advanced selection and design processes. The course provides a wide practical knowledge in rocket and missile issues and technologies. 

The course is right around the corner in August, so if you are interested, do not delay.

And, as always, if want to see the full list of courses offered by ATI, you can find that, and other interesting information at www.aticourses.com

Rockets and Launch Vehicles

As I looked at the title of the upcoming ATI course called Rockets and Launch Vehicles, the first question I asked myself was “What is the difference between a rocket and a launch vehicle?  With the help of google, I learned that all launch vehicles are rockets, but not all rockets are launch vehicles.  A […]

As I looked at the title of the upcoming ATI course called Rockets and Launch Vehicles, the first question I asked myself was “What is the difference between a rocket and a launch vehicle?  With the help of google, I learned that all launch vehicles are rockets, but not all rockets are launch vehicles.  A rocket that is powerful enough to send people, satellites, or spacecrafts into space is called a Launch Vehicle.  So, those things you would build and shoot into the sky as a kid are rockets, but they are not launch vehicles.

Rockets, including launch vehicles, take off by burning fuel, which produces a gas byproduct.  That escaping gas produces the force that creates the thrust to power the rocket upward.

Most launch vehicles need multiple stages to produce enough thrust get a spacecraft into space.  These stages usually sit on top of each other, but there also some designs which have the stages parallel to each other; it all depends on the goals of the mission.  The first stage, the stage closest to the ground, is usually the largest.  Its purpose is to lift the spacecraft above the earth’s atmosphere to a height of about 150,000 feet.  The purpose of the second stage, the stage closest to the spacecraft, is to get the spacecraft to achieve orbital velocity.  Usually, when a stage has used up all of its fuel, it serves no additional purpose, so it is jettisoned.

The Space Launch System is a launch vehicle getting a lot of attention and a lot of funding today,  The SLS is the Launch Vehicle which will be used for the NASA Artemis missions which will first return to the moon, and then explore beyond.  The mission of the first Artemis flight, Artemis I, will be to test the SLS launch vehicle using an uncrewed Orion Spacecraft.  This launch will be occurring in March 2022.

If you would like to learn more about rockets and launch vehicles, consider taking the upcoming ATI course Rockets and Launch Vehicles.  You can read more about this course, and register for it here.

And, as always, you learn about other upcoming ATI courses at the ATI homepage www.aticourses.com

Game Changer

For decades, the state of the art in missile technology has been Ballistic Missiles.   A Ballistic missile follows a ballistic trajectory to deliver its warhead, or warheads, onto a predetermined target.  The missile is put into orbit by a rocket, and the remainder of its flight is unpowered.  The missile simply falls like a rock […]

For decades, the state of the art in missile technology has been Ballistic Missiles.   A Ballistic missile follows a ballistic trajectory to deliver its warhead, or warheads, onto a predetermined target.  The missile is put into orbit by a rocket, and the remainder of its flight is unpowered.  The missile simply falls like a rock on a highly predictable approach.  Due to the nature of its flight, Ballistic Missiles can easily be countered by Anti-Ballistic Missiles.  The ABM can intercept and destroy the Ballistic Missile at any point during its flight.  Many countries have mastered the technology of Ballistic Missiles, and Anti-Ballistic Missile Defense.  It is what drove the Cold War.

In recent years, however, we have been introduced to a new missile technology.  Hypersonic Missiles have changed the art of war as we know it.  Hypersonic missiles travel at least five times the speed of sound, and they can fly much lower to the ground than conventional Ballistic Missiles.  These hypersonic missiles are more of a threat because they are highly maneuverable.  Due to their speed and their maneuverability, they are difficult, if not impossible, to detect by traditional anti-ballistic missile defense systems.  And, due to their immense kinetic energy, they are even more destructive to the target that they are directed toward.  Hypersonic missiles are a game changer. 

Russia, China, North Korea, and the US have all tested hypersonic missiles.  When they become operational and get incorporated into military arsenals, it will be truly significant for both aggressors and target countries. 

This is truly the way of the future in Rocket and Missile technology.  Scientists and engineers need to be familiar with this new type of missile.

If you would like to learn more about Rocket and Missile Fundamentals including the Hypersonic Missile technology, consider enrolling in ATI’s upcoming course Rocket and Missile Fundamentals.  The instructor has recently added a unit discussing Hypersonic Missiles.

As always, a complete listing of ATI’s courses can be found here.

You’re Going To The Moon, Alice

Mankind has always been fascinated with exploring the Moon, and that will probably always be the case.  At first, in the time leading up to the famous first moon landing in 1969, the goal was simply to reach the moon, and spend a short time looking around, and return to earth safely.  Now, 50 years […]

Mankind has always been fascinated with exploring the Moon, and that will probably always be the case.  At first, in the time leading up to the famous first moon landing in 1969, the goal was simply to reach the moon, and spend a short time looking around, and return to earth safely.  Now, 50 years later, the goal is more ambitious since technology can support so much more.  The first objective today is to reach the moon, and stay there.  The next goal would be to use the moon as a landing pad to support exploration of things beyond the moon, most notably Mars.  The NASA Artemis Missions will be the way these objectives are accomplished.  I am not sure about you, but this mission snuck up on me, and I am learning about it now. 

The Artemis Mission is comprised of six projects which together will allow NASA to accomplish its goals of reaching the moon, staying on the moon for long term exploration, and getting closer to ultimate goal of being able to send men (and women) beyond the moon.  The six projects include:

Ground Systems – Upgrading Earth ground systems to support the larger rockets which will be needed

Space Launch System – The new and more powerful rocket that will launch man toward the moon and beyond

Orion – The spacecraft that will bring astronauts to the moon’s orbit, and return them to earth from the moon’s orbit

Gateway – The outpost spacecraft which will orbit the moon and be living quarters for the astronauts when they are not on the moon surface

Lunar Landers – The spacecraft which will transfer astronauts between the Gateway and the moon Surface, and

Space Suits – The new and improved suits that the astronauts will need to carry out their mission.

The timeline for this mission has three major milestones, namely, the three Artemis missions, Artemis I, Artemis II, and Artemis III.

Artemis I – an unmanned flight to test the Space Launch System and Orion, scheduled for 2021

Artemis II – a manned flight to test the Space Launch System and Orion, scheduled for 2022

Artemis III – A manned flight to the moon that will return man to the moon.

This is a truly ambitious mission, and an even more ambitious schedule.

ATI offers a plethora of courses which relate to Space exploration.  Check out our list of Space related courses here.    If you are interested in the legal aspects of Space exploration, you can register for our upcoming Astropolitics class here

Although the author thinks Space Exploration is exciting and important, and I fully endorse all of the goals of the Artemis Mission, I can’t help but wonder why the Government is not spending at least as much money on exploration of the deep oceans.  I would challenge the US to start investing more money in Ocean Exploration, but not at the expense of Space Exploration.  Both of these are important.  I am curious what readers think about this issue, please leave your comments below.

And, if you are interested in Ocean Exploration, ATI has a few courses which may be of interest to you too.  Please check out our full list of offerings here.

And if you simply want to learn more about the Artemis Mission, you can go to the NASA Artemis site that describes the mission in more detail. 

Hand-Held Radar Detects Jet?

An ATI Staff Member who has not taken any ATI Radar Courses yet found a story in her Inbox about a highway officer in Texas who was operating a hand-held radar to catch speeders.  As you can read in the following copy of the letter, the officer purportedly locked onto a USMC F/A-18 Hornet Jet.  […]

An ATI Staff Member who has not taken any ATI Radar Courses yet found a story in her Inbox about a highway officer in Texas who was operating a hand-held radar to catch speeders.  As you can read in the following copy of the letter, the officer purportedly locked onto a USMC F/A-18 Hornet Jet.  The story purports that the Jet detected energy from the hand held radar, and automatic tactical systems on the Jet nearly fired on the radar/officer, but the pilot overrode those automatic systems preventing a catastrophic mishap.

Although this story is humorous, it also demonstrates that the writer, and some readers, are not familiar with how hand-held radars work, and how the Tactical Systems on the USMC Jet work.  In fact, a Snopes article gives an excellent explanation of why this story, although humorous and entertaining, is not factual, and could not have actually occurred.  A fascinating explanation of the fallacies associated with this story can be found at   https://www.snopes.com/fact-check/police-radar-missile/

So, in order to better recognize articles like this for what they are, please consider learning more about Radar Systems.  ATI offers 78 courses dealing with Radar, Missiles, and Defense, but our most popular Radar courses are Radar 101, Radar 201, and Radar Principles.  More information on all of these courses can be found on the ATI web page at the following links.

https://aticourses.com/training_classes/radar-101/
https://aticourses.com/training_classes/radar-201-d223/
https://aticourses.com/training_classes/radar-principles/

Booster Rocket Pioneers

The rockets that hurl the Navstar satellites into orbit are direct descendents of the highly destructive Chinese “fire arrows” built and launched by Chinese military engineers 750 years ago. The earliest Chinese rockets were slender tubes stuffed with gunpowder and fastened to long flat sticks that jutted out behind the rocket to promote stable flight. […]
The rockets that hurl the Navstar satellites into orbit are direct descendents of the highly destructive Chinese “fire arrows” built and launched by Chinese military engineers 750 years ago. The earliest Chinese rockets were slender tubes stuffed with gunpowder and fastened to long flat sticks that jutted out behind the rocket to promote stable flight. In 1232 they were launched in large quantities on the outskirts of Peking, when special Chinese rocket brigades successfully pushed back Mongol cavalrymen. And, in 1249, they were used with great effect by the Moors in their military campaign along the Iberian Peninsula. Near the beginning of the nineteenth century, Englishman William Congreve concocted superior powder blends and moved the stabilizing stick to the center of the rocket for improved accuracy. In 1807 the British blasted Copenhagen with 25,000 Congreve rockets. nine years later, when they bombarded Fort McHenry , they inadvertently provided “the rocket’s red glare,” which helped inspire America’s National Anthem. In 1903 a lonely Russian schoolteacher, Konstantin Tsiolkovsky , correctly concluded that rockets fueled with liquid hydrogen and liquid oxygen would be considerably more efficient than the simpler solid-fueled rockets then in use. He also devised a concept for stacking rockets one atop the other to yield the enormous speeds necessary for successful interplanetary travel. Twenty-three years later Dr. Robert Goddard knelt on the frozen ground in his Aunt Effie’s cabbage patch at Auburn, Massachusetts, and casually used a blowtorch to ignite the world’s first liquid-fueled rocket. Goddard is today revered for his expansive expertise, but during his lifetime his contemporaries criticized him unmercifully because he had once dared to mention the possibility of sending a small flash powder to impact the moon. Years later, when one of his liquid-fueled rockets reached its design altitude of 2,000 feet, a banner headline wryly commented: “Moon Rocket Misses Target by 237,799 Miles!” The rockets built by the Goddard team were all handcrafted machines, but Germany’s rocketeers , working under the direction of Werner von Braun, constructed liquid-fueled rockets in mass-¼production quantities. When World War II fizzled to a halt, many of the German scientists came to Los Alamos to help American’s military and space-age rocketeers . In 1961, when President Kennedy courageously announced that the United States would conquer the moon, America’s rocketeers had not yet orbited a single astronaut. The Saturn V moon rocket they later developed for the mission, was the pinnacle of the rocket maker’s art. But it was expendable; NASA’s space shuttle is a “reusable” booster. It delivers payloads weighing as much as 50,000 pounds and brings others back to earth for refurbishment and repair, gently landing — as TV Newsman Edwin Newman once observed: “like a butterfly with sore feet.”

Rocket Propulsion Fundamentals

Rocket Propulsion Fundamentals White hot combustion by-products blasted rearward with blinding speed generate the rocket’s propulsive force that that hurls a rocket skyward. Pressure inside the rocket combustion chamber pushes in all directions to form balanced pairs of opposing forces which nullify one another, except where the hole for the exhaust nozzle is placed. Here […]

Rocket Propulsion Fundamentals

White hot combustion by-products blasted rearward with blinding speed generate the rocket’s propulsive force that that hurls a rocket skyward. Pressure inside the rocket combustion chamber pushes in all directions to form balanced pairs of opposing forces which nullify one another, except where the hole for the exhaust nozzle is placed. Here the pressure escapes, causing an unbalanced force at the opposite side of the combustion chamber that pushes the rocket up towards its orbital destination. Both rockets and jets are based on the same principle that causes a toy balloon, carelessly released, to swing in kamikaze spirals around the dining room. A jet sucks its oxygen from the surrounding air, but a rocket carries its own supply of oxidizer on board. This oxidizer can be stored in a separate tank, mixed with the fuel, or chemically embedded in oxygen-rich compounds. A rocket usually has two separate tanks, one containing the fuel, the other containing the oxidizer. The two fluids are pumped or pushed under pressure into a small combustion chamber above the exhaust nozzle, where burning takes place to create a thrust. A solid rocket rocket is like a slender tube filled with gunpowder; the fuel and oxidizer are mixed together in a rubbery cylindrical slug called the grain. Solid propellants are not pumped into a separate combustion chamber. Instead, burning takes place along the entire length of the cylinder. Consequently, the tank walls must be built strong enough to withstand the combustion pressure. Rocket design decisions are dominated by the desire to produce the maximum possible velocity when the propellants are burned. A rocket’s velocity can be increased in two principal ways: by using propellants with a high efficiency and by making the rocket casing and its engines as light as design constraints permit. Unfortunately, efficient propellants tend to have some rather undesirable physical and chemical properties. Liquid oxygen is a good oxidizer, but it will freeze all lubricants and crack most seals. Hydrogen is a good fuel but it can spark devastating explosions. Fluorine is even better but it is so reactive it can even cause metals to burn. Miniaturized components, special fabrication techniques and high strength alloys can all be used to shave excess weight. But there are limits beyond which further weight reductions are impractical. The solution is to use staging techniques whereby a series of progressively smaller rockets are stacked one atop the other. Such a multistage rocket cuts down its own weight as it flies along by discarding empty tanks and heavy engines. However, orbiting even a small payload with a multistage rocket requires an enormous booster. The Saturn moon rocket, for example, outweighed the Apollo capsule it carried into space by a factor of 60 to 1.

Rocket Thrust Equation and Launch Vehicles

The fundamental principles of propulsion and launch vehicle physics by Robert A. Nelson A satellite is launched into space on a rocket, and once there it is inserted into the operational orbit and is maintained in that orbit by means of thrusters onboard the satellite itself. This article will summarize the fundamental principles of rocket […]

The fundamental principles of propulsion and launch vehicle physics

by Robert A. Nelson

A satellite is launched into space on a rocket, and once there it is inserted into the operational orbit and is maintained in that orbit by means of thrusters onboard the satellite itself. This article will summarize the fundamental principles of rocket propulsion and describe the main features of the propulsion systems used on both launch vehicles and satellites.

The law of physics on which rocket propulsion is based is called the principle of momentum. According to this principle, the time rate of change of the total momentum of a system of particles is equal to the net external force. The momentum is defined as the product of mass and velocity. If the net external force is zero, then the principle of momentum becomes the principle of conservation of momentum and the total momentum of the system is constant. To balance the momentum conveyed by the exhaust, the rocket must generate a momentum of equal magnitude but in the opposite direction and thus it accelerates forward.

The system of particles may be defined as the sum of all the particles initially within the rocket at a particular instant. As propellant is consumed, the exhaust products are expelled at a high velocity. The center of mass of the total system, subsequently consisting of the particles remaining in the rocket and the particles in the exhaust, follows a trajectory determined by the external forces, such as gravity, that is the same as if the original particles remained together as a single entity. In deep space, where gravity may be neglected, the center of mass remains at rest.

The configuration of a chemical rocket engine consists of the combustion chamber, where the chemical reaction takes place, and the nozzle, where the gases expand to create the exhaust. An important characteristic of the rocket nozzle is the existence of a throat. The velocity of the gases at the throat is equal to the local velocity of sound and beyond the throat the gas velocity is supersonic. Thus the combustion of the gases within the rocket is independent of the surrounding environment and a change in external atmospheric pressure cannot propagate upstream.

The thrust of the rocket is given by the theoretical equation :

F = lm(dot) ve + ( pe – pa ) Ae

This equation consists of two terms. The first term, called the momentum thrust, is equal to the product of the propellant mass flow rate m(dot) and the exhaust velocity ve with a correction factor l for nonaxial flow due to nozzle divergence angle. The second term is called the pressure thrust. It is equal to the difference in pressures pe and pa of the exhaust velocity and the ambient atmosphere, respectively, acting over the area Ae of the exit plane of the rocket nozzle. The combined effect of both terms is incorporated into the effective exhaust velocity c. Thus the thrust is also written

F = m(dot) c

where an average value of c is used, since it is not strictly constant.

The exhaust exit pressure is determined by the expansion ratio given by

which is the ratio of the area of the nozzle exit plane Ae and the area of the throat At . As the expansion ratio e increases, the exhaust exit pressure pe decreases.

The thrust is maximum when the exit pressure of the exhaust is equal to the ambient pressure of the surrounding environment, that is, when pe = pa. This condition is known as optimum expansion and is achieved by proper selection of the expansion ratio. Although optimum expansion makes the contribution of the pressure thrust zero, it results in a higher value of exhaust velocity ve such that the increase in momentum thrust exceeds the reduction in pressure thrust.

A conical nozzle is easy to manufacture and simple to analyze. If the apex angle is 2a , the correction factor for nonaxial flow is

= ½ (1 + cos a )

The apex angle must be small to keep the loss within acceptable limits. A typical design would be a = 15° , for which l = 0.9830. This represents a loss of 1.7 percent. However, conical nozzles are excessively long for large expansion ratios and suffer additional losses caused by flow separation. A bell-shaped nozzle is therefore superior because it promotes expansion while reducing length.

ROCKET PROPULSION PARAMETERS

The specific impulse Isp of a rocket is the parameter that determines the overall effectiveness of the rocket nozzle and propellant. It is defined as the ratio of the thrust and the propellant weight flow rate, or

Isp = F / m(dot) g = c / g

where g is a conventional value for the acceleration of gravity (9.80665 m/s2 exactly). Specific impulse is expressed in seconds.

Although gravity has nothing whatever to do with the rocket propulsion chemistry, it has entered into the definition of specific impulse because in past engineering practice mass was expressed in terms of the corresponding weight on the surface of the earth. By inspection of the equation, it can be seen that the specific impulse Isp is physically equivalent to the effective exhaust velocity c, but is rescaled numerically and has a different unit because of division by g. Some manufacturers now express specific impulse in newton seconds per kilogram, which is the same as effective exhaust velocity in meters per second.

Two other important parameters are the thrust coefficient CF and the characteristic exhaust velocity c*. The thrust coefficient is defined as

CF = F / At pc = m(dot) c / At pc

where F is the thrust, At is the throat area, and pc is the chamber pressure. This parameter is the figure of merit of the nozzle design. The characteristic exhaust velocity is defined as

c* = At pc / m(dot) = c / CF

This parameter is the figure of merit of the propellant. Thus the specific impulse may be written

Isp = CF c* / g

which shows that the specific impulse is the figure of merit of the nozzle design and propellant as a whole, since it depends on both CF and c*. However, in practice the specific impulse is usually regarded as a measure of the efficiency of the propellant alone.

LAUNCH VEHICLE PROPULSION SYSTEMSM

In the first stage of a launch vehicle, the exit pressure of the exhaust is equal to the sea level atmospheric pressure 101.325 kPa (14.7 psia) for optimum expansion. As the altitude of the rocket increases along its trajectory, the surrounding atmospheric pressure decreases and the thrust increases because of the increase in pressure thrust. However, at the higher altitude the thrust is less than it would be for optimum expansion at that altitude. The exhaust pressure is then greater than the external pressure and the nozzle is said to be underexpanded. The gas expansion continues downstream and manifests itself by creating diamond-shaped shock waves that can often be observed in the exhaust plume.

The second stage of the launch vehicle is designed for optimum expansion at the altitude where it becomes operational. Because the atmospheric pressure is less than at sea level, the exit pressure of the exhaust must be less and thus the expansion ratio must be greater. Consequently, the second stage nozzle exit diameter is larger than the first stage nozzle exit diameter.

For example, the first stage of a Delta II 7925 launch vehicle has an expansion ratio of 12. The propellant is liquid oxygen and RP-1 (a kerosene-like hydrocarbon) in a mixture ratio (O/F) of 2.25 at a chamber pressure of 4800 kPa (700 psia) with a sea level specific impulse of 255 seconds. The second stage has a nozzle expansion ratio of 65 and burns nitrogen tetroxide and Aerozene 50 (a mixture of hydrazine and unsymmetrical dimethyl hydrazine) in a mixture ratio of 1.90 at a chamber pressure of 5700 kPa (830 psia), which yields a vacuum specific impulse of 320 seconds.

In space, the surrounding atmospheric pressure is zero. In principle, the expansion ratio would have to be infinite to reduce the exit pressure to zero. Thus optimum expansion is impossible, but it can be approximated by a very large nozzle diameter, such as can be seen on the main engines of the space shuttle with e = 77.5. There is ultimately a tradeoff between increasing the size of the nozzle exit for improved performance and reducing the mass of the rocket engine.

In a chemical rocket, the exhaust velocity, and hence the specific impulse, increases as the combustion temperature increases and the molar mass of the exhaust products decreases. Thus liquid oxygen and liquid hydrogen are nearly ideal chemical rocket propellants because they burn energetically at high temperature (about 3200 K) and produce nontoxic exhaust products consisting of gaseous hydrogen and water vapor with a small effective molar mass (about 11 kg/kmol). The vacuum specific impulse is about 450 seconds. These propellants are used on the space shuttle, the Atlas Centaur upper stage, the Ariane-4 third stage, the Ariane-5 core stage, the H-2 first and second stages, and the Long March CZ-3 third stage.

SPACECRAFT PROPULSION SYSTEMS

The spacecraft has its own propulsion system that is used for orbit insertion, station keeping, momentum wheel desaturation, and attitude control. The propellant required to perform a maneuver with a specified velocity increment Dv is given by the “rocket equation”

D m = m0 [ 1 – exp(- Dv / Isp g) ]

where m0 is the initial spacecraft mass. This equation implies that a reduction in velocity increment or an increase in specific impulse translates into a reduction in propellant.

In the case of a geostationary satellite, the spacecraft must perform a critical maneuver at the apogee of the transfer orbit at the synchronous altitude of 35,786 km to simultaneously remove the inclination and circularize the orbit. The transfer orbit has a perigee altitude of about 200 km and an inclination roughly equal to the latitude of the launch site. To minimize the required velocity increment, it is thus advantageous to have the launch site as close to the equator as possible.

For example, in a Delta or Atlas launch from Cape Canaveral the transfer orbit is inclined at 28.5° and the velocity increment at apogee is 1831 m/s; for an Ariane launch from Kourou the inclination is 7° and the velocity increment is 1502 m/s; while for a Zenit flight from the Sea Launch platform on the equator the velocity increment is 1478 m/s. By the rocket equation, assuming a specific impulse of 300 seconds, the fraction of the separated mass consumed by the propellant for the apogee maneuver is 46 percent from Cape Canaveral, 40 percent from Kourou, and 39 percent from the equator. As a rule of thumb, the mass of a geostationary satellite at beginning of life is on the order of one half its mass when separated from the launch vehicle.

Before performing the apogee maneuver, the spacecraft must be reoriented in the transfer orbit to face in the proper direction for the thrust. This task is sometimes performed by the launch vehicle at spacecraft separation or else must be carried out in a separate maneuver by the spacecraft itself. In a launch from Cape Canaveral, the angle through which the satellite must be reoriented is about 132°.

Once on station, the spacecraft must frequently perform a variety of stationkeeping maneuvers over its mission life to compensate for orbital perturbations. The principal perturbation is the combined gravitational attractions of the sun and moon, which causes the orbital inclination to increase by nearly one degree per year. This perturbation is compensated by a north-south stationkeeping maneuver approximately once every two weeks so as to keep the satellite within 0.05° of the equatorial plane. The average annual velocity increment is about 50 m/s, which represents 95 percent of the total stationkeeping fuel budget. Also, the slightly elliptical shape of the earth’s equator causes a longitudinal drift, which is compensated by east-west stationkeeping maneuvers about once a week, with an annual velocity increment of less than 2 m/s, to keep the satellite within 0.05° of its assigned longitude.

In addition, solar radiation pressure caused by the transfer of momentum carried by light and infrared radiation from the sun in the form of electromagnetic waves both flattens the orbit and disturbs the orientation of the satellite. The orbit is compensated by an eccentricity control maneuver that can sometimes be combined with east-west stationkeeping. The orientation of the satellite is maintained by momentum wheels supplemented by magnetic torquers and thrusters. However, the wheels must occasionally be restored to their nominal rates of rotation by means of a momentum wheel desaturation maneuver in which a thruster is fired to offset the change in angular momentum.

Geostationary spacecraft typical of those built during the 1980s have solid propellant rocket motors for the apogee maneuver and liquid hydrazine thrusters for stationkeeping and attitude control. The apogee kick motor uses a mixture of HTPB fuel and ammonium perchlorate oxidizer with a specific impulse of about 285 seconds. The hydrazine stationkeeping thrusters operate by catalytic decomposition and have an initial specific impulse of about 220 seconds. They are fed by the pressure of an inert gas, such as helium, in the propellant tanks. As propellant is consumed, the gas expands and the pressure decreases, causing the flow rate and the specific impulse to decrease over the mission life. The performance of the hydrazine is enhanced in an electrothermal hydrazine thruster (EHT), which produces a hot gas mixture at about 1000 ° C with a lower molar mass and higher enthalpy and results in a higher specific impulse of between 290 and 300 seconds.

For example, the Ford Aerospace (now Space Systems/Loral) INTELSAT V satellite has a Thiokol AKM that produces an average thrust of 56 kN (12,500 lbf) and burns to depletion in approximately 45 seconds. On-orbit operations are carried out by an array of four 0.44 N (0.1 lbf) thrusters for roll control, ten 2.0 N (0.45 lbf) thrusters for pitch and yaw control and E/W stationkeeping, and two 22.2 N (5.0 lbf) thrusters for repositioning and reorientation. Four 0.3 N (0.07 lbf) EHTs are used for N/S stationkeeping. The nominal mass of the spacecraft at beginning of life (BOL) is 1005 kg and the dry mass at end of life (EOL) is 830 kg. The difference of 175 kg represents the mass of the propellant for a design life of 7 years.

Satellites launched in the late 1980s and 1990s typically have an integrated propulsion system that use a bipropellant combination of monomethyl hydrazine as fuel and nitrogen tetroxide as oxidizer. The specific impulse is about 300 seconds and fuel margin not used for the apogee maneuver can be applied to stationkeeping. Also, since the apogee engine is restartable, it can be used for perigee velocity augmentation and supersynchronous transfer orbit scenarios that optimize the combined propulsion capabilities of the launch vehicle and the spacecraft.

.

For example, the INTELSAT VII satellite, built by Space Systems/Loral, has a Marquardt 490 N apogee thruster and an array of twelve 22 N stationkeeping thrusters manufactured by Atlantic Research Corporation with a 150:1 columbium nozzle expansion ratio and a specific impulse of 235 seconds. For an Ariane launch the separated mass in GTO is 3610 kg, the mass at BOL is 2100 kg, and the mass at EOL is 1450 kg. The mission life is approximately 17 years.

The Hughes HS-601 satellite has a similar thruster configuration. The mass is approximately 2970 kg at launch, 1680 kg at BOL, and 1300 kg for a nominal 14 year mission.

An interesting problem is the estimation of fuel remaining on the spacecraft at any given time during the mission life. This information is used to predict the satellite end of life. There are no “fuel gauges” so the fuel mass must be determined indirectly. There are three principal methods. The first is called the “gas law” method, which is based on the equation of state of an ideal gas. The pressure and temperature of the inert gas in the propellant tanks is measured by transducers and the volume of the gas is computed knowing precisely the pressure and temperature at launch. The volume of the remaining propellant can thus be deduced and the mass determined from the known density as a function of temperature. Corrections must be applied for the expansion of the tanks and the propellant vapor pressure. The second method is called the “bookkeeping” method. In this method the thruster time for each maneuver is carefully measured and recorded. The propellant consumed is then calculated from mass flow rate expressed in terms of the pressure using an empirical model. The third method is much more sophisticated and is based on the measured dynamics of the spacecraft after a stationkeeping maneuver to determine its total mass. In general, these three independent methods provide redundant information that can be applied to check one another.

NEW TECHNOLOGIES

Several innovative technologies have substantially improved the fuel efficiency of satellite stationkeeping thrusters. The savings in fuel can be used to increase the available payload mass, prolong the mission life, or reduce the mass of the spacecraft.

The first of these developments is the electric rocket arcjet technology. The arcjet system uses an electric arc to superheat hydrazine fuel, which nearly doubles its efficiency. An arcjet thruster has a specific impulse of over 500 seconds. Typical thrust levels are from 0.20 to 0.25 N. The arcjet concept was developed by the NASA Lewis Research Center in Cleveland and thrusters have been manufactured commercially by Primex Technologies, a subsidiary of the Olin Corporation.

AT&T’s Telstar 401 satellite, launched in December 1993 (and subsequently lost in 1997 due to an electrical failure generally attributed to a solar flare) was the first satellite to use arcjets. The stationkeeping propellant requirement was reduced by about 40 percent, which was critical to the selection of the Atlas IIAS launch vehicle. Similar arcjet systems are used on INTELSAT VIII and the Lockheed Martin A2100 series of satellites. INTELSAT VIII, for example, has a dual mode propulsion system incorporating a bipropellant liquid apogee engine that burns hydrazine and oxidizer for orbit insertion and four arcjets that use monopropellant hydrazine in the reaction control subsystem for stationkeeping.

Electrothermal hydrazine thrusters continue to have applications on various geostationary satellites and on some small spacecraft where maneuvering time is critical. For example, EHTs are used on the IRIDIUM satellites built by Lockheed Martin.

The most exciting development has been in the field of ion propulsion. The propellant is xenon gas. Although the thrust is small and on the order of a few millinewtons, the specific impulse is from 2000 to 4000 seconds, which is about ten to twenty times the efficiency of conventional bipropellant stationkeeping thrusters. Also, the lower thrust levels have the virtue of minimizing attitude disturbances during stationkeeping maneuvers.

The xenon ion propulsion system, or XIPS (pronounced “zips”), is a gridded ion thruster developed by Hughes. This system is available on the HS-601 HP (high power) and HS-702 satellite models and allows for a reduction in propellant mass of up to 90 percent for a 12 to 15 year mission life. A typical satellite has four XIPS thrusters, including two primary thrusters and two redundant thrusters.

Xenon atoms, an inert monatomic gas with the highest molar mass (131 kg/kmol), are introduced into a thruster chamber ringed by magnets. Electrons emitted by a cathode knock off electrons from the xenon atoms and form positive xenon ions. The ions are accelerated by a pair of gridded electrodes, one with a high positive voltage and one with a negative voltage, at the far end of the thrust chamber and create more than 3000 tiny beams. The beams are neutralized by a flux of electrons emitted by a device called the neutralizer to prevent the ions from being electrically attracted back to the thruster and to prevent a space charge from building up around the satellite.

The increase in kinetic energy of the ions is equal to the work done by the electric field, so that

½ m v2 = q V

where q, m, and v are the charge, mass, and velocity of the ions and V is the accelerating voltage, equal to the algebraic difference between the positive voltage on the positive grid and the negative voltage on the neutralizer. The charge to mass ratio of xenon ions is 7.35 ´ 105 C/kg.

The HS-601 HP satellite uses 13-centimeter diameter XIPS engines to perform north-south stationkeeping and to assist the spacecraft’s gimballed momentum wheel for roll and yaw control. The accelerating voltage is about 750 volts and the ions have a velocity of 33,600 m/s. The specific impulse is 3400 seconds with a mass flow rate of 0.6 mg/s and 18 mN of thrust. Each ion thruster operates for approximately 5 hours per day and uses 500 W from the available 8 kW total spacecraft power.

The HS-702 spacecraft has higher power 25-centimeter thrusters to perform all stationkeeping maneuvers and to complement the four momentum wheels arranged in a tetrahedron configuration for attitude control. The accelerating voltage is 1200 volts, which produces an ion beam with a velocity of 42,500 m/s. The specific impulse is 4300 seconds, the mass flow rate is 4 mg/s, and the thrust is 165 mN. Each HS-702 ion thruster operates for approximately 30 minutes per day and requires 4.5 kW from the 10 to 15 kW solar array. The stationkeeping strategy maintains a tolerance of ± 0.005° that allows for the collocation of several satellites at a single orbital slot.

The HS-702 satellite has a launch mass of up to 5200 kg and an available payload mass of up to 1200 kg. The spacecraft can carry up to 118 transponders, comprising 94 active amplifiers and 24 spares. A bipropellant propulsion system is used for orbit acquisition, with a fuel capacity of 1750 kg. The XIPS thrusters need only 5 kg of xenon propellant per year, a fraction of the requirement for conventional bipropellant or arcjet systems. The HS-702 also has the option of using XIPS thrusters for orbit raising in transfer orbit to further reduce the required propellant mass budget.

The first commercial satellite to use ion propulsion was PAS-5, which was delivered to the PanAmSat Corporation in August 1997. PAS-5 was the first HS-601 HP model, whose xenon ion propulsion system, together with gallium arsenside solar cells and advanced battery performance, permitted the satellite to accommodate a payload twice as powerful as earlier HS-601 models while maintaining a 15 year orbital life. Four more Hughes satellites with XIPS technology were in orbit by the end of 1998. In addition, Hughes also produced a 30-centimeter xenon ion engine for NASA’s Deep Space 1 spacecraft, launched in October 1998.

Another type of ion thruster is the Hall effect ion thruster. The ions are accelerated along the axis of the thruster by crossed electric and magnetic fields. A plasma of electrons in the thrust chamber produces the electric field. A set of coils creates the magnetic field, whose magnitude is the most difficult aspect of the system to adjust. The ions attain a speed of between 15,000 and 20,000 m/s and the specific impulse is about 1800 seconds. This type of thruster has been flown on several Russian spacecraft.

SUMMARY

The demand for ever increasing satellite payloads has motivated the development of propulsion systems with greater efficiency. Typical satellites of fifteen to twenty years ago had solid apogee motors and simple monopropellant hydrazine stationkeeping thrusters. Electrically heated thrusters were designed to increase the hydrazine performance and the principle was further advanced by the innovation of the arcjet thruster. Bipropellant systems are now commonly used for increased performance and versatility.

The future will see a steady transition to ion propulsion. The improvements in fuel efficiency permit the savings in mass to be used for increasing the revenue-generating payloads (with attendant increase in solar arrays, batteries, and thermal control systems to power them), extending the lifetimes in orbit, or reducing the spacecraft mass to permit a more economical launch vehicle.

Author

Dr. Robert A. Nelson, P.E. is president of Satellite Engineering Research Corporation, a satellite engineering consulting firm in Bethesda, Maryland, a Lecturer in the Department of Aerospace Engineering at the University of Maryland and Technical Editor of Via Satellite magazine.