Thursday, 9 November 2017

Science



Mars Mission
India's Mars orbiter mission has left Earth's orbit after performing a manoeuvre to put it on its way to orbit the red planet.
The spacecraft fired its main engine for more than 20 minutes to reach the correct velocity to leave the Earth's orbit, the Bangalore-based Indian Space Research Organisation said. It said that all systems on board the spacecraft were performing normally.
India launched its first spacecraft bound for Mars on 5 November, a complex mission that it hopes will demonstrate and advance technologies for space travel.
The 1.3-tonne orbiter Mangalyaan, which means "Mars craft" in Hindi, must travel 485m miles over 300 days to reach an orbit around Mars next September.
If the mission is successful, India will become the fourth space programme to visit the red planet after the Soviet Union, the US and Europe.
Some have questioned the price tag for a country of 1.2 billion people still dealing with widespread hunger and poverty. But the government defended the Mars mission, and its $1bn space programme in general, by noting its importance in providing hi-tech jobs for scientists and engineers and practical applications in solving problems on Earth.
Decades of space research have allowed India to develop satellite, communications and remote sensing technologies that are helping to solve everyday problems at home, from forecasting where fish can be caught by fishermen to predicting storms and floods.
The orbiter will gather images and data that will help in determining how Martian weather systems work and what happened to the large quantities of water that may have once existed on Mars.
Experts say the data will improve understanding about how planets form, what conditions might make life possible and where else in the universe it might exist.
The orbiter is expected to have at least six months to investigate the planet's landscape and atmosphere. At its closest point, it will be 227 miles from the planet's surface, and its furthest point will be nearly 50,000 miles away.
Since you’re here …
… we have a small favour to ask. More people are reading the Guardian than ever but advertising revenues across the media are falling fast. And unlike many news organisations, we haven’t put up a paywall – we want to keep our journalism as open as we can. So you can see why we need to ask for your help. The Guardian’s independent, investigative journalism takes a lot of time, money and hard work to produce. But we do it because we believe our perspective matters – because it might well be your perspective, too.
I appreciate there not being a paywall: it is more democratic for the media to be available for all and not a commodity to be purchased by a few. I’m happy to make a contribution so others with less means still have access to information. Thomasine F-R.
If everyone who reads our reporting, who likes it, helps fund it, our future would be much more secure. For as little as $1, you can support the Guardian – and it only takes a minute. Thank you.

*  How can a spacecraft leave orbit?

In order to leave orbit, a spacecraft needs to be going fast enough to break free of gravity. A huge push is needed to do that. Either that push was given to a ship as it was launched or it is given to a ship already in orbit. To push a ship that is already in orbit farther from the planet, thrusters should be fired at the highest point in the orbit. Generally, a ship will go into higher and higher orbits until it intersects with its destination.
Spacecraft can go from planet to planet that way. Even if a ship from the Earth leaves Earth orbit, it is still in orbit around the Sun. Huge amounts of energy are needed to push a ship fast enough to break free from the Sun's gravitational pull.
To leave orbit and travel towards the body it is orbiting, a ship only needs to slow down (by retroburning or aerobraking) and wait for gravity to pull the ship in.
**
Matter Invisible
Bypass the light wave in such a way that the Matter in between doesnot become an obstacle and light doesnot appear on the matter.

Speed Increase of a space craft
Make free a matter or space craft from Gravitational force of The Earth and create pressure to create force.

·       If gravity isn't a force, how does it accelerate objects? (Advanced)


Einstein said there is no such thing as a gravitational force. Mass is not attracting mass over a distance. Instead, it's curving spacetime. If there's no force, then how do you explain acceleration due to gravity? Objects should accelerate only when acted upon by a force; otherwise they should maintain a constant velocity. A few of the explanations I've found online refer to equivalence and the thought experiment of a man standing on Earth experiencing the same g-force as a man in a rocket being accelerated in space. I understand why those conditions are the same, but I fail to see how that explains a brick falling from a building accelerating at 9.8 m/s2.  Also, in that thought experiment a force is being exerted (the thrust of the rocket).

This is perhaps the most common question about general relativity. If gravity isn't a force, how does it accelerate objects?

General relativity says that energy (in the form of mass, light, and whatever other forms it comes in) tells spacetime how to bend, and the bending of spacetime tells that energy how to move. The concept of "gravity" is then that objects are falling along the bending of spacetime. The path that objects follow is called a "geodesic". Let's begin by looking at the bending side of things, and then we'll come back to look at geodesics.

The amount of bending that is induced by an object is directly related to that object's energy (typically, the most important part of its energy is its mass energy, but there can be exceptions). The Sun's mass is the biggest contribution to bending in our solar system. So much so, that it dwarfs the bending of spacetime by the Earth to the extent that to a very good approximation, we can just consider the Earth to be massless as it travels around the Sun (we call this the test particle limit). Similarly, when you're standing on the Earth, the Earth's mass dominates the bending of spacetime over your own, and so you can treat yourself as a massless test particle for all intents and purposes. However, truth be told, you warp the spacetime around you just a teensy tiny bit, and that does have an impact upon the earth in response.

Now, let's get back to those geodesics. A body undergoing geodesic motion feels no forces acting upon itself. It is just following what it feels to be a "downward slope through spacetime" (this is how the bending affects the motion of an object). The particular geodesic an object wants to follow is dependent upon its velocity, but perhaps surprisingly, not its mass (unless it is massless, in which case its velocity is exactly the speed of light). There are no forces acting upon that body; we say this body is in freefall. Gravity is not acting as a force. (Technically, if the body is larger than a point, it can have tidal forces acting upon it, which are forces that occur because of a differential in the gravitational effect between the two ends of the body, but we'll ignore those.)

OK, so let's look a little deeper into these geodesic things. What do they look like? Standing on the surface of the Earth, if we throw a ball into the air, it will trace out a parabola through space as it rises and then falls back down to Earth. This is the geodesic that it follows. It turns out that given the appropriate definition, this path is the equivalent of a straight line through four-dimensional spacetime, given the bending of spacetime. How does this relate to what we think of as the acceleration due to gravity?

Let us choose a coordinate system based on our location on the Earth. We'll say that I'm at the origin, and define that we throw the ball up in the air at time t = 0 (this is essentially giving a name to the location, nothing more). We can describe the position of the ball in spacetime in this coordinate system using an appropriate parameter (that we call an "affine parameter"). As the ball moves through spacetime, its position in spacetime is given by appropriate functions of this parameter. We can rewrite things slightly, to relate its position in space to its position in time. Then, when we look at this trajectory, it appears that the object is accelerating towards the earth, giving rise to the idea that gravity is acting as a force.

What is really happening, however, is that the object's motion in our coordinate system is described by the geodesic equation. If you want some maths, this equation looks like the following:

(geodesic equation) (image courtesy of http://en.wikipedia.org/wiki/Geodesic_equation#Affine_geodesics)

Here, x (with superscript Greek indices) describes the position of the ball in our coordinate system. The indices indicate whether we're talking about the x,y,z or time coordinate. The parameter t that the derivatives are being taken with respect to is the affine parameter; in this case, it is known as the "proper time" of the object (for slowly moving objects, we can think of t as the time coordinate in our coordinate system). The first term in this equation is the acceleration of the object in our coordinate system. The second term describes the effect of gravity. The thing that looks like part of a hangman's game is called a connection symbol. It encodes all of the effects of the bending of space time (as well as information about our choice of coordinate system). There are actually sixteen terms here: it's written in a convention called Einstein summation convention. This shows that the effects of the bending of spacetime change the acceleration of an object, based on its velocity through not only space but also through time.

If there is no curvature to spacetime, then the connection symbols are all zero, and we see that an object moves with zero acceleration (constant velocity) unless acted upon by an external force (which would replace the zero on the right-hand side of this equation). (Again, there are some technicalities: this is only true in a Cartesian coordinate system; in something like polar coordinates, the connection symbols may not be vanishing, but they're just describing the vagaries of the coordinate system in that case.)

If there is some bending to spacetime, then the connection symbols are not zero, and all of a sudden, there appears to be an acceleration. It is this curvature of spacetime that gives rise to what we interpret as gravitational acceleration. Note that there is no mass in this equation - it doesn't matter what the mass of the object is, they all follow the same geodesic (so long as it's not massless, in which case things are a little different).

So, what good is this geodesic description of the force of gravity? Can't we just think of gravity as a force and be done with it?

It turns out that there are two cases where this description of the effect of gravity gives vastly different results compared to the concept of gravity as a force. The first is for objects moving very very fast, close to the speed of light. Newtonian gravity doesn't correctly account for the effect of the energy of the object in this case. A particularly important example is for exactly massless particles, such as photons (light). One of the first experimental confirmations of general relativity was that light can be deflected by a mass, such as the sun. Another effect related to light is that as light travels up through the earth's gravitational field, it loses energy. This was actually predicted before general relativity, by considering conservation of energy with a radioactive particle in the earth's gravitational field. However, although the effect was discovered, it had no description in terms of Newtonian gravity.

The second case in which the effect of gravity vastly differs is in the realm of extremely strong gravitational fields, such as those around black holes. Here, the effect of gravity is so severe that not even light can escape from the gravitational pull of such an object. Again, this effect was calculated in Newtonian gravity by thinking about the escape velocity of a body, and contemplating what happens when it gets larger than the speed of light. Surprisingly, the answer you arrive at is exactly the same as in general relativity. However, as light is massless, you once again do not have a good description of this effect in terms of Newtonian gravity, which tells you that there has to be a more complete theory.

So, to summarize, general relativity says that matter bends spacetime, and the effect of that bending of spacetime is to create a generalized kind of force that acts on objects. However, it isn't a force as such that acts on the object, but rather just the object following its geodesic path through spacetime.
****                                                     ****                                                                 ****

Galaxies
Narrow horizontalOn a dark night, we can often see a band of light stretching across the sky. This band is the Milky Way galaxy -- a gigantic collection of stars, gas and dust. Far beyond the Milky Way, there are billions of other galaxies -- some similar to our own and some very different -- scattered throughout space to the very limits of the observable universe.
This portrait of Stephan’s Quintet, also known as Hickson Compact Group 92, was taken by the new Wide Field Camera 3 (WFC3) aboard NASA’s Hubble Space Telescope. NGC 7319, at top right, is a barred spiral with distinct spiral arms that follow nearly 180 degrees back to the bar. Continuing clockwise, the next galaxy appears to have two cores, but it is actually two galaxies, NGC 7318A and NGC 7318B. NGC 7317, at bottom left, is a normal-looking elliptical galaxy that is less affected by the interactions. Sharply contrasting with these galaxies is the dwarf galaxy NGC 7320 at upper left. Bursts of star formation are occurring in the galaxy"s disk, as seen by the blue and pink dots.

Types of Galaxies

Astronomers classify galaxies into three major categories. Spiral galaxies look like flat disks with bulges in their centers and beautiful spiral arms. Elliptical galaxies are redder, more rounded, and often longer in one direction than in the other, like a football. Galaxies that appear neither disk-like nor rounded are classified as irregular galaxies.

Spiral Galaxies

Spiral galaxies usually consist of three components: a flat disk, an ellipsoidally formed bulge and a halo. The disk contains a lot of interstellar gas and dust, and most of the stars in the galaxy. The gas, dust and stars in the disk rotate in the same direction around the galactic center at hundreds of kilometres per second and are often arranged in striking spiral patterns. The bulge is located at the centre of the disk and consists of an older stellar population with little interstellar matter. The near-spherical halo surrounds the disk, and is thought to contain copious amounts of dark matter: matter that acts gravitationally like "normal" matter but that can't be seen! Astronomers infer the presence of this dark matter by the motions of stars and gas in the disk of the galaxy, as well as older stellar populations in the halo like globular clusters. The young stars in the disk are classified as stellar population I, and the old bulge and halo stars as population II.
Narrow horizontalhttp://curious.astro.cornell.edu/images/galaxies/eso-m100.jpg

Astronomers classify spiral galaxies according to their appearance by using the Hubble scheme. Those with pronounced bar structures in their centers are called "barred spirals" and are classified "SB" (examples are given in brackets). Galaxies with conspicuous bulges and tightly wound spiral arms are called "Sa" (Sombrero galaxy) or "SBa" (NGC 3185). Galaxies with prominent bulges and pronounced spiral arms are classified as "Sb" (M31, M81) or "SBb" (M95, NGC 4725). Other spirals with loose spiral arms and a small bulge are classified as "Sc" (M33, M74, M100) or "SBc" (M83, M109).

There are some galaxies like M84, M85 and NGC 5866 that are disk galaxies without any spiral structure. These galaxies are called "S0" or lenticular galaxies. Though the origin of lenticular galaxies is still debated the most plausible explanation to date is that the gas and stars that would reside in the galaxy disk have been stripped by interactions with the hot gas in clusters and groups of galaxies. From their appearance and their stellar contents, they look more like ellipticals rather than spirals and have often been misclassified due to this fact. For instance, misclassification has occured for both the Messier object examples listed above.


* Elliptical Galaxies

Narrow horizontalElliptical galaxies are ellipsoidal agglomerations of stars, which usually do not contain much interstellar matter. Photometric studies of elliptical galaxies suggest that they are triaxial (all the three axes of the ellipsoid are of different sizes). Unlike spiral galaxies, ellipticals have little or no global angular momentum, so that different stars orbit the center in different directions and there is no pattern of orderly rotation. Normally, elliptical galaxies contain very little or no interstellar gas and dust and consist of old population II stars only. Elliptical galaxies are classified according to the Hubble scheme into classes "E0" to "E7", in increasing order of ellipticity. Thus E0 galaxies appear round like M89 while E6 galaxies like M110 and NGC 3377 are almost cigar shaped.

This image of NGC 1132 was taken with Hubble’s Advanced Camera for Surveys. This elliptical galaxy reveals the final result of what may have been a group of galaxies that merged together in the recent past. Another possibility is that the galaxy formed in isolation as a "lone wolf" in a universe ablaze with galaxy groups and clusters.

The largest galaxies in the universe are giant elliptical galaxies. They contain a trillion stars or more and span as much as two million light years - about 20 times the width of the Milky Way. These giant ellipticals are often found in the hearts of galaxy clusters. For example the giant elliptical galaxy M87 is found in the heart of the Virgo Cluster.

Elliptical galaxies also constitute some of the smallest galaxies in the universe. These galaxies are called dwarf elliptical galaxies and dwarf spheroids. Relative to normal ellipticals they are very faint, and are often found in galaxy clusters or near large spiral galaxies. For instance, there are 9 dwarf spheroids like Leo I which are satellites of our Milky Way galaxy.



Irregular Galaxies

A small percentage of the large galaxies we see nearby fall into neither of the two major categories. This irregular class of galaxies is a miscellaneous class, comprising small galaxies with no identifiable form like the Magellanic clouds (the Large Magellanic Cloud and Small Magellanic Cloud are two satellite galaxies of the Milky Way) and "peculiar" galaxies that appear to be in disarray like NGC 1313. There is no discernable disk in these systems, although they often have copious amounts of gas as well as high rates of star formation. Irregular galaxies are often found to be gravitationally interacting with galaxies nearby, which often accounts for their ragged appearance.

Galaxy Evolution, Interactions and Mergers

Galaxies were once thought of as "island universes" evolving slowly in complete isolation. Today we think that just the opposite is true: gravitational interactions of galaxies with each other, and even the coalescence of two galaxies into one, or mergers, are commonplace in the Universe! We see striking examples of merging galaxies in the local Universe, such as NGC 2207 and its companion IC 2163 and the Mice. These interacting systems often sport long tidal tails of gas and stars, a result of the mutual gravitational pull of each system. During the merger, the gas in each galaxy disk flow to the galaxy centers, becomes very dense, and forms stars at an alarming rate. This inflowing material also feeds the supermassive black holes at the galaxy centres, which heat up the infalling material to millions of degrees and eject some of it along powerful jets. All of these mechanisms make merging galaxies very bright, such as Arp 220, are among the most luminous objects in the local Universe.
Narrow horizontalArp 273, a photogenic group of interacting galaxies, lies in the constellation Andromeda and is roughly 300 million light-years away from Earth. This Hubble image shows a tenuous tidal bridge of material between the two galaxies that are separated by tens of thousands of light-years from each other. The larger of the spiral galaxies, known as UGC 1810, has a disk that is tidally distorted into a rose-like shape by the gravitational tidal pull of the companion galaxy below it, known as UGC 1813. The smaller, nearly edge-on companion shows distinct signs of intense star formation at its nucleus, perhaps triggered by the encounter with the companion galaxy.

What do mergers leave behind? Both observations of actual systems and simulations of merging galaxies on a computer suggest that merging spirals create elliptical galaxies. The gas in the progenitor spiral galaxies is used up in making stars which subsequently eject heavier elements and dust from the system, and the collision is forceful enough to randomize the orbits of the stars in the incoming disks into a spheroidal shape. Different types of galaxies are therefore intimately linked by galaxy evolution and mergers: spiral galaxies evolve into elliptical galaxies, and irregular galaxies are galaxies in the process of becoming one or the other!

Galaxy coalescence doesn't only happen between two large galaxies: in fact, most large galaxies are constantly swallowing up the smaller, dwarf galaxies that surround them. Our Milky Way is no exception to this rule: it is currently ripping appart our nearest neighbor, the Sagittarius Dwarf!


Supernovae

In a sense, stars are like people: they are born, they live and they die. A star "lives" by fusing lighter elements into heavier ones in its central regions.

The pressure generated by this "combustion" holds the star up against the enormous gravitational force that its outer layers extert on the stellar core. The supply of elements that the star can fuse is limited, and when this runs out the star "dies": its properties change rapidly and violently, and a new astronomical object is created. Supernovae represent the most catastrophic (and picturesque!) of these stellar deaths.

Anatomy of a Supernova

Stars of all masses spend the majority of their lives fusing hydrogen nuclei into helium nuclei: we call this stage the main sequence. When all of the hydrogen in the central regions of a star is converted into helium, the star will begin to "burn" helium into carbon. However, the helium in the stellar core will eventually run out as well; so in order to survive, a star must be hot enough to fuse progressively heavier elements, as the lighter ones become exhausted one by one. Stars heavier than about 5 times the mass of the Sun can do this with no problem: they burn hydrogen, and then helium, and then carbon, oxygen, silicon, and so on... until they attempt to fuse iron. Iron is special in that it is the lightest element in the periodic table that doesn't release energy when you attempt to fuse it together. In fact, instead of giving you energy, you end up with less energy than you started with! This means that instead of generating additional pressure to hold up the now extended outer layers of the aging star, the iron fusion actually takes thermal energy from the stellar core. Thus, there is nothing left to combat the ever-present force of gravity from these outer layers. The result: collapse! The lack of radiation pressure generated by the iron-fusing core causes the outer layers to fall towards the centre of the star. This implosion happens very, very quickly: it takes about 15 seconds to complete. During the collapse, the nuclei in the outer parts of the star are pushedWhat happens next depends on the mass of the star. Stars with masses between about 5 and 8 times the mass of our Sun form neutron stars during the implosion: the nuclei in the central regions are pushed close enough together to form a very dense neutron core. The outer layers bounce off this core, and a catastrophic explosion ensues: this is the visible part of the supernova. Stars with masses greater than about 10 times the mass of the Sun meet a very different fate. The collapse of the outer regions of the star is so forceful that not even a neutron star can support itself against the pressure of the infalling material. In fact, no physical force is strong enough to counter the collapse: the supernova creates a black hole, or a region of spacetime that is so small and so dense that not even light can escape from its clutches. In this case, the details of how the ensuing explosion actually occurs have still to be worked out. Observationally, supernovae are found by patiently observing the sky and looking for bright objects where there were none before. At its peak luminosity, the supernova resulting from a single star may be bright enough to outshine an entire galaxy. very close together, so close that elements heavier than iron are formed.

A delicate sphere of gas, photographed by NASA’s Hubble Space Telescope, floats serenely in the depths of space. The pristine shell, or bubble, is the result of gas that is being shocked by the expanding blast wave from a supernova. Called SNR 0509-67.5 (or SNR 0509 for short), the bubble is the visible remnant of a powerful stellar explosion in the Large Magellanic Cloud (LMC), a small galaxy about 160,000 light-years from Earth.

A Cosmic Cycle...

 

Supernovae play a fundamental role in a great cosmic recycling program. We believe that almost all of the elements in the Universe that are heavier than hydrogen and helium are created either in the centres of stars during their lifetimes or in the supernova explosions that mark the demise of larger stars. Supernovae then disperse this newly synthesized material in the interstellar neighbourhood. From this material a new, enriched generation of stars will form, and the cycle begins anew. This is how we think that the heavy elements in the Sun came to be. Since the planets in the solar system formed from leftover material in a disk around the proto-Sun, all of the heavy elements in the Earth (including those in humans!) must have come from the same source. This means that in the most literal sense, we are stardust!

 

 

 

 

Warp drive
From Wikipedia, the free encyclopedia
This article is about the fictional propulsion system. For the actual model of spacetime, see Alcubierre drive. For the street in Virginia, see Warp Drive.
Warp drive is a faster-than-light (FTL) spacecraft propulsion system in many science fiction works, most notably Star Trek. A spacecraft equipped with a warp drive may travel at speeds greater than that of light by many orders of magnitude. In contrast to some other FTL technologies such as a jump drive or hyper drive, the warp drive does not permit instantaneous travel between two points but involves a measurable passage of time which is pertinent to the concept. Spacecraft at warp velocity theoretically continue to interact with objects in "normal space". The general concept of "warp drive" was introduced by John W. Campbell in his 1931 novel Islands of Space.[1]
Einstein's theory of special relativity states that energy and mass are interchangeable, thus, speed of light travel is impossible for material objects that weigh more than photons. The problem of a material object exceeding light speed is that an infinitely increasing amount of kinetic energy is required to attempt moving as fast as a massless photon. This problem can theoretically be solved by warping space to move an object instead of increasing the kinetic energy of the object to do so.[2]

Contents

Star Trek

The Original Series: Establishing a background

Warp drive is one of the fundamental features of the Star Trek franchise; in the first pilot episode of Star Trek: The Original Series, "The Cage", it is referred to as a "hyperdrive"/"time warp" drive combination, and it is stated that the "time barrier" has been broken, allowing a group of stranded interstellar travelers to return to Earth far sooner than would have otherwise been possible. The light speed time barrier shouldn't be confused with time dilation which occurs when approaching very fast speeds. Warp drive technology avoids time dilation.
The episode "Metamorphosis", also from The Original Series, establishes a backstory for the invention of warp drive on Earth, in which Zefram Cochrane discovered the "space warp". Cochrane is repeatedly referred to afterwards, but the exact details of the first warp trials were not shown until the second Star Trek: The Next Generation movie, Star Trek: First Contact. The movie depicts Cochrane as having first operated warp drive on Earth in 2063 (two years after the date speculated by the first edition of the Star Trek Chronology). By using a matter/antimatter reactor to create plasma, and by sending this plasma through warp coils, he created a warp bubble which he could use to move a craft into subspace, thus allowing it to exceed the speed of light. This successful first trial led directly to first contact with the Vulcans.

Enterprise: Leading up to The Original Series

Later on, a prequel series titled Star Trek: Enterprise describes the warp engine technology as a "Gravimetric Field Displacement Manifold" (Commander Tucker's tour, "Cold Front"), and describes the device as being powered by a matter/anti-matter reaction which powers the two separate nacelles (one on each side of the ship) to create a displacement field.[citation needed]
The episode also firmly establishes that many other civilizations had warp drive before humans; First Contact co-writer Ronald D. Moore suggested Cochrane's drive was in some way superior to forms which existed beforehand, and was gradually adopted by the galaxy at large.[3]
Enterprise, set in 2151 and onwards, follows the voyages of the first human ship capable of traveling at warp factor 5.2, which under the old warp table formula (the cube of the warp factor times the speed of light), is about 140 times the speed of light (i.e., 5.2 cubed). In the series pilot episode "Broken Bow", Capt. Archer equates warp 4.5 to "Neptune and back [from Earth] in six minutes" (which would correspond to a distance of 547 light-minutes or 66 au, consistent with Neptune being a minimum of 29 au distant from Earth).[citation needed]

The Next Generation onwards

Only three stories in the original Star Trek series involved the Enterprise traveling beyond Warp 10 ( Warp 11, briefly, as a result of Nomad's "correction of inefficiencies" in the antimatter control system in "The Changeling"; Warp 11 again in "By Any Other Name" after the Kelvans modify the Enterprise's engines for greater sustained speed to make the trip from the Milky Way Galaxy to the Andromeda Galaxy; and Warp 14.1 in "That Which Survives" after the ship was put through a Kalandan transporter, beamed parsecs away from where it had been, and reassembled slightly out of phase). In The Next Generation, such stories were rare, and usually involved a malfunction in (or alien interference with) a starship's engines. A new warp scale was drawn up, with Warp Factor 10 set as an unattainable maximum (according to the new scale, reaching or exceeding Warp 10 required an infinite amount of energy). This is described in some technical manuals as "Eugene's limit", in homage to creator/producer Gene Roddenberry. Warp 8 in the original series was the "Never Exceed" speed for the hulls and engines of Constitution-class starships, equivalent to the aircraft VNE V-speed. Warp 6 was the VNO "Normal Operation" maximum safe cruising speed for that vessel class.[4] The Warp 14.1 incident was the result of runaway engines which brought the hull within seconds of structural failure before power was disengaged.[5]
The limit of 10 did not entirely stop warp inflation. By the mid-24th century, the Enterprise-D could travel at Warp 9.8 at "extreme risk", while normal maximum operating speed was Warp 9.6 and the maximum rated cruise was Warp 9.2. According to the Deep Space Nine Tech Manual, during the Dominion War, Galaxy-class starships were refitted with newer technology including modifications which increased their maximum speed to Warp 9.9.
In the episode "Where No One Has Gone Before" the Enterprise-D was shown to exceed Warp 10, traveling 2.7 million light-years from their home galaxy in a matter of minutes (though the ship's extreme velocity was due to the influence of an alien being and could not be achieved by starship engines). The Intrepid-class starship Voyager has a maximum sustainable cruising speed of Warp 9.975; the Enterprise-E can go even faster, with a maximum velocity of Warp 9.999[citation needed]. In the alternative future depicted in "All Good Things...", the series finale of The Next Generation, the "future" Enterprise-D travels at Warp 13, although it is never established whether this is truly "above" Warp 10, or simply the result of another reconfiguration of the warp scale.

Warp velocities

Warp drive velocity in Star Trek is generally expressed in "warp factor" units, which—according to the Star Trek Technical Manuals—correspond to the magnitude of the warp field. Achieving warp factor 1 is equal to breaking the light barrier, while the actual velocity corresponding to higher factors is determined using an ambiguous formula. Several episodes of the original series placed the Enterprise in peril by having it travel at high warp factors; at one point in "That Which Survives" the Enterprise traveled at a warp factor of 14.1. In the Star Trek: The Next Generation episode "The Most Toys" the crew of Enterprise-D discovers that the android Data may have been stolen while on board another ship, Jovis. At this point the Jovis, which has a maximum warp factor of 3, has had a 23-hour head start, which the Enterprise-D figures puts her anywhere within a 0.102 light year radius of her last known position. However, the velocity (in present dimensional units) of any given warp factor is rarely the subject of explicit expression, and travel times for specific interstellar distances are not consistent through the various series.
According to the Star Trek episode writer's guide for The Original Series, warp factors are converted to multiples of c with the cubic function v = w3c, where w is the warp factor, v is the velocity, and c is the speed of light. Accordingly, "warp 1" is equivalent to the speed of light, "warp 2" is 8 times the speed of light, "warp 3" is 27 times the speed of light, etc.
https://upload.wikimedia.org/wikipedia/en/thumb/4/4b/Warptable.gif/250px-Warptable.gif
Michael Okuda's new warp scale
For Star Trek: The Next Generation and the subsequent series, Star Trek artist Michael Okuda devised a formula based on the original one but with important differences; for warp factors 1 through 9, v = w10/3c. In the half-open interval from warp 9 to warp 10, the exponent of w increases toward infinity. Thus, in the Okuda scale, warp velocities approach warp 10 asymptotically.
There is no exact formula for this interval because the quoted velocities are based on a hand-drawn curve; what can be said is that at velocities greater than warp 9, the form of the warp function changes because of an increase in the exponent of the warp factor w. Due to the resultant increase in the derivative, even minor changes in the warp factor eventually correspond to a greater than exponential change in velocity. In the episode "Threshold", Tom Paris breaks the warp 10 threshold, but travel beyond the threshold is later discovered to be unacceptably hazardous to biological life.

Slingshot effect

The "slingshot effect" is first depicted in "Tomorrow Is Yesterday" (1967) as a method of time travel. The procedure involves traveling at a high warp velocity in the direction of a star, on a precisely calculated "slingshot" path; if successful, the ship is caused to travel to a desired point, past or future. The same technique is used in the episode "Assignment: Earth" (1968) for historic research. The term "time warp" was first used in "The Naked Time" (1966) when a previously untried cold-start intermix of matter and antimatter threw the Enterprise back three days in time. The term was later used in Star Trek IV in describing the slingshot effect. The technique was mentioned as a viable method of time travel in the TNG episode "Time Squared" (1989).
This "slingshot" effect has been explored in theoretical physics: it is hypothetically possible to slingshot oneself "around" the event horizon of a black hole. As a result of the black hole's extreme gravitation, time would pass at a slower rate near the event horizon, relative to the outside universe; the traveler would experience the passage of only several minutes or hours, while hundreds of years would pass in 'normal' space.

Warp core

A primary component of the warp drive method of propulsion in the Star Trek universe is the "gravimetric field displacement manifold", more commonly referred to as a warp core. It is a fictional reactor that taps the energy released in a matter-antimatter annihilation to provide the energy necessary to power a starship's warp drive, allowing faster-than-light travel. Starship warp cores generally also serve as powerplants for other primary ship systems.
When matter and antimatter come into contact, they annihilate—both matter and antimatter are converted directly and entirely into enormous quantities of energy, in the form of subnuclear particles and electromagnetic radiation (specifically, mesons and gamma rays). In the Star Trek universe, fictional "dilithium crystals" are used to regulate this reaction. These crystals are described as being non-reactive to anti-matter when bombarded with high levels of radiation.
Usually, the reactants are deuterium, which is an isotope of hydrogen, and antideuterium (its antimatter counterpart). In The Original Series and in-universe chronologically subsequent series, the warp core reaction chamber is often referred to as the "dilithium intermix chamber" or the "matter/antimatter reaction chamber", depending upon the ship's intermix type. The reaction chamber is surrounded by powerful magnetic fields to contain the anti-matter. If the containment fields ever fail, the subsequent interaction of the antimatter fuel with the container walls would result in a catastrophic release of energy, with the resultant explosion capable of utterly destroying the ship. Such "warp core breaches" are used as plot devices in many Star Trek episodes. An intentional warp core breach can also be deliberately created, as one of the methods by which a starship can be made to self-destruct.
The mechanisms that provide a starship's propulsive force are the "warp nacelles", one (or more) cylindrical pods that are offset from the hull of the ship by large pylons; the nacelles generate the actual 'warp bubble' that surrounds the ship, and destruction of one or both nacelles will cripple the ship, and possibly cause a warp-core breach.

Real-world theories and science

https://upload.wikimedia.org/wikipedia/commons/thumb/a/a4/Warp_Requirements_for_10m_OD_Sphere.jpg/220px-Warp_Requirements_for_10m_OD_Sphere.jpg
Warp requirements for 10m OD sphere.
In 1994, physicist Miguel Alcubierre formulated a theoretical solution, called the Alcubierre drive, for faster-than-light travel which models the warp drive concept. Subsequent calculations found that such a model would require prohibitive amounts of negative energy or mass.[6]
In 2012, NASA researcher Harold White hypothesized that by changing the shape of the warp drive, much less negative mass and energy could be used, though the energy required ranges from the mass of Voyager 1 to the mass of the observable universe, or many orders of magnitude greater than anything currently possible by modern technology. NASA engineers have begun preliminary research into such technology.[7]

See also

  • When Stephen Hawking guest starred on the Star Trek: The Next Generation episode "Descent", he was taken on a guided tour of the set. Pausing in front of the warp core set piece, he remarked: "I'm working on that."[8]

References

1.       
  J. Gardiner, "Warp Drive - From Imagination to Reality", Journal of the British Interplanetary Society, vol. 61, p. 353-357 (2008)
   Moore, Ronald D. (1997-10-07). Memory Alpha:AOL chats/Ronald D. Moore/ron063.txt. memory-alpha.org, 7 October 1997. Retrieved from http://memory-alpha.org/wiki/Memory_Alpha:AOL_chats/Ronald_D._Moore/ron063.txt.
   Gene Roddenberry: The Making of Star Trek
   Ford, Lawrence H.; Roman, Thomas A. (2000-01-01). "Negative Energy: Wormholes and Warp Drive". Scientific American.
   Moskowitz, Clara (2012-09-17). "CBS News: Scientists say "warp drive" spaceships could be feasible". Retrieved 2012-09-22.
8.       Shatner, William; Walter, Chip (2002). I'm Working on That: A Trek From Science Fiction to Science Fact. Simon & Schuster. ISBN 0-671-04737-X.


Anti-gravity
From Wikipedia, the free encyclopedia
Anti-gravity also known as non-gravitational field is an idea of creating a place or object that is free from the force of gravity. It does not refer to the lack of weight under gravity experienced in free fall or orbit, or to balancing the force of gravity with some other force, such as electromagnetism or aerodynamic lift. Anti-gravity is a recurring concept in science fiction, particularly in the context of spacecraft propulsion. Examples are the gravity blocking substance "Cavorite" in H. G. Wells' The First Men in the Moon and the Spindizzy machines in James Blish's Cities in Flight.
In Newton's law of universal gravitation, gravity was an external force transmitted by unknown means. In the 20th century, Newton's model was replaced by general relativity where gravity is not a force but the result of the geometry of spacetime. Under general relativity, anti-gravity is impossible except under contrived circumstances.[1][2][3] Quantum physicists have postulated the existence of gravitons, massless elementary particles that transmit gravitational force, but the possibility of creating or destroying these is unclear.
"Anti-gravity" is often used colloquially to refer to devices that look as if they reverse gravity even though they operate through other means, such as lifters, which fly in the air by moving air with electromagnetic fields.[4][5]

Contents

Hypothetical solutions

Gravity shields

https://upload.wikimedia.org/wikipedia/commons/thumb/9/95/New_boston_babson_monument.JPG/220px-New_boston_babson_monument.JPG
A monument at Babson College dedicated to Roger Babson for research into anti-gravity and partial gravity insulators
In 1948 successful businessman Roger Babson (founder of Babson College) formed the Gravity Research Foundation to study ways to reduce the effects of gravity.[6] Their efforts were initially somewhat "crankish", but they held occasional conferences that drew such people as Clarence Birdseye known for his frozen-food products and Igor Sikorsky, inventor of the helicopter. Over time the Foundation turned its attention away from trying to control gravity, to simply better understanding it. The Foundation nearly disappeared after Babson's death in 1967. However, it continues to run an essay award, offering prizes of up to $4,000. As of 2017, it is still administered out of Wellesley, Massachusetts, by George Rideout, Jr., son of the foundation's original director.[7] Winners include California astrophysicist George F. Smoot, who later won the 2006 Nobel Prize in physics.

General relativity research in the 1950s

General relativity was introduced in the 1910s, but development of the theory was greatly slowed by a lack of suitable mathematical tools. Although it appeared that anti-gravity was outlawed under general relativity.
It is claimed the US Air Force also ran a study effort throughout the 1950s and into the 1960s.[8] Former Lieutenant Colonel Ansel Talbert wrote two series of newspaper articles claiming that most of the major aviation firms had started gravity control propulsion research in the 1950s. However, there is little outside confirmation of these stories, and since they take place in the midst of the policy by press release era, it is not clear how much weight these stories should be given.
It is known that there were serious efforts underway at the Glenn L. Martin Company, who formed the Research Institute for Advance Study.[9][10] Major newspapers announced the contract that had been made between theoretical physicist Burkhard Heim and the Glenn L. Martin Company. Another effort in the private sector to master understanding of gravitation was the creation of the Institute for Field Physics, University of North Carolina at Chapel Hill in 1956, by Gravity Research Foundation trustee, Agnew H. Bahnson.
Military support for anti-gravity projects was terminated by the Mansfield Amendment of 1973, which restricted Department of Defense spending to only the areas of scientific research with explicit military applications. The Mansfield Amendment was passed specifically to end long-running projects that had little to show for their efforts.
Under general relativity, gravity is the result of following spatial geometry (change in the normal shape of space) caused by local mass-energy. This theory holds that it is the altered shape of space, deformed by massive objects, that causes gravity, which is actually a property of deformed space rather than being a true force. Although the equations cannot normally produce a "negative geometry", it is possible to do so by using "negative mass". The same equations do not, of themselves, rule out the existence of negative mass.
Both general relativity and Newtonian gravity appear to predict that negative mass would produce a repulsive gravitational field. In particular, Sir Hermann Bondi proposed in 1957 that negative gravitational mass, combined with negative inertial mass, would comply with the strong equivalence principle of general relativity theory and the Newtonian laws of conservation of linear momentum and energy. Bondi's proof yielded singularity free solutions for the relativity equations.[11] In July 1988, Robert L. Forward presented a paper at the AIAA/ASME/SAE/ASEE 24th Joint Propulsion Conference that proposed a Bondi negative gravitational mass propulsion system.[12]
Bondi pointed out that a negative mass will fall toward (and not away from) "normal" matter, since although the gravitational force is repulsive, the negative mass (according to Newton's law, F=ma) responds by accelerating in the opposite of the direction of the force. Normal mass, on the other hand, will fall away from the negative matter. He noted that two identical masses, one positive and one negative, placed near each other will therefore self-accelerate in the direction of the line between them, with the negative mass chasing after the positive mass.[11] Notice that because the negative mass acquires negative kinetic energy, the total energy of the accelerating masses remains at zero. Forward pointed out that the self-acceleration effect is due to the negative inertial mass, and could be seen induced without the gravitational forces between the particles.[12]
The Standard Model of particle physics, which describes all presently known forms of matter, does not include negative mass. Although cosmological dark matter may consist of particles outside the Standard Model whose nature is unknown, their mass is ostensibly known – since they were postulated from their gravitational effects on surrounding objects, which implies their mass is positive. The proposed cosmological dark energy, on the other hand, is more complicated, since according to general relativity the effects of both its energy density and its negative pressure contribute to its gravitational effect.

Fifth force

Under general relativity any form of energy couples with spacetime to create the geometries that cause gravity. A longstanding question was whether or not these same equations applied to antimatter. The issue was considered solved in 1960 with the development of CPT symmetry, which demonstrated that antimatter follows the same laws of physics as "normal" matter, and therefore has positive energy content and also causes (and reacts to) gravity like normal matter (see gravitational interaction of antimatter).
For much of the last quarter of the 20th century, the physics community was involved in attempts to produce a unified field theory, a single physical theory that explains the four fundamental forces: gravity, electromagnetism, and the strong and weak nuclear forces. Scientists have made progress in unifying the three quantum forces, but gravity has remained "the problem" in every attempt. This has not stopped any number of such attempts from being made, however.
Generally these attempts tried to "quantize gravity" by positing a particle, the graviton, that carried gravity in the same way that photons (light) carry electromagnetism. Simple attempts along this direction all failed, however, leading to more complex examples that attempted to account for these problems. Two of these, supersymmetry and the relativity related supergravity, both required the existence of an extremely weak "fifth force" carried by a graviphoton, which coupled together several "loose ends" in quantum field theory, in an organized manner. As a side effect, both theories also all but required that antimatter be affected by this fifth force in a way similar to anti-gravity, dictating repulsion away from mass. Several experiments were carried out in the 1990s to measure this effect, but none yielded positive results.[13]
In 2013 CERN looked for an antigravity effect in an experiment designed to study the energy levels within antihydrogen. The antigravity measurement was just an "interesting sideshow" and was inconclusive.[14]

General-relativistic "warp drives"

There are solutions of the field equations of general relativity which describe "warp drives" (such as the Alcubierre metric) and stable, traversable wormholes. This by itself is not significant, since any spacetime geometry is a solution of the field equations for some configuration of the stress–energy tensor field (see exact solutions in general relativity). General relativity does not constrain the geometry of spacetime unless outside constraints are placed on the stress–energy tensor. Warp-drive and traversable-wormhole geometries are well-behaved in most areas, but require regions of exotic matter; thus they are excluded as solutions if the stress–energy tensor is limited to known forms of matter. Dark matter and dark energy are not understood enough at this present time to make general statements regarding their applicability to a warp-drive.

Breakthrough Propulsion Physics Program

During the close of the twentieth century NASA provided funding for the Breakthrough Propulsion Physics Program (BPP) from 1996 through 2002. This program studied a number of "far out" designs for space propulsion that were not receiving funding through normal university or commercial channels. Anti-gravity-like concepts were investigated under the name "diametric drive". The work of the BPP program continues in the independent, non-NASA affiliated Tau Zero Foundation.[15]

Empirical claims and commercial efforts

There have been a number of attempts to build anti-gravity devices, and a small number of reports of anti-gravity-like effects in the scientific literature. None of the examples that follow are accepted as reproducible examples of anti-gravity.

Gyroscopic devices

https://upload.wikimedia.org/wikipedia/commons/thumb/f/fa/H_W_Wallace_force_field_figure_4.png/350px-H_W_Wallace_force_field_figure_4.png
A "kinemassic field" generator from U.S. Patent 3,626,605: Method and apparatus for generating a secondary gravitational force field
Gyroscopes produce a force when twisted that operates "out of plane" and can appear to lift themselves against gravity. Although this force is well understood to be illusory, even under Newtonian models, it has nevertheless generated numerous claims of anti-gravity devices and any number of patented devices. None of these devices have ever been demonstrated to work under controlled conditions, and have often become the subject of conspiracy theories as a result. A famous example is that of Professor Eric Laithwaite of Imperial College, London, in the 1974 address to the Royal Institution.[16]
Another "rotating device" example is shown in a series of patents granted to Henry Wallace between 1968 and 1974. His devices consist of rapidly spinning disks of brass, a material made up largely of elements with a total half-integer nuclear spin. He claimed that by rapidly rotating a disk of such material, the nuclear spin became aligned, and as a result created a "gravitomagnetic" field in a fashion similar to the magnetic field created by the Barnett effect.[17][18][19] No independent testing or public demonstration of these devices is known.
In 1989, it was reported that a weight decreases along the axis of a right spinning gyroscope.[20] A test of this claim a year later yielded null results.[21] A recommendation was made to conduct further tests at a 1999 AIP conference.[22]

Thomas Townsend Brown's gravitator

In 1921, while still in high school, Thomas Townsend Brown found that a high-voltage Coolidge tube seemed to change mass depending on its orientation on a balance scale. Through the 1920s Brown developed this into devices that combined high voltages with materials with high dielectric constants (essentially large capacitors); he called such a device a "gravitator". Brown made the claim to observers and in the media that his experiments were showing anti-gravity effects. Brown would continue his work and produced a series of high-voltage devices in the following years in attempts to sell his ideas to aircraft companies and the military. He coined the names Biefeld–Brown effect and electrogravitics in conjunction with his devices. Brown tested his asymmetrical capacitor devices in a vacuum, supposedly showing it was not a more down to earth electrohydrodynamic effect generated by high voltage ion flow in air.
Electrogravitics is a popular topic in ufology, anti-gravity, free energy, with government conspiracy theorists and related websites, in books and publications with claims that the technology became highly classified in the early 1960s and that it is used to power UFOs and the B-2 bomber.[23] There is also research and videos on the internet purported to show lifter-style capacitor devices working in a vacuum, therefore not receiving propulsion from ion drift or ion wind being generated in air.[23][24]
Follow-up studies on Brown's work and other claims have been conducted by R. L. Talley in a 1990 US Air Force study, NASA scientist Jonathan Campbell in a 2003 experiment,[25] and Martin Tajmar in a 2004 paper.[26] They have found that no thrust could be observed in a vacuum and that Brown's and other ion lifter devices produce thrust along their axis regardless of the direction of gravity consistent with electrohydrodynamic effects.

Gravitoelectric coupling

In 1992, the Russian researcher Eugene Podkletnov claimed to have discovered, whilst experimenting with superconductors, that a fast rotating superconductor reduces the gravitational effect.[27] Many studies have attempted to reproduce Podkletnov's experiment, always to negative results.[28][29][30][31]
Ning Li and Douglas Torr, of the University of Alabama in Huntsville proposed how a time dependent magnetic field could cause the spins of the lattice ions in a superconductor to generate detectable gravitomagnetic and gravitoelectric fields in a series of papers published between 1991 and 1993.[32][33][34] In 1999, Li and her team appeared in Popular Mechanics, claiming to have constructed a working prototype to generate what she described as "AC Gravity." No further evidence of this prototype has been offered.[35][36]
Douglas Torr and Timir Datta were involved in the development of a "gravity generator" at the University of South Carolina.[37] According to a leaked document from the Office of Technology Transfer at the University of South Carolina and confirmed to Wired reporter Charles Platt in 1998, the device would create a "force beam" in any desired direction and that the university planned to patent and license this device. No further information about this university research project or the "Gravity Generator" device was ever made public.[38]

Göde Award

The Institute for Gravity Research of the Göde Scientific Foundation has tried to reproduce many of the different experiments which claim any "anti-gravity" effects. All attempts by this group to observe an anti-gravity effect by reproducing past experiments have been unsuccessful thus far. The foundation has offered a reward of one million euros for a reproducible anti-gravity experiment.[39]

See also

References

1.       
  Peskin, M and Schroeder, D.; An Introduction to Quantum Field Theory (Westview Press, 1995) ISBN 0-201-50397-2
   Wald, Robert M. (1984). General Relativity. Chicago: University of Chicago Press. ISBN 0-226-87033-2.
   Polchinski, Joseph (1998). String Theory, Cambridge University Press. A modern textbook
   Thompson, Clive (August 2003). "The Antigravity Underground". Wired. Archived from the original on 18 August 2010. Retrieved 23 July 2010.
   "On the Verge of Antigravity". About.com. Retrieved 23 July 2010.
   Mooallem, J. (October 2007). "A curious attraction". Harper's Magazine. 315 (1889): 84–91.
   List of winners
   Goldberg, J. M. (1992). US air force support of general relativity: 1956–1972. In, J. Eisenstaedt & A. J. Kox (Ed.), Studies in the History of General Relativity, Volume 3 Boston, Massachusetts: Center for Einstein Studies. ISBN 0-8176-3479-7
   Mallan, L. (1958). Space satellites (How to book 364). Greenwich, CT: Fawcett Publications, pp. 9–10, 137, 139. LCCN 58-001060
   Clarke, A. C. (1957). "The conquest of gravity". Holiday. 22 (6): 62.
   Bondi, H. (1957). "Negative mass in general relativity". Reviews of Modern Physics. 29 (3): 423–428. Bibcode:1957RvMP...29..423B. doi:10.1103/revmodphys.29.423.
   Forward, R. L. (1990). "Negative matter propulsion". Journal of Propulsion and Power. 6 (1): 28–37. doi:10.2514/3.23219.; see also commentary Landis, G.A. (1991). "Comments on Negative Mass Propulsion". Journal of Propulsion and Power. 7 (2): 304. doi:10.2514/3.23327.
   Supergravity and the Unification of the Laws of Physics, by Daniel Z. Freedman and Peter van Nieuwenhuizen, Scientific American, February 1978
   Tau Zero Foundation
   "Eric LAITHWAITE Gyroscope Levitation". Rex research. rexresearch.com. Retrieved 23 October 2010.
   U.S. Patent 3,626,606
   U.S. Patent 3,626,605
   U.S. Patent 3,823,570
   Nitschke, J. M. & Wilmath, P. A. (1990). "Null result for the weight change of a spinning gyroscope". Physics Review Letters. 64 (18): 2115–2116. Bibcode:1989PhRvL..63.2701H. doi:10.1103/PhysRevLett.64.2115. Retrieved 5 January 2014.
   Iwanaga, N. (1999). "Reviews of some field propulsion methods from the general relativistic standpoint". AIP Conference Proceedings. 458: 1015–1059..
   Thompson, Clive (August 2003). "The Antigravity Underground". Wired Magazine.
   Thomas Valone, Electrogravitics II: Validating Reports on a New Propulsion Methodology, Integrity Research Institute, page 52-58
   Thompson, Clive (August 2003). "The Antigravity Underground". Wired Magazine.
   Tajmar, M. (2004). "Biefeld-Brown Effect: Misinterpretation of Corona Wind Phenomena". AIAA Journal. 42 (2): 315–318. Bibcode:2004AIAAJ..42..315T. doi:10.2514/1.9095.
   Podkletnov, E; Nieminen, R (10 December 1992). "A possibility of gravitational force shielding by bulk YBa2Cu3O7−x superconductor". Physica C. 203 (3–4): 441–444. Bibcode:1992PhyC..203..441P. doi:10.1016/0921-4534(92)90055-H. Retrieved 29 April 2014.
   N. Li; D. Noever; T. Robertson; R. Koczor; et al. (August 1997). "Static Test for a Gravitational Force Coupled to Type II YBCO Superconductors". Physica C. 281 (2–3): 260–267. Bibcode:1997PhyC..281..260L. doi:10.1016/S0921-4534(97)01462-7.
   Woods, C., Cooke, S., Helme, J., and Caldwell, C., "Gravity Modification by High Temperature Superconductors," Joint Propulsion Conference, AIAA 2001–3363, (2001).
   Hathaway, G., Cleveland, B., and Bao, Y., "Gravity Modification Experiment using a Rotating Superconducting Disc and Radio Frequency Fields," Physica C, 385, 488–500, (2003).
   Tajmar, M., and de Matos, C.J., "Gravitomagnetic Field of a Rotating Superconductor and of a Rotating Superfluid," Physica C, 385(4), 551–554, (2003).
   Li, Ning; Torr, DG (1 September 1992). "Gravitational effects on the magnetic attenuation of superconductors". Physical Review. B46: 5489–5495. Bibcode:1992PhRvB..46.5489L. doi:10.1103/PhysRevB.46.5489. Retrieved 6 March 2014.
   Li, Ning; Torr, DG (15 January 1991). "Effects of a gravitomagnetic field on pure superconductors". Physical Review. D43: 457–459. Bibcode:1991PhRvD..43..457L. doi:10.1103/PhysRevD.43.457. Retrieved 6 March 2014.
   Li, Ning; Torr, DG (August 1993). "Gravitoelectric-electric coupling via superconductivity". Foundations of Physics Letters. 6 (4): 371–383. Bibcode:1993FoPhL...6..371T. doi:10.1007/BF00665654. Retrieved 6 March 2014.
   Wilson, Jim (1 October 2000). "Taming Gravity". Popular Mechanics. HighBeam Reseatch. Retrieved 5 January 2014.
   Cain, Jeanette. "Gravity Conquered?". light-science.com. Archived from the original on 6 July 2013. Retrieved 5 January 2014.
   Platt, Charles (3 June 1998). "Breaking the Law of Gravity". Wired. Retrieved 1 May 2014.
39.  "The Göde award - One Million Euro to overcome gravity". Institute of Gravity Research. Retrieved 2 January 2014.

Further reading

  • Cady, W. M. (15 September 1952). "Thomas Townsend Brown: Electro-Gravity Device" (File 24-185). Pasadena, CA: Office of Naval Research. Public access to the report was authorized on 1 October 1952.


Spacecraft propulsion
From Wikipedia, the free encyclopedia
https://upload.wikimedia.org/wikipedia/commons/thumb/b/b1/Shuttle_Main_Engine_Test_Firing.jpg/220px-Shuttle_Main_Engine_Test_Firing.jpg
A remote camera captures a close-up view of a Space Shuttle Main Engine during a test firing at the John C. Stennis Space Center in Hancock County, Mississippi.
Spacecraft propulsion is any method used to accelerate spacecraft and artificial satellites. There are many different methods. Each method has drawbacks and advantages, and spacecraft propulsion is an active area of research. However, most spacecraft today are propelled by forcing a gas from the back/rear of the vehicle at very high speed through a supersonic de Laval nozzle. This sort of engine is called a rocket engine.
All current spacecraft use chemical rockets (bipropellant or solid-fuel) for launch, though some (such as the Pegasus rocket and SpaceShipOne) have used air-breathing engines on their first stage. Most satellites have simple reliable chemical thrusters (often monopropellant rockets) or resistojet rockets for orbital station-keeping and some use momentum wheels for attitude control. Soviet bloc satellites have used electric propulsion for decades, and newer Western geo-orbiting spacecraft are starting to use them for north-south stationkeeping and orbit raising. Interplanetary vehicles mostly use chemical rockets as well, although a few have used ion thrusters and Hall effect thrusters (two different types of electric propulsion) to great success.

Contents

Requirements

Further information: Escape velocity
Artificial satellites must be launched into orbit and once there they must be placed in their nominal orbit. Once in the desired orbit, they often need some form of attitude control so that they are correctly pointed with respect to Earth, the Sun, and possibly some astronomical object of interest.[1] They are also subject to drag from the thin atmosphere, so that to stay in orbit for a long period of time some form of propulsion is occasionally necessary to make small corrections (orbital stationkeeping).[2] Many satellites need to be moved from one orbit to another from time to time, and this also requires propulsion.[3] A satellite's useful life is usually over once it has exhausted its ability to adjust its orbit.
Spacecraft designed to travel further also need propulsion methods. They need to be launched out of the Earth's atmosphere just as satellites do. Once there, they need to leave orbit and move around.
For interplanetary travel, a spacecraft must use its engines to leave Earth orbit. Once it has done so, it must somehow make its way to its destination. Current interplanetary spacecraft do this with a series of short-term trajectory adjustments.[4] In between these adjustments, the spacecraft simply falls freely along its trajectory. The most fuel-efficient means to move from one circular orbit to another is with a Hohmann transfer orbit: the spacecraft begins in a roughly circular orbit around the Sun. A short period of thrust in the direction of motion accelerates or decelerates the spacecraft into an elliptical orbit around the Sun which is tangential to its previous orbit and also to the orbit of its destination. The spacecraft falls freely along this elliptical orbit until it reaches its destination, where another short period of thrust accelerates or decelerates it to match the orbit of its destination.[5] Special methods such as aerobraking or aerocapture are sometimes used for this final orbital adjustment.[6]
https://upload.wikimedia.org/wikipedia/commons/thumb/d/de/Ssunsail.jpg/220px-Ssunsail.jpg
Artist's concept of a solar sail
Some spacecraft propulsion methods such as solar sails provide very low but inexhaustible thrust;[7] an interplanetary vehicle using one of these methods would follow a rather different trajectory, either constantly thrusting against its direction of motion in order to decrease its distance from the Sun or constantly thrusting along its direction of motion to increase its distance from the Sun. The concept has been successfully tested by the Japanese IKAROS solar sail spacecraft.
Spacecraft for interstellar travel also need propulsion methods. No such spacecraft has yet been built, but many designs have been discussed. Because interstellar distances are very great, a tremendous velocity is needed to get a spacecraft to its destination in a reasonable amount of time. Acquiring such a velocity on launch and getting rid of it on arrival will be a formidable challenge for spacecraft designers.[8]

Effectiveness

When in space, the purpose of a propulsion system is to change the velocity, or v, of a spacecraft. Because this is more difficult for more massive spacecraft, designers generally discuss momentum, mv. The amount of change in momentum is called impulse.[9] So the goal of a propulsion method in space is to create an impulse.
When launching a spacecraft from Earth, a propulsion method must overcome a higher gravitational pull to provide a positive net acceleration.[10] In orbit, any additional impulse, even very tiny, will result in a change in the orbit path.
The rate of change of velocity is called acceleration, and the rate of change of momentum is called force. To reach a given velocity, one can apply a small acceleration over a long period of time, or one can apply a large acceleration over a short time. Similarly, one can achieve a given impulse with a large force over a short time or a small force over a long time. This means that for maneuvering in space, a propulsion method that produces tiny accelerations but runs for a long time can produce the same impulse as a propulsion method that produces large accelerations for a short time. When launching from a planet, tiny accelerations cannot overcome the planet's gravitational pull and so cannot be used.
Earth's surface is situated fairly deep in a gravity well. The escape velocity required to get out of it is 11.2 kilometers/second. As human beings evolved in a gravitational field of 1g (9.8 m/s²), an ideal propulsion system would be one that provides a continuous acceleration of 1g (though human bodies can tolerate much larger accelerations over short periods). The occupants of a rocket or spaceship having such a propulsion system would be free from all the ill effects of free fall, such as nausea, muscular weakness, reduced sense of taste, or leaching of calcium from their bones.
The law of conservation of momentum means that in order for a propulsion method to change the momentum of a space craft it must change the momentum of something else as well. A few designs take advantage of things like magnetic fields or light pressure in order to change the spacecraft's momentum, but in free space the rocket must bring along some mass to accelerate away in order to push itself forward. Such mass is called reaction mass.
In order for a rocket to work, it needs two things: reaction mass and energy. The impulse provided by launching a particle of reaction mass having mass m at velocity v is mv. But this particle has kinetic energy mv²/2, which must come from somewhere. In a conventional solid, liquid, or hybrid rocket, the fuel is burned, providing the energy, and the reaction products are allowed to flow out the back, providing the reaction mass. In an ion thruster, electricity is used to accelerate ions out the back. Here some other source must provide the electrical energy (perhaps a solar panel or a nuclear reactor), whereas the ions provide the reaction mass.[10]
When discussing the efficiency of a propulsion system, designers often focus on effectively using the reaction mass. Reaction mass must be carried along with the rocket and is irretrievably consumed when used. One way of measuring the amount of impulse that can be obtained from a fixed amount of reaction mass is the specific impulse, the impulse per unit weight-on-Earth (typically designated by I sp {\displaystyle I_{\text{sp}}} ). The unit for this value is seconds. Because the weight on Earth of the reaction mass is often unimportant when discussing vehicles in space, specific impulse can also be discussed in terms of impulse per unit mass. This alternate form of specific impulse uses the same units as velocity (e.g. m/s), and in fact it is equal to the effective exhaust velocity of the engine (typically designated v e {\displaystyle v_{e}} ). Confusingly, both values are sometimes called specific impulse. The two values differ by a factor of gn, the standard acceleration due to gravity 9.80665 m/s² ( I sp g n = v e {\displaystyle I_{\text{sp}}g_{\mathrm {n} }=v_{e}} ). A rocket with a high exhaust velocity can achieve the same impulse with less reaction mass. However, the energy required for that impulse is proportional to the exhaust velocity, so that more mass-efficient engines require much more energy, and are typically less energy efficient. This is a problem if the engine is to provide a large amount of thrust. To generate a large amount of impulse per second, it must use a large amount of energy per second. So high-mass-efficient engines require enormous amounts of energy per second to produce high thrusts. As a result, most high-mass-efficient engine designs also provide lower thrust due to the unavailability of high amounts of energy.

Methods

Propulsion methods can be classified based on their means of accelerating the reaction mass. There are also some special methods for launches, planetary arrivals, and landings.

Reaction engines

Main article: Reaction engine
A reaction engine is an engine which provides propulsion by expelling reaction mass, in accordance with Newton's third law of motion. This law of motion is most commonly paraphrased as: "For every action there is an equal, and opposite, reaction".
Examples include both duct engines and rocket engines, and more uncommon variations such as Hall effect thrusters, ion drives and mass drivers. Duct engines are obviously not used for space propulsion due to the lack of air; however some proposed spacecraft have these kinds of engines to assist takeoff and landing.

Delta-v and propellant

https://upload.wikimedia.org/wikipedia/commons/thumb/f/f6/Tsiolkovsky_rocket_equation.svg/220px-Tsiolkovsky_rocket_equation.svg.png
Rocket mass ratios versus final velocity, as calculated from the rocket equation
Exhausting the entire usable propellant of a spacecraft through the engines in a straight line in free space would produce a net velocity change to the vehicle; this number is termed 'delta-v' ( Δ v {\displaystyle \Delta v} ).
If the exhaust velocity is constant then the total Δ v {\displaystyle \Delta v} of a vehicle can be calculated using the rocket equation, where M is the mass of propellant, P is the mass of the payload (including the rocket structure), and v e {\displaystyle v_{e}} is the velocity of the rocket exhaust. This is known as the Tsiolkovsky rocket equation:
Δ v = v e ln ( M + P P ) . {\displaystyle \Delta v=v_{e}\ln \left({\frac {M+P}{P}}\right).}
For historical reasons, as discussed above, v e {\displaystyle v_{e}} is sometimes written as
v e = I sp g 0 {\displaystyle v_{e}=I_{\text{sp}}g_{0}}
where I sp {\displaystyle I_{\text{sp}}} is the specific impulse of the rocket, measured in seconds, and g 0 {\displaystyle g_{0}} is the gravitational acceleration at sea level. For a high delta-v mission, the majority of the spacecraft's mass needs to be reaction mass. Because a rocket must carry all of its reaction mass, most of the initially-expended reaction mass goes towards accelerating reaction mass rather than payload. If the rocket has a payload of mass P, the spacecraft needs to change its velocity by Δ v {\displaystyle \Delta v} , and the rocket engine has exhaust velocity ve, then the reaction mass M which is needed can be calculated using the rocket equation and the formula for I sp {\displaystyle I_{\text{sp}}} :
M = P ( e Δ v v e − 1 ) . {\displaystyle M=P\left(e^{\frac {\Delta v}{v_{e}}}-1\right).}
For Δ v {\displaystyle \Delta v} much smaller than ve, this equation is roughly linear, and little reaction mass is needed. If Δ v {\displaystyle \Delta v} is comparable to ve, then there needs to be about twice as much fuel as combined payload and structure (which includes engines, fuel tanks, and so on). Beyond this, the growth is exponential; speeds much higher than the exhaust velocity require very high ratios of fuel mass to payload and structural mass. For a mission, for example, when launching from or landing on a planet, the effects of gravitational attraction and any atmospheric drag must be overcome by using fuel. It is typical to combine the effects of these and other effects into an effective mission delta-v. For example, a launch mission to low Earth orbit requires about 9.3–10 km/s delta-v. These mission delta-vs are typically numerically integrated on a computer.
Some effects such as Oberth effect can only be significantly utilised by high thrust engines such as rockets; i.e., engines that can produce a high g-force (thrust per unit mass, equal to delta-v per unit time).

Power use and propulsive efficiency

For all reaction engines (such as rockets and ion drives) some energy must go into accelerating the reaction mass. Every engine will waste some energy, but even assuming 100% efficiency, to accelerate an exhaust the engine will need energy amounting to
1 2 m ˙ v e 2 {\displaystyle {\frac {1}{2}}{\dot {m}}v_{e}^{2}} [11]
This energy is not necessarily lost- some of it usually ends up as kinetic energy of the vehicle, and the rest is wasted in residual motion of the exhaust.
https://upload.wikimedia.org/wikipedia/commons/thumb/2/2b/Rocket_propulsion_efficiency.svg/220px-Rocket_propulsion_efficiency.svg.png
Due to energy carried away in the exhaust, the energy efficiency of a reaction engine varies with the speed of the exhaust relative to the speed of the vehicle, this is called propulsive efficiency
Comparing the rocket equation (which shows how much energy ends up in the final vehicle) and the above equation (which shows the total energy required) shows that even with 100% engine efficiency, certainly not all energy supplied ends up in the vehicle - some of it, indeed usually most of it, ends up as kinetic energy of the exhaust.
The exact amount depends on the design of the vehicle, and the mission. However, there are some useful fixed points:
  • if the I sp {\displaystyle I_{\text{sp}}} is fixed, for a mission delta-v, there is a particular I sp {\displaystyle I_{\text{sp}}} that minimises the overall energy used by the rocket. This comes to an exhaust velocity of about ⅔ of the mission delta-v (see the energy computed from the rocket equation). Drives with a specific impulse that is both high and fixed such as Ion thrusters have exhaust velocities that can be enormously higher than this ideal for many missions.
  • if the exhaust velocity can be made to vary so that at each instant it is equal and opposite to the vehicle velocity then the absolute minimum energy usage is achieved. When this is achieved, the exhaust stops in space [1] and has no kinetic energy; and the propulsive efficiency is 100%- all the energy ends up in the vehicle (in principle such a drive would be 100% efficient, in practice there would be thermal losses from within the drive system and residual heat in the exhaust). However, in most cases this uses an impractical quantity of propellant, but is a useful theoretical consideration. Anyway, the vehicle has to move before the method can be applied.
  • Some drives (such as VASIMR or electrodeless plasma thruster) actually can significantly vary their exhaust velocity. This can help reduce propellant usage or improve acceleration at different stages of the flight. However the best energetic performance and acceleration is still obtained when the exhaust velocity is close to the vehicle speed. Proposed ion and plasma drives usually have exhaust velocities enormously higher than that ideal (in the case of VASIMR the lowest quoted speed is around 15000 m/s compared to a mission delta-v from high Earth orbit to Mars of about 4000 m/s).
    It might be thought that adding power generation capacity is helpful, and although initially this can improve performance, this inevitably increases the weight of the power source, and eventually the mass of the power source and the associated engines and propellant dominates the weight of the vehicle, and then adding more power gives no significant improvement.
    For, although solar power and nuclear power are virtually unlimited sources of energy, the maximum power they can supply is substantially proportional to the mass of the powerplant (i.e. specific power takes a largely constant value which is dependent on the particular powerplant technology). For any given specific power, with a large v e {\displaystyle v_{e}} which is desirable to save propellant mass, it turns out that the maximum acceleration is inversely proportional to v e {\displaystyle v_{e}} . Hence the time to reach a required delta-v is proportional to v e {\displaystyle v_{e}} . Thus the latter should not be too large.

    Energy

    https://upload.wikimedia.org/wikipedia/commons/thumb/5/50/Average_propulsive_efficiency_of_rockets.png/330px-Average_propulsive_efficiency_of_rockets.png
    Plot of instantaneous propulsive efficiency (blue) and overall efficiency for a vehicle accelerating from rest (red) as percentages of the engine efficiency
    In the ideal case m 1 {\displaystyle m_{1}} is useful payload and m 0 − m 1 {\displaystyle m_{0}-m_{1}} is reaction mass (this corresponds to empty tanks having no mass, etc.). The energy required can simply be computed as
    1 2 ( m 0 − m 1 ) v e 2 {\displaystyle {\frac {1}{2}}(m_{0}-m_{1})v_{\text{e}}^{2}}
    This corresponds to the kinetic energy the expelled reaction mass would have at a speed equal to the exhaust speed. If the reaction mass had to be accelerated from zero speed to the exhaust speed, all energy produced would go into the reaction mass and nothing would be left for kinetic energy gain by the rocket and payload. However, if the rocket already moves and accelerates (the reaction mass is expelled in the direction opposite to the direction in which the rocket moves) less kinetic energy is added to the reaction mass. To see this, if, for example, v e {\displaystyle v_{e}} =10 km/s and the speed of the rocket is 3 km/s, then the speed of a small amount of expended reaction mass changes from 3 km/s forwards to 7 km/s rearwards. Thus, although the energy required is 50 MJ per kg reaction mass, only 20 MJ is used for the increase in speed of the reaction mass. The remaining 30 MJ is the increase of the kinetic energy of the rocket and payload.
    In general:
    d ( 1 2 v 2 ) = v d v = v v e d m m = 1 2 [ v e 2 − ( v − v e ) 2 + v 2 ] d m m {\displaystyle d\left({\frac {1}{2}}v^{2}\right)=vdv=vv_{\text{e}}{\frac {dm}{m}}={\frac {1}{2}}\left[v_{\text{e}}^{2}-\left(v-v_{\text{e}}\right)^{2}+v^{2}\right]{\frac {dm}{m}}}
    Thus the specific energy gain of the rocket in any small time interval is the energy gain of the rocket including the remaining fuel, divided by its mass, where the energy gain is equal to the energy produced by the fuel minus the energy gain of the reaction mass. The larger the speed of the rocket, the smaller the energy gain of the reaction mass; if the rocket speed is more than half of the exhaust speed the reaction mass even loses energy on being expelled, to the benefit of the energy gain of the rocket; the larger the speed of the rocket, the larger the energy loss of the reaction mass.
    We have
    Δ ϵ = ∫ v d ( Δ v ) {\displaystyle \Delta \epsilon =\int v\,d(\Delta v)}
    where ϵ {\displaystyle \epsilon } is the specific energy of the rocket (potential plus kinetic energy) and Δ v {\displaystyle \Delta v} is a separate variable, not just the change in v {\displaystyle v} . In the case of using the rocket for deceleration; i.e., expelling reaction mass in the direction of the velocity, v {\displaystyle v} should be taken negative. The formula is for the ideal case again, with no energy lost on heat, etc. The latter causes a reduction of thrust, so it is a disadvantage even when the objective is to lose energy (deceleration).
    If the energy is produced by the mass itself, as in a chemical rocket, the fuel value has to be v e 2 / 2 {\displaystyle \scriptstyle {v_{\text{e}}^{2}/2}} , where for the fuel value also the mass of the oxidizer has to be taken into account. A typical value is v e {\displaystyle v_{\text{e}}} = 4.5 km/s, corresponding to a fuel value of 10.1 MJ/kg. The actual fuel value is higher, but much of the energy is lost as waste heat in the exhaust that the nozzle was unable to extract. The required energy E {\displaystyle E} is
    E = 1 2 m 1 ( e Δ v v e − 1 ) v e 2 {\displaystyle E={\frac {1}{2}}m_{1}\left(e^{\frac {\Delta v}{v_{\text{e}}}}-1\right)v_{\text{e}}^{2}}
    Conclusions:
    • for Δ v v e {\displaystyle \Delta v\ll v_{e}} we have E ≈ 1 2 m 1 v e Δ v {\displaystyle E\approx {\frac {1}{2}}m_{1}v_{\text{e}}\Delta v}
    • for a given Δ v {\displaystyle \Delta v} , the minimum energy is needed if v e = 0.6275 Δ v {\displaystyle v_{\text{e}}=0.6275\Delta v} , requiring an energy of
      E = 0.772 m 1 ( Δ v ) 2 {\displaystyle E=0.772m_{1}(\Delta v)^{2}} .
      In the case of acceleration in a fixed direction, and starting from zero speed, and in the absence of other forces, this is 54.4% more than just the final kinetic energy of the payload. In this optimal case the initial mass is 4.92 times the final mass.
      These results apply for a fixed exhaust speed.
      Due to the Oberth effect and starting from a nonzero speed, the required potential energy needed from the propellant may be less than the increase in energy in the vehicle and payload. This can be the case when the reaction mass has a lower speed after being expelled than before – rockets are able to liberate some or all of the initial kinetic energy of the propellant.
      Also, for a given objective such as moving from one orbit to another, the required Δ v {\displaystyle \Delta v} may depend greatly on the rate at which the engine can produce Δ v {\displaystyle \Delta v} and maneuvers may even be impossible if that rate is too low. For example, a launch to Low Earth orbit (LEO) normally requires a Δ v {\displaystyle \Delta v} of ca. 9.5 km/s (mostly for the speed to be acquired), but if the engine could produce Δ v {\displaystyle \Delta v} at a rate of only slightly more than g, it would be a slow launch requiring altogether a very large Δ v {\displaystyle \Delta v} (think of hovering without making any progress in speed or altitude, it would cost a Δ v {\displaystyle \Delta v} of 9.8 m/s each second). If the possible rate is only g {\displaystyle g} or less, the maneuver can not be carried out at all with this engine. The power is given by P = 1 2 m a v e = 1 2 F v e {\displaystyle P={\frac {1}{2}}mav_{\text{e}}={\frac {1}{2}}Fv_{\text{e}}} where F {\displaystyle F} is the thrust and a {\displaystyle a} the acceleration due to it. Thus the theoretically possible thrust per unit power is 2 divided by the specific impulse in m/s. The thrust efficiency is the actual thrust as percentage of this. If, e.g., solar power is used, this restricts a {\displaystyle a} ; in the case of a large v e {\displaystyle v_{\text{e}}} the possible acceleration is inversely proportional to it, hence the time to reach a required delta-v is proportional to v e {\displaystyle v_{\text{e}}} ; with 100% efficiency: for Δ v ≪ v e {\displaystyle \Delta v\ll v_{\text{e}}} we have t ≈ m v e Δ v 2 P {\displaystyle t\approx {\frac {mv_{\text{e}}\Delta v}{2P}}} Examples: power, 1000 W; mass, 100 kg; Δ v {\displaystyle \Delta v} = 5 km/s, v e {\displaystyle v_{\text{e}}} = 16 km/s, takes 1.5 months. power, 1000 W; mass, 100 kg; Δ v {\displaystyle \Delta v} = 5 km/s, v e {\displaystyle v_{\text{e}}} = 50 km/s, takes 5 months. Thus v e {\displaystyle v_{\text{e}}} should not be too large. Power to thrust ratio The power to thrust ratio is simply:[11] P F = 1 2 m ˙ v 2 m ˙ v = 1 2 v {\displaystyle {\frac {P}{F}}={\frac {{\frac {1}{2}}{{\dot {m}}v^{2}}}{{\dot {m}}v}}={\frac {1}{2}}v} Thus for any vehicle power P, the thrust that may be provided is: F = P 1 2 v = 2 P v {\displaystyle F={\frac {P}{{\frac {1}{2}}v}}={\frac {2P}{v}}} Example Suppose a 10,000 kg space probe will be sent to Mars. The required Δ v {\displaystyle \Delta v} from LEO is approximately 3000 m/s, using a Hohmann transfer orbit. For the sake of argument, assume the following thrusters are options to be used: Engine Effective exhaust velocity (km/s) Specific impulse (s) Fuel mass (kg) Energy required (GJ) Energy per kg of propellant Minimum[a] power/thrust Power generator mass/thrust[b] Solid rocket 1 100 190,000 95 500 kJ 0.5 kW/N N/A Bipropellant rocket 5 500 8,200 103 12.6 MJ 2.5 kW/N N/A Ion thruster 50 5,000 620 775 1.25 GJ 25 kW/N 25 kg/N 1.          Assuming 100% energetic efficiency; 50% is more typical in practice. 2.        Assumes a specific power of 1 kW/kg Observe that the more fuel-efficient engines can use far less fuel; their mass is almost negligible (relative to the mass of the payload and the engine itself) for some of the engines. However, note also that these require a large total amount of energy. For Earth launch, engines require a thrust to weight ratio of more than one. To do this with the ion or more theoretical electrical drives, the engine would have to be supplied with one to several gigawatts of power, equivalent to a major metropolitan generating station. From the table it can be seen that this is clearly impractical with current power sources. Alternative approaches include some forms of laser propulsion, where the reaction mass does not provide the energy required to accelerate it, with the energy instead being provided from an external laser or other beam-powered propulsion system. Small models of some of these concepts have flown, although the engineering problems are complex and the ground based power systems are not a solved problem. Instead, a much smaller, less powerful generator may be included which will take much longer to generate the total energy needed. This lower power is only sufficient to accelerate a tiny amount of fuel per second, and would be insufficient for launching from Earth. However, over long periods in orbit where there is no friction, the velocity will be finally achieved. For example, it took the SMART-1 more than a year to reach the Moon, whereas with a chemical rocket it takes a few days. Because the ion drive needs much less fuel, the total launched mass is usually lower, which typically results in a lower overall cost, but the journey takes longer. Mission planning therefore frequently involves adjusting and choosing the propulsion system so as to minimise the total cost of the project, and can involve trading off launch costs and mission duration against payload fraction. Rocket engines Main article: Rocket engine SpaceX's Kestrel engine is tested Most rocket engines are internal combustion heat engines (although non combusting forms exist). Rocket engines generally produce a high temperature reaction mass, as a hot gas. This is achieved by combusting a solid, liquid or gaseous fuel with an oxidiser within a combustion chamber. The extremely hot gas is then allowed to escape through a high-expansion ratio nozzle. This bell-shaped nozzle is what gives a rocket engine its characteristic shape. The effect of the nozzle is to dramatically accelerate the mass, converting most of the thermal energy into kinetic energy. Exhaust speed reaching as high as 10 times the speed of sound at sea level are common. Rocket engines provide essentially the highest specific powers and high specific thrusts of any engine used for spacecraft propulsion. Ion propulsion rockets can heat a plasma or charged gas inside a magnetic bottle and release it via a magnetic nozzle, so that no solid matter need come in contact with the plasma. Of course, the machinery to do this is complex, but research into nuclear fusion has developed methods, some of which have been proposed to be used in propulsion systems, and some have been tested in a lab. See rocket engine for a listing of various kinds of rocket engines using different heating methods, including chemical, electrical, solar, and nuclear. Electromagnetic propulsion This test engine accelerates ions using electrostatic forces Main article: Electrically powered spacecraft propulsion Rather than relying on high temperature and fluid dynamics to accelerate the reaction mass to high speeds, there are a variety of methods that use electrostatic or electromagnetic forces to accelerate the reaction mass directly. Usually the reaction mass is a stream of ions. Such an engine typically uses electric power, first to ionize atoms, and then to create a voltage gradient to accelerate the ions to high exhaust velocities. The idea of electric propulsion dates back to 1906, when Robert Goddard considered the possibility in his personal notebook.[12] Konstantin Tsiolkovsky published the idea in 1911. For these drives, at the highest exhaust speeds, energetic efficiency and thrust are all inversely proportional to exhaust velocity. Their very high exhaust velocity means they require huge amounts of energy and thus with practical power sources provide low thrust, but use hardly any fuel. For some missions, particularly reasonably close to the Sun, solar energy may be sufficient, and has very often been used, but for others further out or at higher power, nuclear energy is necessary; engines drawing their power from a nuclear source are called nuclear electric rockets. With any current source of electrical power, chemical, nuclear or solar, the maximum amount of power that can be generated limits the amount of thrust that can be produced to a small value. Power generation adds significant mass to the spacecraft, and ultimately the weight of the power source limits the performance of the vehicle. Current nuclear power generators are approximately half the weight of solar panels per watt of energy supplied, at terrestrial distances from the Sun. Chemical power generators are not used due to the far lower total available energy. Beamed power to the spacecraft shows some potential. 6 kW Hall thruster in operation at the NASA Jet Propulsion Laboratory. Some electromagnetic methods: Ion thrusters (accelerate ions first and later neutralize the ion beam with an electron stream emitted from a cathode called a neutralizer) Electrostatic ion thruster Field-emission electric propulsion Hall effect thruster Colloid thruster Electrothermal thrusters (electromagnetic fields are used to generate a plasma to increase the heat of the bulk propellant, the thermal energy imparted to the propellant gas is then converted into kinetic energy by a nozzle of either physical material construction or by magnetic means) DC arcjet microwave arcjet Helicon Double Layer Thruster Electromagnetic thrusters (ions are accelerated either by the Lorentz Force or by the effect of electromagnetic fields where the electric field is not in the direction of the acceleration) Plasma propulsion engine Magnetoplasmadynamic thruster Electrodeless plasma thruster Pulsed inductive thruster Pulsed plasma thruster Variable specific impulse magnetoplasma rocket (VASIMR) Mass drivers (for propulsion) In electrothermal and electromagnetic thrusters, both ions and electrons are accelerated simultaneously, no neutralizer is required. Without internal reaction mass See also: Zero-propellant maneuver NASA study of a solar sail. The sail would be half a kilometer wide. The law of conservation of momentum is usually taken to imply that any engine which uses no reaction mass cannot accelerate the center of mass of a spaceship (changing orientation, on the other hand, is possible). But space is not empty, especially space inside the Solar System; there are gravitation fields, magnetic fields, electromagnetic waves, solar wind and solar radiation. Electromagnetic waves in particular are known to contain momentum, despite being massless; specifically the momentum flux density P of an EM wave is quantitatively 1/c^2 times the Poynting vector S, i.e. P = S/c^2, where c is the velocity of light. Field propulsion methods which do not rely on reaction mass thus must try to take advantage of this fact by coupling to a momentum-bearing field such as an EM wave that exists in the vicinity of the craft. However, because many of these phenomena are diffuse in nature, corresponding propulsion structures need to be proportionately large.[original research?] There are several different space drives that need little or no reaction mass to function. A tether propulsion system employs a long cable with a high tensile strength to change a spacecraft's orbit, such as by interaction with a planet's magnetic field or through momentum exchange with another object.[13] Solar sails rely on radiation pressure from electromagnetic energy, but they require a large collection surface to function effectively. The magnetic sail deflects charged particles from the solar wind with a magnetic field, thereby imparting momentum to the spacecraft. A variant is the mini-magnetospheric plasma propulsion system, which uses a small cloud of plasma held in a magnetic field to deflect the Sun's charged particles. An E-sail would use very thin and lightweight wires holding an electric charge to deflect these particles, and may have more controllable directionality. As a proof of concept, NanoSail-D became the first nanosatellite to orbit Earth.[14][full citation needed] There are plans to add them[clarification needed] to future Earth orbit satellites, enabling them to de-orbit and burn up once they are no longer needed. Cubesail will be the first mission to demonstrate solar sailing in low Earth orbit, and the first mission to demonstrate full three-axis attitude control of a solar sail.[15] Japan also launched its own solar sail powered spacecraft IKAROS in May 2010. IKAROS successfully demonstrated propulsion and guidance and is still flying today. A satellite or other space vehicle is subject to the law of conservation of angular momentum, which constrains a body from a net change in angular velocity. Thus, for a vehicle to change its relative orientation without expending reaction mass, another part of the vehicle may rotate in the opposite direction. Non-conservative external forces, primarily gravitational and atmospheric, can contribute up to several degrees per day to angular momentum,[16] so secondary systems are designed to "bleed off" undesired rotational energies built up over time. Accordingly, many spacecraft utilize reaction wheels or control moment gyroscopes to control orientation in space.[17] A gravitational slingshot can carry a space probe onward to other destinations without the expense of reaction mass. By harnessing the gravitational energy of other celestial objects, the spacecraft can pick up kinetic energy.[18] However, even more energy can be obtained from the gravity assist if rockets are used. Planetary and atmospheric propulsion A successful proof of concept Lightcraft test, a subset of beam-powered propulsion. Launch-assist mechanisms Main article: Space launch The conceptual ocean-located Quicklauncher, a light-gas gun–based space gun There have been many ideas proposed for launch-assist mechanisms that have the potential of drastically reducing the cost of getting into orbit. Proposed non-rocket spacelaunch launch-assist mechanisms include: Skyhook (requires reusable suborbital launch vehicle, not engineeringly feasible using presently available materials) Space elevator (tether from Earth's surface to geostationary orbit, cannot be built with existing materials) Launch loop (a very fast enclosed rotating loop about 80 km tall) Space fountain (a very tall building held up by a stream of masses fired from its base) Orbital ring (a ring around Earth with spokes hanging down off bearings) Electromagnetic catapult (railgun, coilgun) (an electric gun) Rocket sled launch Space gun (Project HARP, ram accelerator) (a chemically powered gun) Beam-powered propulsion rockets and jets powered from the ground via a beam High-altitude platforms to assist initial stage Airbreathing engines Main article: Jet engine Studies generally show that conventional air-breathing engines, such as ramjets or turbojets are basically too heavy (have too low a thrust/weight ratio) to give any significant performance improvement when installed on a launch vehicle itself. However, launch vehicles can be air launched from separate lift vehicles (e.g. B-29, Pegasus Rocket and White Knight) which do use such propulsion systems. Jet engines mounted on a launch rail could also be so used. On the other hand, very lightweight or very high speed engines have been proposed that take advantage of the air during ascent: SABRE - a lightweight hydrogen fuelled turbojet with precooler[19] ATREX - a lightweight hydrogen fuelled turbojet with precooler[20] Liquid air cycle engine - a hydrogen fuelled jet engine that liquifies the air before burning it in a rocket engine Scramjet - jet engines that use supersonic combustion Shcramjet - similar to a scramjet engine, however it takes advantage of shockwaves produced from the aircraft in the combustion chamber to assist in increasing overall efficiency. Normal rocket launch vehicles fly almost vertically before rolling over at an altitude of some tens of kilometers before burning sideways for orbit; this initial vertical climb wastes propellant but is optimal as it greatly reduces airdrag. Airbreathing engines burn propellant much more efficiently and this would permit a far flatter launch trajectory, the vehicles would typically fly approximately tangentially to Earth's surface until leaving the atmosphere then perform a rocket burn to bridge the final delta-v to orbital velocity. Planetary arrival and landing A test version of the MARS Pathfinder airbag system When a vehicle is to enter orbit around its destination planet, or when it is to land, it must adjust its velocity. This can be done using all the methods listed above (provided they can generate a high enough thrust), but there are a few methods that can take advantage of planetary atmospheres and/or surfaces. Aerobraking allows a spacecraft to reduce the high point of an elliptical orbit by repeated brushes with the atmosphere at the low point of the orbit. This can save a considerable amount of fuel because it takes much less delta-V to enter an elliptical orbit compared to a low circular orbit. Because the braking is done over the course of many orbits, heating is comparatively minor, and a heat shield is not required. This has been done on several Mars missions such as Mars Global Surveyor, Mars Odyssey and Mars Reconnaissance Orbiter, and at least one Venus mission, Magellan. Aerocapture is a much more aggressive manoeuver, converting an incoming hyperbolic orbit to an elliptical orbit in one pass. This requires a heat shield and much trickier navigation, because it must be completed in one pass through the atmosphere, and unlike aerobraking no preview of the atmosphere is possible. If the intent is to remain in orbit, then at least one more propulsive maneuver is required after aerocapture—otherwise the low point of the resulting orbit will remain in the atmosphere, resulting in eventual re-entry. Aerocapture has not yet been tried on a planetary mission, but the re-entry skip by Zond 6 and Zond 7 upon lunar return were aerocapture maneuvers, because they turned a hyperbolic orbit into an elliptical orbit. On these missions, because there was no attempt to raise the perigee after the aerocapture, the resulting orbit still intersected the atmosphere, and re-entry occurred at the next perigee. A ballute is an inflatable drag device. Parachutes can land a probe on a planet or moon with an atmosphere, usually after the atmosphere has scrubbed off most of the velocity, using a heat shield. Airbags can soften the final landing. Lithobraking, or stopping by impacting the surface, is usually done by accident. However, it may be done deliberately with the probe expected to survive (see, for example, Deep Impact (spacecraft)), in which case very sturdy probes are required. Table of methods Below is a summary of some of the more popular, proven technologies, followed by increasingly speculative methods. Four numbers are shown. The first is the effective exhaust velocity: the equivalent speed that the propellant leaves the vehicle. This is not necessarily the most important characteristic of the propulsion method; thrust and power consumption and other factors can be. However: if the delta-v is much more than the exhaust velocity, then exorbitant amounts of fuel are necessary (see the section on calculations, above) if it is much more than the delta-v, then, proportionally more energy is needed; if the power is limited, as with solar energy, this means that the journey takes a proportionally longer time The second and third are the typical amounts of thrust and the typical burn times of the method. Outside a gravitational potential small amounts of thrust applied over a long period will give the same effect as large amounts of thrust over a short period. (This result does not apply when the object is significantly influenced by gravity.) The fourth is the maximum delta-v this technique can give (without staging). For rocket-like propulsion systems this is a function of mass fraction and exhaust velocity. Mass fraction for rocket-like systems is usually limited by propulsion system weight and tankage weight. For a system to achieve this limit, typically the payload may need to be a negligible percentage of the vehicle, and so the practical limit on some systems can be much lower. Testing Spacecraft propulsion systems are often first statically tested on Earth's surface, within the atmosphere but many systems require a vacuum chamber to test fully. Rockets are usually tested at a rocket engine test facility well away from habitation and other buildings for safety reasons. Ion drives are far less dangerous and require much less stringent safety, usually only a large-ish vacuum chamber is needed. Famous static test locations can be found at Rocket Ground Test Facilities Some systems cannot be adequately tested on the ground and test launches may be employed at a Rocket Launch Site. Speculative methods Artist's conception of a warp drive design A variety of hypothetical propulsion techniques have been considered that mostly require a deeper understanding of the properties of space, particularly inertial frames and the quantum vacuum. To date, such methods are highly speculative and include: Black hole starship Differential sail Faster-than-light Field propulsion RF resonant cavity thruster Gravitational shielding Diametric drive Disjunction drive Pitch drive & bias drive Photon rocket Photonic laser thruster Quantum vacuum thruster Reactionless drive Abraham—Minkowski drive Alcubierre drive Heim drive Woodward effect A NASA assessment of its Breakthrough Propulsion Physics Program divides such proposals into those that are non-viable for propulsion purposes, those that are of uncertain potential, and those that are theoretically not impossible.[37] See also Spaceflight portal In-space propulsion technologies Index of aerospace engineering articles Lists of rockets Ozone depletion by rocket launches Stochastic electrodynamics Anti-gravity Artificial gravity Notes 1.      ^ With things moving around in orbits and nothing staying still, the question may be quite reasonably asked, stationary relative to what? The answer is for the energy to be zero (and in the absence of gravity which complicates the issue somewhat), the exhaust must stop relative to the initial motion of the rocket before the engines were switched on. It is possible to do calculations from other reference frames, but consideration for the kinetic energy of the exhaust and propellant needs to be given. In Newtonian mechanics the initial position of the rocket is the centre of mass frame for the rocket/propellant/exhaust, and has the minimum energy of any frame. References 1.          Hess, M.; Martin, K. K.; Rachul, L. J. (February 7, 2002). "Thrusters Precisely Guide EO-1 Satellite in Space First". NASA. Archived from the original on 2007-12-06. Retrieved 2007-07-30.     Phillips, Tony (May 30, 2000). "Solar S'Mores". NASA. Archived from the original on June 19, 2000. Retrieved 2007-07-30.     Olsen, Carrie (September 21, 1995). "Hohmann Transfer & Plane Changes". NASA. Archived from the original on 2007-07-15. Retrieved 2007-07-30.     Staff (April 24, 2007). "Interplanetary Cruise". 2001 Mars Odyssey. NASA. Archived from the original on August 2, 2007. Retrieved 2007-07-30.     Doody, Dave (February 7, 2002). "Chapter 4. Interplanetary Trajectories". Basics of Space Flight. NASA JPL. Retrieved 2007-07-30.     Hoffman, S. (August 20–22, 1984). "A comparison of aerobraking and aerocapture vehicles for interplanetary missions". AIAA and AAS, Astrodynamics Conference. Seattle, Washington: American Institute of Aeronautics and Astronautics. pp. 25 p. Archived from the original on September 27, 2007. Retrieved 2007-07-31.     Anonymous (2007). "Basic Facts on Cosmos 1 and Solar Sailing". The Planetary Society. Archived from the original on July 3, 2007. Retrieved 2007-07-26.     Rahls, Chuck (December 7, 2005). "Interstellar Spaceflight: Is It Possible?". Physorg.com. Retrieved 2007-07-31.     Zobel, Edward A. (2006). "Summary of Introductory Momentum Equations". Zona Land. Archived from the original on September 27, 2007. Retrieved 2007-08-02.     Benson, Tom. "Guided Tours: Beginner's Guide to Rockets". NASA. Retrieved 2007-08-02.     equation 19-1 Rocket propulsion elements 7th edition- Sutton     Choueiri, Edgar Y. (2004). "A Critical History of Electric Propulsion: The First 50 Years (1906–1956)". Journal of Propulsion and Power. 20 (2): 193–203. doi:10.2514/1.9245.     Drachlis, Dave (October 24, 2002). "NASA calls on industry, academia for in-space propulsion innovations". NASA. Archived from the original on December 6, 2007. Retrieved 2007-07-26.     NASA's Nanosail-D Becomes the First Solar Sail Spacecraft to Orbit the Earth | Inhabitat - Green Design Will Save the World     "Space Vehicle Control". University of Surrey. Retrieved 8 August 2015.     King-Hele, Desmond (1987). Satellite orbits in an atmosphere: Theory and application. Springer. ISBN 978-0-216-92252-5.     Tsiotras, P.; Shen, H.; Hall, C. D. (2001). "Satellite attitude control and power tracking with energy/momentum wheels" (PDF). Journal of Guidance, Control, and Dynamics. 43 (1): 23–34. Bibcode:2001JGCD...24...23T. doi:10.2514/2.4705. ISSN 0731-5090.     Dykla, J. J.; Cacioppo, R.; Gangopadhyaya, A. (2004). "Gravitational slingshot". American Journal of Physics. 72 (5): 619–000. Bibcode:2004AmJPh..72..619D. doi:10.1119/1.1621032.     Anonymous (2006). "The Sabre Engine". Reaction Engines Ltd. Archived from the original on 2007-02-22. Retrieved 2007-07-26.     Harada, K.; Tanatsugu, N.; Sato, T. (1997). "Development Study on ATREX Engine". Acta Astronautica. 41 (12): 851–862. Bibcode:1997AcAau..41..851T. doi:10.1016/S0094-5765(97)00176-8.     ESA Portal – ESA and ANU make space propulsion breakthrough     Hall effect thrusters have been used on Soviet/Russian satellites for decades.     A Xenon Resistojet Propulsion System for Microsatellites[dead link] (Surrey Space Centre, University of Surrey, Guildford, Surrey)     Alta - Space Propulsion, Systems and Services - Field Emission Electric Propulsion     RD-701     Google Translate     RD-0410     Young Engineers' Satellite 2 Archived 2003-02-10 at the Wayback Machine.     Gnom Archived 2010-01-02 at the Wayback Machine.     NASA GTX Archived November 22, 2008, at the Wayback Machine.     The PIT MkV pulsed inductive thruster     Pratt & Whitney Rocketdyne Wins $2.2 Million Contract Option for Solar Thermal Propulsion Rocket Engine (Press release, June 25, 2008, Pratt & Whitney Rocketdyne)[dead link]     "Operation Plumbbob". July 2003. Retrieved 2006-07-31.     Brownlee, Robert R. (June 2002). "Learning to Contain Underground Nuclear Explosions". Retrieved 2006-07-31.     PSFC/JA-05-26:Physics and Technology of the Feasibility of Plasma Sails, Journal of Geophysical Research, September 2005     MagBeam 37.   Millis, Marc (June 3–5, 2005). "Assessing Potential Propulsion Breakthroughs" (PDF). New Trends in Astrodynamics and Applications II. Princeton, NJ. External links NASA Breakthrough Propulsion Physics project Different Rockets Earth-to-Orbit Transportation Bibliography Spaceflight Propulsion - a detailed survey by Greg Goebel, in the public domain Johns Hopkins University, Chemical Propulsion Information Analysis Center Tool for Liquid Rocket Engine Thermodynamic Analysis Smithsonian National Air and Space Museum's How Things Fly website   Space Shuttle main engine From Wikipedia, the free encyclopedia   (Redirected from Space Shuttle Main Engine) "SSME" redirects here. For the services field, see Service science, management and engineering. RS-25 Space Shuttle Main Engine test firing (the bright area at the bottom of the picture is a Mach disk) Country of origin United States First flight April 12, 1981 (STS-1) Manufacturer Rocketdyne Associated L/V Space Shuttle Space Launch System Predecessor HG-3 Status Out of service since STS-135, in testing for SLS Liquid-fuel engine Propellant Liquid oxygen / Liquid hydrogen Cycle Staged combustion Configuration Nozzle ratio 69:1[1] Performance Thrust (vac.) 512,300 lbf (2,279 kN)[1] Thrust (SL) 418,000 lbf (1,860 kN)[1] Chamber pressure 2,994 psi (20.64 MPa)[1] Isp (vac.) 452.3 seconds (4.436 km/s)[1] Isp (SL) 366 seconds (3.59 km/s)[1] Dimensions Length 168 inches (4.3 m) Diameter 96 inches (2.4 m) References References [2][3] Notes Data is for RS-25D at 109% throttle. The Aerojet Rocketdyne RS-25, otherwise known as the Space Shuttle main engine (SSME),[4] is a liquid-fuel cryogenic rocket engine that was used on NASA's Space Shuttle and is planned to be used on its successor, the Space Launch System. Designed and manufactured in the United States by Rocketdyne, the RS-25 burns cryogenic liquid hydrogen and liquid oxygen propellants, with each engine producing 1,859 kN (418,000 lbf) of thrust at liftoff. Although the RS-25 can trace its heritage back to the 1960s, concerted development of the engine began in the 1970s, with the first flight, STS-1, occurring on April 12, 1981. The RS-25 has undergone several upgrades over its operational history to improve the engine's reliability, safety, and maintenance load. The engine produces a specific impulse (Isp) of 452 seconds (4.43 km/s) in a vacuum, or 366 seconds (3.59 km/s) at sea level, has a mass of approximately 3.5 tonnes (7,700 pounds), and is capable of throttling between 67% and 109% of its rated power level in one-percent increments. The RS-25 operates at temperatures ranging from −253 °C (−423 °F) to 3300 °C (6000 °F).[1] The Space Shuttle used a cluster of three RS-25 engines mounted in the stern structure of the orbiter, with fuel being drawn from the external tank. The engines were used for propulsion during the entirety of the spacecraft's ascent, with additional thrust being provided by two solid rocket boosters and the orbiter's two AJ10-190 orbital maneuvering system engines. Following each flight, the RS-25 engines were removed from the orbiter, inspected, and refurbished before being reused on another mission. Contents 1 Components 1.1 Turbopumps 1.1.1 Oxidizer system 1.1.2 Fuel system 1.2 Powerhead 1.2.1 Preburners 1.2.2 Main combustion chamber 1.3 Nozzle 1.4 Controller 1.4.1 Main valves 1.5 Gimbal 1.6 Helium system 2 History 2.1 Development 2.2 Space Shuttle program 2.2.1 Upgrades 2.2.2 Engine throttle/output 2.2.3 Incidents 2.3 After Shuttle 2.3.1 Constellation 2.3.2 Space Launch System 3 See also 4 Notes 5 References 6 External links Components RS-25 schematic Fuel flow Oxidizer flow RS-25 propellant flow The RS-25 engine consists of various pumps, valves, and other components which work in concert to produce thrust. Fuel (liquid hydrogen) and oxidizer (liquid oxygen) from the Space Shuttle's external tank entered the orbiter at the umbilical disconnect valves and from there flowed through the orbiter's main propulsion system (MPS) feed lines; whereas in the Space Launch System (SLS), fuel and oxidizer from the rocket's core stage will flow directly into the MPS lines. Once in the MPS lines, the fuel and oxidizer each branch out into separate paths to each engine (three on the Space Shuttle, four on the SLS). In each branch, prevalves then allow the propellants to enter the engine.[5][6] Once in the engine, the propellants flow through low-pressure fuel and oxidizer turbopumps (LPFTP and LPOTP), and from there into high-pressure turbopumps (HPFTP and HPOTP). From these HPTPs the propellants take different routes through the engine. The oxidizer is split into four separate paths: to the oxidizer heat exchanger, which then splits into the oxidizer tank pressurization and pogo suppression systems; to the low pressure oxidizer turbopump (LPOTP); to the high pressure oxidizer preburner, from which it is split into the HPFTP turbine and HPOTP before being reunited in the hot gas manifold and sent on to the main combustion chamber (MCC); or directly into the main combustion chamber (MCC) injectors. Meanwhile, fuel flows through the main fuel valve into regenerative cooling systems for the nozzle and MCC, or through the chamber coolant valve. Fuel passing through the MCC cooling system then passes back through the LPFTP turbine before being routed either to the fuel tank pressurization system or to the hot gas manifold cooling system (from where it passes into the MCC). Fuel in the nozzle cooling and chamber coolant valve systems is then sent via pre-burners into the HPFTP turbine and HPOTP before being reunited again in the hot gas manifold, from where it passes into the MCC injectors. Once in the injectors, the propellants are mixed and injected into the main combustion chamber where they are ignited. The burning propellant mixture is then ejected through the throat and bell of the engine's nozzle, the pressure of which creates the thrust.[5] Turbopumps Oxidizer system The low-pressure oxidizer turbopump (LPOTP) is an axial-flow pump which operates at approximately 5,150 rpm driven by a six-stage turbine powered by high-pressure liquid oxygen from the high-pressure oxidizer turbopump (HPOTP). It boosts the liquid oxygen's pressure from 0.7 to 2.9 MPa (100 to 420 psi), with the flow from the LPOTP then being supplied to the HPOTP. During engine operation, the pressure boost permits the high-pressure oxidizer turbine to operate at high speeds without cavitating. The LPOTP, which measures approximately 450 by 450 mm (18 by 18 in), is connected to the vehicle propellant ducting and supported in a fixed position by being mounted on the launch vehicle's structure.[5] The HPOTP consists of two single-stage centrifugal pumps (a main pump and a preburner pump) mounted on a common shaft and driven by a two-stage, hot-gas turbine. The main pump boosts the liquid oxygen's pressure from 2.9 to 30 MPa (420 to 4,350 psi) while operating at approximately 28,120 rpm, giving a power output of 23,260 hp (17.34 MW). The HPOTP discharge flow splits into several paths, one of which drives the LPOTP turbine. Another path is to, and through, the main oxidizer valve and enters the main combustion chamber. Another small flow path is tapped off and sent to the oxidizer heat exchanger. The liquid oxygen flows through an anti-flood valve that prevents it from entering the heat exchanger until sufficient heat is present for the heat exchanger to utilize the heat contained in the gases discharged from the HPOTP turbine, converting the liquid oxygen to gas. The gas is sent to a manifold and then routed to pressurize the liquid oxygen tank. Another path enters the HPOTP second-stage preburner pump to boost the liquid oxygen's pressure from 30 to 51 MPa (4,300 psia to 7,400 psia). It passes through the oxidizer preburner oxidizer valve into the oxidizer preburner, and through the fuel preburner oxidizer valve into the fuel preburner. The HPOTP measures approximately 600 by 900 mm (24 by 35 in). It is attached by flanges to the hot-gas manifold.[5] The HPOTP turbine and HPOTP pumps are mounted on a common shaft. Mixing of the fuel-rich hot gases in the turbine section and the liquid oxygen in the main pump can create a hazard and, to prevent this, the two sections are separated by a cavity that is continuously purged by the engine's helium supply during engine operation. Two seals minimize leakage into the cavity; one seal is located between the turbine section and the cavity, while the other is between the pump section and cavity. Loss of helium pressure in this cavity results in automatic engine shutdown.[5] Fuel system The low-pressure fuel turbopump (LPFTP) is an axial-flow pump driven by a two-stage turbine powered by gaseous hydrogen. It boosts the pressure of the liquid hydrogen from 30 to 276 psia (0.2 to 1.9 MPa) and supplies it to the high-pressure fuel turbopump (HPFTP). During engine operation, the pressure boost provided by the LPFTP permits the HPFTP to operate at high speeds without cavitating. The LPFTP operates at around 16,185 rpm, and is approximately 450 by 600 mm (18 by 24 in) in size. It is connected to the vehicle propellant ducting and is supported in a fixed position by being mounted to the launch vehicle's structure.[5] The HPFTP is a three-stage centrifugal pump driven by a two-stage hot-gas turbine. It boosts the pressure of the liquid hydrogen from 1.9 to 45 MPa (276 to 6,515 psia), and operates at approximately 35,360 rpm with a power of 71,140 hp. The discharge flow from the turbopump is routed to, and through, the main valve and is then split into three flow paths. One path is through the jacket of the main combustion chamber, where the hydrogen is used to cool the chamber walls. It is then routed from the main combustion chamber to the LPFTP, where it is used to drive the LPFTP turbine. A small portion of the flow from the LPFTP is then directed to a common manifold from all three engines to form a single path to the liquid hydrogen tank to maintain pressurization. The remaining hydrogen passes between the inner and outer walls of the hot-gas manifold to cool it and is then discharged into the main combustion chamber. A second hydrogen flow path from the main fuel valve is through the engine nozzle (to cool the nozzle). It then joins the third flow path from the chamber coolant valve. This combined flow is then directed to the fuel and oxidizer preburners. The HPFTP is approximately 550 by 1,100 mm (22 by 43 in) in size and is attached to the hot-gas manifold by flanges.[5] Powerhead The large silver pipe across the top carries fuel from the low-pressure fuel turbopump (not visible) to the high pressure fuel turbopump (HPFTP, silver drum at lower left). The top of the HPFTP is bolted to part of the hot gas manifold (black, with brown diagonal pipe) and above that is the fuel preburner (also black, with brown pipe entering at right).[7] Preburners The oxidizer and fuel preburners are welded to the hot-gas manifold. The fuel and oxidizer enter the preburners and are mixed so that efficient combustion can occur. The augmented spark igniter is a small combination chamber located in the center of the injector of each preburner. The two dual-redundant spark igniters, which are activated by the engine controller, are used during the engine start sequence to initiate combustion in each preburner. They are turned off after approximately three seconds because the combustion process is then self-sustaining. The preburners produce the fuel-rich hot gases that pass through the turbines to generate the power needed to operate the high-pressure turbopumps. The oxidizer preburner's outflow drives a turbine that is connected to the HPOTP and to the oxidizer preburner pump. The fuel preburner's outflow drives a turbine that is connected to the HPFTP.[5] The speed of the HPOTP and HPFTP turbines depends on the position of the corresponding oxidizer and fuel preburner oxidizer valves. These valves are positioned by the engine controller, which uses them to throttle the flow of liquid oxygen to the preburners and, thus, control engine thrust. The oxidizer and fuel preburner oxidizer valves increase or decrease the liquid oxygen flow, thus increasing or decreasing preburner chamber pressure, HPOTP and HPFTP turbine speed, and liquid oxygen and gaseous hydrogen flow into the main combustion chamber, which increases or decreases engine thrust. The oxidizer and fuel preburner valves operate together to throttle the engine and maintain a constant 6.03:1 propellant mixture ratio.[2] The main oxidizer and main fuel valves control the flow of liquid oxygen and liquid hydrogen into the engine and are controlled by each engine controller. When an engine is operating, the main valves are fully open.[5] Main combustion chamber Each engine main combustion chamber (MCC) receives fuel-rich hot gas from a hot-gas manifold cooling circuit. The gaseous hydrogen and liquid oxygen enter the chamber at the injector, which mixes the propellants. A small augmented-spark igniter-chamber is located in the center of the injector, and this dual-redundant igniter is used during the engine start sequence to initiate combustion. The igniters are turned off after approximately three seconds because the combustion process is self-sustaining. The main injector and dome assembly are welded to the hot-gas manifold, and the MCC is also bolted to the hot-gas manifold.[5] The MCC comprises a structural shell made of Inconel 718 which is lined with a copper-silver-zirconium alloy called NARloy-Z, developed specifically for the RS-25 in the 1970s. Around 390 channels are machined into the liner wall to carry liquid hydrogen through the liner to provide MCC cooling, as the temperature in the combustion chamber reaches 3300 °C (6000 °F) during flight – higher than the boiling point of iron.[8][9] Nozzle The nozzles of Space Shuttle Columbia's three RS-25s following the landing of STS-93 The engine's nozzle is 121 in (3.1 m) long with a diameter of 10.3 in (0.26 m) at its throat and 90.7 in (2.30 m) at its exit.[10] The nozzle is a bell-shaped extension bolted to the main combustion chamber, referred to as a de Laval nozzle. The RS-25 nozzle has an unusually large expansion ratio (about 77.5:1) for the chamber pressure.[11] At sea level, a nozzle of this ratio would normally undergo flow separation of the jet from the nozzle, which would cause control difficulties and could even mechanically damage the vehicle. However, to aid the engine's operation Rocketdyne engineers varied the angle of the nozzle walls from the theoretical optimum for thrust, reducing it near the exit. This raises the pressure just around the rim to an absolute pressure between 4.6 and 5.7 psi (32 and 39 kPa), and prevents flow separation. The inner part of the flow is at much lower pressure, around 2 psi (14 kPa) or less.[12] The inner surface of each nozzle is cooled by liquid hydrogen flowing through brazed stainless steel tube wall coolant passages. On the Space Shuttle, a support ring welded to the forward end of the nozzle is the engine attach point to the orbiter-supplied heat shield. Thermal protection is necessary because of the exposure portions of the nozzles experience during the launch, ascent, on-orbit and entry phases of a mission. The insulation consists of four layers of metallic batting covered with a metallic foil and screening.[5] Controller A Block II RS-25D main engine controller Each engine is equipped with a main engine controller (MEC), an integrated computer which controls all of the engine's functions (through the use of valves) and monitors its performance. Built by Honeywell Aerospace, each MEC originally comprised two redundant Honeywell HDC-601 computers,[13] later upgraded to a system composed of two doubly redundant Motorola 68000 (M68000) processors (for a total of four M68000s per controller).[14] Having the controller installed on the engine itself greatly simplifies the wiring between the engine and the launch vehicle, because all the sensors and actuators are connected directly to only the controller, each MEC then being connected to the orbiter's general purpose computers (GPCs) or the SLS's avionics suite via its own engine interface unit (EIU).[15] Using a dedicated system also simplifies the software and thus improves its reliability. Two independent dual-CPU computers, A and B, form the controller; giving redundancy to the system. The failure of controller system A automatically leads to a switch-over to controller system B without impeding operational capabilities; the subsequent failure of controller system B would provide a graceful shutdown of the engine. Within each system (A and B), the two M68000s operate in "lock-step", thereby enabling each system to detect failures by comparing the signal levels on the buses of the two M68000 processors within that system. If differences are encountered between the two buses, then an interrupt is generated and control turned over to the other system. Because of subtle differences between M68000s from Motorola and the second source manufacturer TRW, each system uses M68000s from the same manufacturer (for instance system A would have two Motorola CPUs while system B would have two CPUs manufactured by TRW). Memory for block I controllers was of the plated-wire type, which functions in a manner similar to magnetic core memory and retains data even after power is turned off.[16] Block II controllers used conventional CMOS static RAM.[14] The controllers were designed to be tough enough to survive the forces of launch, and proved to be extremely resilient to damage. During the investigation of the Challenger accident the two MECs (from engines 2020 and 2021), recovered from the seafloor, were delivered to Honeywell Aerospace for examination and analysis. One controller was broken open on one side, and both were severely corroded and damaged by marine life. Both units were disassembled and the memory units flushed with deionized water. After they were dried and vacuum baked, data from these units was retrieved for forensic examination.[17] Main valves To control the engine's output, the MEC operates five hydraulically actuated propellant valves on each engine; the oxidizer preburner oxidizer, fuel preburner oxidizer, main oxidizer, main fuel, and chamber coolant valves. In an emergency, the valves can be fully closed by using the engine's helium supply system as a backup actuation system.[5] In the Space Shuttle the main oxidizer and fuel bleed valves were used after shutdown to dump any residual propellant, with residual liquid oxygen venting through the engine and residual liquid hydrogen venting through the liquid hydrogen fill and drain valves. After the dump was completed, the valves closed and remain closed for the remainder of the mission.[5] A coolant control valve is mounted on the combustion chamber coolant bypass duct of each engine. The engine controller regulates the amount of gaseous hydrogen allowed to bypass the nozzle coolant loop, thus controlling its temperature. The chamber coolant valve is 100% open before engine start. During engine operation, it is 100% open for throttle settings of 100 to 109% for maximum cooling. For throttle settings between 65 and 100%, its position ranged from 66.4 to 100% open for reduced cooling.[5] Gimbal External video SSME gimbal test SSME gimbal test Each engine is installed with a gimbal bearing, a universal ball and socket joint which is bolted to the launch vehicle by its upper flange and to the engine by its lower flange. It represents the thrust interface between the engine and the launch vehicle, supporting 7,480 lb (3,390 kg) of engine weight and withstanding over 500,000 lb (230,000 kg) of thrust. As well as providing a means to attach the engine to the launch vehicle, the gimbal bearing allows the engine to be pivoted (or "gimballed") around two axes of freedom with a range of ±10.5°.[18] This motion allows the engine's thrust vector to be altered, thus steering the vehicle into the correct orientation. The bearing assembly is approximately 290 by 360 mm (11 by 14 in), has a mass of 105 lb (48 kg), and is made of titanium alloy.[7] The low-pressure oxygen and low-pressure fuel turbopumps were mounted 180° apart on the orbiter's aft fuselage thrust structure. The lines from the low-pressure turbopumps to the high-pressure turbopumps contain flexible bellows that enable the low-pressure turbopumps to remain stationary while the rest of the engine is gimbaled for thrust vector control, and also to prevent damage to the pumps when loads were applied to them. The liquid-hydrogen line from the LPFTP to the HPFTP is insulated to prevent the formation of liquid air.[5] Helium system In addition to fuel and oxidizer systems, the launch vehicle's main propulsion system is also equipped with a helium system consisting of ten storage tanks in addition to various regulators, check valves, distribution lines, and control valves. The system is used in-flight to purge the engine and provides pressure for actuating engine valves within the propellant management system and during emergency shutdowns. During entry, on the Space Shuttle, any remaining helium was used to purge the engines during reentry and for repressurization.[5] History Development Play media RS-25 testing at Stennis Space Center The history of the RS-25 traces back to the 1960s when NASA's Marshall Space Flight Center and Rocketdyne were conducting a series of studies on high-pressure engines, developed from the successful J-2 engine used on the S-II and S-IVB upper stages of the Saturn V rocket during the Apollo program. The studies were conducted under a program to upgrade the Saturn V engines, which produced a design for a 350,000 lbf upper-stage engine known as the HG-3.[19] As funding levels for Apollo wound down the HG-3 was cancelled as well as the upgraded F-1 engines already being tested.[20] It was the design for the HG-3 that would form the basis for the RS-25.[21] Meanwhile, in 1967, the US Air Force funded a study into advanced rocket propulsion systems for use during Project Isinglass, with Rocketdyne asked to investigate aerospike engines and Pratt & Whitney (P&W) to research more efficient conventional de Laval nozzle-type engines. At the conclusion of the study, P&W put forward a proposal for a 250,000 lbf engine called the XLR-129, which used a two-position expanding nozzle to provide increased efficiency over a wide range of altitudes.[22][23] In January 1969 NASA awarded contracts to General Dynamics, Lockheed, McDonnell Douglas, and North American Rockwell to initiate early development of the Space Shuttle.[24] As part of these 'Phase A' studies, the involved companies selected an upgraded version of the XLR-129, developing 415,000 lbf, as the baseline engine for their designs.[22] This design can be found on many of the planned Shuttle versions right up to the final decision. However since NASA was interested in pushing the state of the art in every way they decided to select a much more advanced design in order to "force an advancement of rocket engine technology".[11][22] They called for a new design based on a high-pressure combustion chamber running around 3000 psi, which increases the performance of the engine. Development began in 1970, when NASA released a request for proposal for 'Phase B' main engine concept studies, requiring development of a throttleable, staged combustion, de Laval-type engine.[11][22] The request was based on the then-current design of the Space Shuttle which featured two reusable stages, the orbiter and a manned fly-back booster, and required one engine which would be able to power both vehicles via two different nozzles (12 booster engines with 550,000 lbf sea level thrust each and 3 orbiter engines with 632,000 lbf vacuum thrust each).[11] Rocketdyne, P&W and Aerojet General were selected to receive funding although, given P&W's already-advanced development (demonstrating a working 350,000 lbf concept engine during the year) and Aerojet General's prior experience in developing the 1,500,000 lbf M-1 engine, Rocketdyne was forced to put a large amount of private money into the design process to allow the company to catch up to its competitors.[22] By the time the contract was awarded, budgetary pressures meant that the shuttle's design had changed to its final orbiter, external tank and two boosters configuration, and so the engine was only required to power the orbiter during ascent.[11] During the year-long 'Phase B' study period, Rocketdyne was able to make use of their experience developing the HG-3 engine to design their SSME proposal, producing a prototype by January 1971. The engine made use of a new Rocketdyne-developed copper-zirconium alloy (called NARloy-Z), and was tested on February 12, 1971, producing a chamber pressure of 3172 psi. The three participating companies submitted their engine development bids in April 1971, with Rocketdyne being awarded the contract on July 13, 1971—although work did not begin on engine development until March 31, 1972, due to a legal challenge from P&W.[11][22] Following the awarding of the contract, a preliminary design review was carried out in September 1972, followed by a critical design review in September 1976 after which the engine's design was set and construction of the first set of flight-capable engines began. Final review of all the Space Shuttle's components, including the engines, was conducted in 1979. The design reviews operated in parallel with several test milestones, initial tests consisting of individual engine components which identified shortcomings with various areas of the design, including the HPFTP, HPOTP, valves, nozzle and fuel preburners. The individual engine component tests were followed by the first test of a complete engine (0002) on March 16, 1977. NASA specified that, prior to the Shuttle's first flight, the engines must have undergone at least 65,000 seconds of testing, a milestone that was reached on March 23, 1980, with the engine having undergone 110,253 seconds of testing by the time of STS-1 both on test stands at Stennis Space Center and installed on the Main Propulsion Test Article (MPTA). The first set of engines (2005, 2006 and 2007) were delivered to Kennedy Space Center in 1979 and installed on Columbia, before being removed in 1980 for further testing and reinstalled on the orbiter. The engines, which were of the first manned orbital flight (FMOF) configuration and certified for operation at 100% rated power level (RPL), were operated in a twenty-second flight readiness firing on February 20, 1981, and, after inspection, declared ready for flight.[11] Space Shuttle program See also: List of space shuttle missions Space Shuttle Atlantis's three RS-25D main engines at liftoff during STS-110 Play media SSME startup and shutdown sequences Each Space Shuttle had three RS-25 engines, installed in the aft structure of the Space Shuttle orbiter in the Orbiter Processing Facility prior to the orbiter being transferred to the Vehicle Assembly Building. If necessary the engines could be changed on the pad. The engines, drawing propellant from the Space Shuttle external tank (ET) via the orbiter's main propulsion system (MPS), were ignited at T−6.6 seconds prior to liftoff (with each ignition staggered by 120 ms[25]), which allowed their performance to be checked prior to ignition of the Space Shuttle Solid Rocket Boosters (SRBs), which committed the shuttle to the launch.[26] At launch, the engines would be operating at 100% RPL, throttling up to 104.5% immediately following liftoff. The engines would maintain this power level until around T+40 seconds, where they would be throttled back to around 70% to reduce aerodynamic loads on the shuttle stack as it passed through the region of maximum dynamic pressure, or max Q.[note 1][22][25] The engines would then be throttled back up until around T+8 minutes, at which point they would be gradually throttled back down to 67% to prevent the stack exceeding 3 g of acceleration as it became progressively lighter due to propellant consumption. The engines were then shut down, a procedure known as main engine cutoff (MECO), at around T+8.5 minutes.[22] After each flight the engines would be removed from the orbiter and transferred to the Space Shuttle Main Engine Processing Facility (SSMEPF), where they would be inspected and refurbished in preparation for reuse on a subsequent flight.[27] A total of 46 reusable RS-25 engines, each costing around US$40 million, were flown during the Space Shuttle program, with each new or overhauled engine entering the flight inventory requiring flight qualification on one of the test stands at Stennis Space Center prior to flight.[25][28][29] Upgrades Flight history of the Space Shuttle Main Engines Over the course of the Space Shuttle program, the RS-25 went through a series of upgrades, including combustion chamber changes, improved welds and turbopump changes in an effort to improve the engine's performance and reliability and so reduce the amount of maintenance required after use. As a result, several versions of the RS-25 were used during the program:[9][22][24][25][30][31][32][33][34] FMOF (first manned orbital flight) – Certified for 100% rated power level (RPL). Used for the orbital flight test missions STS-1—STS-5 (engines 2005, 2006 and 2007). Phase I – Used for missions STS-6—STS-51-L, the Phase I engine offered increased service life and was certified for 104% RPL. Phase II (RS-25A) – First flown on STS-26, the Phase II engine offered a number of safety upgrades and was certified for 104% RPL & 109% full power level (FPL) in the event of a contingency. Block I (RS-25B) – First flown on STS-70, the Block I engines offered improved turbopumps featuring ceramic bearings, half as many rotating parts and a new casting process reducing the number of welds. Block I improvements also included a new, two-duct powerhead (rather than the original design, which featured three ducts connected to the HPFTP and two to the HPOTP), which helped improve hot gas flow, and an improved engine heat exchanger. Block IA (RS-25B) – First flown on STS-73, the Block IA engine offered main injector improvements. Block IIA (RS-25C) – First flown on STS-89, the Block IIA engine was an interim model used whilst certain components of the Block II engine completed development. Changes included a new large throat main combustion chamber (which had originally been recommended by Rocketdyne in 1980), improved low pressure turbopumps and certification for 104.5% RPL to compensate for a 2 seconds (0.020 km/s) reduction in specific impulse (original plans called for the engine to be certified to 106% for heavy International Space Station payloads, but this was not required and would have reduced engine service life). A slightly modified version first flew on STS-96. Block II (RS-25D) – First flown on STS-104, the Block II upgrade included all of the Block IIA improvements plus a new high pressure fuel turbopump. This model was ground-tested to 111% FPL in the event of a contingency abort, and certified for 109% FPL for use during an intact abort. Engine throttle/output The most obvious effects of the upgrades the RS-25 received through the Space Shuttle program were the improvements in engine throttle. Whilst the FMOF engine had a maximum output of 100% RPL, Block II engines could throttle as high as 109% or 111% in an emergency, with usual flight performance being 104.5%. These increases in throttle level made a significant difference to the thrust produced by the engine:[7][25] Sea level Vacuum 100% thrust 1,670 kN (380,000 lbf) 2,090 kN (470,000 lbf) 104.5% thrust 1,750 kN (390,000 lbf) 2,170 kN (490,000 lbf) 109% thrust 1,860 kN (420,000 lbf) 2,280 kN (510,000 lbf) Specifying power levels over 100% may seem nonsensical, but there was a logic behind it. The 100% level does not mean the maximum physical power level attainable, rather it was a specification decided on during engine development—the expected rated power level. When later studies indicated the engine could operate safely at levels above 100%, these higher levels became standard. Maintaining the original relationship of power level to physical thrust helps reduce confusion, as it created an unvarying fixed relationship so that test data (or operational data from past or future missions) can be easily compared. If the power level was increased, and that new value was said to be 100%, then all previous data and documentation would either require changing, or cross-checking against what physical thrust corresponded to 100% power level on that date.[11] Engine power level affects engine reliability, with studies indicating the probability of an engine failure increasing rapidly with power levels over 104.5%, which was why power levels above 104.5% were retained for contingency use only.[30] Incidents This shuttle control panel is set to select the abort to orbit (ATO) option, as used in the STS-51-F mission. After orbit was achieved, the mission continued normally and the orbiter returned to Earth with the crew. Recovered power-head of one of Columbia's main engines. Columbia was lost on re-entry, from a suspected heat shield failure. During the course of the Space Shuttle program, a total of 46 RS-25 engines were used (with one extra RS-25D being built but never used). During the 135 missions, for a total of 405 individual engine-missions,[28] Pratt & Whitney Rocketdyne reports a 99.95% reliability rate, with the only in-flight SSME failure occurring during Space Shuttle Challenger's STS-51-F mission.[2] The engines, however, did suffer from a number of pad failures (redundant set launch sequencer aborts, or RSLSs) and other issues during the course of the program: STS-41-D Discovery – No. 3 engine caused an RSLS shutdown at T−4 seconds due to loss of redundant control on main engine valve, stack rolled back and engine replaced.[35] STS-51-F Challenger – No. 2 engine caused an RSLS shutdown at T−3 seconds due to a coolant valve malfunction.[36][37] STS-51-F Challenger – No. 1 engine (2023) shutdown at T+5:43 due to faulty temperature sensors, leading to an abort to orbit (although the mission objectives and length were not compromised by the ATO).[25][37] STS-55 Columbia – No. 3 engine caused an RSLS shutdown at T−3 seconds due to a leak in its liquid-oxygen preburner check valve.[38] STS-51 Discovery – No. 2 engine caused an RSLS shut down at T−3 seconds due to a faulty hydrogen fuel sensor.[39] STS-68 Endeavour – No. 3 engine (2032) caused an RSLS shutdown at T−1.9 seconds when a temperature sensor in its HPOTP exceeded its redline.[40] STS-93 Columbia – At T+5 seconds, an electrical short disabled one primary and one secondary controller on two of the three engines. In addition, a 0.1-inch diameter, 1-inch long gold-plated pin, used to plug an oxidizer post orifice, came loose inside an engine's main injector and impacted the engine nozzle inner surface, rupturing three hydrogen cooling lines. The resulting three breaches caused a leak resulting in a premature engine shutdown due to increased propellant consumption.[41] After Shuttle The 6 RS-25Ds used during STS-134 and STS-135 in storage at Kennedy Space Center Constellation During the period preceding final Space Shuttle retirement, various plans for the remaining engines were proposed, ranging from them all being kept by NASA, to them all being given away (or sold for US$400,000–800,000 each) to various institutions such as museums and universities.[42] This policy followed changes to the planned configurations of the Constellation program's Ares V cargo-launch vehicle and Ares I crew-launch vehicle rockets, which had been planned to use the RS-25 in their first and second stages respectively.[43] Whilst these configurations had initially seemed worthwhile, as they would use then-current technology following the shuttle's retirement in 2010, the plan had several drawbacks:[43] The engines would not be reusable, as they would be permanently attached to the discarded stages. Each engine would have to undergo a test firing prior to installation and launch, with refurbishment required following the test. It would be expensive, time-consuming, and weight-intensive to convert the ground-started RS-25D to an air-started version for the Ares I second stage. Following several design changes to the Ares I and Ares V rockets, the RS-25 was to be replaced with a single J-2X engine for the Ares I second stage and six modified RS-68 engines (which was based on both the SSME and Apollo-era J-2 engine) on the Ares V core stage; this meant that the RS-25 would be retired along with the space shuttle fleet.[43] In 2010, however, NASA was directed to halt the Constellation program, and with it development of the Ares I and Ares V, instead focusing on building a new heavy lift launcher.[44] Space Launch System See also: Space Launch System NASA's SLS reference configuration from February 2011 Following the retirement of the Space Shuttle, NASA announced on September 14, 2011, that it would be developing a new launch vehicle, known as the Space Launch System (SLS), to replace the shuttle fleet.[45] The design for the SLS features the RS-25 on its core stage, with different versions of the rocket being installed with between three and five engines.[46][47] The initial flights of the new launch vehicle will make use of flown Block II RS-25D engines, with NASA keeping the remaining such engines in a "purged safe" environment at Stennis Space Center, "along with all of the ground systems required to maintain them."[48][49] In addition to the RS-25Ds, the SLS program will make use of the Main Propulsion Systems from the three remaining orbiters for testing purposes (having been removed as part of the orbiters' decommissioning), with the first two launches (SLS-1 and SLS-2) possibly making use of the MPS hardware from Space Shuttles Atlantis and Endeavour in their core stages.[47][49][50] The SLS's propellants will be supplied to the engines from the rocket's core stage, which will consist of a modified Space Shuttle external tank with the MPS plumbing and engines at its aft, and an interstage structure at the top.[6] Once the remaining RS-25Ds are used up, they are to be replaced with a cheaper, expendable version, currently designated the RS-25E.[6] This engine may be based on one or both of two single-use variants which were studied in 2005, the RS-25E (referred to as the 'Minimal Change Expendable SSME') and the even more simplified RS-25F (referred to as the 'Low Cost Manufacture Expendable SSME'), both of which were under consideration in 2011.[32][51] 2015 tests A test series was planned to determine RS-25 engine performance with: the new engine controller unit; lower liquid oxygen temperatures; greater inlet pressure due to the taller SLS core stage liquid oxygen tank and higher vehicle acceleration; and, more nozzle heating due to the four-engine configuration and its position in-plane with the SLS booster exhaust nozzles. New ablative insulation and heaters were to be tested during the series.[52][better source needed] 9 January 28 May 11 June - 500 seconds 17 July - 535 seconds 13 August 27 August Following these series of tests, four more engines are planned to enter a new test cycle.[53][better source needed] 2017 tests A new series of tests designed to evaluate performance in SLS use cases was initiated in 2017.[54][better source needed] See also Shuttle-C Notes External video STS-49 Flight Readiness Firing Time-lapse video of STS-135 SSME installation RS-25 Engine Test for SLS on 28 May 2015 RS-25 Engine controller system test on 27 July 2017 1.        1.      The level of throttle was initially set to 65%, but, following review of early flight performance, this was increased to a minimum of 67% to reduce fatigue on the MPS. The throttle level was dynamically calculated based on initial launch performance, generally being reduced to a level around 70%. References  This article incorporates public domain material from websites or documents of the National Aeronautics and Space Administration. 1.          Aerojet Rocketdyne, RS-25 Engine (accessed July 22, 2014)     "Space Shuttle Main Engine" (PDF). Pratt & Whitney Rocketdyne. 2005. Archived from the original (PDF) on February 8, 2012. Retrieved November 23, 2011.     Wade, Mark. "SSME". Encyclopedia Astronautica. Retrieved October 27, 2011.     "RS-25 Engine".     "Main Propulsion System (MPS)" (PDF). Shuttle Press Kit.com. Boeing, NASA & United Space Alliance. October 6, 1998. Archived from the original (PDF) on 2012-02-04. Retrieved December 7, 2011.     Chris Bergin (September 14, 2011). "SLS finally announced by NASA – Forward path taking shape". NASASpaceflight.com. Retrieved December 14, 2011.     "Space Shuttle Main Engine Orientation" (PDF). Boeing/Rocketdyne. June 1998. Retrieved December 12, 2011.     "NASA Relies on Copper for Shuttle Engine". Discover Copper Online. Copper Development Association. 1992. Retrieved January 19, 2012.     Steve Roy (August 2000). "Space Shuttle Main Engine Enhancements". NASA. Retrieved December 7, 2011.     R.A. O'Leary and J. E. Beck (1992). "Nozzle Design". Threshold. Pratt & Whitney Rocketdyne. Archived from the original on March 16, 2008.     Robert E. Biggs (May 1992). "Space Shuttle Main Engine: The First Ten Years". In Stephen E. Doyle. History of Liquid Rocket Engine Development in the United States 1955–1980. AAS History Series. American Astronautical Society. pp. 69–122. ISBN 978-0-87703-350-9. Retrieved December 12, 2011.     "Nozzle Design". March 16, 2009. Retrieved November 23, 2011.     "Computers in the Space Shuttle Avionics System". Computers in Spaceflight: The NASA Experience. NASA. July 15, 2005. Retrieved November 23, 2011.     "The future of the shuttle's computers". NASA. July 15, 2005. Retrieved November 23, 2011.     "Space Shuttle Main Engine Controllers". NASA. April 4, 2004. Retrieved December 8, 2011.     RM Mattox & JB White (November 1981). "Space Shuttle Main Engine Controller" (PDF). NASA. Retrieved December 15, 2011.     "The Cause of the Accident". Report of the Presidential Commission on the Space Shuttle Challenger Accident. NASA. June 6, 1986. Retrieved December 8, 2011.     Jim Dumoulin (August 31, 2000). "Main Propulsion System". NASA. Retrieved January 16, 2012.     Mark Wade. "HG-3". Encyclopedia Astronautica. Retrieved December 13, 2011.     title=(NASA-CR-138312) F-LA TASK ASSIGNMENT PROGRAM final Report, (Rocketdyne)     "MSFC Propulsion Center of Excellence is Built on Solid Foundation". NASA. 1995. Retrieved December 13, 2011.     David Baker (April 2011). NASA Space Shuttle. Owners' Workshop Manuals. Haynes Publishing. ISBN 978-1-84425-866-6.     Dwayne Day (April 12, 2010). "A bat outta Hell: the ISINGLASS Mach 22 follow-on to OXCART". The Space Review. Retrieved January 8, 2012.     Fred H. Jue. "Space Shuttle Main Engine: 30 Years of Innovation" (PDF). Boeing. Retrieved November 27, 2011.     Wayne Hale & various (January 17, 2012). "An SSME-related request". NASASpaceflight.com. Retrieved January 17, 2012.     "Countdown 101". NASA. September 17, 2009. Retrieved January 8, 2012.     John Shannon (June 17, 2009). "Shuttle-Derived Heavy Lift Launch Vehicle" (PDF).     "SSME Flight Experience" (JPEG). Pratt & Whitney Rocketdyne. November 2010.     Chris Bergin (December 3, 2007). "Constellation transition – phased retirement plan for the SSME set". NASASpaceflight.com. Retrieved January 23, 2012.     "Report of the SSME Assessment Team" (PDF). NASA. January 1993. Retrieved November 27, 2011.     F. Jue and F. Kuck (July 2002). "Space Shuttle Main Engine (SSME) Options for the Future Shuttle". American Institute of Aeronautics and Astronautics. Archived from the original (DOC) on October 9, 2007. Retrieved November 27, 2011.     Ryan Crierie (November 13, 2011). "Reference Spacecraft Engines". Retrieved January 8, 2012.     "The Roar of Innovation". NASA. November 6, 2002. Archived from the original on November 8, 2002. Retrieved December 7, 2011.     "MSFC and Exploration: Our Path Forward" (PPT). NASA. September 2005.     Mike Mullane (February 3, 2007). Riding Rockets: The Outrageous Tales of a Space Shuttle Astronaut. Scribner. ISBN 0-7432-7682-5.     Jim Dumoulin (June 29, 2001). "51-F". NASA. Retrieved January 16, 2012.     Ben Evans (2007). Space Shuttle Challenger: Ten Journeys into the Unknown. Warwickshire, United Kingdom: Springer-Praxis. ISBN 978-0-387-46355-1.     Jim Dumoulin (June 29, 2001). "STS-55". NASA. Retrieved January 16, 2012.     Jim Dumoulin (June 29, 2001). "STS-51". NASA. Retrieved January 16, 2012.     Jim Dumoulin (June 29, 2001). "STS-68". NASA. Retrieved January 16, 2012.     Ben Evans (August 30, 2005). Space Shuttle Columbia: Her Missions and Crews. Springer Praxis. ISBN 978-0-387-21517-4.     Dunn, Marcia (January 15, 2010). "Recession Special: NASA Cuts Space Shuttle Price". ABC News. Archived from the original on January 18, 2010.     D Harris & C Bergin (December 26, 2008). "Return to SSME – Ares V undergoes evaluation into potential switch". NASASpaceflight.com. Retrieved December 15, 2011.     "Obama signs Nasa up to new future". BBC News. October 11, 2010.     "NASA Announces Design For New Deep Space Exploration System". NASA. Archived from the original on September 21, 2011. Retrieved December 14, 2011.     Chris Bergin (October 4, 2011). "SLS trades lean towards opening with four RS-25s on the core stage". NASASpaceflight.com. Retrieved December 14, 2011.     Chris Bergin (January 13, 2012). "SSME family prepare for SLS core stage role following Shuttle success". NASASpaceflight.com. Retrieved January 16, 2012.     Carreau, Mark (March 29, 2011). "NASA Will Retain Block II SSMEs". Aviation Week. Archived from the original on April 20, 2011. Retrieved March 30, 2011.     Chris Bergin (January 22, 2012). "Engineers begin removing orbiter MPS components for donation to SLS". NASASpaceflight.com. Retrieved January 23, 2012.     Chris Bergin (September 20, 2011). "PRCB managers recommend Atlantis and Endeavour become SLS donors". NASASpaceflight.com. Retrieved December 14, 2011.     P. McConnaughey; et al. (February 2011). "NASA Technology Area 1: Launch Propulsion Systems" (PDF). NASA. Retrieved January 23, 2012.     RS-25 Engine Fires Up for Third Test in Series, Kim Henry, Marshall Space Flight Center, in SpaceDaily.com, 17 June 2015, accessed 18 June 2015     "Pedal to the Metal – RS-25 Engine Revs Up Again". NASA.       Ionocraft From Wikipedia, the free encyclopedia This article has multiple issues. Please help improve it or discuss these issues on the talk page. (Learn how and when to remove these template messages) This article may be written from a fan's point of view, rather than a neutral point of view. (March 2014)   This article includes a list of references, but its sources remain unclear because it has insufficient inline citations. (November 2009) An ionocraft or ion-propelled aircraft (commonly known as a lifter or hexalifter) is a device that uses an electrical electrohydrodynamic (EHD) phenomenon to produce thrust in the air without requiring any combustion or moving parts. The term "ionocraft" dates back to the 1960s, an era in which EHD experiments were at their peak. In its basic form, it simply consists of two parallel conductive electrodes, one in the form of a fine wire and another which may be formed out of wire grid, tubes or foil skirts with a smooth round surface. When such an arrangement is powered by high voltage (in the range of a few kilovolts), it produces thrust. The ionocraft forms part of the EHD thruster family, but is a special case in which the ionisation and accelerating stages are combined into a single stage. The device is a popular science fair project for students.[citation needed] It is also popular among anti-gravity or so-called "electrogravitics" proponents, due to the research of Thomas Townsend Brown, who built these devices in the 1920s and incorrectly believed that he had found a way to modify gravity using electric fields. The term "lifter" is an accurate description because it is not an anti-gravity device; rather, it produces lift using the same basic principle as a rocket, i.e. from the equal but opposite force upward generated by the driving force downward, specifically by driving the ionized air downward in the case of the ionocraft. Much like a rocket or a jet engine (it can actually be much more thrust efficient than a jet engine[1]), the force that an ionocraft generates is consistently oriented along its own axis, regardless of the surrounding gravitational field. Claims of the device also working in a vacuum have been disproved.[2] Ionocraft require many safety precautions due to the high voltage required for their operation; nevertheless, a large subculture has grown up around this simple EHD thrusting device and its physics are now known to a much better extent. Contents 1 Description 2 Construction 2.1 Components 2.1.1 Corona wire 2.1.2 Air gap 2.1.3 Collector 3 Mechanism 4 See also 5 References 6 Sources 7 External links Description An ionocraft is a propulsion device based on ionic air propulsion that works without moving parts, uses only electrical energy, and is able to lift its own weight, not including its own power supply. The principle of ionic wind propulsion with corona-generated charged particles has been known from the earliest days of the discovery of electricity with references dating back to 1709 in a book titled Physico-Mechanical Experiments on Various Subjects by Francis Hauksbee. Its use for propulsion was given serious thought by Major Alexander Prokofieff de Seversky who contributed much to its basic physics and construction variations in 1960. In fact, it was Major de Seversky himself who in 1964 coined[non-primary source needed] the term Ionocraft in his (U.S. Patent 3,130,945). There are also designs by the American experimenter Thomas Townsend Brown, such as his 1960 patents for "Elektrokinetic Apparatus". Brown spent most of his life trying to develop what he thought was an anti-gravity effect, which he named the Biefeld–Brown effect. Since Brown's devices produce thrust along their axis regardless of the direction of gravity and do not work in a vacuum, the effect he identified has been attributed to electrohydrodynamics instead of anti-gravity.[3][4] Construction Typical ionocraft construction A simple ionocraft derivative, also known as a lifter, can be easily constructed by anyone with a minimal amount of technical knowledge. The model in its simplest form has the shape of an equilateral triangle with sides generally between 10 and 30 cm. They consist of three parts, the corona wire (or emitting wire), the air gap (or dielectric fluid), and the foil skirt (collector). The electrical polarities of the emitting and collecting electrodes can be reversed. All of this is usually supported by a lightweight balsawood or other electrically isolating frame so that the corona wire is supported at a fixed distance above the foil skirt, generally at 1 mm per kilo-volt. The corona wire and foil should be as close as possible to achieve a saturated corona current condition which results in the highest production of thrust. However the corona wire should not be too close to the foil skirt as it will tend to arc in a spectacular show of tiny lightning bolts which has a twofold effect: It degrades the thrust as it is shorting the device and there is current flow through the arc instead of the ions that do the lifting It can destroy the power supply or burn the balsa structure of the lifter. Components Corona wire The corona wire is usually, but not necessarily, connected to the positive terminal of the high voltage power supply. In general, it is made from a small gauge bare conductive wire. While copper wire can be used, it does not work quite as well as stainless steel. Similarly, thinner wire such as 50 gauge tends to work well compared to more common, larger sizes such as 30 gauge, as the stronger electric field around the smaller diameter wire results in better ionisation and a larger corona current. 44 gauge stainless steel wire is also known to work very well and has been used with great success by hobbyists. Due to anomalies in construction adding supports made from 3D printed staggered mesh (to prevent surface arc-over) from SLS in resin can work if atmospheric humidity is low enough. The corona wire is so called because of its tendency to emit a purple corona-like glow while in use. This is simply a side effect of ionization. Excessive corona is to be avoided, as too much means the electrodes are dangerously close and may arc at any moment, not to mention the associated health hazards due to excess inhalation of ozone and NOx produced by the corona. Air gap The air gap is simply that, a gap of free flowing air between the two electrodes that make up the structure of an ionocraft. The air gap is a vital necessity to the functioning of this device as it is the dielectric used during operation. Best results have been observed with an air gap of 1 mm to every kV.[citation needed] Collector The collector may take various shapes, as long as it results in a smooth equipotential surface underneath the corona wire. Variations of this include a wire mesh, parallel conductive tubes, or a foil skirt with a smooth round edge. The foil skirt collector is the most popular for small models, and is usually, but not necessarily, connected to the negative side of the power supply. It is usually conveniently made from cheap, lightweight aluminum foil. The foil skirt is named simply because it is shaped much like a skirt, and is made from aluminum foil. It is by far the most fragile part, and must not be crumpled to work properly. Any sharp edges on the skirt will degrade the performance of the thruster, as this will generate ions of opposite polarity to those within the thrust mechanism. Reversing the polarities of the corona wire with that of the foil does not alter the direction of motion. Thrust will be produced regardless of whether the ions are positive or negative. For positive corona polarity, nitrogen ions are the main charge carriers, whilst for negative polarity, oxygen ions will be the main carriers and ozone production will be higher. The slight difference in their ion mobility, results in slightly higher thrust for the positive corona polarity case.[citation needed] Mechanism The generated thrust can be explained in terms of electrokinetics or, in modern terms, electrohydrodynamics propulsion and can be derived through a modified use of the Child-Langmuir equation.[5] A generalized one-dimensional treatment gives the equation: F = I d k {\displaystyle F={\frac {Id}{k}}} where F is the resulting force, measured in dimension M L T−2 I is the current flow of electric current, measured in dimension I. d is the air gap distance, measured in dimension L. k is the ion mobility coefficient of air, measured in dimension M−1 T2 I (Nominal value 2·10−4 m2 V−1 s−1). In its basic form, the ionocraft is able to produce forces great enough to lift about a gram of payload per watt,[6] so its use is restricted to a tethered model. Ionocraft capable of payloads in the order of a few grams usually need to be powered by power sources and high voltage converters weighing a few kilograms, so although its simplistic design makes it an excellent way to experiment with this technology, it is unlikely that a fully autonomous ionocraft will be made with the present construction methods. Further study in electrohydrodynamics, however, show that different classes and construction methods of EHD thrusters and hybrid technology (mixture with lighter-than-air techniques), can achieve much higher payload or thrust-to-power ratios than those achieved with the simple lifter design. Practical limits can be worked out using well defined theory and calculations.,[7][8] Thus, a fully autonomous EHD thruster is theoretically possible. When the ionocraft is turned on, the corona wire becomes charged with high voltage, usually between 20 and 50 kV. The user must be extremely careful not to touch the device at this point, as it can give a nasty shock. At extremely high current, well over the amount usually used for a small model, contact could be fatal. When the corona wire is at approximately 30 kV, it causes the air molecules nearby to become ionised by stripping the electrons away from them. As this happens, the ions are strongly repelled away from the anode but are also strongly attracted towards the collector, causing the majority of the ions to begin accelerating in the direction of the collector. These ions travel at a constant average velocity termed the drift velocity. Such velocity depends on the mean free path between collisions, the external electric field, and on the mass of ions and neutral air molecules. The fact that the current is carried by a corona discharge (and not a tightly-confined arc) means that the moving particles are diffusely spread out into an expanding ion cloud, and collide frequently with neutral air molecules. It is these collisions that create a net movement. The momentum of the ion cloud is partially imparted onto the neutral air molecules that it collides with, which, being neutral, do not eventually migrate back to the second electrode. Instead they continue to travel in the same direction, creating a neutral wind. As these neutral molecules are ejected from the ionocraft, there are, in agreement with Newton's Third Law of Motion, equal and opposite forces, so the ionocraft moves in the opposite direction with an equal force. There are hundreds of thousands of molecules per second ejected from the device, so the force exerted is comparable to a gentle breeze. Still, this is enough to make a light balsa model lift its own weight. The resulting thrust also depends on other external factors including air pressure and temperature, gas composition, voltage, humidity, and air gap distance. The air gap is very important for the function of this device. Between the electrodes there is a mass of air, consisting of neutral air molecules, which gets in the way of the moving ions. This air mass is impacted repeatedly by excited particles moving at high drift velocity. This creates resistance, which must be overcome. The barrage of ions will eventually either push the whole mass of air out of the way, or break through to the collector where electrons will be reattached, making it neutral again. The end result of the neutral air caught in the process is to effectively cause an exchange in momentum and thus generate thrust. The heavier and denser the gas, the higher the resulting thrust. Recent research suggests electrohydrodynamic propulsion is more energy efficient (thrust per unit power) than other means of propulsion, generating up to 100N of thrust per kilowatt of power, compared to 2 N/kW for jet engines.[9] This is mainly due to the much lower air speed of an ionocraft vs a jet engine, as power requirement per unit mass of payload drops with air velocity. However this also means the ionocraft needs a much wider surface area to lift the same payload. See also Hall effect thruster Ion thruster Magnetoplasmadynamic thruster Plasma actuator Wingless Electromagnetic Air Vehicle References 1.          Massachusetts Institute of Technology (2013, April 3). Ionic thrusters generate efficient propulsion in air. ScienceDaily Quote: "...In their experiments, they found that ionic wind produces 110 newtons of thrust per kilowatt, compared with a jet engine's 2 newtons per kilowatt..."     "Ion Propulsion" (PDF).     Thompson, Clive (August 2003). "The Antigravity Underground". Wired Magazine.     Tajmar, M. (2004). "Biefeld-Brown Effect: Misinterpretation of Corona Wind Phenomena". AIAA Journal. 42 (2): 315. Bibcode:2004AIAAJ..42..315T. doi:10.2514/1.9095.     "Electrokinetic devices in air" (PDF). Retrieved 2013-04-25.     Lifter efficiency relation to ion velocity "J L Naudin’s Lifter-3 pulsed HV 1.13g/Watt" Archived 2014-08-08 at the Wayback Machine.     Full analysis & design solutions for EHD Thrusters at saturated corona current conditions     Granados, Victor H.; Pinheiro, Mario J.; Sa, Paulo A. (July 2016). "Electrostatic propulsion device for aerodynamics applications". Physics of Plasmas. Retrieved 27 April 2017. 9.        Barrett, Stephen R.H.; Masuyama, Kento (5 March 2013). "On the performance of electrohydrodynamic propulsion". Proceedings of the Royal Society. Retrieved 3 April 2013. Sources Talley, R .L., "Twenty First Century Propulsion Concept". PLTR-91-3009, Final Report for the period Feb 89 to July 90, on Contract FO4611-89-C-0023, Phillips Laboratory, Air Force Systems Command, Edwards AFB, CA 93523-5000, 1991. Tajmar, M., "Experimental Investigation of 5-D Divergent Currents as a Gravity-Electromagnetism Coupling Concept". Proceedings of the Space Technology and Applications International Forum (STAIF-2000), El-Genk editor, AIP Conference Proceedings 504, American Institute of Physics, New York, pp. 998–1003, 2000. Tajmar, M., "The Biefeld-Brown Effect: Misinterpretation of Corona Wind Phenomena". AIAA Journal, Vol 42, pp 315–318 2004. DR Buehler, Exploratory Research on the Phenomenon of the Movement of High Voltage Capacitors. Journal of Space Mixing, 2004 FX Canning, C Melcher, E Winet, Asymmetrical Capacitors for Propulsion. 2004. GV Stephenson The Biefeld Brown Effect and the Global Electric Circuit. AIP Conference Proceedings, 2005.[dead link]   Electrode From Wikipedia, the free encyclopedia   (Redirected from Electrodes) For the Pokémon, see Electrode (Pokémon). This article needs additional citations for verification. Please help improve this article by adding citations to reliable sources. Unsourced material may be challenged and removed. (January 2010) (Learn how and when to remove this template message) Electrodes used in arc welding An electrode is an electrical conductor used to make contact with a nonmetallic part of a circuit (e.g. a semiconductor, an electrolyte, a vacuum or air). The word was coined by William Whewell at the request of the scientist Michael Faraday from the Greek words elektron, meaning amber (from which the word electricity is derived), and hodos, a way.[1][2] The electrophore, invented by Johan Wilcke, was an early version of an electrode used to study static electricity.[3] Contents 1 Anode and cathode in electrochemical cells 1.1 Primary cell 1.2 Secondary cell 2 Other anodes and cathodes 3 Welding electrodes 4 Alternating current electrodes 5 Uses 6 Chemically modified electrodes 7 See also 8 References Anode and cathode in electrochemical cells Configuration of the electrode An electrode in an electrochemical cell is referred to as either an anode or a cathode (words that were coined by William Whewell at Faraday's request).[1] The anode is now defined as the electrode at which electrons leave the cell and oxidation occurs (indicated by a minus symbol, "−"), and the cathode as the electrode at which electrons enter the cell and reduction occurs (indicated by a plus symbol, "+"). Each electrode may become either the anode or the cathode depending on the direction of current through the cell. A bipolar electrode is an electrode that functions as the anode of one cell and the cathode of another cell. Primary cell A primary cell is a special type of electrochemical cell in which the reaction cannot be reversed,[4] and the identities of the anode and cathode are therefore fixed. The anode is always the negative electrode. The cell can be discharged but not recharged. Secondary cell A secondary cell, for example a rechargeable battery, is a cell in which the chemical reactions are reversible. When the cell is being charged, the anode becomes the positive (+) and the cathode the negative (−) electrode. This is also the case in an electrolytic cell. When the cell is being discharged, it behaves like a primary cell, with the anode as the negative and the cathode as the positive electrode. Other anodes and cathodes In a vacuum tube or a semiconductor having polarity (diodes, electrolytic capacitors) the anode is the positive (+) electrode and the cathode the negative (−). The electrons enter the device through the cathode and exit the device through the anode. Many devices have other electrodes to control operation, e.g., base, gate, control grid. In a three-electrode cell, a counter electrode, also called an auxiliary electrode, is used only to make a connection to the electrolyte so that a current can be applied to the working electrode. The counter electrode is usually made of an inert material, such as a noble metal or graphite, to keep it from dissolving. Welding electrodes In arc welding, an electrode is used to conduct current through a workpiece to fuse two pieces together. Depending upon the process, the electrode is either consumable, in the case of gas metal arc welding or shielded metal arc welding, or non-consumable, such as in gas tungsten arc welding. For a direct current system, the weld rod or stick may be a cathode for a filling type weld or an anode for other welding processes. For an alternating current arc welder, the welding electrode would not be considered an anode or cathode. Alternating current electrodes For electrical systems which use alternating current, the electrodes are the connections from the circuitry to the object to be acted upon by the electric current but are not designated anode or cathode because the direction of flow of the electrons changes periodically, usually many times per second. Uses Electrodes are used to provide current through nonmetal objects to alter them in numerous ways and to measure conductivity for numerous purposes. Examples include: Electrodes for fuel cells Electrodes for medical purposes, such as EEG (for recording brain activity), ECG (recording heart beats), ECT (electrical brain stimulation), defibrillator (recording and delivering cardiac stimulation) Electrodes for electrophysiology techniques in biomedical research Electrodes for execution by the electric chair Electrodes for electroplating Electrodes for arc welding Electrodes for cathodic protection Electrodes for grounding Electrodes for chemical analysis using electrochemical methods Inert electrodes for electrolysis (made of platinum) Membrane electrode assembly Chemically modified electrodes Chemically modified electrodes are electrodes that have their surfaces chemically modified to change the electrode's physical, chemical, electrochemical, optical, electrical, and transportive properties. These electrodes are used for advanced purposes in research and investigation.[5] See also Working electrode Reference electrode Gas diffusion electrode Battery Redox Reaction Cathodic protection Galvanic cell Anion vs. Cation Electron versus hole Electrolyte Electron microscope Noryl Tafel equation Hot cathode Cold cathode Electrolysis Reversible charge injection limit References Wikimedia Commons has media related to Electrodes. 1.          Weinberg, Steven (2003). The Discovery of Subatomic Particles Revised Edition. Cambridge University Press. pp. 81–. ISBN 978-0-521-82351-7. Retrieved 18 February 2015.     Faraday, Michael (1834). "On Electrical Decomposition". Philosophical Transactions of the Royal Society. Archived from the original on 2010-01-17. Retrieved 2010-01-17. In this article Faraday coins the words electrode, anode, cathode, anion, cation, electrolyte, and electrolyze.     Whitaker, Harry (2007). Brain, mind and medicine : essays in eighteenth-century neuroscience. New York, NY: Springer. p. 140. ISBN 0387709673.     Sivasankar (2008). Engineering Chemistry. Tata McGraw-Hill Education. ISBN 9780070669321.   Durst, R., Baumner, A., Murray, R., Buck, R., & Andrieux, C., "Chemically modified electrodes: Recommended terminology and definitions (PDF)", IUPAC, 1997, pp 1317–1323.

বেদের পুনরুদ্ধার ( পার্ট - ৪ )

বেদের পুনরুদ্ধার (পার্ট - ৪) যারা নিয়মিত এই লেখা ধারাবাহিক ভাবে পড়ছেন তারা দেখেছেন এর আগে চাঁদ - পৃথিবীর দুরত্ব, pi এর মান, এমন কী পিথাগোর...