精華區beta StarTrek 關於我們 聯絡資訊
====================================================================== Relativity and FTL Travel by Jason W. Hinson ([email protected]) ---------------------------------------------------------------------- Part III: A Bit About General Relativity ====================================================================== This is PART III of the &quotRelativity and FTL Travel" FAQ. It is an &quotoptional reading" part of the FAQ in that the FTL discussion in PART IV does not assume that the reader has read the information discussed below. If your only interest in this FAQ is the consideration of FTL travel with relativity in mind, then you may only want to read PART I and PART IV. In this part, we take a look at general relativity. The discussion is rather lengthy, but I hope you will find it straight forward and easy to follow. The subject of GR is quite new to this FAQ, and your comments on the usefulness, ease of reading, etc. for this part of the FAQ would be appreciated. For more information about this FAQ (including copyright information and a table of contents for all parts of the FAQ), see the &quotRelativity and FTL Travel--Introduction to the FAQ" portion which should be distributed with this document. Contents of PART III: 5. Introduction to General Relativity 5.1 Reasoning for its Existence 5.2 The &quotNew Inertial Frame" 5.3 Manifolds, Geodesics, Curvature, and Local Flatness 5.4 The Invariant Interval 5.5 A Bit About Tensors 5.6 The Metric Tensor and the Stress-Energy Tensor 5.7 Applying these Concepts to Gravity 5.7.1 The Basic Idea 5.7.2 Some Notes on the Physics and the Math 5.7.3 First Example: Back to SR 5.7.4 Second Example: Stars and Black Holes 5.8 Experimental Support for GR 5. Introduction to General Relativity Thus far, we have confined our talks to the realm of what is known as Special Relativity (or SR). In this section I will introduce a few of the main concepts in General Relativity (or GR). The difference between the two is basically that GR deals with how relativity applies to gravitation. As it turns out, our concept of how gravity works must be changed because of relativity, and GR explains the new concept of gravity. It is called &quotGeneral" relativity because if you look at General Relativity in the case where there is little or no gravity, you get Special Relativity (SR is a special case of GR). Now, GR is a heavily mathematical theory, and while I will try to simply give the reader some understanding of the physical notions underlining the theory, some mathematics will inevitably come into play. I will, however, try to give simple, straight-forward explanations of where the math comes from and how it helps explain the theory. I will start by discussing why we might even think that gravity and relativity are related in the first place. This will lead us to change our concepts of space and time in the presence of gravity. To discuss this new concept of space-time, we will need to introduce the idea of mathematical constructs known as Tensors. The two tensors we will talk about in specific are called the Metric Tensor and the Stress-Energy Tensor. Once we have discussed these concepts, we will look at how it all comes together to produce the basic ideas behind the theory of general relativity. We will also consider a couple of examples to illustrate the use of the theory. Finally, we will mention some of the experimental evidence which supports general relativity. 5.1 Reasoning for its Existence To start off our discussion, I want to indicate why one would reason that gravity and relativity are connected. While I could start with a somewhat unbelievable thought experiment to explain the first point I want to make, perhaps it will be better if I just tell you about actual experimental evidence. We thus start by considering an experiment in which a light beam is emitted from Earth and rises in the atmosphere to some point where the light is detected. When one performs this experiment, one finds that the energy of the light decreases as it rises. So, what does this have to do with our view of relativity and gravity? Well, let's reason through the situation: First, we note that the energy of light is related to its frequency. (If you think of light as a wave with crests and troughs, and if you could make note of the crests and troughs as they passed you, then you could calculate the frequency of the wave as 1/dt, where dt is the time between the point when one crest passes you and the point when the next crest passes you.) So, if the energy of the light decreases (and thus its frequency decreases), then dt (the time between crests) must increase. Let's then consider a frame of reference sitting stationary on the Earth. We will look at a space-time diagram in this frame which shows the paths that two crests would take as the light travels away from the Earth. In Diagram 5-1 I have drawn indications of the paths the two crests might take. The diagram shows distance above the Earth as distance in the positive x direction, so as time goes on, the two crests rise (move in the positive x direction) and eventually meet the detector. Now, we don't know what the gravity of the Earth might do to the light. We thus want to generalize our diagram by allowing for the possibility that the paths of the crests might be influenced in some unknown way by gravity. So, I have drawn a haphazard path for the two crests marked with question marks. The actual paths don't matter for our argument, but what does matter is this: whatever gravity does to the light, it must act the same way on both crests. Therefore, the two haphazard paths are drawn the same way. Diagram 5-1 t # = detector's path | # | ? | ? # | second ? # dt-final | crest ? ? | ? ? # | ? ? # | ? ? # ? ? first # dt-initial | ? crest # | ? # ------------?------------------#------> x (distance above surface) | # | # As we see in the diagram, because gravity acts the same way on both crests, the time between them when they leave the surface (dt- initial) is the same as the time between them when they are detected (dt-final). Thus, our diagram does not predict that the energy of the light should change, but experimental evidence shows it does. According to special relativity, this frame of reference we have drawn is an inertial frame (we are ignoring the Earth's motion around the sun here) which should explain the geometry of the situation, but does not. That indicates that SR must be changed in light of gravity. However, we have yet to show that SR must be completely thrown out. What if there were another way to define an inertial frame such that it's geometry would explain the above situation and other situations which occur in the presence of a gravitational field? That is what we will consider next. 5.2 The &quotNew Inertial Frame" Before starting this section, I want to mention something to the reader: in the end, when gravity is concerned, we will not be able to find a single inertial frame of reference which will correctly explain the geometry of all situations. This will be the actual deathblow to special relativity. At the beginning of the discussion in this section, it will look as if the situation is hopeful, and that by defining a proper inertial frame, SR will be saved. However, later in this section, we will see where this all falls apart, and I wanted the reader to realize this from the beginning. Now, rather than consider a frame of reference which is sitting stationary on the surface of the Earth, let's consider one which is freely falling in Earth's gravity near the surface of the Earth. Why would we want to consider this frame? Okay, let's first address that question. In special relativity, an inertial frame was one which moved at a constant velocity. Thus, an accelerating frame was not an inertial frame in SR. Consider, then, a frame of reference inside a spaceship which is accelerating at a constant rate. The ship we are considering will be far away from any gravitational fields. If you were to release an object inside that ship, the object would continue to travel at the velocity it had when you let it go. However, the ship would continue to accelerate past that velocity, and the bottom of the ship would soon catch up to the object you released. So, if you are on this ship, and you release an object, then to you the object seems to accelerate at a constant rate toward the bottom of the ship. Notice that the rate of acceleration the object seems to have is that of the ship, regardless of the object's mass or composition. This is very similar to the situation in which you release an object while sitting stationary on the Earth's surface. All such objects near the Earth's surface accelerate towards the Earth at a constant rate, regardless of their mass or composition. So we can argue that the frame of reference which is sitting stationary on the Earth's surface seems to have the properties of a non-inertial frame (like that of the accelerating ship). Now, think of the object being released in the above situation. Once you release the object, it continues on at the velocity it had when you released it. It continues at a constant velocity. It is in an inertial frame. If this situation is quite similar to the one on Earth, then we might argue that an object released near the surface of the Earth--an object in free-fall--is also in an inertial frame. For another important argument, consider a point we mentioned above--gravity creates the same rate of acceleration for all objects released at a given point with a given initial velocity. This fact is what distinguishes gravity from all other forces in nature. With the other three forces (electromagnetism, the strong nuclear force, and the weak nuclear force) the motion of an object in the presence of the force depends on the composition of the object. For example, electromagnetism doesn't act on neutral particles, but does act on charged ones. However, when we consider gravity, the path taken by an object which is released with a given velocity in a gravitational field does not depend on the composition of the object. Thus, if you are in a freely falling frame of reference (one in which you are only being acted on by gravity), then any object you release will follow the same trajectory that you are following. It will not move with respect to you, and it will seem to you as if no force is acting on it. So, the freely falling frame looks, again, like an inertial frame of reference. Finally, let's consider the &quotlight rising in the presence of Earth's gravity" experiment. As it turns out (though I won't go into the proof) if the light is detected while it is still relatively close to the Earth, and we consider the experiment in a frame of reference which is freely falling near the Earth's surface, then in that frame, the light does not loose energy. Thus, in the freely falling frame of reference, Diagram 5-1 can correctly depicts the geometry of the situation. And so, things are looking deceptively hopeful. At this point it looks as if we could simply consider free falling frames as inertial, and the space-time diagrams we have drawn throughout our discussions would thus work just fine in the presence of gravity, as long as we understand that they are drawn in free falling frames. However, there is a problem. To illustrate why, consider the accelerating ship we were discussing earlier, but let the ship be very, very tall. No matter how tall the ship is, an object dropped at the top of the ship will accelerate at the same rate as an object dropped at the bottom of the ship. However, general gravitational fields don't work this way. Objects in a weaker gravitational field (further from the Earth, for example) accelerate at a different rate than those in a stronger field. Now, as long as you are close to the surface of the Earth, you won't notice the different acceleration rates for objects dropped at different heights. However, if you drop one object close to the surface of the Earth and the other object far above the first, then they will accelerate at different rates. If you consider the frame of reference of one of the objects, the other object will be accelerating in that frame. Thus, while our previous discussion would have us call both of these frames inertial, one frame is accelerating in the other frame of reference. Similarly, consider dropping two objects from different sides of the Earth. Because they will each fall towards the center of the Earth, they will be accelerating in different physical directions. Thus, they will each be accelerating in the other object's frame. And so, we note that a free falling frame seems much like an inertial frame as long as you are close to the origin of the frame; however, if you consider a point further away, the frame does not represents the inertial frame at that far away point. Not only that, but if you consider a frame which starts far from the Earth, then that frame will eventually fall into an area with a stronger gravitational field. Thus, as time goes on, the frame of reference changes from representing an inertial frame far from the Earth to representing another inertial frame close to the Earth. So, the extent to which the free falling frame represents an inertial frame at the point it was originally dropped depends on how long you consider that frame. In other words, the free falling frame only represents a good inertial frame for a limited time. In the end, we see that free falling frames can be considered as inertial frames only over a small distance and for a small period of time. We call them &quotlocal" inertial frames (&quotlocal" meaning in space as well as time). It is similar to noting that locally on a the surface of a sphere, a plane closely represents a good coordinate system for the surface of the sphere. However, globally--as you extend that plane--it stops being a good coordinate system for the curved surface of the sphere. Similarly in relativity, there is no way to define a single, rigid frame of reference which has the properties of an inertial frame everywhere within a gravitational field. In special relativity, such frames existed, but with gravity involved, we must rethink the situation. We will now continue this rethinking process by discussing concepts which can be used to describe space-time in the presence of gravity. We will begin by discussing some general ideas which will help us explain the geometry of space-time. 5.3 Manifolds, Geodesics, Curvature, and Local Flatness Before we discuss space-time in the presence of gravity, we need to understand some basic geometric concepts which we will use. We will develop these concepts by considering normal, spatial geometry which can be fully grasped using common sense. Applying these concepts to space-time becomes less intuitive (in part because we still aren't that used to thinking of time as just another dimension); therefore, developing them using normal spatial geometry will be beneficial. First, we introduce the term &quotmanifold&quot. Basically, for our purposes, you can think of a manifold as a fancy term for a space. The space around us that you are used to thinking of can be called a three dimensional manifold. The surface of a sheet of paper is a two dimensional manifold, as is the surface of a cylinder or the surface of a sphere. Next, we look at a particular type of path on a manifold. This path is called a geodesic, and it is essentially the path which takes the shortest distance between two points on the manifold. On a piece of paper (a flat manifold) the shortest distance between two points is found by following the path of a straight line. However, for a sphere, the shortest distance between two points would be traveled by following a curve known as a great circle. If you imagine cutting a sphere directly in half and then putting it back together, then the cut mark on the surface of the sphere would be a great circle. If you move along the surface of a sphere between two points, then the shortest path you could take would lie on a great circle. Thus, a great circle on a sphere is basically equivalent to a line on a flat manifold--they are both geodesics on their respective manifolds. Similarly, on any other manifold there would be a path to follow between two points such that you would travel the shortest distance. Such a path is a geodesic on that manifold. Next, we introduce the concept of the curvature of a manifold. When we discuss this concept, we are talking about an intrinsic property of the geometry of a manifold. To demonstrate what I mean, let's consider the surface of a cylinder. You can create such a surface by taking a flat sheet of paper and rolling it up. While the two dimensional surface will then look curved in our three dimensional perspective, the geometry of the surface is exactly the same as the geometry of the flat sheet of paper from which it was made. If you were a two dimensional creature confined to live on the two dimensional surface of the cylinder, then you could not perform an experiment which would prove that your geometry was that of a three dimensional cylinder rather than a flat sheet of paper. Thus, though a cylinder looks curved from our three dimensional perspective, it has no intrinsic curvature to its geometry. On the other hand, consider a sphere. You cannot bend a flat sheet of paper around a sphere without crumpling the paper. The geometry on the surface of a sphere will then be different from the geometry of a flat sheet of paper. To distinctly show this, let's consider a couple of two dimensional creatures who are confined to the surface of a sphere. Say that they stand next to one another on the two dimensional surface and begin walking parallel to one another. As they continue to walk, each will continue in what seems to him to be a straight line. If they do this--if each of them believes that they are following a straight line from one step to the next--then they will follow the path of a geodesic on the sphere. As we said earlier, this means that they will each follow a great circle. But if they each follow a great circle on the surface of a sphere, then they will eventually come towards one another and meet. Now, they started out moving on parallel paths, and they each believed that they were walking in a straight line, but their paths eventually came together. This would not be the case if they performed this experiment on a flat sheet of paper. Thus, creatures who are confined to live on the two dimensional surface of a sphere could tell that the geometry of their space was different from the geometry of a flat piece of paper. That intrinsic difference is due to the curvature of the sphere's surface. This, then, is what we want to note about curvature: The curvature of a manifold as in intrinsic property of the geometry of the manifold itself. It is intrinsic because it is part of the manifold, regardless of whether the manifold is considered in a higher dimensions. In fact, just because a manifold may looked &quotcurved" in a higher dimension, that doesn't mean that its intrinsic geometry is different from that of a flat manifold (i.e. it's geometry can still be flat--like the cylinder). Thus, the test of whether a manifold is curved does not have anything to do with higher dimensions, but with experiments that could be performed by beings confined on that manifold. (Specifically, if two parallel lines do not remain parallel when extended on the manifold, then the manifold possesses curvature). This is important to us in our discussion of space-time in the presence of gravity. It means that the curvature of the four dimensional manifold of space-time in which we live can be understood without having to worry about or even speculate on the existence of any other dimensions. As a final note in this introduction to manifolds, I want to mention a bit about local flatness. Note that even though a manifold can be curved, on a small enough portion of that manifold, it is fairly flat. For example, we can represent a city on our curved Earth by using a flat map. The map will be a very good representation of the city because it is a very small piece of the curved manifold. Earlier I mentioned that over a small enough piece of space-time in the presence of gravity, you can define a frame of reference which is still very similar to an inertial reference frame in special relativity. This gives an indication as to why the geometry of space- time in special relativity is that of a flat manifold, while with general relativity, space-time is said to be curved in the presence of gravity. Later we will see how the concepts discussed here will help us in explaining gravity and relativity. Next, however, we want to discuss another property of manifolds which itself will tell us everything we want to know about a particular manifold. We will call this property the invariant interval. 5.4 The Invariant Interval Here we will basically be discussing distances on manifolds, and what we can learn about a manifold based on how we calculate distances on that manifold. We start by discussing the le &quotL" (L is positive if we move east). Next, we need to move north or south on the sphere to reach P. The distance we move north or south to reach P will be called &quotH" (H is positive if we move north). That gives us our coordinate system. Every point on the sphere can now be represented by an L-H coordinate pair. The &quotgrid" on the surface of the sphere which represents this coordinate system would be made of latitude and longitude lines such as those on a globe. Next, we need to figure out what infinitesimal distance (ds) would be associated with moving a small distance in L (dL) and a small distance in H (dH). For the sake of time, I'll just give the answer here (Note, R is the radius of the sphere): ds^2 = (1/R^2)*dH^2 + [cos(H/R)/R]^2*dL^2 Remember what this represents. If you start at some point (L,H) on the sphere, and you move a small distance in L (dL) and a small distance in H (dH) then the shortest distance along the sphere between your first position and your second position would be ds. Note that this distance depends on your L position (because of the &quotsin(L/R)" part of the equation). (This is an interesting point because as soon as you start moving from one position to the next, the equation for ds becomes slightly different. We basically think of this difference as negligible as long as dL is very small, but, in fact, the equation is only correct when dL is truly &quotinfinitesimal&quot. Such concepts are generally covered in calculus, and for our purposes, we will just claim that the equation is practically true as long as dL is very small.) So now, we come to an important point in this section. What if I told you that I could find another coordinate system on the sphere using two independent coordinates (a and b) such that the invariant interval on the sphere would be given by the following: ds^2 = da^2 + db^2? (Note: by &quotindependent coordinates" I mean that you can always change your position in one coordinate independent of any change in the other.) Here I'll try to show that my claim cannot be true, because it would imply that the sphere and a flat sheet of paper have the same geometry, regardless of how I try to define &quota" and &quotb" on the sphere. First, what if I draw a normal grid on a flat piece of paper and label the axes &quota" and &quotb&quot. &quotBig deal," you might say, &quotyou could just as easily label them 'L' and 'H', which were the coordinate you really did use on the sphere." AH, but here is the difference between the two labelings. The invariant interval along the flat sheet of paper would be da^2 + db^2 and dL^2 + dH^2 for the two labelings, respectively. In the second case, we obviously see that the geometry of the sphere is different from the geometry of the flat grid (because the invariant interval on the sphere is different from the &quotdL^2 + dH^2" invariant interval on the flat grid). However, I have claimed that the invariant interval on the sphere using my new a-b system is &quotda^2 + db^2&quot. That would make it's physical geometry the same as that of the flat sheet of paper--which cannot be the case. Considering this example, let's make some general points: First, consider some manifold, M1. On M1, we have some coordinate system, S1. Next we consider two very-nearby points on M1 (call the points P and Q). If we know the distance between P and Q along each of the coordinates (like dx and dy, for example), then we can find some function for ds (the shortest distance on M1 between the very-nearby points) using the coordinates in S1. Now, consider a second manifold, M2. If a coordinate system, S2, can be defined on that manifold such that ds has the same functional form in S2 as it did using the S1 coordinate system on M1, then the geometry of the two manifolds must be identical. This indicates that the geometry of a manifold is completely determined if one knows the form of the invariant interval using a particular coordinate system on that manifold. And, there you have it. In fact, starting with the form of the invariant interval in some coordinate system on a manifold, we can determine the curvature of the manifold, the path of a geodesic on the manifold, and everything we need to know about the manifold's geometry. Now, the mathematics used to describe these properties involves geometric constructs known as tensors. In fact, the invariant interval on a manifold is directly related to a tensor known as the metric tensor on the manifold, and we will discuss this a bit later. First, I want to give a very brief introduction to tensors in general. 5.5 A Bit About Tensors In this section I will introduce just a few basic ideas which will give the reader a feeling for what tensors are. This is simply meant to provide a minimum amount of information to those who do not know about tensors. Basically, a tensor is a geometrical entity which is identified by its various components. To give a solid example, I note that a vector is a type of tensor. In an x-y coordinate system, a vector has one component which points in the x direction (its x component) and another component which points in the y direction (its y component). If you consider a vector defined in three dimensional space, then it will also have a z component as well. Similarly a tensor in general is defined in a particular space which has some number of dimensions. The number of dimensions of the space is also called the number of dimensions of the tensor. Note that vectors have a component for each individual (one) dimension, and they are called tensors of rank 1. For other tensors, you have to use two of the dimensions in order to specify one component of the tensor. In x-y space, such a tensor would have an xx component, an xy component, a yx component, and a yy component. In three-space, it would also have components for xz, zx, yz, zy, and zz. Since you have to specify two of the dimensions for each component of such a tensor, it is called a tensor of rank 2. Similarly, you can have third rank tensors (which have components for xxx, xxy, ...), fourth rank tensors, and so on. So that you aren't confused, I want to explicitly note that the dimensionality of a tensor (the number of dimensions of the space in which the tensor is defined) is independent of the rank of the tensor (the amount of those dimensions that have to be used to specify each component of the tensor). In any dimensional space, we can have a tensor of rank 0 (just a number by itself, because it is not associated in any way with any of the dimensions), a tensor of rank 1 (like a vector--it has a component for every one dimension you can specify), a tensor of rank 2 (it has a component for every pair of dimensions you can specify), etc. Now we look at a very important property of tensors. In fact, it is the property which really defines whether a set of components make up a tensor. This property involves the question of how the tensor's components change when you change the coordinate system you are using for the space in which the tensor is defined. So, let's consider an example in two dimensional space where you go from some coordinate system (call the coordinates x and y) to some other coordinate system (call these coordinates x' and y'). There will be some sort of relationship between the two systems. For example, say we start at some point in this space such that our coordinates are x,y and x',y' (depending on which coordinate system you are using). Now, say we move an &quotinfinitesimal distance" in x (using the first coordinate system). Call that distance dx. When we do so, we will have changed or x' position (using the second coordinate system) by some infinitesimal amount, dx'. Also, we will have changed our y' position by some amount dy'. We can use these concepts of infinitesimal changes to define some relationships between the two systems. We can answer the question &quothow does x' change when x changes at this point" by noting the ratio, dx'/dx. Similarly we can write dx/dx' to denote how much x changes with changes in x' at some point, and dy'/dx denotes how y' changes with changes in x. All together there are four of these ratios which denote how the x' and y' coordinates change with changes in x and y: dx'/dx, dx'/dy, dy'/dx, and dy'/dy. Similarly, there are four more to denote how x and y change with changes in x' and y': dx/dx', dx/dy', dy/dx', and dy/dy'. In general the values of these ratios will depend on where you are, so each ratio is a function of x and y (or x' and y', if you like). Now, we have these ratios which help us relate one coordinate system to another. If we have a tensor defined in this space, then we must be able to use those ratios to find out how the tensor's components themselves change when we go from considering them in one coordinate system to considering them in the other. Let's consider a tensor of rank 1 (a vector) in a two dimensional space. Let the vector, call it V, have an x component (V_x) and a y component (V_y). Then, the rules for finding the x' and y' components of the vector at some point are the following: V_x' = dx'/dx V_x + dx'/dy V_y and V_y' = dy'/dx V_x + dy'/dy V_y. That is the way in which this type of first rank tensor must transform from one coordinate system to another. Note that we can write the above equations by using the following: V_a = SUM(b = x,y) [da'/db V_b] In that expression, &quota" can be either x or y (so we actually have two equations). Also, the right side of the equation is a summation where the first term in the summation is found by letting b = x, and the second term is found by letting b = y. Further, we could make this expression more general by noting that it will be true for a space with higher dimensions when we let &quota" be any one of those dimensions and let the sum with b extend over all the dimensions. The fact that the physical components of a vector do actually transform this way is what makes the vector a tensor. However, we should note that not all types of vectors transform this way. To show this is so, first we will consider a function which has a value at every point in x-y space. Call the function f(x,y). Such a function is a 0 rank tensor, because at any point in the space, it has some single, numerical value (it does not have components for x and y like a vector does--you can't ask &quotwhat's its value in the x direction", or &quotwhat's its value in the y direction", because it has only a single number at any point). Note that if we change to another coordinate system, the value of f at some physical point in the space will not change. Because it has no x or y component, it is invariant when you change coordinate systems, as are all 0 rank tensors. This is the way all 0 rank tensors must transform when you change coordinate systems--they must be invariant. Now, back to the point that there are other types of vectors which do not transform as discussed earlier. Let's take the above function at some point and ask &quothow does it change with small changes in x?" If the function changes by an amount df when we move to another x location a distance dx away, then we can write the expression df/dx do tell how f changes with x. We can do the same in y and have the expression df/dy. Then We could define a vector (call it G) which has an x component (G_x) equal to df/dx at every point in x and y, while it has a y component (G_y) equal to df/dy at every point. Now, what if we do this same procedure in the x'-y' coordinate system. We will end up with the x' and y' components of the G vector such that G_x' = df'/dx' and G_y' = df'/dy'. Because of the way this vector is defined, it turns out that it transforms as follows: G_x' = dx/dx' G_x + dy/dx' G_y and G_y' = dx/dy' G_x + dy/dy' G_y As before, we can rewrite these two equations as follows: G_a' = SUM(b = x, y) [db/da' G_b] Note that we are using ratios like db/da' rather than da'/db (which we used earlier). That means that this is a different type of vector (because it transforms in a different way). The vector we discussed earlier (V) is called a contravariant vector, and the fact that it transforms as discussed earlier is what defines it as that type of vector. The G vector is called a covariant vector, and it is defined as such because of the way it transforms. Usually, we express which type of vector we have by the way we denote its components. For contravariant vectors, we denote their components by putting their indexes (the x or the y) in superscripts: x y V and V (or V^{x} and V^{y}), While we denote the components of covariant vectors by putting their indices in subscripts: G and G (or G_x and G_y) x y With this notation, the two different transformations begin to take on an easy to remember form. See if you can't figure out how the &quotupper" indices and the &quotlower" indices match up on both sides of the two transformation equations when they are written as follows: a' da' b V = SUM(b = x,y) -- V db and db G = SUM(b = x,y) -- V a' da' b Notice that the subscript (or superscript) on one side remains &quotupper" (or &quotlower") in the ratio on the other side. Also, note that the summation is always over the index which is repeated on the right side, once in an &quotupper" position and once in a &quotlower" position. This basic &quotformula" helps to produce equations for all transformation in tensor analyses (note this in the next part of this section). It is interesting to note that in the normal spatial coordinates we are used to using (Cartesian coordinates), db/da' = da'/db, and there is no distinction between covariant and contravariant vectors. However, in other systems, the difference is there and must be considered. Finally, we note that with higher rank tensors, they are also defined by the way they transform from one coordinate system to another. For example, consider a second rank tensor, U. It could be that both of its indices are associated with the contravariant type of transformation (note: the following actually denotes four equations because a'b' can be set to x'x', x'y', y'x', or y'y'): a'b' da' db' xx da' db' xy da' db' yx da' db' yy U = -- * -- U + -- * -- U + -- * -- U + -- * -- U dx dx dx dy dy dx dy dy [ da' db' ce ] = SUM(c & e vary over all dimensions) [ -- * -- U ] [ dc de ] Or they could both be associated with covariant the type of transformation: [ dc de ] U = SUM(c,e) [ -- * -- U ] a'b' [ da' db' ce ] Or it could be a mix of the two: a' [ da' de c ] U = SUM(c,e) [ -- * -- U ] b' [ dc db' e ] And that about ends our discussion on tensors. To sum up, they are geometric entities which have components denoted by some number of indices. Each index can be any of the dimensions in which the tensor is defined, and the number of indices needed to specify a component of a tensor is called the tensor's rank. We are familiar with 0 and 1 rank tensors (numbers--or &quotscalars&quot--and vectors). Finally, the way one transforms a tensor from one coordinate system to another depends on the type of tensor, and it (in fact) defines the tensor itself. Each index of a vector will transform in either a contravariant way or a covariant way. These are the basic ideas behind tensors, and they allow us to define some very powerful mathematics. If you are familiar with the usefulness of vectors, then you have touched the surface of the usefulness of tensors in general. In the following section, we will look at two particular tensors, and we will see that they can be quite useful. 5.6 The Metric Tensor and the Stress-Energy Tensor Now that we have had a glimpse at tensors, let's consider a couple that will be important to us. The first is called the metric tensor. I mentioned a couple of sections ago that this tensor is related to the invariant interval for a certain coordinate system on a given manifold. So, let's go back and look at a the two specific invariant intervals which we introduced. First, in normal, x-y, Cartesian coordinates, we have the following: ds^2 = dx^2 + dy^2 Second, on the surface of a sphere, using the L-H coordinate system which we defined, we have this: ds^2 = (1/R^2)*dL^2 + [cos(L/R)/R]^2*dH^2 Now, let's make this more general by considering an arbitrary, two dimensional manifold and an arbitrary coordinate system on that manifold. Let's call the coordinates &quota" and &quotb&quot. Now, in general, the invariant interval on this manifold is defined in terms of the square of that interval (ds^2). The equation for ds^2 involves the infinitesimal distances da and db in second order combinations. By second order combinations, I mean, for example, da^2 or da*db. Thus, in general, the invariant interval will have the following form (note: the g components are generally formulas of a and b): ds^2 = g *da^2 + g *da*db + g *db*da + g *db^2 aa ab ba bb In that equation you see the four components of the metric tensor in this two dimensional, a-b coordinate system. They are the &quotg's" in the equation. For our x-y coordinate system, we have g = 1, g = 0, g = 0, g = 1 xx xy yx yy For our L-H coordinate system, we have g = (1/R^2), g = 0, g = 0, g = [cos(L/R)/R]^2 LL LH HL HH So, we can construct the invariant interval if we know the metric tensor for a coordinate system on a manifold. Now, remember that we said that the form of the invariant interval for a particular coordinate system tells us everything there is to know about the manifold for which those coordinates are valid. So, now we see that all we need to know is the form of the metric tensor. Once we know g, we know the geometry of the manifold. Using tensor analysis, we can take the metric tensor and find an equation for geodesics on the manifold. We can use it to find out all about the curvature of the manifold. We can even use it to find the dot product (we will discuss this a bit later) of two vectors in the a particular coordinate system. Another thing the metric allows us to do is something generally called &quotraising" or &quotlowering" indices. Basically, if you consider a tensor with a contravariant index (which transforms in a particular way as discussed earlier), then there is another corresponding tensor which has a covariant index (and vice versa). For example, consider the tensor A^{a}, which has a contravariant index, a. There is a corresponding covariant tensor, A_a, which can be found using the metric of the space (and coordinate system) we are dealing with. Here is an example how you find it (finding A_x when you know A^{x}) for a coordinate system with some arbitrary coordinates, x and y: x y A = g A + g A x xx xy For a general space and coordinate system, you can write this rule as follows (remember, &quota" can be any one dimension in the space, so this represents a number of equations): b A = SUM(b varies over all dimensions) g A a ab Similarly, if you know the covariant form of A (A_a) you can find the contravariant form by using the following: a ab A = SUM(b varies over all dimensions) g A b But that equation involves the contravariant form of the metric (g^{ab}). In the invariant interval, the metric is expressed in its covariant form (g_ab). It is therefore important for the reader to remember as we discuss various metrics below, that for all of them we have ab 1 g = --- if a = b g ab ab g = 0 if a doesn't = b Thus, using the metric tensor, one can &quotraise" or &quotlower" any index of a tensor. Remember, what one is really doing is finding a form of that tensor which transforms in a different way. With this example of how the metric can be used, we will end our discussion of this tensor. To sum up, the metric tensor on a manifold is a very important entity which not only tells us all about the manifold's geometry, but which also provides a very powerful tool which allows us to deal with that geometry mathematically. The second tensor we want to mention is the stress-energy tensor. I don't want to get to deep into a discussion of the stress-energy tensor, but the reader should know a couple of key points. With the stress-energy tensor, we see our first example of a tensor explicitly defined in four dimensional space-time (though later we will look at the metric tensor defined in 4-d space-time). The stress-energy tensor (T) is also a tensor of rank 2 (like the metric tensor), which gives it 16 components in 4 dimensions. Sometimes we express such a tensor in the form of a matrix as follows: +- -+ | tt tx ty tz | | T T T T | | | | xt xx xy xz | ab | T T T T | T = | | | yt yx yy yz | | T T T T | | | | zt zx zy zz | | T T T T | +- -+ There you can see the 16 different components. Now, each of these components tell us something about the distribution and &quotflow" of energy and momentum in a region. More precisely, T contains information about all the stresses and pressures and momenta in a region. For example, The &quottt" component of the stress-energy tensor would be the density of the energy in the region (the amount of energy--including mass energy--per unit volume). As to why the stress-energy tensor is important to us, that will be discussed further in a bit. However, here we can note the following in order to pull us back towards our discussion of relativity and gravity: In Newtonian physics, gravity was caused by the density of mass in an area. However, in SR we find that mass is just a form of energy, and so we might think that the &quottt" component of the stress-energy tensor would be the right thing to look at when it comes to gravity. However, if we write a rule using one component of a tensor, then because the value of that component will depend on your coordinate system (or frame of reference in space-time) then the rule will also be frame-dependent. In short gravity would not be an invariant theory, and it would require a preferred frame if we based it only on the &quottt" component of T. However, if we use all the components of a tensor to form our theory, then (as it turns out) the theory can be made frame-independent. Einstein thus considered the possibility that the whole stress-energy tensor would need to play a part as the source of gravity. Add to this some insight on curved manifolds and you end up with general relativity, as we will see. 5.7 Applying these Concepts to Gravity Now that we have discussed manifolds and their properties along with some of the basic concepts of tensors, let's see how all of this applies to relativity and gravitation. First, I will go over the main ideas which lead us from what we have discussed so far to a general relativistic theory. After that, I want to mention a few notes on the physics and the mathematics we will be using given the concepts we have gone over. Next, we will go back and looking again at special relativity while applying a bit of our new knowledge. This will show that GR is indeed general, because when applied to space-time without the presence of gravity it will explain a special case--special relativity. Finally, we will look quickly at a specific application of the GR concepts to a space-time in which there is a gravitational field. This application will focus on a particular class of stars and black holes. 5.7.1 The Basic Idea Lets get started with the basic ideas which combine the concepts we have discussed to produce GR. Here I will simply state the main ideas without an explanation of their application. You will get some feel for their application in our two examples to follow. So, here are the main claims of GR which involve the concepts we have discussed. First, the space-time in which we live is a four dimensional manifold. On that manifold there is a metric tensor (or just &quota metric") which describes the geometry of space-time. The metric can be used to find geodesics on the space-time manifold, and when an object goes from one point in space-time to another point in space-time (note: these are not just two points in space, but two potric can be used to find the invariant interval between two space-time points. That interval (recall) can generally be expressed as ds^2 = SUM(a & b vary over space and time dimensions) g *da*db ab Second, consider a vector in our four dimensional space. Such a vector (usually called a four-vector) has four components, three relating to space and one relating to tim