0804.3742/Analysis-of-time.tex
1: \begin{flushright}
2: \begin{tabular}{p{10cm}}
3: \emph{An analysis of the concept of time was my solution.} \smallskip \\
4: Albert Einstein, ``How I created the theory of relativity''
5: \end{tabular}
6: \end{flushright}
7: 
8: \bigskip
9: 
10: Philosophers of science thrive in the analysis of scientific concepts. If we want to draw a lesson on how to make a good analysis of the physical concept of time, we should take a look at the works of (broadly speaking) natural philosophers---from V b.C.\ to XXI a.D.---\emph{or should we?} What do philosophers of science do when they make an analysis?
11: 
12: A conceptual analysis can be interpreted variously. It may refer to the meticulous \emph{logical dissection} of a concept into constituent parts or to the \emph{physical examination} of a concept regarding its physical origin and nature. If, as we suggest, these two activities are not one and the same, how do they differ? And what is their connection with the activities of scientists and philosophers of science?
13: 
14: In his enlightening book, Reichenbach \citeyear{Reichenbach:1951} finds the source of ``philosophical error'' in speculative philosophy in ``certain \emph{extralogical motives}'' (p.~27). These are compounded in the speculative philosopher's psychological make-up, ``a problem which deserves more attention than is usually paid to it'' (p.~36). Reichenbach describes the fall of speculative (i.e.\ traditional) philosophy and the rise of scientific philosophy, a philosophy to be guided by modern science and its tools. Unlike speculative philosophy, the goal of scientific philosophy is not a one-person explanation of the natural world but the elucidation of scientific thought; its object of study is science itself. Or in the words of Reichenbach, ``Scientific philosophy\ldots leaves the explanation of the universe entirely to the scientist; it constructs the theory of knowledge by the analysis of the results of science'' (p.~303). Identifying scientific
15: philosophy with modern philosophy of science, we can preliminarily conclude that, whereas the \emph{raw material} of science is the natural world, the \emph{raw material} of philosophy of science is (as its name correctly suggests) the scientific enterprise.
16: 
17: As distinct as science and philosophy of science seem to be, we still find ourselves at a loss to tell which of the two disciplines we should pursue if we want to better understand the physical concept of time. A reading of Reichen\-bach suggests that we should embrace philosophy of science, because it is through it, and not normal science, that we gain \emph{clarification} of scientific concepts (p.~123). Indeed, Reichenbach traces the rise of the full-blown ``professional philosopher of science'' to the incipient attempts of scientists who would unsystematically ponder the nature of their fields between the lines of their technical works. But has this evolutionary development from enquiring scientist to professional philosopher of science truly yielded a species better adapted to the clarification of conceptual scientific puzzles? Or has the speciation product, becoming overspecialized, shifted instead to a separate and altogether different habitat, a new place where it can not only survive but even thrive?
18: 
19: If by ``philosophy of science'' we understand the natural extension of the ordinary work of the scientist that comes from asking not only ``how do I calculate this?'' but also---or instead---``what is this, and why should I calculate it?'' then philosophy of science is a useful and integral part of science. Progress in the scientific worldview would not be possible without it. But philosophy of science as a discipline in its own right, carried out ``professionally'' by philosophers because, as Reichenbach put it, ``scientific research does not leave a man time enough to do the work of logical analysis, and\ldots conversely logical analysis demands a concentration which does not leave time for scientific work'' (p.~123)---this is another matter altogether.
20: 
21: Does the meticulous logical dissection of scientific concepts produce useful results without limit? Or is there instead a fuzzy boundary carried beyond which logical analysis stops clarifying anything at all? Reichenbach goes on to admit that philosophy of science, ``aiming at clarification rather than discovery may even impede scientific productivity.'' May it not impede, when carried too far, not just scientific productivity but even clarity itself, thus turning against its original goal?
22: 
23: With the professionalization of the conceptual examination naturally falling on the enquiring scientist, the link with the natural world as the motivator of scientific concepts and \emph{informer of their analysis} was severed in professional philosophy of science, for, as Reichenbach put it, the professional philosopher of science \emph{has no time to deal with nature}. What is then his working method? His starting point is established science and his methods scientific, but his logical analyses roam largely unconstrained by feedback from the physical world. That is to say, philosophy of science freezes the science of the day and, armed with the tools of the same science, looks away from the physical world and proceeds to dissect the concepts of science in as many different ways, into as many possible constituents, as \emph{logic and frozen science} will allow. Thus, philosophers of science become---in the words of one of them---``professional distinction drawers'' \cite[p.~2]{Earman:1989}, because, without \emph{fresh feedback} from the natural world, progress (appropriately understood) must necessarily be made from the differentiation of frozen scientific material into new, abstract forms.
24: 
25: Whether these new forms are useful to the scientific enquiry, whether they actually clarify it or only succeed in making of it a worse conceptual shambles, is not the measure by which philosophical fruitfulness is assayed. Inevitably, having severed the link with nature in the name of professionalism, philosophers judge their works, \emph{de facto} as well as in the words of the same renowned philosopher of science above, ``by one of the most reliable yardsticks of fruitful philosophizing, namely, How many discussions does it engender? How many dissertation topics does it spin off?'' \cite[p.~xi]{Earman:1989}. To a non-philosopher, this is an astounding confession, for one would have thought that philosophers were not just after \emph{more discussion} but \emph{solutions}.
26: 
27: In this way, the discussions of scientific philosophy become similar to those of speculative philosophy, which it thought to have superseded. Stuck in a field of hardened scientific lava and without any objective, natural standard to assay their merits and discard the unfit, the debates of scientific philosophy too are apt to perpetuate themselves through millennia, to last from everlasting to everlasting. To be sure, the debates of scientific philosophy, unlike those of traditional philosophy, are capable of growing in complexity on a par with the development of scientific tools---but whether this is for better or for worse, one does not dare say.
28: 
29: Can we expect the resolution of a conceptual scientific problem out of the arguments of professional scientific philosophy? Hardly, because its yardstick of success inevitably demands of it that it become argument for argument's sake (cf.\ Earman's remarks). Scientific puzzles have always been solved by more and better scientific investigations, by individuals who dared to give a fresh new look at the physical world around them. For this to be possible, the link with the natural world must not be severed. The establishment of new scientific knowledge is then apt to banish some philosophical debates that look rather rusted and obsolete in the light of the new knowledge. (Are spheres the only acceptable shape for the locus of the heavenly bodies? Is the centre of the universe the only acceptable position of the earth? How can Achilles ever overtake the tortoise?) But scientific philosophy drinks in the waters of science---insatiably: for every banished debate, an untold number of others will arise on the basis of the new science, and so on forevermore.
30: 
31: A telltale indication of scientific philosophy as argument for argument's sake can be found in its conspicuously bellicose language, i.e.\ the language typically used by philosophers as they, intellectual enemies, battle each other---till they draw blood (see below) but not to death: it is essential for the well-being of philosophy that all arguments be kept alive. The scientific philosopher's everyday vocabulary is fraught with all sorts of words allusive of violent conflict, so much so that at times one gets the impression to be reading an account of war. One hears\footnote{All
32: italicized words from genuine texts by professional philosophers of science.} of \emph{battlegrounds}, where \emph{enemies}, unfurling myriad \emph{pro-} and \emph{anti-ism} flags, \emph{deploy} \emph{attacks} on the \emph{opponent's} \emph{position}, and \emph{defend} their own with \emph{counterattacks}; of \emph{struggles} in which, amidst the \emph{smoke of battle}, \emph{rivals} seek \emph{victory} over \emph{defeat} by dealing each other \emph{blows} and \emph{undermining} each other's \emph{ground}. Not without a shock, we catch an epitomic glimpse of the scientific philosopher's bellicose mindset in Earman's \citeyear{Earman:1989} book on spacetime ontology, where he seeks to ``\emph{revitalize} what has become an insular and \emph{bloodless} philosophical discussion'' (p.~3, italics added). One may be inclined to excuse such expressions as mere figures of speech, but when these repeat themselves constantly, one is led to presume an underlying reason of being.
33: 
34: The reason behind this custom is now apparent: locked within the inert confines of a frozen science, yet seeking development and progress, the focus of scientific philosophy shifts in practice to the dynamic and constantly evolving views of other philosophers. As a result, argument is built upon argument in a viciously upward spiral, which only a fresh confrontation with the \emph{bare bones of nature's raw material} can collapse back to the ground. A sustained reading of scientific philosophy is thus apt to lead a scientist (besides despair) to the following reflection: philosophy---or \emph{philoversy}? The love of knowledge, or the love of war/confusion?\footnote{The etymological association is particularly apt. \emph{Wers-}, meaning ``to mix up, to lead to confusion,'' is the original Indo-European root of the modern English ``war.'' Cf.\ German \emph{verwirren}, meaning ``to confuse, to perplex,'' originally from the Proto-Germanic root \emph{werra-}. From the first reference below we learn that ``Cognates suggest the original sense was `to bring into confusion.' There was no common Germanic word for `war' at the dawn of historical times.'' (See ``war'' in Online Etymology Dictionary, http://www.etymonline.com; and American Heritage Dictionary, http://www.bartleby.com/61)}
35: 
36: This situation reminds us of Wittgenstein's words: 
37: \begin{quote} A person caught in a philosophical confusion is like a man in a room
38: who wants to get out but doesn't know how. He tries the window but it is too high. He tries the chimney but it is too narrow. And if he would only \emph{turn around}, he would see that the door has been open all the time!\footnote{Quoted from \cite[p.~51]{Malcolm:1972}.} 
39: \end{quote} 
40: The symbolic door is in actuality not so easy to find, but the direction in which we should look to find it is clear: the way out of the philosophical house is the way that originally brought us in, namely, the study of nature. Wittgenstein's metaphorical ``turn around'' turns out to be singularly apt.
41: 
42: To sum up with an exemplifying question: if Einstein had been a philosopher, as he is sometimes called, would he have discovered special relativity through his physical analysis of time, or would he have instead remained scientifically frozen in an endless logical analysis of Newtonian absolute time and space?
43: 
44: All this is not to say that professional philosophy of science \emph{never} helps science or clarifies anything. At times, in the hands of exceptional, scientifically inspired philosophers, professional though they may be, it does; but on a whole, in the hands of the average professional philosopher, it does not.
45: 
46: If the analyses of scientific philosophy go astray due to being logical dissections uninformed by fresh experience, scientists should know better. Scientists, after all, deal with the physical world. This is at any rate true of the ideal scientist, for it is not today uncommon to come across the variety that finds more inspiration in ``self-consistent'' mathematical abstractions than in nature, and has thus more in common with scientific philosophers than scientists. But what about the scientist who investigates the physical world and works in the spirit in which physics has been traditionally carried out for most of its history?\footnote{In connection with space and time, most notably by Riemann and Einstein. In his famous essay, Riemann \citeyear{Riemann:1873}, a mathematician, talks as a physicist when he points out that ``the properties which distinguish [physical] space from other conceivable triply extended magnitudes are only to be deduced from experience,'' and is even wary of assuming the validity of experience ``beyond the limits of observation, on the side both of the infinitely great and of the infinitely small'' (pp.~1--2, online ref.). As for Einstein, his continual use of simple thought experiments during the first half of his life as a starting point for the theoretical concepts of physics is well known.} What does an analysis mean to him? To him, an analysis does not mean definition, postulation, or logical dissection (into what?) but examination. Examination of a concept's \emph{physical} meaning, of its \emph{physical} root and nature. The content of his analysis is not to be found primarily in his head, but in the world around him. This is not to say that an analysis of this type does not rely on the scientist's conceptual tools of thought. It necessarily does, but only in conjunction with a keen observation of nature: his elemental theoretical concepts are
47: developed directly from observations and experiments---science's raw
48: material---not off the top of his secluded head. One particular advantage of proceeding thus is that the scientist in question knows from the start what he is talking \emph{about}: the ontology of his foundational concepts is, by and large, uncontroversial.
49: 
50: This being said, we proceed to a geometric analysis of the concept of time. What can be expected from such an analysis? First and foremost, an understanding of relativity theory on the basis of the raw material of simple measurements and on physical (as opposed to mathematical or philosophical) ideas. Secondly, an understanding of relativity built progressively as a natural concatenation of steps, where by ``natural'' we mean that, at every roadblock, we let physics (and, when necessary, psychology) suggest the way forward. In particular, we let only our physical needs suggest what kind of mathematical concepts we need, and we develop these naturally, not as formal, hard-to-swallow definitions without
51: any apparent reason of being.
52: 
53: An expected upside of this way of proceeding is a minimization of mathematical baggage, but---as every upside surely must have a downside---the sceptical reader may ask, are we perhaps trading depth of understanding for mathematical lightness? Breaking the above-insinuated conservation law of upsides and downsides, we are here unexpectedly confronted by an upside without a downside, but instead conducive to a further upside: it is \emph{because} mathematics is used only in its due measure, as physics demands and no more, that we end up gaining understanding of relativity theory
54: instead of, as is more common, ending up clouded in confusion. The fact that mathematics is in physics only a tool can be ignored only at the physicist's peril.\footnote{Mathematics as practiced by the mathematician, i.e.\ as a formal system empty of semantic meaning, is another matter altogether and not the subject of this criticism.}
55: 
56: The dire consequences of the abuse of mathematical baggage in physics have also been noticed by other people pondering these issues. Physicist Robert Geroch writes: 
57: \begin{quote} 
58: One sometimes hears expressed the view that some sort of uncertainty principle operates in the interaction between mathematics and physics: the greater the mathematical care used to formulate a concept, the less the physical insight to be gained from that formulation. It is not difficult to imagine how such a  viewpoint could come to be popular. It is often the case that the essential physical ideas of a discussion are smothered by
59: mathematics through excessive definitions, concern over irrelevant generality, etc. Nonetheless, one can make a case that mathematics as mathematics, if used thoughtfully, is almost always useful---and occasionally essential---to progress in theoretical physics.
60: \cite[p.~vii]{Geroch:1985} 
61: \end{quote}
62: 
63: In conclusion, for an insightful physical analysis, mathematics in
64: due measure and professional philosophy not at all.
65: 
66: \section{Clock readings as spatiotemporal basis}
67: 
68: \subsection{Events and particles as geometric objects}\label{Events-and-particles}
69: 
70: As we survey the macroscopic world in connection with classical and relativistic theories, two key concepts come to our attention: that of an event and that of a particle. An event is, quite simply, a phenomenon, anything that we  register as happening: a flash of lightning, the sound of a horn, the touch of a
71: hand on the shoulder, etc.
72: 
73: Is an event as such a geometric concept? It certainly is a human-made concept, for what we call an event never appears alone and differentiated from the rest of a jumble of impressions. But it certainly is not a geometric concept. We can ask of it a pair of test questions: does an event have a shape? Does the concept of size apply to it? The answer to both questions is no. Not before we idealize it.
74: 
75: The idealization of an event made in physics is one of the most basic yet extreme cases of geometric idealization. Going much farther than having, for example, the more tangible earth crust idealized as an even spherical surface, here we have an indefinite physical phenomenon squeezed into the tightest of all geometric objects, a point.
76: 
77: Does now a point-event have a shape, and can we speak of its size? The first question is the trickier of the two because, of all geometric objects, a point is by far the most abstract---an extreme case. Unlike all other geometric objects, the shape of a point is \emph{irrelevant} because its size (understood in any way) is zero. This answers the second question above: the concept of size does apply to a point, namely, it is zero. This is not at all the same as saying that it is meaningless to speak of its size, as it is to speak of the size of a raw event. Due to their utmost geometric abstractness, points are strictly speaking impossible to visualize, although they are normally thought of and depicted as dots of some irrelevant shape (circular, square, spherical, etc.).
78: 
79: How does this translate to the physically inspired point-events? By geometrically idealizing physical phenomena as points, the former become endowed with two seemingly physical properties: a point-event is something that does not take space and does not have duration. But with such extreme features, how can point-events be measured (or localized) as such? After all, if they do not take space and do not have duration, they must not exist.
80: 
81: At this difficult junction, counting comes to our aid. The quality of being point-like can be characterized analytically by means of numbers: a point is simply a set of real numbers. In classical physics, a point-event is identified with the set of four numbers $(x,y,z;t)$, and the whole of all possible point-events forms a four-dimensional continuum called space-time---with a hyphen, since $t$ is special and given separately. The same idea can be transferred to relativity theory by loosening the restriction above, according to which time $t$ is given independently of the other three numbers. In relativity theory, a point-event is identified with the set of four numbers $(x^1,x^2,x^3,x^4)$, and the whole of all possible point-events forms a four-dimensional continuum called spacetime---without a hyphen, since no element of the set $(x^1,x^2,x^3,x^4)$ can be separately singled out as special and, in particular, representing time.
82: 
83: The sets of four numbers $(x,y,z;t)$ and $(x^1,x^2,x^3,x^4)$ are the
84: coordinates of a point-event, but what physical meaning shall we attribute to them? If coordinates are to have any meaning, it must be given them operationally: \emph{how} are these numbers attributed to a point-event?
85: 
86: Since the space-time coordinate $t$, symbolizing absolute time in $(x,y,z;t)$, has in fact no operational meaning, we focus only on the general spacetime coordinates $(x^1,x^2,x^3,x^4)$. To localize a spacetime event, we resort to an arbitrary operational method, whereby we specify the way in which the four numbers are measured. Consider as an example Ohanian's \citeyear[pp.~222--223]{Ohanian:1976} astrogator coordinates used  by a hypothetical astronavigator. To determine the spacetime coordinates of his spaceship, he identifies four different stars (red, yellow, blue, and white), measures the angles $\theta^\alpha$ between any four different pairs of stars, and associates each number with a spacetime coordinate, $\theta^\alpha =: x^\alpha$ ($\alpha=1,2,3,4$). 
87: 
88: In spacetime, no coordinate is differentiated from the others; each shares the same status with the other three. But so that two distinct events (e.g.\ two distinct phenomena inside the spaceship) will be tagged with different numbers, at least one of the stars (but possibly all four) must be in relative motion with respect to any of the other three. Ohanian calls this coordinate the ``time coordinate.'' However, this name may be misleading for, if all the stars are moving relative to one another, which one is to be the ``time coordinate''? If the four coordinates are operationally equivalent and any of them can be taken as the ``time coordinate,'' then why give one a special name without reason? The ghost of Newtonian time keeps haunting relativity. 
89: 
90: The establishment of a connection between time and a coordinate, say, $x^4$, is only possible in special cases. For example, if spacetime is separable into a spatial part and a temporal part, i.e.\ if its quadratic form $Q$ can be written as
91: \begin{equation}
92: 	Q=g_{ab}(x^1,x^2,x^3,x^4)\D x^a \D x^b + g_{44}(x^1,x^2,x^3) (\D x^4)^2, 
93: \end{equation}
94: and if we choose comoving coordinates such that $x^1=x^2=x^3=\mathrm{constant}$ (only one moving star with respect to the ship), then $\D s^2=\mathrm{constant}\ (\D x^4)^2$. When these requirements are fulfilled, the differential $\D x^4$ of the fourth coordinate is directly proportional to the clock separation $\D s$ (see Sections \ref{Clocks-and-line-element} and \ref{Problem-cosmology}). 
95: 
96: Whenever spacetime coordinates are chosen without establishing at the same time their operational meaning (e.g.\ spherical coordinates), these remain physically meaningless until a connection is established between them and a measurable parameter such as $s$. For example, once the metric field is known and the geodesic equation solved, one can find a relation $x^\alpha=x^\alpha(s)$.
97: 
98: Given two sets of coordinates $(x^1,x^2,x^3,x^4)$ and $(y^1,y^2,y^3,y^4)$, we can find analytic transformations $x^\alpha=x^\alpha(y)$ and $y^\alpha=y^\alpha(x)$ that take one set of coordinate values to another ($x$ represents the whole set of coordinate values, and likewise for $y$).
99: 
100: The next concept essential to classical and relativistic theories is that of a (massive or massless) particle. Like events but less dramatically so, particles are also idealized geometrically as points. Unlike events, however, particles are taken to be moving points both in classical and relativistic theories. An event and a particle are connected as follows: the history of a particle is a sequence of events.
101: 
102: In classical physics, the history of a particle is given by the set of equations $\{x=x(t),y=y(t),z=z(t)\}$, where $t$ has the role of a (coordinate) parameter. These three equations determine a particle's classical worldline. In relativistic physics, on the other hand, the history of a particle is given by the set of equations $x^\alpha=x^\alpha(u)$, where $u$ is a one-dimensional (non-coordinate) parameter. These four equations determine a particle's relativistic worldline. Now, if possible and if so wanted, one could solve for $u$ (e.g.\ $u=u(x^4)$) and use this to rewrite the above set of four equations as a set of three, $x^a=x^a(x^4)$ ($a=1,2,3$), but in so doing the separation between space and time has not taken place---there is no reason to take in general $x^4$ as a time coordinate in relativity theory.
103: 
104: In classical physics, a particle of mass $m$ satisfies the equations of motion
105: \begin{equation}
106: m\frac{\D^2 x^a}{\D t^2}=F^a,
107: \end{equation}
108: where the classical force $F^a$ is of the form $F^a(x,\mathrm{d}x/\mathrm{d}t)$. These three equations determine the history of a particle, or its worldline, if one is given (i) an event $(x^1,x^2,x^3;t)$ that belongs to it and (ii) a direction in space-time at the said event, i.e.\ the ratios between the differentials $\mathrm{d}x^a$ and $\mathrm{d}t$. Can this idea of an equation of motion also be applied in relativity theory? We now demand that the history of a massive particle, or its worldline, be determined once one is given (i) an event $(x^1,x^2,x^3,x^4)$ that belongs to it and (ii) a direction in spacetime at the said event, i.e.\ the ratios between the differentials $\mathrm{d}x^\alpha$ and $\D u$. This demand can be met if the differential equations for the worldline of a particle are of the form 
109: \begin{equation}\label{equation-of-motion}
110: \frac{\mathrm{d}^2x^\alpha}{\mathrm{d}u^2}=X^\alpha,	
111: \end{equation}
112: where $X^\alpha=X^\alpha(x,\mathrm{d}x/\mathrm{d}u)$. Here $\mathrm{d}u$ appears only as a differential. What is $X^\alpha$ equal to and what is its physical meaning? We return to this issue later on, when the time is ripe for it.
113: 
114: \subsection{The observer as psychological agent}
115: We saw that physical phenomena are geometrically idealized as point-events, and that matter is geometrically idealized as point-particles. In particular, now a human observer is geometrically idealized as a moving point-particle that makes observations along its worldline. Conventionally, the so-called observer does not need to be human, but is embodied by a set of measuring devices moving along a worldline. However, there exists one essential side to the observer that only a human observer can provide, namely, its subjective experience. According to this experience, events can be laid down in a certain order along the human observer's worldline, who thus creates in his mind the \emph{past}, the \emph{present}, and the \emph{future}.
116: 
117: We saw earlier that none of the four coordinates of a relativistic event $(x^1,x^2,x^3,x^4)$ could in general be isolated as time. The isolation of one of these four coordinates as universal time would unavoidably lead to an absolute ordering of events as in classical physics (classical absolute time), but relativity theory does not claim the existence of a natural order for an arbitrary pair of point-events. When at all possible, a pair of events can only be ordered with respect to the worldline of a given observer---he is actually the one who does the ordering. For those pairs of events that cannot possibly be contained on the observer's worldline due to restrictions of his motion, no ordering relation is possible at all. All these events belong together in the relativistic \emph{four-dimensional present}, whereas those than can be ordered along the observer's worldline belong in the relativistic \emph{four-dimensional past} or \emph{future}. 
118: 
119: Also the classical worldview can be understood from this temporal perspective. As the constraint imposed by the three-dimensional hypersurface limiting the motion of the observer is relaxed, the forbidden four-dimensional present becomes an allowed three-dimensional present, together with a wider four-dimensional past and future (Figure \ref{past-present-future}). Classical physics, then, appears to be a special case of relativistic physics as far as the dimensions of the present are concerned, but, on the other hand, relativistic physics is also more limiting than classical physics, because in it the present is a forbidden area. The restrictive nature of the relativistic present is one of the reasons why classical physics continues to be preferred for most of our daily needs, since in it the allowed present is understood more widely.
120: 
121: \begin{figure}
122: \centering
123: \includegraphics[width=\linewidth]{past-present-future}
124: \caption[Past, present, and future]{In a relativistic world, an
125: observer determines the past, the present, and the future, as he orders events on his worldline into these three categories according to his subjective experience. The present is formed by the set of all events that lie outside the physical boundary imposed by restrictions on his motion. The classical conception of an absolute present arises when the boundary is collapsed onto a three-dimensional plane \mbox{$t=\mathrm{constant}$.} Event $D$ belongs to both the relativistic and classical present, while event $C$ belongs to the relativistic present but to the classical future.} 
126: \label{past-present-future}
127: \end{figure}
128: 
129: \subsection{Clocks and the line element} \label{Clocks-and-line-element}
130: The foundations of classical and relativistic physics can be laid down on the basis of \emph{clock measurements}. Ideally, a clock should progress at a constant pace; this requirement is realized in practice by searching for natural processes with respect to which a great many other natural processes are in synchronicity, thereby simplifying their study. As an ideal clock can only be determined by comparison with other clocks, its basic property may be given as follows: two ideal clocks carried along by an observer along his worldline remain synchronous---their ticking ratio remains constant---for any worldline $\mathcal{C}$ joining any pair of events $A$ and $B$.
131: 
132: Consider now an observer who carries an ideal clock along his worldline $\mathcal{C}$ from point-event $A$ to point-event $B$. At $A$ he records the \emph{clock reading} $s_A$ and on reaching $B$ records the new \emph{clock reading} $s_B$. These clock readings (not time instants) are the results of physical measurements. We now say that the \emph{separation} between events $A$ and $B$ along worldline $\mathcal{C}$ is \mbox{$\Delta s=s_B-s_A$.} When event $B$ is close to event $A$, by which we mean operationally that the observer ``stops his watch quickly after $A$,''\footnote{We may also say that $A$ and $B$ are close if their coordinate values differ by a small amount.} a good approximation to $\Delta s$ is furnished by the separation $\D s$ along a straight worldline element; in symbols, $\Delta s \approx \D s$.
133: 
134: The worldline-element separation $\mathrm{d}s$ as here presented (in terms of clock measurements) gives enough foundation to lead, via different choices, to the concepts of either Newton's classical physics, Einstein's special- and general-relativistic physics, and presumably to any other geometric theory of space and time. Here we shall deal with classical and relativistic theories on the basis of $\mathrm{d}s$. 
135: 
136: It is noteworthy, however, that $\mathrm{d}s$, as a temporal correlation between two events, or two ``nows,'' is not a geometric concept. There is nothing geometric in the (psychological) identification of the present (when the initial and final observations are made), and there is nothing geometric in an event, since, as we saw, these are not originally points but raw phenomena. The activity comprised by a clock measurement and the values it produces, however, \emph{can} lead to a geometric representation. Such a geometrization and its main results are in fact the essence of this chapter. These observations lead then to a provoking thought: what kind of theory may be arrived at by following an opposite, no-geometry direction from $\mathrm{d}s$? What kind of metageometric theory, describing what physics of time, would this be? (Figure~\ref{line-element}).
137: 
138: \begin{figure}
139: \centering
140: \[
141: \begin{psmatrix}[colsep=3.6cm,rowsep=1.2cm]
142:        &               & {\small \mathrm{classical\ physics}} \\
143:      {\small?} & {\small\mathrm{d}s}   & {\small \mathrm{relativistic\ physics}} \\
144:        &               & {\small \mathrm{other\ geometric\ theories}}
145: \psset{arrows=->,labelsep=3pt,nodesep=3pt}
146: \ncline{2,2}{2,1}\ncput*{{\small \mathrm{no\ geometry}}}
147: \ncline{2,2}{2,3}\ncput*{{\small \mathrm{geometry}}}
148: \ncline{2,2}{1,3}\ncput*{{\small \mathrm{geometry}}}
149: \ncline{2,2}{3,3}\ncput*{{\small \mathrm{geometry}}} \end{psmatrix} \]
150: \caption[Line element as starting point]{The line element
151: $\mathrm{d}s$ can act as a starting point for the development of
152: classical and relativistic theories, and presumably for any other
153: geometric theory of space and time, when we choose to geometrize it
154: according to tradition. But what kind of metageometric theory, describing what physics, would one arrive at if one followed the opposite, no-geometry direction?}
155: \label{line-element}
156: \end{figure}
157: 
158: A geometric picture emerges when events become \emph{point}-events, an observer a \emph{point}-particle, an observer's history a world-\emph{line}, $\Delta s$ the \emph{length} of the worldline between the joined point-events, and $\mathrm{d}s$ the length of the differential \emph{arrow} $\D x$ between point-events $A$ at $x$ and $B$ at $x+\mathrm{d}x$, i.e.\ when we start to view $\mathrm{d}s$ in terms of coordinate points and coordinate arrows. In fact, the geometrization of the differential $\D s$ leads us---naturally enough---to differential geometry. What else do we know about $\D s$? How does it depend on the coordinates of point-events $A$ and $B$? 
159: 
160: To move forward, some choices must be made. The main stopovers on the road to Einstein's geometric theory of gravitation have, in fact, already been called into attention in two public lectures: by Riemann in 1854, ``On the foundations which lie at the bases of geometry,'' and by Einstein in 1922, ``How I created the theory of relativity''---it is through their steps that we shall progress. 
161: 
162: The first clue is older than relativity itself and comes from Riemann's famous 1854 lecture. As he sets himself the task of determining the length of a line in a manifold, he writes:
163: \begin{quote} The problem consists then in establishing a
164: mathematical expression for the length of a line\ldots I shall
165: confine myself in the first place to lines in which the ratios of
166: the quantities $\mathrm{d}x$ of the respective variables vary
167: continuously. We may then conceive these lines broken up into
168: elements, within which the ratios of the quantities $\mathrm{d}x$
169: may be regarded as constant; and the problem is then reduced to
170: establishing for each point a general expression for the linear
171: element $\mathrm{d}s$ starting from that point, an expression which
172: will thus contain the quantities $x$ and the quantities
173: $\mathrm{d}x$. I shall suppose, secondly, that the length of the
174: linear element, to the first order, is unaltered when all the points
175: of this element undergo the same infinitesimal displacement, which
176: implies at the same time that if all the quantities $\mathrm{d}x$
177: are increased in the same ratio, the linear element will also vary
178: in the same ratio. On these assumptions, \emph{the linear element may be
179: any homogeneous function of the first degree of the quantities $\mathrm{d}x$}\ldots \cite[p.~5, online ref., italics added]{Riemann:1873}
180: \end{quote} 
181: Following Riemann, we set the line element $\mathrm{d}s$ to be a function of the coordinates of $A$ and $B$,
182: \begin{equation}
183:     \mathrm{d}s=f(x,\mathrm{d}x),
184: \end{equation} 
185: where $f(x,\mathrm{d}x)$ is a homogeneous function of first degree in the differentials $\mathrm{d}x$ , i.e.\ $f(x,k\mathrm{d}x)=kf(x,\mathrm{d}x)$ ($k>0$ can be a function of $x$).\footnote{We have found that also Synge \citeyear{Synge:1964} has understood relativity chronometrically based on the Riemannian hypothesis.}  When the equation of worldline $\mathcal{C}$ is given by $x^\alpha=x^\alpha(u)$, where $u$ is a parameter whose value grows from the past to the future, we can then write
186: \begin{equation}
187:     \mathrm{d}s=f\left(x,\frac{\mathrm{d}x}{\D u}\D u\right)=
188:     f\left(x,\frac{\mathrm{d}x}{\D u}\right)\D u,
189: \end{equation}
190: and therefore the separation $\Delta s$ between $A$ and $B$ is given by
191: \begin{equation}
192:     \Delta s=\int^B_{\begin{smallmatrix}A\\\mathcal{C}
193:     \end{smallmatrix}}\mathrm{d}s=
194:     \int_{u_A}^{u_B}f\left(x,\frac{\mathrm{d}x}{\mathrm{d}u}\right)\mathrm{d}u.
195: \end{equation}
196: 
197: \subsection{Invariance of the line element}\label{sec:Invariance-line-element}
198: Talking about his personal struggle to understand gravitation geometrically, Einstein \citeyear{Einstein:1982} relates during his Kyoto lecture discussing with Marcel Grossmann whether ``the problem could be solved using Riemann theory, in other words, by using the concept of the invariance of the line elements'' (p.~47). This is a key issue, whose study now allows us to move on.
199: 
200: Event separation $\Delta s$ and its differential approximation $\D s$ are the results of physical measurements with a clock made by the observer along his worldline; $\D s$ is an absolute spacetime property of this worldline, namely, its extension between events $A$ and $B$. Because $\D s$ is here meant only in this classical sense, its invariance is a simple matter.\footnote{In the present context, the physical scale (seconds, minutes, hours) or dimension used to express $\D s$ is inconsequential to its invariance. A measurement result expressed as $60\ \mathrm{s}$, $1\ \mathrm{min}$, etc., or in a different dimension, say $2\pi\ \mathrm{rad}$, makes reference to the same objective physical extension---this is something we can metaphorically point at, in the same way that ``book'' and ``Buch'' refer to the same physical object. (Cf.\ comments by Eakins and Jaroszkiewicz \citeyear[p.~10]{Eakins/Jaroszkiewicz:2004}). Beyond the tentative, exploratory remarks at the end of Chapter~\ref{ch:ESE}, a quantum-mechanical discussion of invariance (in particular, that of $\D s$) falls outside the scope of this thesis. For a discussion of quantum experiments involving two different inertial frames, see \cite[pp.~50--54]{Jaroszkiewicz:2007}.} Having proposed that the line element is a function of the coordinates, $\D s=f(x,\D x)$, the \emph{physical} raw material encoded as $\D s$ now demands that its \emph{mathematical} expression $f(x,\D x)$ be an invariant in the mathematical sense of the word. That is, the value of $f(x,\D x)$ must be independent of the set of coordinates chosen by the observer on his worldline, i.e.\ $f(x,\D x)=f'(y,\D y)$, where $y^\alpha=y^\alpha(x)$.
201: 
202: The need to deal with physical invariance leads us to search for mathematical objects that are meaningful regardless of the coordinate system in which they may be expressed. \emph{Tensor equations} (not just tensors) are such objects; if a tensor equation holds in one coordinate system, it holds in any coordinate system. Tensor equations have this property because tensor transformations between coordinate systems are linear and homogeneous.
203: 
204: Events $A$ at $x$ and $B$ at $x+\mathrm{d}x$ determine a differential displacement vector whose components are $\mathrm{d}x^\alpha$ and whose tail is fixed at $A$. In another coordinate system, the components of this differential displacement vector are $\mathrm{d}y^\alpha$ and can be obtained by differentiation of the coordinate transformation $y^\alpha=y^\alpha(x)$. We find
205: \begin{equation}
206: \mathrm{d}y^\alpha=\frac{\partial y^\alpha}{\partial x^\beta}\mathrm{d}x^\beta,
207: \end{equation}
208: which is a linear and homogeneous coordinate transformation of the differential with coefficients $\partial y^\alpha/\partial x^\beta$. The set of differentials $\mathrm{d}x^\alpha$, together with the way they transform from one coordinate system to another, gives us a prototype for a first type of vector (and, in general, tensor) called \emph{contravariant}.\footnote{Although contravariant vectors and tensors are inspired in differential displacement vectors (their components are small), the components of contravariant vectors and tensors do not in general need to be small. For example, the components $\mathrm{d}x^\alpha/\mathrm{d}u$ of the derivative along a worldline represent a unbounded contravariant vector.}
209: 
210: Following the contravariant vector prototype, we can build other sets of magnitudes attached to $A$ that transform in the same way. The magnitudes $T^{\alpha\beta}$, transforming linearly and homogeneously in $T^{\alpha\beta}$ according to
211: \begin{equation}
212:     T'^{\alpha\beta}=\frac{\partial x'^\alpha}{\partial x^\gamma}\frac{\partial x'^\beta}{\partial x^\delta} T^{\gamma\delta},
213: \end{equation}
214: constitute a second-order \emph{contravariant tensor}. And similarly for higher-order contravariant tensors.
215: 
216: A zeroth-order tensor is a scalar that transforms according to the identity relation, $T'=T$, i.e.\ it is an invariant. If we now take a zeroth-order tensor to be a function of the coordinates $f(x)$, we must have that \mbox{$f(x)=f'(x')$.} How does the gradient of $f(x)$ transform when expressed in another coordinate system? Using the chain rule, we have
217: \begin{equation}
218:     \frac{\partial f'(x')}{\partial x'^\alpha}=
219:     \frac{\partial f(x)}{\partial x'^\alpha}=
220:     \frac{\partial x^\beta}{\partial x'^\alpha} \frac{\partial f(x)}{\partial x^\beta}.
221: \end{equation}
222: It turns out that a gradient transforms in a manner opposite to that of contravariant vectors (and tensors). The gradient then serves as a prototype for a second type of vector (and, in general, tensor) called \emph{covariant}. Historically, this name actually comes from a shortening and reshuffling of ``variation co-gradient,'' i.e.\ variation in the same manner as the gradient, and the name contravariant comes from ``variation contra-gradient,'' i.e.\ variation against the manner of the gradient.
223: 
224: Like before, following the covariant-gradient prototype, we can have other sets of magnitudes attached to $A$ that transform in the same way. The magnitudes $T_ {\alpha\beta}$, transforming linearly and homogeneously in $T_{\alpha\beta}$ according to
225: \begin{equation}
226:     T'_{\alpha\beta}=\frac{\partial x^\gamma}{\partial x'^\alpha}\frac{\partial x^\delta}{\partial x'^\beta} T_{\gamma\delta},
227: \end{equation}
228: constitute a second-order \emph{covariant tensor}. And similarly for higher-order covariant tensors, and also for mixtures of both types. Finally, when a tensor is attached not only to a single event, but to every event along a worldline or to every event in the four-dimensional spacetime continuum, we have a spacetime field called \emph{tensor field}.
229: 
230: \subsection{The Newtonian line element}
231: So far, we know two chief properties of $\D s$. One is that it represents a classical measurement of an absolute spacetime property and is therefore a physical invariant; this means that $f(x,\D x)$ must be a mathematical invariant. The other is that $f(x,\D x)$ is homogeneous with respect to all components $\D x^\alpha$. Within these requirements there is still freedom to choose the form of $f(x,\D x)$, but what choices lead to physically relevant results? We proceed in what one may call a natural way. Let us try to separate the dependence of $f(x,\D x)$ on $x$ and $\D x$ using the mathematical invariance and homogeneity of this function. Both requirements can be fulfilled by considering an \emph{inner product}. 
232: 
233: We propose as a first case to be analysed an inner product of two
234: vectors, a \emph{linear form} 
235: \begin{equation}\label{linear-form}
236:     \D s=f(x,\D x)=g_\alpha(x)\D x^\alpha.
237: \end{equation}
238: This function is clearly homogeneous of first degree with respect to $\D x^\alpha$ and produces an invariant number. Moreover, because $g_\alpha(x)\D x^\alpha$ is invariant and $\D x^\alpha$ is a general contravariant vector, we must have that $g_\alpha(x)$ is a covariant vector.\footnote{The invariance of the inner product means that $g'_\alpha\D x'^\alpha=g_\beta \D x^\beta$, and the fact that $\D x^\alpha$ is a contravariant vector means that it transforms as $\D x'^\alpha=(\partial x'^\alpha/\partial x^\beta)\D x^\beta$. Together these two relations lead to $[(\partial x'^\alpha/\partial x^\beta)g'_\alpha - g_\beta]\D x^\beta=0$. On account of $\D x^\beta$ being an arbitrary contravariant vector, the bracketed expression must be null, which means that $g_\alpha$ transforms as a covariant vector.} But what is the physical meaning of $g_\alpha(x)$?
239: 
240: The simplest option is to choose the covariant vector after the covariant prototype,  
241: \begin{equation}\label{covariant-prototype}
242: g_\alpha(x)=\frac{\partial t(x)}{\partial x^\alpha},
243: \end{equation}
244: where $t(x)$ is an invariant function of the spacetime coordinates $x$. The separation between events $A$ and $B$ becomes
245: \begin{equation}
246:     s_B-s_A =
247:     \int_A^B \D s = \int^B_{\begin{smallmatrix}A\\\mathcal{C}
248:     \end{smallmatrix}} \frac{\partial t(x)}{\partial x^\alpha}\D x^\alpha = \int^B_{\begin{smallmatrix}A\\\mathcal{C}
249:     \end{smallmatrix}} \D t(x) = t_B-t_A.
250: \end{equation}
251: In other words, because $\D t(x)$ is an exact differential, the separation between events $A$ and $B$ does not depend on the worldline $\mathcal{C}$ travelled. Because we know experimentally that this result is false, we can conclude that there must be something wrong with our choice of Eq.~(\ref{covariant-prototype}), and presumably even with the linear form (\ref{linear-form}). On the positive side, we learn that this choice leads to the interpretation of $s$ as \emph{absolute classical time} $t$. This is one of the essential differences between classical and relativistic physics; in the latter, $\D s$ is not an exact differential and the separation $\Delta s$ between events depends on the worldline travelled.
252: 
253: \subsection{The Riemann-Einstein line element}
254: The second simplest choice for $f(x,\D x)$ satisfying the same two requirements set above consists of an inner product of a second-order covariant tensor and two differential displacement vectors, that is, a \emph{quadratic form} 
255: \begin{equation}\label{line-element-eq}
256:     \D s= f(x,\D x)=\sqrt{-g_{\alpha\beta}(x)\D x^\alpha \D x^\beta}.
257: \end{equation}
258: The square root is necessary to meet the requirement of homogeneity of first degree with respect to the differential displacements, whereas the reason for the minus sign will become apparent later on.
259: 
260: On the basis of a similar procedure as above (see previous footnote), we can now deduce something about the tensorial nature of $g_{\alpha\beta}(x)$, but not as much as before. Although $\D x^\alpha$ is a general contravariant tensor, since it appears in the symmetric product $\D x^\alpha \D x^\beta$, we can now only deduce that $g_{\alpha\beta}+g_{\beta\alpha}$ must be a second-order covariant tensor, but we cannot conclude anything about the tensorial nature of $g_{\alpha\beta}$ itself on the basis of the invariance of the inner product. However, if we assume that $g_{\alpha\beta}(x)$ is symmetric---and this we shall do---then we can conclude that $g_{\alpha\beta}(x)$ is a second-order covariant tensor.\footnote{Einstein considered the possibility of an asymmetric metric tensor in the latter part of his life in the context of attempts at a unified theory of electromagnetism and gravity.}
261: 
262: Given then a symmetric second-order covariant tensor $g_{\alpha\beta}(x)$ that is, moreover, non-singular ($\mathrm{det}(g_{\alpha\beta})\neq 0$), the invariant quadratic form 
263: \begin{equation}
264: Q=g_{\alpha\beta}(x)\D x^\alpha \D x^\beta	
265: \end{equation}
266: determines a four-dimensional continuum called \emph{Riemann-Einstein spacetime}. The line element (\ref{line-element-eq}) of Riemann-Einstein spacetime, which is based on this quadratic form, is the foundation of relativity theory.
267: 
268: A challenging question, however, remains: what is the physical meaning of $g_{\alpha\beta}(x)$? To sketch a solution, we now need to take a look at the foundations of Riemannian metric geometry. In connection with the search for a new description of gravitation, the necessity of this stopover was related by Einstein during his Kyoto lecture:
269: \begin{quote}
270: What should we look for to describe our problem? This problem was unsolved  until 1912, when I hit upon the idea that the surface theory of Karl Friedrich Gauss might be the key to this mystery. I found that Gauss' surface coordinates were very meaningful for understanding this problem. Until then I did not know that Bernhard Riemann [who was a student of Gauss] had discussed the foundation of geometry deeply\ldots I found that the foundations of geometry had deep physical meaning. \cite[p.~47]{Einstein:1982}
271: \end{quote}
272: 
273: \subsection{From Gaussian surfaces to Riemannian space}
274: Metric geometry---geometry that deals essentially with magnitudes---has length as its basic concept. In Euclidean metric geometry, this concept always refers to the finite distance between two space points. In Riemannian metric geometry, on the other hand, it refers to the distance between two very (not necessarily infinitesimally) close points. In other words, the passage from finite Euclidean geometry to differential Riemannian geometry is achieved by approximating the finite distance $\Delta l$ between two points with the differential distance $\D l$.
275: 
276: We start by examining Cartesian coordinates (straight lines meeting at right angles) in a three-dimensional Euclidean space. On the basis of Pythagoras' theorem, the squared distance between the points $(y^1,y^2,y^3)$ and $(y^1+\D y^1,y^2+\D y^2,y^3+\D y^3)$ is given by 
277: \begin{equation}
278:       \D l^2=(\D y^1)^2+(\D y^2)^2+(\D y^3)^2.	
279: \end{equation}
280: In terms of general coordinates $x^1,x^2,x^3$ (curved lines meeting at oblique angles), because $\D y^i=(\partial y^i/\partial x^a)\D x^a$ ($i,a=1,2,3)$, we obtain 
281: \begin{equation}
282:       \D l^2=h_{ab}(x)\D x^a \D x^b,
283: \end{equation}
284: where
285: \begin{equation}
286:     h_{ab}(x)=\sum_{i=1}^3 \frac{\partial y^i}{\partial x^a}
287:     \frac{\partial y^i}{\partial x^b}
288: \end{equation}
289: is symmetric with respect to $a$ and $b$ by inspection. Again, because $h_{ab}(x)$ is assuredly symmetric, the squared length $\D l^2$ is an invariant, and $\D x^a$ is a general vector, we can conclude (based on the results of the previous section) that $h_{ab}(x)$ is a (symmetric) second-order covariant tensor.\footnote{If $h_{ab}(x)=0$ when $a\neq b$, the coordinate lines meet at right angles. If, in addition, $h_{ab}(x)=1$ when $a=b$, the coordinates are Cartesian.} This tensor determines the metric structure, or metric, of three-dimensional Euclidean space, and is therefore called the \emph{metric tensor}.
290: 
291: The next step towards Riemannian space is to consider with Gauss a two-dimensional \emph{surface} $y^i=y^i(x^1,x^2)$ ($i=1,2,3$), where $x^1$ and $x^2$ are parameters, \emph{embedded} in a three-dimensional Euclidean space. Now the squared distance between neighbouring points on the surface is given by 
292: \begin{equation}
293: \D l^2=(\D y^1)^2+(\D y^2)^2+(\D y^3)^2=\gamma_{ab}(x)\D x^a \D x^b,	
294: \end{equation}
295: where
296: \begin{equation}
297:     \gamma_{ab}(x)=\sum_{i=1}^3 \frac{\partial y^i}{\partial x^a}
298:     \frac{\partial y^i}{\partial x^b} \qquad (a,b=1,2).
299: \end{equation}
300: This symmetric second-order covariant tensor is called the \emph{surface-induced metric}, i.e.\ the metric that the surrounding three-dimensional Euclidean space induces on the surface. What happens if we now \emph{remove} the space surrounding the surface altogether?
301: 
302: Although it was Gauss who first developed the idea of a surface embedded in a higher-dimensional Euclidean space, it was Riemann who took the seemingly nonsensical step of removing the embedding space and taking what remained seriously: a two-dimensional \emph{non-Euclidean space}, a two-dimensional continuum determined by the invariant quadratic form 
303: \begin{equation}\label{two-dim-Q}
304: Q=\gamma_{ab}(x)\D x^a \D x^b \qquad (a,b=1,2),	
305: \end{equation}
306: where $x^1$ and $x^2$ are Gauss's surface coordinates and $\gamma_{ab}(x)$ is a second-order covariant tensor. We assume this new tensor inherits its metric role in this new space from Gauss's case, and that it, too, is symmetric. What remains, then, is a two-dimensional Riemannian \emph{space}, a space one can inhabit but not look at from the outside because it has no outside. For example, what is normally thought as a spherical surface is, in fact, only a Gaussian picture (embedded in Euclidean space) of what is truly a non-Euclidean elliptical plane.
307: 
308: If $\gamma_{ab}(x)$ is the (symmetric) metric tensor of two-dimensional Riemannian space, does this mean that its quadratic form gives a meaningful idea of the squared distance between two neighbouring points of this space? As it stands, it does not, because there is no reason to assume that the quadratic form (\ref{two-dim-Q}) is positive definite,\footnote{The quadratic form is positive definite if it is always positive except when all the components of the differential displacement $\D x^a$ are null, i.e.\ when the two points in question coincide with each other.} as it was for Euclidean space and Gaussian surfaces embedded in it. In Riemannian space, the quadratic form can be negative for some differential displacements $\D x^a$ (i.e.\ in certain directions from a point), positive for others, and null for others; in consequence, the concept of distance must be introduced separately. This can be done with the help of an indicator $\epsilon$, which can have the values $+1$ or $-1$, such that 
309: \begin{equation}
310: \D l^2=\epsilon \gamma_{ab}(x)\D x^a \D x^b	\geq 0
311: \end{equation}
312: for all $\D x^a$. The equality holds for a null displacement, but can also hold for a non-null one, i.e.\ the distance between two points can be null even when they are not coincident.
313: 
314: Indicators can also be used in a different way. The coordinates can be so chosen that the quadratic form $\gamma_{ab}(x)\D x^a \D x^b$ can be brought into the simpler one
315: \begin{equation}
316: Q=\epsilon_1(\D x'^1)^2+\epsilon_2(\D x'^2)^2, 	
317: \end{equation}
318: \emph{at one point at a time} but not globally.\footnote{To do this, write $\gamma_{ab}(x)\D x^a \D x^b$ ($a=1,2$) in explicit form, complete squares, and make the change of variables $ x'^1=\sqrt{\epsilon_1\gamma_{11}}x^1+(\gamma_{12}/\sqrt{\epsilon_1\gamma_{11}}x^2)$, $x'^2=\sqrt{\epsilon_2[\gamma_{22}-(\gamma_{12})^2/\gamma_{11}]}x^2$ to obtain $\epsilon_1(\D x'^1)^2+\epsilon_2(\D x'^2)^2$. The quadratic form thus obtained is valid only at a point at a time, because for this method to work $\gamma_{ab}(x)$ must be kept constant. The indicators should be chosen so as to avoid the appearance of imaginary roots, thus keeping the transformations real.} For a Riemannian space with a positive-definite quadratic form, we have the case $\epsilon_1=\epsilon_2=+1$, leading to a locally Cartesian quadratic form; it  describes a space whose geometry is locally Euclidean. For a Riemannian space with an indefinite quadratic form, we have the rest of the cases, leading to a pseudo-Cartesian quadratic form; they describe a space whose geometry is locally pseudo-Euclidean. The pair $(\epsilon_1,\epsilon_2)$ is called the space signature and it is an invariant, i.e.\ it cannot be changed by means of a (real) coordinate transformation.
319: 
320: With the introduction of a stand-alone two-dimensional Riemannian space, we have banished three-dimensional Euclidean space. But as we next add a dimension to take into account a three-dimensional space, should Euclidean space be resurrected again? If we want to conserve the chance of describing new physics through the idea of Riemannian space, we should not. We should now consider a three-dimensional Riemannian space (i) which is not embedded in any higher-dimensional Euclidean space, (ii) whose geometry is by nature non-Euclidean, and (iii) which is determined by the invariant indefinite quadratic form $h_{ab}(x)\D x^a \D x^b$ ($a,b=1,2,3$). Here $h_{ab}(x)$ is a symmetric second-order covariant metric tensor, and the quadratic form in question is locally Cartesian or locally pseudo-Cartesian, i.e.\ locally of the form 
321: \begin{equation}
322: Q=\epsilon_1(\D x^1)^2+\epsilon_2(\D x^2)^2+\epsilon_3(\D x^3)^2,	
323: \end{equation}
324: with $\epsilon_1,\epsilon_2,\epsilon_3$ each being $+1$ or $-1$. Finally, the concept of distance is attached to this three-dimensional space via the use of indicator $\epsilon$ as before:\footnote{In physical spacetime, $\D l^2$ is not an invariant, but the spatial part of the invariant $\D s^2$ (when it is possible to make the space-time split).} 
325: \begin{equation}
326: \D l^2=\epsilon h_{ab}(x)\D x^a \D x^b.
327: \end{equation}
328: 
329: \section{Riemann-Einstein physical spacetime}
330: 
331: \subsection{Towards local physical geometry}
332: Having studied Gaussian surfaces and Riemannian space to suit our needs, we now turn to physical spacetime, resuming where we left earlier.
333: 
334: We proposed the form of the line element $\D s$ of Riemann-Einstein spacetime in Eq.~(\ref{line-element-eq}) to be
335: \begin{equation}
336: \D s=\sqrt{-g_{\alpha\beta}(x)\D x^\alpha \D x^\beta}. \nonumber	
337: \end{equation}
338: We can now make physical sense of this expression, including an elucidation of the minus sign, by approaching the line element from the perspective of Riemannian geometry. Because $\D s$ represents clock-reading separations and these are positive real numbers, $\D s$ is always real and positive. Its square $\D s^2$ must then also be real and positive.\footnote{The case of a photon is dealt with later.} On the other hand, we saw that the quadratic form $g_{\alpha\beta}(x)\D x^\alpha \D x^\beta$ of a Riemannian (now four-dimensional) space is not positive definite. However, we learnt how to solve this problem (in trying to get a non-negative length $\D l^2$) using an indicator $\epsilon$. It is then natural to propose
339: \begin{equation}\label{squared-line-element}
340: \D s^2=\epsilon g_{\alpha\beta}(x)\D x^\alpha \D x^\beta
341: \end{equation}
342: in order to ensure that $\D s^2$ is never negative.
343: 
344: As an aside, we also find the geometric meaning of $g_{\alpha\beta}(x)$: $g_{\alpha\beta}(x)$ is the \emph{metric tensor}, or metric, of Riemann-Einstein four-dimensional spacetime. As a Riemannian metric, $g_{\alpha\beta}$ links contravariant and covariant vectors, $T^\alpha$ and $T_\alpha$, which have otherwise no connection; the invariant inner product between two such vectors $T^\alpha T_\alpha$ is the same as the invariant quadratic form $g_{\alpha\beta}T^\alpha T^\beta$, whereby we obtain the identification 
345: \begin{equation}
346: g_{\alpha\beta}T^\beta=:T_\alpha.	
347: \end{equation}
348: We conclude that, from the point of view of Riemannian geometry, contravariant and covariant vectors (and, in general, tensors) are essentially the same. 
349: 
350: This interpretation of $g_{\alpha\beta}(x)$ comes as an anticlimactic answer to the challenging question we posed earlier, for at the back of our minds we already knew that $g_{\alpha\beta}(x)$ was the metric of spacetime. This mathematical, geometric meaning, then, is not really what we were pursuing when we asked for the meaning of $g_{\alpha\beta}(x)$; it is a more physical meaning, if any, that we are still after.
351: 
352: As we now compare Eq.~(\ref{line-element-eq}) with Eq.~(\ref{squared-line-element}), we must conclude that for any observer moving along his worldline and actually measuring the \emph{positive} real-number separation between two events with his clock, $\epsilon$ must be equal to $-1$. In other words, an observer with a clock can only move in directions determined by differential displacements $\D x^\alpha$ such that $g_{\alpha\beta}(x)\D x^\alpha \D x^\beta$ is negative; other directions are physically forbidden.
353: 
354: According to the direction of a differential displacement $\D x^\alpha$, and in keeping with the usual names, we can now say that\footnote{The same classification can be generalized to include any contravariant vector $T^\alpha$. An example of another timelike (i.e.\ allowed) contravariant vector is the tangent vector $\D x^\alpha/\D u$ to a worldline.}
355: \begin{itemize}
356: 	\item[(i)] $\D x^\alpha$ is timelike (i.e.\ allowed) if $g_{\alpha\beta}(x)\D x^\alpha \D x^\beta < 0$ ($\epsilon = -1)$,
357: 	\item[(ii)] $\D x^\alpha$ is spacelike (i.e.\ forbidden) if $g_{\alpha\beta}(x)\D x^\alpha \D x^\beta > 0$ ($\epsilon = +1)$, and 
358: 	\item[(iii)]$\D x^\alpha$ is lightlike (i.e.\ allowed only for photons) if $g_{\alpha\beta}(x)\D x^\alpha \D x^\beta = 0$.
359: \end{itemize}
360: Regarding item (iii), in relativity theory the hypothesis is made that the differential displacements characterizing the history of a photon satisfy the relation $\D s=0$. Strictly speaking, this hypothesis is outside the boundaries of the operational meaning of $\D s$, since the clock to be carried along by a photon to make the necessary measurements is a material object which can never reach the speed of light. One may, however, understand this photon-history hypothesis as the limiting case of increasing velocities of material bodies (see Section \ref{absolute-derivative-sec}).
361: 
362: Because Riemann-Einstein spacetime is locally pseudo-Cartesian, coordinates can be found such that its indefinite quadratic form takes the form 
363: \begin{equation}
364: Q=\epsilon_1(\D x^1)^2+\epsilon_2(\D x^2)^2+\epsilon_3(\D x^3)^2+\epsilon_4(\D x^4)^2	
365: \end{equation}
366: at one point at a time. What are all the possible different combinations of $\epsilon_i$? Because no separation between three space coordinates and one time coordinate has as yet been made---and any such separation would at this point be artificial---all four coordinates stand on an equal footing. It is then immaterial the signs of which $\epsilon_i$ are being changed in any one combination. The five different possible combinations are listed in Table \ref{epsilon-table}.
367: 
368: \begin{table}
369: \centering
370: \begin{tabular}{|c|c|c|c|c|}
371: \hline
372: Signature & $\epsilon_1$ & $\epsilon_2$ & $\epsilon_3$ & $\epsilon_4$ \\
373: \hline \hline
374: $(+,+,+,+)$  & $+1$ & $+1$ & $+1$ & $+1$ \\ \hline
375: $(+,+,+,-)$  & $+1$ & $+1$ & $+1$ & $-1$ \\ \hline
376: $(+,+,-,-)$  & $+1$ & $+1$ & $-1$ & $-1$ \\ \hline
377: $(+,-,-,-)$  & $+1$ & $-1$ & $-1$ & $-1$ \\ \hline
378: $(-,-,-,-)$  & $-1$ & $-1$ & $-1$ & $-1$ \\ \hline
379: \end{tabular}
380: \caption[Choices of local spacetime geometry]{Five different possibilities for the signature of Riemann-Einstein spacetime. Each of these choices leads to a different local physical geometry for spacetime.}
381: \label{epsilon-table}
382: \end{table}
383: 
384: Which of the five possibilities should we choose? And on what grounds? No answer makes itself naturally available to us---we have reached a roadblock. \emph{This is a critical stage of this analysis}. It is time to recapitulate.
385: 
386: \subsection{Local physical geometry from psychology}
387: The present analysis of time has so far been based, first and foremost, on a \emph{physical component} consisting of clock measurements $\D s$; and secondly, on the other side of the equation, on a \emph{mathematical component} consisting of the geometric description of spacetime via its characteristic invariant quadratic form $g_{\alpha\beta}(x)\D x^\alpha \D x^\beta$. Is there anything missing from this analysis of time? Should we consider something else besides physics and mathematics? We must; if we are to move forwards in a natural, non-axiomatic way, we are forced to also take into account a \emph{psychological component} regarding the conception of time: the human notions of past, present (or being), and future.
388: 
389: The past and the future include all the allowed directions that an observer can follow, and so, recalling that we can only measure positive values of $\D s$ and that Eq.~(\ref{line-element-eq}) carries a minus sign, we have that
390: \begin{itemize}
391: 	\item[(i)] $g_{\alpha\beta}(x)\D x^\alpha \D x^\beta < 0$ ($\epsilon=-1$) for the past and the future, 
392: 	\item[(ii)] $g_{\alpha\beta}(x)\D x^\alpha \D x^\beta > 0$ ($\epsilon=+1$) for the present, and 
393: 	\item[(iii)] $g_{\alpha\beta}(x)\D x^\alpha \D x^\beta = 0$ for the boundary surface between the present and the past and future, i.e.\ the lightcone (see below).
394: \end{itemize}
395: 
396: What choice of indicators $\epsilon_i$ do the conceptions of past, present, and future lead to? Choices $(+,+,+,+)$ and $(-,-,-,-)$ can be discarded as soon as we notice that they lead to quadratic forms of the type
397: \begin{equation}
398: Q=\pm[(\D x^1)^2+(\D x^2)^2+(\D x^3)^2+(\D x^4)^2],	
399: \end{equation}
400: which are always positive and do not allow the conceptions of the past and the future, or always negative and do not allow the conception of the present. Of the remaining options, which should we choose? And again, on what grounds?
401: 
402: While the existence of the present on the one hand, and the past and future on the other, acted as a first guideline, it is now the distinction between the past and the future that acts as the next one. We perceive the past becoming the future as a serial, unidirectional, or one-dimensional affair through the present, and so it appears that this process of becoming should be pictured locally by means of the progression of a single coordinate parameter and not of several of them. Given one coordinate parameter $x$ that we may locally associate with the psychological conception of the unidirectional flow of time, we may then connect the past with decreasing values of $x$ and the future with increasing values of $x$. 
403: 
404: Because we associated the past and the future with negative values of the invariant quadratic form, we may now discard options $(+,+,-,-)$ and $(+,-,-,-)$, which introduce, respectively, two and three (coordinate) parameters locally associated with the past and the future (prefixed with negative signs)---a so-called embarrassment of riches. Why? Because there do not exist two or three different ways in which we feel it is possible to go from the past to the future---e.g.\ is there a way through the present but perhaps also another that shortcuts it, as if past and future shared one or more direct connections? This idea does not conform with the seriality of the psychological conception of time, and we therefore discard these two options. We are left with option $(+,+,+,-)$, which we adopt.\footnote{If we had not prefixed Eq.~(\ref{line-element-eq}) with a minus sign, we would have been forced to choose option $(+,-,-,-)$, with the positive coordinate now being associated with the past and the future. This would have been a workable alternative too, although we would now have to carry along three minus signs instead of one.}
405: 
406: The choice of the invariant quadratic form with signature $(+,+,+,-)$ leads to the following classification:
407: \begin{itemize}
408: 	\item[(i)] $(\D x^1)^2+(\D x^2)^2+(\D x^3)^2-(\D x^4)^2 < 0$ for the past and future,
409: 	\item[(ii)] $(\D x^1)^2+(\D x^2)^2+(\D x^3)^2-(\D x^4)^2 > 0$ for the present, and 
410: 	\item[(iii)] $(\D x^1)^2+(\D x^2)^2+(\D x^3)^2-(\D x^4)^2 = 0$ for the lightcone.
411: \end{itemize}
412: Moreover, through the increase and decrease of coordinate $x^4$, we can locally differentiate the past from the future, thus: 
413: \begin{itemize}
414: 	\item[(i-a)] $(\D x^1)^2+(\D x^2)^2+(\D x^3)^2-(\D x^4)^2 < 0 \quad \& \quad \D x^4<0$ for the past, and
415: 	\item[(i-b)] $(\D x^1)^2+(\D x^2)^2+(\D x^3)^2-(\D x^4)^2 < 0 \quad \& \quad \D x^4>0$ for the future.
416: \end{itemize}
417: In addition, we can also identify 
418: \begin{itemize}
419: 	\item[(iii-a)] $(\D x^1)^2+(\D x^2)^2+(\D x^3)^2-(\D x^4)^2 = 0 \quad \&\quad \D x^4 < 0$ with the branch of the lightcone that differentiates the past from the present, and 
420:       \item[(iii-b)] $(\D x^1)^2+(\D x^2)^2+(\D x^3)^2-(\D x^4)^2 = 0 \quad \& \quad \D x^4 > 0$ with the branch of the lightcone that differentiates the future from the present.
421: \end{itemize}
422: 
423: Now, why should physics take into account these aspects of the psychological notion of time? Why should physics be partly based on them? Should physics also take into account the properties of other psychological conceptions, such as hunger, happiness, pain, and disdain? Not really, not in any specific way. What is then special about time? It is this: the psychological notion of time as a stream that drags us from past to future through the ubiquitous present is the deepest rooted feature of human consciousness and, as a result, physics has since its beginnings described the changing world guided by this form of human experience, yet without succeeding (or even trying) to explain it. From Galilean dynamics to quantum gravity, a picture of time as ``$t$''is all physics has so far managed: an abstract mathematical parameter that by itself cannot tell us why the past is gone, the present all-pervading, and the future unknown and out of reach.  
424: 
425: In classical physics, the mathematical depiction of time as ``$t$'' is simple enough for the psychological property of things-in-flow to be superimposed on it in a straightforward way: past, present, and future are taken as universal notions, each associated with different positions along the absolute parametric timeline. This is not to say that the human experience of time was \emph{explained} by $t$, but that it was easy enough to project ourselves onto that timeline and imagine ourselves to flow along it. Classically, then, as well as in conventional expositions of relativity theory with its block-universe picture, the human categories of past, present, and future remain physically ignored.
426: 
427: In this presentation of relativity, however, we notice that in this theory an allowance must be made for human temporal notions in a somewhat more explicit way. On the one hand, from the subjective point of view of the ordering of events along an observer's worldline, we again find that past, present, and future are ignored by physics, if we now replace the absolute parametric timeline given by $t$ by the relative parametric timeline given by clock readings $s$. On the other hand, however, relativistic physics is not just about the history of \emph{one} observer's worldline; it is a description of any and every worldline. In other words, relativistic physics is about \emph{absolute spacetime properties}, and we find that in the construction of the absolute property of spacetime signature $(+,+,+,-)$, it needs to explicitly acknowledge, though minimally, the psychological conceptions of past, present, and future for the problem to be unlocked in a meaningful way. 
428: 
429: The sceptical reader, having heard that physics ought to be independent of humans and external to them, may ask: but if the conceptions of past, present, and future are psychological, why make them a part of physics? Or then, if they are part of physics, why stress they are of psychological origin? Because we want psychology, \emph{where relevant}, to inform physics, so that physics, \emph{where possible}, will be reconciled with psychology. The pursuit of this twofold connection, which one may call \emph{psychophysics}, requires that we seriously acknowledge the factualness of conscious experience and that we identify those features of it that are relevant to physics. In order to do this, we must start by taking ourselves seriously and not dismiss, in the pursuit of the ideal of ``objectivity,'' human experience as an irrelevant delusion.
430: 
431: When it comes to time, that we should take ourselves seriously means that we should not consider ourselves the victims of an illusion in which \emph{we feel} that time flows irreversibly, with the past known and irretrievable, the present a permanent and all-pervading ``now,'' and the future unknown and out of reach, while \emph{in fact} time is reversible because the deterministic equations of physics make no dynamical distinction between $t$ and $-t$, and past, present, and future are all on an equal footing because spacetime theories lay them out as a homogeneous block. If physics says the earthly carousel and the heavenly bodies in their orbits can equally well run backwards in time, but our raw experience of time says that this is impossible, which should we trust? Is the physics of time correct as it is and we delusional, or could we learn a better physics of time by taking ourselves more seriously? 
432: 
433: That we should do so is not a forgone conclusion but an attitude we propose. Against such an attitude, one may observe with Dennett \citeyear{Dennett:1991} that ``nature doesn't build epistemic engines'' (p.~382). That is, why should we trust our minds to capture something essential about the nature of time that physical theories have not yet grasped, if our minds did not arise to know the world but only to improve our chances of staying alive? Regarding colour perception, for example, the context in which Dennett makes his observation, this is quite true; the perception of red does not represent any basic property of the world. But when it comes to time, we are inclined to give the mind the benefit of the doubt. Unlike any other feeling, time is an inalienable experiential product of human consciousness and, while physics desperately needs \emph{some} concept of time, so far it has only managed to offer a caricature of it, grasping barely its serial aspect in a mathematical parameter and no more. Because the conception of time is a consciousness-related idea, to understand it, we need to pay heed to ourselves.
434: 
435: This presentation of relativity takes a small step in a positive direction. Through it we learn that, despite appearances, relativity does not straightout hold the spatiotemporal universe to be a completely undifferentiated block, because in its choice of local physical geometry (or global signature) there is a mild recognition that (i) past and future are different from the present (first discard of signatures), and (ii) echoing ``$t$'' in classical physics, that past and future are serially connected (second discard of signatures).
436: 
437: However, this understanding of relativity does not achieve nearly enough. How are past, present, and future qualitatively different from each other? Why is past-present-future a serial chain? And, more significantly, what about the present, that permanent now in which we constantly find ourselves? How does physics explain the \emph{present moment}? Can it single it as special in any way? Our consciousness does, and so distinctly so that we feel we are nothing but the present. If we decide to heed ourselves when it comes to time, we should consider this an enigma. An enigma neither classical nor relativistic physics can solve. But what about quantum physics? Can it succeed where its predecessor theories fail? Can quantum physics do better in accounting for what is otherwise the human delusion that the present is paramount? And what is the role of geometric tools of thought in this search? We return to this issue in Chapter~\ref{ch:MQM}.
438: 
439: \subsection{Spacelike separation and orthogonality from clocks}
440: Up to now, we have used an ideal clock to measure separations between events that can be reached by an observer travelling along his worldline, i.e.\ the separations measured have always been in directions given by timelike displacement vectors $\D x^\alpha$. Can we also measure separations between events along directions given by spacelike displacement vectors? And if so, what kind of experimental setup do we need to do this?
441: 
442: We examine two neighbouring events $A$ and $B$ joined by a spacelike displacement vector $\D x_{AB}$, with $A$ situated between events $P$ and $Q$ and all three events on the observer's timelike worldline. Events $P$ and $Q$ are determined experimentally by the application of the radar method: $P$ is an event such that a light signal (a photon) emitted from it is reflected at $B$ and received by the observer at $Q$; both timelike separations $\D s_{PA}$ and $\D s_{AQ}$ are experimentally determined with the observer's clock. Looking at the diagram of Figure \ref{spacelike-separation}, we can write down the system of equations
443: \begin{equation}\left\{ \begin{array}{rclc}
444: 	   \D x_{AQ}&=&k \D x_{PA}& \qquad (k>0) \\
445:    \D x_{PB}-\D x_{PA}&=&\D x_{AB}&\\
446:    \D x_{AQ}-\D x_{AB}&=&\D x_{BQ}& 
447:    \end{array}\right..
448: \end{equation}
449: Using the last two equations to express the lightlike
450: separations $\D s^2_{PB}=0$ and $\D s^2_{BQ}=0$ in terms of the two
451: timelike and one spacelike displacement vectors, and subsequently
452: using the first equation, we find 
453: \begin{equation}
454: \D s^2_{AB}=\D s_{PA}\D s_{AQ}.\footnote{In more detail, we have (1) $\D s^2_{PB}=-\D s^2_{PA}+\D s^2_{AB}+2g_{\alpha\beta}\D x^\alpha_{PA} d x^\beta_{AB}=0$ and (2) $\D s^2_{BQ}=-k^2\D s^2_{PA}+\D s^2_{AB}-2g_{\alpha\beta}(k\D x^\alpha_{PA}) d x^\beta_{AB}=0$. Now $k$(1)+(2) gives $\D s^2_{AB}=k\D s^2_{PA}$, from which follows
455: the desired result.}	
456: \end{equation}
457: Spacelike separations can then be measured with
458: the simple addition of emission, reflection, and reception of
459: photons to the original clock-on-worldline type of measurements, i.e.\
460: spacelike separations can be expressed in terms of timelike ones.
461: 
462: \begin{figure}
463: \centering
464: \includegraphics[width=70mm]{spacelike-separation}
465: \caption{Measurement of a spacelike separation by means of a clock
466: and a reflected photon. Displacement vectors can be detached from
467: their original positions due to the fact that they are small.}
468: \label{spacelike-separation}
469: \end{figure}
470: 
471: We find, moreover, that if events $A$, $B$, $P$, and $Q$ can be experimentally so arranged as to have $\D s_{PA}=\D s_{AQ}$, then
472: the $k$-factor is unity and $\D x_{AQ}=\D x_{PA}$. In that case, $\D s^2_{BQ}=0$ becomes the bilinear form 
473: \begin{equation}\label{orthogonality-eq}
474: g_{\alpha\beta}\D x^\alpha_{AB}\D x^\beta_{AQ}=0,	
475: \end{equation}
476: where $\D x^\alpha_{AB}$ is spacelike and $\D x^\alpha_{AQ}$ is timelike. What is the geometric meaning of this result? 
477: 
478: We know that, in Euclidean space, whose quadratic form is positive definite, the cosine of the angle between vectors $\overrightarrow{AB}$ and $\overrightarrow{AQ}$ is given by the dot product of the corresponding unit vectors:
479: \begin{equation}
480: \cos(\theta)=\hat u_{AB}\cdot \hat u_{AQ}=u_{AB}^1u_{AQ}^1+u_{AB}^2
481: u_{AQ}^2+u_{AB}^3 u_{AQ}^3.	
482: \end{equation}
483: Generalizing for a Riemannian space with a \emph{positive-definite} quadratic form, we have 
484: \begin{equation}
485: \cos(\theta)=h_{ab}\left(\frac{\D x^a_{AB}}{\D l_{AB}}\right)\left(\frac{\D x^b_{AQ}}{\D l_{AQ}}\right).	
486: \end{equation}
487: In consequence, also in Riemannian space we say that two vectors are orthogonal when their inner product is null. Although in the case of an indefinite quadratic form it is not sensible to define an acute or obtuse angle with a cosine-type relation, we do nevertheless accept the concept of a straight angle, i.e.\ of orthogonality, from the positive definite case, and characterize two orthogonal vectors in Riemann-Einstein spacetime by the relation of Eq.~(\ref{orthogonality-eq}). As a result, the experimental determination of spacelike separations supplies a method for the experimental determination of orthogonality between displacement vectors: Eq.~(\ref{orthogonality-eq}) means that spacelike displacement vector $\D x^\alpha_{AB}$ and timelike displacement vector $\D x^\beta_{AQ}$ are \emph{orthogonal} in an \emph{experimental test} where $\D s_{AB}=\D s_{AQ}$ and where events $B$ and $Q$ are joined by the worldline of a photon (Figure \ref{orthogonality-time-space}). This concept of orthogonality can be generalized for any pair of contravariant vectors $T^\alpha$ \mbox{and $U^\alpha$.}
488: 
489: \begin{figure} 
490: \centering
491: \includegraphics[height=30mm]{orthogonality-time-space}
492: \caption{Experimental test to determine the orthogonality of a
493: spacelike vector $\D x_{AB}$ and a timelike vector $\D
494: x_{AQ}$. The two vectors are orthogonal if $\D
495: s_{AB}=\D s_{AQ}$, where events $B$ and $Q$ are joined
496: by the worldline of a photon.} 
497: \label{orthogonality-time-space}
498: \end{figure}
499: 
500: Can now two spacelike displacement vectors $\D x_{AB}$ and $\D
501: x_{AC}$ be orthogonal? And if so, under what conditions? By means of
502: what experimental test can we verify their orthogonality? Because now we have two spacelike displacement vectors, we use the radar method twice to measure $\D s^2_{AB}$ and $\D s^2_{AC}$ in terms of timelike separations. Noting that $\D x_{BC}=\D x_{AC}-\D x_{AB}$ (Figure \ref{orthogonality-space-space}), we find 
503: \begin{equation}
504:       g_{\alpha\beta}\D x^\alpha_{BC}\D x^\beta_{BC}-\D s^2_{AC}-\D s^2_{AB} =
505:       -2 g_{\alpha\beta}\D x^\alpha_{AC}\D x^\beta_{AB}.	
506: \end{equation}
507: The left-hand side can only be null and orthogonality possible when $\D x_{BC}$ is spacelike and thus $g_{\alpha\beta}\D x^\alpha_{BC}\D x^\beta_{BC}=\D s^2_{BC}>0$. In
508: that case, the condition for the orthogonality of $\D x_{AB}$ and $\D x_{AC}$ is 
509: \begin{equation}
510: \D s^2_{BC}=\D s^2_{AC}+\D s^2_{AB}.	
511: \end{equation}
512: This is Pythagoras' theorem in spacetime!
513: 
514: \begin{figure} 
515: \centering
516: \includegraphics[height=27mm]{orthogonality-space-space}
517: \caption{Experimental test to determine the orthogonality of two
518: spacelike vectors $\D x_{AB}$ and $\D x_{AC}$. The two
519: vectors are orthogonal if they are joined by a spacelike vector $\D
520: x_{BC}$ and $\D s^2_{BC}=\D s^2_{AC}+\D s^2_{AB}$, that is,
521: when $\D x_{BC}$, $\D x_{AC}$, and $\D x_{AB}$
522: satisfy Pythagoras' theorem in spacetime.} 
523: \label{orthogonality-space-space}
524: \end{figure}
525: 
526: Two timelike vectors, however, can never be orthogonal. This fact leads to a counterintuitive (from a Euclidean point of view) triangle inequality in spacetime, with consequences reaching farther than one would expect. To show first that two timelike vectors $T^\alpha$ and $U^\alpha$ cannot be orthogonal, we investigate their behaviour at a point (chosen to be the origin) after recasting them in pseudospherical coordinates $r,\theta,\phi,\chi$ ($r>0$):
527: \begin{equation} \left\{
528: \begin{array}{ccl}
529:     T^1&=&r\sin(\theta)\cos(\phi)\sinh(\chi)\\
530:     T^2&=&r\sin(\theta)\sin(\phi)\sinh(\chi)\\
531:     T^3&=&r\cos(\theta)\sinh(\chi)\\
532:     T^4&=&r\cosh(\chi)
533: \end{array} \right.
534: \end{equation}
535: and
536: \begin{equation}\left\{
537: \begin{array}{ccl}
538:     U^1&=&\tilde r\sin(\tilde\theta)\cos(\tilde\phi)\sinh(\tilde\chi)\\
539:     U^2&=&\tilde r\sin(\tilde\theta)\sin(\tilde\phi)\sinh(\tilde\chi)\\
540:     U^3&=&\tilde r\cos(\tilde\theta)\sinh(\tilde\chi)\\
541:     U^4&=&\tilde r\cosh(\tilde\chi)
542: \end{array}\right.
543: \end{equation}
544: The quadratic form of $T^\alpha$ at the origin gives 
545: \begin{equation}
546: (T^1)^2+(T^2)^2+(T^3)^2-(T^4)^2=-r^2<0,	
547: \end{equation}
548: which assures us that $T^\alpha$ is timelike at that event; and similarly for $U^\alpha$. After using several trigonometric- and hyperbolic-function identities,\footnote{The trigonometric identities needed are $2\sin(x)\sin(y)=\cos(x-y)-\cos(x+y)$, $2\cos(x)\cos(y)=\cos(x-y)+\cos(x+y)$, and $\cos(2x)=\cos^2(x)-\sin^2(x)$. The hyperbolic identities needed are $2\sinh(x)\sinh(y)=\cosh(x+y)-\cosh(x-y)$ and $2\cosh(x)\cosh(y)=\cosh(x+y)+\cosh(x-y)$.} we find that the inner product of these two vectors at the origin is
549: \begin{multline}
550: T^1U^1+T^2U^2+T^3U^3-T^4U^4=-r\tilde r[\cosh(\chi-\tilde\chi)\cos^2(\omega/2)-\\ \cosh(\chi+\tilde\chi)\sin^2(\omega/2)]\leq -r\tilde r <0,	
551: \end{multline}
552: where 
553: \begin{equation}
554: \cos(\omega)=\cos(\theta)\cos(\tilde\theta)+\sin(\theta)\sin(\tilde\theta)\cos(\phi-\tilde\phi).
555: \end{equation}
556: This means that 
557: \begin{equation}
558: g_{\alpha\beta}T^\alpha U^\beta \leq -r\tilde r \neq 0,	
559: \end{equation}
560: i.e.\ timelike vectors $T^\alpha$ and $U^\alpha$ cannot be orthogonal.
561: 
562: The triangle inequality in spacetime results now as follows. In Euclidean space, the triangle inequality for the sides of a triangle with vertices $A$, $B$, and $C$ reads $|\overrightarrow{AC}|\leq |\overrightarrow{AB}|+|\overrightarrow{BC}|$, which says that the sum of the lengths of any two sides of a triangle is greater than the length of the remaining side. In spacetime, things are otherwise.
563: 
564: Let $T^\alpha_{AB}$, $T^\alpha_{BC}$, and $T^\alpha_{AC}$ be three timelike vectors forming a triangle with vertices at events $A$, $B$, and $C$. We assume that the three vectors point towards the future and that the vectors characterizing the displacements are small. On the basis of Figure \ref{timelike-triangle}, we have
565: \begin{equation}
566: T^\alpha_{AC}=T^\alpha_{AB}+T^\alpha_{BC}.	
567: \end{equation}
568: We express these vectors using their norms
569: \begin{equation}
570: T=\sqrt{\epsilon g_{\alpha\beta}T^\alpha T^\beta} \qquad (\epsilon=-1)	
571: \end{equation}
572: and unit vectors $\hat u$ to get 
573: \begin{equation}
574: T_{AC}u^\alpha_{AC}=T_{AB}u^\alpha_{AB}+T_{BC}u^\alpha_{BC}.	
575: \end{equation}
576: Because the unit vectors are timelike, we have
577: \begin{equation}
578:       g_{\alpha\beta}u^\alpha_{AB} u^\beta_{AB}=
579:       g_{\alpha\beta}u^\alpha_{BC} u^\beta_{BC}=-1	
580:       \quad \mathrm{and} \quad
581:       g_{\alpha\beta}u^\alpha_{AB} u^\beta_{BC}\leq -1. 
582: \end{equation}
583: As a result, we find
584: \begin{equation}
585: T_{AC}^2=T_{AB}^2+T_{BC}^2-2T_{AB}T_{BC}g_{\alpha\beta}u^\alpha_{AB}u^\beta_{BC},	
586: \end{equation}
587: from where 
588: \begin{equation}
589: T_{AC}^2\geq (T_{AB}+T_{BC})^2. 	
590: \end{equation}
591: It follows that
592: \begin{equation}
593: T_{AC}\geq T_{AB}+T_{BC}	
594: \end{equation}
595: is the \emph{triangle inequality} in spacetime, with equality holding when
596: \mbox{$T^\alpha_{AC}=T^\alpha_{AB}$} or \mbox{$T^\alpha_{AC}=T^\alpha_{BC}$} (no actual triangle).
597: 
598: \begin{figure}
599: \centering
600: \includegraphics[width=42mm]{timelike-triangle} \caption{The
601: counterintuitive triangle inequality $T_{AC}\geq T_{AB}+T_{BC}$ for
602: three timelike vectors in spacetime.} 
603: \label{timelike-triangle}
604: \end{figure}
605: 
606: The triangle inequality for timelike vectors has consequences that extend beyond the character of the geometric structure of spacetime. Notably, it has a bearing on the features of material processes that take place in spacetime, such as the spontaneous decay or induced fission of a massive particle into decay or fission products with an ensuing release of energy.
607: 
608: In classical physics, the collision of material particles is understood in terms of the conservation of two quantities, the total quantity of motion $\sum_i m_iv_i$ and the total kinetic energy $\sum_i m_iv_i^2/2$ of the system of particles. From a strictly mechanical point of view, this is only true of elastic collisions, in which no energy is lost by the colliding bodies in the form of heat. Classical mechanics needs to be complemented by thermodynamics in order to make full sense of energy-dissipating (i.e.\ inelastic) collisions. On the other hand, relativity theory is in this respect self-contained; every aspect of the interaction of massive particles can be explained based on the geometric structure of spacetime where the processes occur.
609: 
610: Taking now the natural step of using clock readings $s$ as a worldline parameter, we recast the worldline equation as $x^\alpha=x^\alpha(s)$. Because $\D s$ is always positive, the covariant vector $\D x^\alpha/\D s$ points in the same direction as $\D x^\alpha$, and so
611: \begin{equation}
612: \frac{\D x^\alpha}{\D s} =: V^\alpha 	
613: \end{equation}
614: is a good representation of the worldline tangent vector, which is called the \emph{four-velocity}. The four-velocity $V^\alpha$ is, in fact, a unit vector. Writing $\D x^\alpha=V^\alpha \D s$ in the quadratic form for $\D s^2$, we find 
615: \begin{equation}
616: V^2=\epsilon g_{\alpha\beta} V^\alpha V^\beta=1.      
617: \end{equation}
618: Because the worldline is that of a material particle, then $\epsilon=-1$, and so
619: \begin{equation}
620: g_{\alpha\beta} V^\alpha V^\beta=-1.	
621: \end{equation}
622: Finally, we attach the mass of the particle $m$ (an invariant scalar) to the four-velocity $V^\alpha$ to get the \emph{four-momentum}
623: \begin{equation}
624: P^\alpha=m V^\alpha.	
625: \end{equation}
626: 
627: In the event of a particle decaying into two subproducts, in relativity theory we only require the conservation of the four-momentum of the total system, namely, 
628: \begin{equation}
629: P^\alpha_1=P^\alpha_2+P^\alpha_3,	
630: \end{equation}
631: which we can graphically represent with three timelike vectors in the form 
632: \begin{equation}
633: P^\alpha_{AC}=P^\alpha_{AB}+P^\alpha_{BC},	
634: \end{equation}
635: as shown in Figure \ref{particle-decay}. Because $V^\alpha$ is a unit vector, $m$ acts as the norm of $P^\alpha$, and we can apply the results of the triangle inequality we found for $T^\alpha$, to get 
636: \begin{equation}
637: m_1>m_2+m_3.	
638: \end{equation}
639: We find that the conservation of the four-momentum in the decay of a particle involves the release of energy, $m_1c^2>m_2c^2+m_3c^2$, as an \emph{inescapable consequence} of the non-Euclidean geometric structure of Riemann-Einstein spacetime, namely, the non-orthogonality of timelike vectors. This positive energy deficit $m_1c^2-(m_2c^2+m_3c^2)$ is normally denoted as $Q$ in textbooks on the topic, but its origin is left unexplained.
640: 
641: \begin{figure}
642: \centering
643: \includegraphics[width=42mm]{particle-decay} \caption{Particle decay
644: ($A$) into two subproducts ($B$ and $C$). Energy is released in these processes as a
645: result of the spacetime triangle inequality.} 
646: \label{particle-decay}
647: \end{figure}
648: 
649: \section{Gravitation as spacetime curvature}
650: In his Kyoto lecture, Einstein describes the train of thought leading him to the general theory of relativity: 
651: \begin{quote}
652: If a man falls freely, he would not feel his weight\ldots A falling man is accelerated\ldots In the accelerated frame of reference, we need a new gravitational field\ldots [I]n the frame of reference with acceleration Euclidean geometry cannot be applied. \cite[p.~47]{Einstein:1982}
653: \end{quote}
654: 
655: If a falling man does not feel his weight yet is accelerated, it seems he must not be subjected to any gravitational \emph{force} or \emph{field} but to some form of interaction (whose effects we see in his acceleration) nonetheless---he must be subjected to a ``new gravitational field.'' What is this new interaction that induces gravitational effects yet is not a force? An answer can be sketched as follows. It has been long since realized that the gravitational force is a special interaction. Galilei and Newton observed that inertial masses $m_i$ and gravitational masses $m_g$ were the same (Newton's principle) and that, therefore, all massive bodies in free fall in a vacuum suffer the same acceleration (Galilei's principle) regardless of their constitution. This pulling on all bodies and light in like fashion insinuates that the ``new gravitational field'' we are after could be some property of space or spacetime in which massive bodies and light move.
656: 
657: Could gravitation, then, be conceived as a property of \emph{space}? Intuitively, as Rindler \citeyear[p.~115]{Rindler:1977} explains, this cannot be just so because the future motion of a body under gravitational interaction depends on its initial velocity, but this information is not included in the description of a \emph{space} geodesic, which is determined only by initial position and direction.\footnote{For example, a free particle confined to a surface follows a geodesic that depends only on its initial position and direction, but is independent of its initial velocity.} However, an account of all necessary initial conditions is included in the specification of a \emph{spacetime} geodesic. The validity of this heuristic reasoning is strengthened when we observe with Einstein that, in the man's accelerated frame of reference, Euclidean geometry cannot be applied beyond a point-event,\footnote{This means that, because the velocity of the (rigid) frame changes constantly, we must attach a different inertial system to each \emph{spacetime} point or event. However, Einstein \citeyear{Einstein:1983b} sometimes reached the same conclusion in a more intuitive way by looking at a rotating frame; here an inhomogeneous distortion of ``rigid bodies'' (measuring rods) ensues in a more easily visualizable way, as even each \emph{space} point becomes an own, separate inertial frame ``on account of the Lorentz contraction'' (p.~33).} and note at the same time the essential role that the geometry of the \emph{four-dimensional}, \emph{non-Euclidean} Riemann-Einstein spacetime has had so far in the development of relativity theory.
658: 
659: Given the geometric setting of relativity theory, we suspect then that the ``new gravitational field'' must be somehow included in the geometry of Riemann-Einstein spacetime. Gravitation, moreover, must be an \emph{absolute property of spacetime}, in the sense that either gravitational \emph{effects} are present or they are not. As has been often remarked, when it comes to \emph{spacetime} properties, there is nothing relative in relativity theory. The prime example of this truth is the spacetime interval $\D s$. In absolute classical physics, on the other hand, gravitation is, in fact, a relative concept: an observer on earth deems it to be a force acting on the falling man, while the falling man deems it as no force at all. Einstein's classical thought experiment would teach us nothing were we not willing to also take into account the relative point of view of the man on earth.
660: 
661: What spacetime property shall we now consider to represent gravitation as an absolute interaction? And what is the role of measurements in this search? On account of a long acquaintance with it (and perhaps also because of the suggestive notation), one's first intention is to take the metric tensor $g_{\alpha\beta}(x)$ to play the vacant role of ``new gravitational field,'' while at the same time finally giving an answer to the still-standing question of its physical meaning. However, the metric tensor is a \emph{rather bad candidate} to act as ``gravitation,'' the absolute spacetime property we are seeking. For, if truth be told, there is nothing absolute in $g_{\alpha\beta}(x)$: it is a tensor field, and that is all very well, but what absolute spacetime property does it represent? In one coordinate system, $g_{\alpha\beta}(x)$ takes certain values; in another, it takes different ones. 
662: 
663: When we introduced tensors at the beginning of this chapter, we observed that their reason of being was to provide mathematical objects meaningful in all coordinate systems. We then identified such mathematical objects, not with the tensors themselves, but with \emph{tensor equations}. Here is then the key to this search. Is there any tensor equation that $g_{\alpha\beta}(x)$ satisfies as such? So far the only absolute equation connected with the metric field has been $\D s^2-\epsilon g_{\alpha\beta}(x)\D x^\alpha \D x^\beta=0$, but this is an equation for the whole quadratic form and for the clock separation $\D s$. Evidently, this elemental relation is tightly connected with clock measurements but has no straightforward connection with gravitation. It is time to give up the illusion that, in relativity theory, it is simply the metric field that has taken up the role of the ``new gravitational field.'' If a name with a strong physical connotation must still be given to the metric field, one should then perhaps call $g_{\alpha\beta}(x)$ the \emph{chronometric field}---and stop at that.
664: 
665: \subsection{Einstein's geodesic hypothesis}
666: Newton's first law of motion, or the law of inertia, states that a free massive particle follows a uniform and rectilinear trajectory. Here ``free'' means free of all external interactions, including the gravitational force.
667: 
668: In relativity, we proceed to include a description of gravitation by means of an extension of Newton's first law of motion called Einstein's \emph{geodesic hypothesis}. According to it, a free massive particle has a timelike worldline that is a geodetic curve, or geodesic, in Riemann-Einstein spacetime, where by ``free'' we now mean free of all external \emph{forces} but not of gravitation, which we want to describe not as an external force but as an absolute geometric property of spacetime. As such, it is impossible to be free of its effects. Einstein's geodesic hypothesis also applies to a free photon, in which case it says that a free photon follows a null geodesic in Riemann-Einstein spacetime.
669: 
670: A geodesic is a line whose extension (length) between two given points is stationary, i.e.\ has an extremal value. This can be a maximum or a minimum. When we consider two points in three-dimensional Euclidean space, a geodesic refers to a spatial path of minimal length, i.e.\ the shortest distance between the two points. In Riemann-Einstein spacetime, on the other hand, our Euclidean intuition fails again: where we expect less we get more (cf.\ triangle inequality and energy surplus in particle decay). In spacetime, a geodesic represents a worldline of maximal separation as measured by clocks. 
671: 
672: Incidentally, this is not, strictly speaking, a maximal \emph{length} in any literal sense, although our geometric intuition leads us to think of it that way. Insightfully, in this regard, Synge writes:
673: \begin{quote}
674: It is indeed a Riemannian \emph{chronometry} rather than a \emph{geometry}, and the word \emph{geometry}, with its dangerous suggestion that we should go
675: about measuring \emph{lengths} with \emph{yardsticks}, might well be
676: abandoned altogether in the present connection were it not for the
677: fact that the crude literal meaning of the word geometry has been
678: transmuted into the abstract mathematical concept of a ``space'' with
679: a ``metric'' in it. \cite[p.~109]{Synge:1964}
680: \end{quote}
681: Here is yet another instance of a rule we have repeatedly observed: the mind turns everything it beholds into its golden standard---geometry.\footnote{What mythical character does this behaviour remind us of? (See Epilogue.)} 
682: 
683: In order to find the equation $x^\alpha=x^\alpha(u)$ satisfied by the geodetic worldline with extremes at events $A$ and $B$, we calculate the variation of the event separation $\Delta s$ and equate it to zero, 
684: \begin{equation}
685: \delta \int_A^B \D s=\delta \int_{u_A}^{u_B} f\left(x,\frac{\D x}{\D u}\right)\D u=0.	
686: \end{equation}
687: Taking the variations of $x^\alpha$ and $x'^\alpha=:\D x^\alpha/\D u$, we find
688: \begin{equation}
689:     \int_{u_A}^{u_B} \left(\frac{\partial f}{\partial x^\alpha}\delta x^\alpha+
690:     \frac{\partial f}{\partial x'^\alpha}\delta x'^\alpha \right) \D u=0.
691: \end{equation}
692: Rewriting $\delta x'^\alpha=\D (\delta x^\alpha)/\D u$ and integrating the second term by parts (\mbox{$\tilde u_\alpha=\partial f/\partial x'^\alpha$,} $\D \tilde v=\D(\delta x^\alpha)$), we get
693: \begin{equation}
694:     \Big\vert_{u_A}^{u_B}\frac{\partial f}{\partial x'^\alpha}\delta x^\alpha(u)+
695:     \int_{u_A}^{u_B} \left[\frac{\partial f}{\partial x^\alpha}-
696:     \frac{\D}{\D u}\left(\frac{\partial f}{\partial x'^\alpha}\right)\right] \delta x^\alpha \D u =0.
697: \end{equation}
698: Because the first term is null on account of $\delta x(u_B)=\delta x(u_A)=0$ (the extremes are fixed), and because $\delta x^\alpha$ is an arbitrary small variation, we find that, at every point of the geodesic, $f(x,x')$ must satisfy the Euler-Lagrange equations
699: \begin{equation}\label{Euler-Lagrange-I}
700: \frac{\D}{\D u}\left(\frac{\partial f}{\partial x'^\alpha} \right)-
701: \frac{\partial f}{\partial x^\alpha}=0,
702: \end{equation}
703: where $f(x,x')=\sqrt{-g_{\alpha\beta} x'^\alpha x'^\beta}$. To avoid having to differentiate a square-root expression, we rewrite Eq.~(\ref{Euler-Lagrange-I}) in the equivalent form
704: \begin{equation}\label{Euler-Lagrange-II}
705: \frac{\D}{\D u}\left(\frac{\partial f^2}{\partial x'^\alpha} \right)-
706: \frac{\partial f^2}{\partial x^\alpha}-
707: \frac{1}{2f^2}\frac{\D f^2}{\D u}
708: \frac{\partial f^2}{\partial x'^\alpha}=0.
709: \end{equation}
710: Replacing $u$ by the natural worldline parameter $s$, we find 
711: \begin{equation}
712: f^2(x,x')=-g_{\alpha\beta} V^\alpha V^\beta=+1 \qquad \mathrm{and} \qquad \frac{\D f^2}{\D s}=0, 
713: \end{equation}
714: to get
715: \begin{equation}\label{Euler-Lagrange-III}
716: \frac{\D}{\D s}\left(\frac{\partial f^2}{\partial x'^\alpha} \right)-
717: \frac{\partial f^2}{\partial x^\alpha}=0.
718: \end{equation}
719: Replacing $f^2(x,x')$ by $-g_{\alpha\beta} x'^\alpha x'^\beta$, we obtain the geodesic equation in the more explicit form
720: \begin{equation}\label{Euler-Lagrange-IV}
721: g_{\delta\beta}\frac{\D x'^\beta}{\D s}+[\alpha\beta,\delta] x'^\alpha x'^\beta=0,
722: \end{equation}
723: where
724: \begin{equation}\label{Christoffel-first-kind}
725: [\alpha\beta,\delta]=\frac{1}{2}\left(
726: \frac{\partial g_{\alpha\delta}}{\partial x^\beta} +
727: \frac{\partial g_{\beta\delta}}{\partial x^\alpha} -
728: \frac{\partial g_{\alpha\beta}}{\partial x^\delta} \right)
729: \end{equation}
730: is called \emph{Christoffel's symbol of the first kind}, which is symmetric with respect to its first two indices. Finally, multiplying Eq.~(\ref{Euler-Lagrange-IV}) by $g^{\gamma\delta}$, we obtain a yet more explicit form of the geodesic equation:
731: \begin{equation}\label{Euler-Lagrange-V}
732: \frac{\D^2 x^\gamma}{\D s^2}+\big\{\begin{smallmatrix}\gamma\\\alpha\beta
733: \end{smallmatrix}\big\} \frac{\D x^\alpha}{\D s} \frac{\D x^\beta}{\D s}=0,
734: \end{equation}
735: where
736: $\big\{\begin{smallmatrix}\gamma\\\alpha\beta\end{smallmatrix}\big\}=g^{\gamma\delta}[\alpha\beta,\delta]$ is called \emph{Christoffel's symbol of the second kind}. In addition, the timelike condition 
737: \begin{equation}
738: g_{\alpha\beta}\frac{\D x^\alpha}{\D s}\frac{\D x^\beta}{\D s}=-1	
739: \end{equation}
740: holds for all events on the geodesic. The solution $x^\alpha=x^\alpha(s)$ to this second-order differential equation can be determined unequivocally when we know $x^\alpha(s)$ and $\D x^\alpha$ at some event on the timelike geodesic, i.e.\ for some clock-reading value $s$.
741: 
742: The geodesic equation (\ref{Euler-Lagrange-V}) gives us now the answer to a question we asked earlier on. On page~\pageref{equation-of-motion}, we had expected that the equation of motion~(\ref{equation-of-motion}) of a particle would be of the form $\D^2 x^\alpha/\D u^2=X^\alpha(x,\D x/\D u)$. Now we have found  that this is indeed the case, with
743: \begin{equation}\label{form-of-X}
744: X^\alpha \left( x,\frac{\D x}{\D s}\right)=
745: -\big\{\begin{smallmatrix}\alpha\\ \beta\gamma\end{smallmatrix}\big\} \frac{\D x^\beta}{\D s} \frac{\D x^\gamma}{\D s},
746: \end{equation}
747: where the $x$-dependence is contained in Christoffel's symbol.
748: 
749: As an example, we can apply the geodesic equation to the case of a spacetime whose quadratic form is 
750: \begin{equation}
751: Q=(\D x^1)^2+(\D x^2)^2+(\D x^3)^2-(\D x^4)^2	
752: \end{equation}
753: everywhere and its metric, therefore, $g_{\alpha\beta}=\mathrm{diag}(1,1,1,-1)=\eta_{\alpha\beta}$. As a result, all Christoffel symbols are null and the geodesic equation reduces to
754: \begin{equation}
755: \frac{\D^2 x^\alpha}{\D s^2}=0,
756: \end{equation}
757: with timelike condition
758: \begin{equation}\label{timelike-condition}
759: \left(\frac{\D x^1}{\D s}\right)^2+\left(\frac{\D x^2}{\D s}\right)^2+
760: \left(\frac{\D x^3}{\D s}\right)^2-\left(\frac{\D x^4}{\D s}\right)^2=-1.
761: \end{equation}
762: The solutions are \emph{straight lines in spacetime} 
763: \begin{equation}
764: x^\alpha(s)=A^\alpha s+B^\alpha,	
765: \end{equation}
766: where $A^\alpha$ and $B^\alpha$ are constants, and $A^\alpha$ satisfies timelike condition (\ref{timelike-condition}). Excepting this last restriction, \label{flat-space} the equation of motion and its solution have the same form as those for a free particle in classical physics, namely, $\D^2 x^a/\D t^2=0$ with solution $x^a(t)=A^a t+B^a$. The correspondence suggests that a spacetime whose quadratic form is pseudo-Cartesian everywhere resembles flat Euclidean space.
767: 
768: So far we have dealt with geodesics of material particles, for which \mbox{$\D s>0$}. How shall we deal with geodesics of photons, for which $\D s=0$? In this case, the use of $s$ as a geodesic parameter is impossible, and we must use a different parameter $u$ together with its corresponding lightlike condition 
769: \begin{equation}
770: 	f^2(x,x')=g_{\alpha\beta}U^\alpha U^\beta=0,
771: \end{equation} 
772: where $U^\alpha=\D x^\alpha/\D u$. On account of it, the simplification $\D f^2/\D s=0$ introduced earlier now becomes $\D f^2/\D u=0$. From here on, the procedure is analogous to the one for a timelike geodesic. The equation of a lightlike geodesic is then
773: \begin{equation}\label{null-geodesic-equation}
774: \frac{\D^2 x^\gamma}{\D u^2}+\big\{\begin{smallmatrix}\gamma\\\alpha\beta\end{smallmatrix}\big\} \frac{\D x^\alpha}{\D u} \frac{\D x^\beta}{\D u}=0,
775: \end{equation}
776: with lightlike condition
777: \begin{equation}
778: g_{\alpha\beta}\frac{\D x^\alpha}{\D u} \frac{\D x^\beta}{\D u}=0.
779: \end{equation}
780: This way of proceeding, however, begs the question as to what kind of parameter $u$ is. How general can it be and what is its relation to $s$? We defer the answer to this question until we have studied the absolute derivative in connection with the generalized law of inertia of relativity theory.
781: 
782: The geodesic equation for material particles and photons can also be derived by means of a different method. We return to the original Euler-Lagrange equation (\ref{Euler-Lagrange-I}) and proceed to find its first integral. In order to avoid confusion with the preceding analysis, we now rename $f$ as $\phi$ in this equation.
783: 
784: Multiplying Eq.~(\ref{Euler-Lagrange-I}) by $x'^\alpha=\D x^\alpha/\D u$ in the sense of an inner product, we find
785: \begin{equation}
786:     x'^\alpha \frac{\D}{\D u}\left(\frac{\partial \phi}{\partial x'^\alpha} \right)-
787:     x'^\alpha \frac{\partial \phi}{\partial x^\alpha}=0,
788: \end{equation}
789: which can be rewritten as
790: \begin{equation}
791:     \frac{\D}{\D u}\left( x'^\alpha \frac{\partial \phi}{\partial x'^\alpha} \right) -
792:     \left( \frac{\partial \phi}{\partial x^\alpha} \frac{\D x^\alpha}{\D u} +
793:            \frac{\partial \phi}{\partial x'^\alpha} \frac{\D x'^\alpha}{\D u}\right)=0.
794: \end{equation}
795: Because the second term is $\D \phi/\D u$, the whole expression can be recast as a total derivative,
796: \begin{equation}
797:     \frac{\D}{\D u}\left( x'^\alpha \frac{\partial \phi}{\partial x'^\alpha}-\phi \right)=0.
798: \end{equation}
799: Integrating once with respect to $u$, we get
800: \begin{equation}\label{first-integral}
801:     x'^\alpha \frac{\partial \phi}{\partial x'^\alpha}-\phi = C\ (\mathrm{constant}).
802: \end{equation}
803: Proposing the solution $\phi=g_{\alpha\beta}x'^\alpha x'^\beta$, we find $\phi=C$. Choosing $u=s$ leads to the earlier results for a \emph{timelike geodesic}; in that case, $C=-1$ (timelike condition). When $C=+1$, we find the spacelike condition $g_{\alpha\beta}x'^\alpha x'^\beta=+1$, and we are led to a ``spacelike geodesic,'' a mathematical concept that has no correspondent in the physical world and we can forthwith forget. When $C=0$, we find the lightlike condition $g_{\alpha\beta}x'^\alpha x'^\beta=0$, and we are led to a \emph{null geodesic}, the worldline followed by a free photon. Finally, because we now have $\phi=-f^2(x,x')$,\footnote{The extra minus sign does not change the final result.} the earlier derivation of the geodesic equation leading to Eq.~(\ref{null-geodesic-equation}) remains the same.
804: 
805: \subsection{The absolute derivative}\label{absolute-derivative-sec}
806: Given the geodesic equation (\ref{null-geodesic-equation}) (in principle applicable to both material particles and photons), let us perform a general parameter transformation $u\mapsto v(u) $ and rewrite the geodesic equation in terms of the new parameter $v$. Using the chain rule to go from $u$ to $v$, we find
807: \begin{equation}
808: \frac{\D^2 x^\gamma}{\D v}+\big\{\begin{smallmatrix}\gamma\\\alpha\beta\end{smallmatrix}\big\} \frac{\D x^\alpha}{\D v} \frac{\D x^\beta}{\D v}=
809: -\frac{\D^2 v}{\D u^2}\left(\frac{\D v}{\D u}\right)^{-2} \frac{\D x^\gamma}{\D v}.
810: \end{equation}
811: This result tells us two things. First, that the geodesic equation retains its simple form only for linear (also called affine) parametric transformations. Returning to the possible relation between the null parameter $u$ and timelike parameter $s$, we learn that, given $s$ is a naturally valid parameter, in order for the geodesic equation of a photon to conserve its simplest form, we must have that $u=as+b$, where $a$ and $b$ are yet physically undetermined constants. Second, it tells us that because the right-hand side is a contravariant vector $U^\gamma(v)$ of the form $f(v)\D x^\gamma/\D v$, then so must be the left-hand side, despite the fact that neither its first term nor its second term in isolation are vectors themselves (their sum must be). As we see next, the second observation opens up a natural (i.e.\ physically inspired) road towards the concept of absolute derivative. We pursue the implications of the first observation later on.
812: 
813: What we wish to do is find a form for the geodesic equation that is explicitly tensorial; to achieve it, we need to remove from sight both the total second derivative of the coordinates (which is not a tensor) and Christoffel's symbol (which is not a tensor either) and recast the equation in terms of a single, explicitly contravariant vector.
814: 
815: Returning to a notation in terms of $u$, in a change of coordinates $x\mapsto x'$ we should then have $U'^\gamma(u)=(\partial x'^\gamma/\partial x^\delta)U^\delta$; explicitly, this is
816: \begin{equation}\label{coordinate-change}
817: \frac{\D^2 x'^\gamma}{\D u}+\big\{\begin{smallmatrix}\gamma\\\alpha\beta\end{smallmatrix}\big\}' \frac{\D x'^\alpha}{\D u} \frac{\D x'^\beta}{\D u}=
818: \frac{\partial x'^\gamma}{\partial x^\delta}\frac{\D^2 x^\delta}{\D u}+\big\{\begin{smallmatrix}\delta\\\alpha\beta\end{smallmatrix}\big\} \frac{\partial x'^\gamma}{\partial x^\delta} \frac{\D x^\alpha}{\D u} \frac{\D x^\beta}{\D u}.
819: \end{equation}
820: Applying the chain rule repeatedly to express the right-hand side in terms of partial derivatives with respect to $x'$ and total derivatives of $x'$ with respect to $u$, and subsequently comparing with the left-hand side, we get
821: \begin{equation}
822: \left( \big\{\begin{smallmatrix}\gamma\\\alpha\beta\end{smallmatrix}\big\}'-
823: \big\{\begin{smallmatrix}\delta\\\epsilon\zeta\end{smallmatrix}\big\}
824: \frac{\partial x'^\gamma}{\partial x^\delta}\frac{\partial x^\epsilon}{\partial x'^\alpha}\frac{\partial x^\zeta}{\partial x'^\beta}-
825: \frac{\partial x'^\gamma}{\partial x^\delta}
826: \frac{\partial^2 x^\delta}{\partial x'^\alpha \partial x'^\beta}\right)
827: \frac{\D x'^\alpha}{\D u}\frac{\D x'^\beta}{\D u}=0.
828: \end{equation}
829: Therefore, Christoffel's symbol of the second kind transforms as follows:
830: \begin{equation}\label{Christoffel-symbol-transformation}
831: \big\{\begin{smallmatrix}\gamma\\\alpha\beta\end{smallmatrix}\big\}'=
832: \big\{\begin{smallmatrix}\delta\\\epsilon\zeta\end{smallmatrix}\big\}
833: \frac{\partial x'^\gamma}{\partial x^\delta}\frac{\partial x^\epsilon}{\partial x'^\alpha}\frac{\partial x^\zeta}{\partial x'^\beta}+
834: \frac{\partial x'^\gamma}{\partial x^\delta}
835: \frac{\partial^2 x^\delta}{\partial x'^\alpha \partial x'^\beta};
836: \end{equation}
837: the appearance of the second term on the right-hand side shows that Christoffel's symbol of the second kind is not a tensor.\footnote{With a similar procedure, one can also show that Christoffel's symbol of the first kind is not a tensor either.}
838: 
839: Next, we take a general contravariant vector $T^\alpha$, which is attached to every point of a worldline $x^\alpha(u)$; this is, in other words, a vector field $T^\alpha[x(u)]$, although its being attached only to a worldline is, at the moment, a more realistic (i.e.\ physical) situation than its being attached to the whole spacetime continuum. Inspired by the   present form of the geodesic equation and our goal of expressing it tensorially, we propose a \emph{derivative extension} of it of the form
840: \begin{equation}\label{derivative-extension-eq}
841: \frac{\delta T^\alpha}{\delta u}=:\frac{\D T^\alpha}{\D u} + \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\} T^\beta \frac{\D x^\gamma}{\D u}.
842: \end{equation}
843: Is this derivative extension of $T^\alpha$ a contravariant vector?
844: 
845: To check the possible tensorial nature of $\delta T^\alpha/\delta u$, we must verify whether
846: \begin{equation}
847: \frac{\delta T'^\alpha}{\delta u} - \frac{\partial x'^\alpha}{\partial x^\beta} \frac{\delta T^\beta}{\delta u} \nonumber
848: \end{equation}
849: is null. Written explicitly, the expression becomes
850: \begin{equation}
851: \frac{\D T'^\alpha}{\D u} - \frac{\partial x'^\alpha}{\partial x^\beta}
852: \frac{\D T^\beta}{\D u} +
853: \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}' T'^\beta \frac{\D x'^\gamma}{\D u} -
854: \frac{\partial x'^\alpha}{\partial x^\beta}
855: \big\{\begin{smallmatrix}\beta\\\gamma\delta\end{smallmatrix}\big\}
856: T^\gamma \frac{\D x^\delta}{\D u}.
857: \end{equation}
858: The first two terms taken together reduce to
859: \begin{equation}
860: \frac{\partial^2 x'^\alpha}{\partial x^\beta \partial x^\gamma}T^\beta
861: \frac{\D x^\gamma}{\D u}\neq 0,
862: \end{equation}
863: showing that $\D T^\alpha/\D u$ alone is not a vector. Using Eq.~(\ref{Christoffel-symbol-transformation}), the third and fourth terms taken together reduce to
864: \begin{equation}
865: \frac{\partial^2 x^\delta}{\partial x'^\epsilon \partial x'^\zeta}
866: \frac{\partial x'^\alpha}{\partial x^\delta}
867: \frac{\partial x'^\epsilon}{\partial x^\beta}
868: \frac{\partial x'^\zeta}{\partial x^\gamma} T^\beta
869: \frac{\D x^\gamma}{\D u}\neq 0,
870: \end{equation}
871: showing that the second term of the derivative extension (\ref{derivative-extension-eq}) is not a vector either. However, when both results are taken together, we find
872: \begin{equation}
873: \left( \frac{\partial^2 x'^\alpha}{\partial x^\beta \partial x^\gamma} +
874: \frac{\partial^2 x^\delta}{\partial x'^\epsilon \partial x'^\zeta}
875: \frac{\partial x'^\alpha}{\partial x^\delta}
876: \frac{\partial x'^\epsilon}{\partial x^\beta}
877: \frac{\partial x'^\zeta}{\partial x^\gamma}\right) T^\beta
878: \frac{\D x^\gamma}{\D u}=0
879: \end{equation}
880: on account of the parenthesized expression being null.\footnote{This can be proved by differentiating both sides of $(\partial x'^\alpha/\partial x^\gamma)
881: (\partial x^\beta/\partial x'^\alpha)=\delta^\beta_{\ \gamma}$ with respect to $x^\delta$.} We conclude that $\delta T^\alpha/\delta u$ is an explicitly contravariant vector and call it the \emph{absolute derivative} of $T^\alpha$.
882: 
883: We can now put the absolute derivative of a vector to good use. We observe that the condition $\delta T^\alpha/\delta u=0$ pictures the \emph{absolute constancy} of the vector $T^\alpha$; $T^\alpha$ is said to be \emph{parallel-transported} along a worldline. These names are inspired in the Euclidean picture, shown in Figure \ref{parallel-transport}, that results from visualizing this situation: a curve in space and an arrow vector $\vec V$ attached to two points $P$ and $Q$ on the curve, such that $\vec V(P)$ points in the same direction as $\vec V(Q)$. This Euclidean image is given by the condition $\D \vec V/\D u=0$, which is only a special case of $\delta T^\alpha/\delta u=0$. The image also holds under conditions of pseudo-Euclidicity as in special relativity, where the Christoffel symbol is null and the absolute derivative reduces to the usual total derivative, $\delta T^\alpha/\delta u=\D T^\alpha/\D u$. The null absolute derivative pictures the parallel transport of a vector along a worldline in a spacetime that is not in general flat.
884: 
885: \begin{figure}
886: \centering
887: \includegraphics[width=60mm]{parallel-transport} 
888: \caption{Euclidean
889: vizualization via space vector $\vec V$ of the parallel
890: transport of a spacetime vector $T^\alpha$ along a worldline. The
891: parallel-transport condition on $\vec V$ is given by the null ordinary
892: derivative, $\D \vec V/\D u=0$, at each point of a space trajectory,
893: while the same condition on $T^\alpha$ is given by a null absolute
894: derivative, $\delta T^\alpha/\delta u=0$, at each point of the
895: wordline.} 
896: \label{parallel-transport}
897: \end{figure}
898: 
899: Because the form of the absolute derivative was inspired in that of the geodesic equation, we should not be surprised to find that, when parallel-transported, $T^\alpha$ satisfies an equation remarkably similar to the geodesic equation, namely,
900: \begin{equation}
901: \frac{\delta T^\alpha}{\delta u}=:\frac{\D T^\alpha}{\D u} + \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\} T^\beta \frac{\D x^\gamma}{\D u}=0.
902: \end{equation}
903: It is now only a small step to recast the geodesic equation in the form we were seeking, one that is explicitly covariant---and not only that but, moreover, one that has both an elegant form and interpretation:
904: \begin{equation}\label{geodesic-equation-final}
905: \frac{\delta}{\delta u}\frac{\D x^\alpha}{\D u}=0.
906: \end{equation}
907: While the elegant form is self-evident, the elegant interpretation is the following: a geodesic tangent vector progresses along its worldline in the direction which it itself points to, and so, necessarily, without changing direction. This is a generalization of the case of a geodesic in the Euclidean plane, which, being a straight line, is traced by the unidirectional, unperturbed progression of its tangent vector. In consequence, a geodesic in Riemann-Einstein spacetime is a straight line which progresses, as it were, by following the natural \emph{intrinsic} shape of spacetime. The usual Euclidean visualization of this situation is, however, somewhat misleading, because by picturing Riemann-Einstein spacetime as an \emph{extrinsically} curved surface (e.g.\ a sphere embedded in three-dimensional Euclidean space), a geodesic tangent vector (the tangent to a maximal circle) sticks out of the surface into the embedding space, thus giving the impression that it does not progress in the direction that it points to but that, rather, some external pressure forces it to turn with the sphere.
908: 
909: We find, furthermore, that the above metaphorical allusions to ``unperturbed progression'' of and ``external pressure'' on a geodesic tangent vector are particularly apt to describe its behaviour, since in question here is a (new) physical picture of gravitation. A free particle (in the general-relativistic sense, i.e.\ not gravitation-free) follows a geodesic in spacetime. Our classical intuition does not mislead us when we expect the free particle to continue along its worldline unperturbed, since there are no external forces to change the direction or rate of its progression; and so it does, as it is carried along in the intrinsic spacetime direction of least resistance on account of its inertia.
910: 
911: This is a generalization of Newton's first law of motion. Because the mass $m$ of a particle is a constant, we can rewrite Eq.~(\ref{geodesic-equation-final}) in terms of its four-momentum $P^\alpha=mV^\alpha$ and timelike parameter $s$ to get
912: \begin{equation}\label{generalized-first-law}
913: \frac{\delta P^\alpha}{\delta s}=0.
914: \end{equation}
915: This means that the four-momentum of a free massive particle progresses unperturbed (i.e.\ is parallel-transported) in spacetime, of which a special case is
916: \begin{equation}
917: \frac{\D \vec p}{\D t}=\vec 0,
918: \end{equation}
919: namely, Newton's first law of motion, or the law of inertia (with the proviso that we interpret $s$ as absolute time $t$).
920: 
921: Finally, we examine the special case of a photon. This is a massless particle whose four-momentum $P^\alpha$ is nevertheless not null, and so for a photon we may simply write\footnote{Choosing $P^\alpha=\theta \D x^\alpha/\D u$, we find that $\theta$ must be absolutely constant, $\delta \theta/\delta u=0$, on account of $\delta P^\alpha/\delta u=0$ and $\delta U^\alpha/\delta u=0$, where $U^\alpha=\D x^\alpha/\D u$. The simplest choice is $\theta=1$. In general, Eq.~(\ref{particle-photon-relation}) becomes $u=(\theta/m)s$.} 
922: \begin{equation}
923: P^\alpha=\frac{\D x^\alpha}{\D u}.	
924: \end{equation}
925: Comparing with the four-momentum $P^\alpha=m \D x^\alpha/\D s$ of a massive particle leads to the suggestion that a photon can be understood as a limiting case of a massive particle whose mass $m$ tends to zero as the interval separation $\D s$ measured by a clock on its worldline tends to zero with it, such that the quotient of $\D s$ to $m$ converges to a non-null limit; this limit is $\D u$. The linear connection between $u$ and $s$ envisioned earlier is thus 
926: \begin{equation}\label{particle-photon-relation}
927: u=\frac{1}{m}s.	
928: \end{equation}
929: 
930: So far, the absolute derivative was developed only for a contravariant vector, but because in the context of Riemannian geometry contravariant and covariant vectors are fundamentally the same, it is reasonable to expect that the absolute derivative can also be a derivative extension of covariant vectors. In order to find its form, we resort once more to the concept of invariance. Let $T_\alpha$ be a covariant vector attached to the worldline \mbox{$x^\alpha=x^\alpha(u)$}, and let $U^\alpha$ be an arbitrary contravariant vector which is parallel-transported along the same worldline, and therefore 
931: \begin{equation}\label{parallel-transport-u}
932: \frac{\D U^\alpha}{\D u}=-\big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\} U^\beta \frac{\D x^\gamma}{\D u}.	
933: \end{equation}
934: Comparing the ordinary rate of change of the invariant overlap between $T_\alpha$ and $U^\alpha$, after applying the chain rule and Eq.~(\ref{parallel-transport-u}), we find
935: \begin{equation}
936: \frac{\D (T_\alpha U^\alpha)}{\D u}= \left( \frac{\D T_\alpha}{\D u}-
937: \big\{\begin{smallmatrix}\beta\\\alpha\gamma\end{smallmatrix}\big\} T_\beta
938: \frac{\D x^\gamma}{\D u} \right) U^\alpha.
939: \end{equation}
940: Because the left-hand side is an invariant and $U^\alpha$ on the right-hand side is an arbitrary contravariant vector, the parenthesized expression, which is an extension of vector $T_\alpha$, must then be a covariant vector. We call 
941: \begin{equation}
942:       \frac{\delta T_\alpha}{\delta u}=:
943:       \frac{\D T_\alpha}{\D u}-\big\{\begin{smallmatrix}\beta\\\alpha\gamma\end{smallmatrix}\big\} T_\beta \frac{\D x^\gamma}{\D u}
944: \end{equation}
945: the absolute derivative of $T_\alpha$.
946: 
947: The same heuristic process that led us to the absolute derivative of a covariant vector can also lead us to the absolute derivative of a tensor of any type and order. For example, by setting up the ordinary derivative of the invariant $T^{\alpha\beta}U_\alpha W_\beta$, where $U_\alpha$ and $W_\beta$ are parallel-transported along the worldline, we obtain, proceeding as before, the absolute derivative of a second-order contravariant tensor
948: \begin{equation}
949: \frac{\delta T^{\alpha\beta}}{\delta u}= 
950: \frac{\D T^{\alpha\beta}}{\D u}+\big\{\begin{smallmatrix}\alpha\\\gamma\delta\end{smallmatrix}\big\} T^{\gamma\beta}\frac{\D x^\delta}{\D u}+
951: \big\{\begin{smallmatrix}\beta\\\gamma\delta\end{smallmatrix}\big\} T^{\alpha\gamma}\frac{\D x^\delta}{\D u}.
952: \end{equation}
953: The manner in which this result is obtained allows us to conclude that the absolute derivative of a general tensor is built out of its ordinary derivative plus one positive term for each of its contravariant indices and one negative term for each of its covariant indices, each index being summed over one at a time.
954: 
955: From its general form follows that the absolute derivative satisfies the linear-combination rule and Leibniz's chain rule. Another useful result also follows, namely, that the absolute derivative of the metric field is null in both its covariant and contravariant versions,\footnote{This can be shown by rewriting $\D g_{\alpha\beta}/\D u$ as $(\partial g_{\alpha\beta}/\partial x^\delta) (\D x^\delta/\D u)$ and considering the identity $[\delta\alpha,\beta]+[\delta\beta,\alpha]=0$ in the expression for the absolute derivative of the metric tensor.} 
956: \begin{equation}
957: \frac{\delta g_{\alpha\beta}}{\delta u}=0 \qquad \mathrm{and} \qquad \frac{\delta g^{\alpha\beta}}{\delta u}=0.	
958: \end{equation}
959: This means that, although the metric field is not a constant in the usual sense, it is a constant in the absolute sense of Riemann-Einstein spacetime. Finally, because an invariant scalar has no indices, we add the convention that its absolute derivative corresponds to its ordinary derivative, 
960: \begin{equation}
961: \frac{\delta T}{\delta u}=:\frac{\D T}{\D u}.	
962: \end{equation}
963: 
964: At this point, the reader might object that our way of proceeding in order to find the absolute derivative as an extension of vectors and general tensors has no real justification behind it. Why should it be what it is and not otherwise? To this we reply simply that there is no certified, royal road to mathematical physics. By pretending that there is and defining mathematical concepts in an axiomatic spirit, we trade learning the informal, freewheeling, and privately conceived physical motivation behind mathematical concepts for the illusion of security offered by authoritative disembodied definitions. Would defining the absolute derivative in a dry axiom, while leaving out all mention of the heuristic method by which we obtained it, have made this concept clearer or more obscure in the present physical context? Let us leave rigorously stated definitions and axioms to pure mathematicians; these are their rightful turf---but physics is not mathematics nor should it strive to be.
965: 
966: \subsection{The covariant derivative}
967: Absolute differentiation applies to tensors that are attached to a worldline in Riemann-Einstein spacetime. And this is all very well because the object of our examination has so far been material particles and photons moving along geodesics in spacetime. However, because the geodesic equation does not give us knowledge of the spacetime geometry but only informs us how particles will move given the metric tensor, it needs the partnership of a field equation. A field equation can fill this gap by providing knowledge of the metric tensor given the distribution of matter in spacetime, which thus acts as the source of spacetime geometry, akin to the way charge and current act as the sources of the electric and magnetic forces.
968: 
969: A field equation linking matter distribution to curvature distribution in spacetime differs from the geodesic equation in that the tensors involved in the former need to be attached to whole regions of spacetime and not just to the worldlines of particles. The absolute derivative, then, will not do as a physically meaningful derivative extension of fields, since the term $\D x^\alpha/\D u$ appearing in it does not have any meaning in the absence of a worldline. This is enough motivation for the development of a different kind of extension of a tensor---for a \emph{spacetime derivative}---as closely related to the absolute derivative as possible, yet without reference to a worldline.
970: 
971: What this spacetime derivative should be is insinuated after a simple rewriting of the absolute derivative as follows:
972: \begin{equation}
973: \frac{\delta T^\alpha}{\delta u}=\left( \frac{\partial T^\alpha}{\partial x^\gamma}+\big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\} T^\beta\right) \frac{\D x^\gamma}{\D u}.
974: \end{equation}
975: Again, because the absolute derivative of $T^\alpha$ on the left-hand side is a contravariant vector and the worldline derivative on the right-hand side is as well, the parenthesized expression---which does not depend on the worldline---must now be a second-order once-covariant once-contravariant tensor. As this expression satisfies all we asked from a derivative extension applicable to fields in spacetime, we take
976: \begin{equation}
977: T^\alpha_{\ \ ;\gamma}=: \frac{\partial T^\alpha}{\partial x^\gamma}
978: +\big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\} T^\beta 
979: \end{equation}
980: as the spacetime derivative we seek. We see that, just like the absolute derivative was a generalization of the total derivative in flat space or flat spacetime, so is the spacetime derivative a generalization of the partial derivative in flat space or flat spacetime. From now on, we call $T^\alpha_{\ \ ;\gamma}$ the \emph{covariant derivative} of $T^\alpha$ in line with a long physical tradition. As justification for this name---less descriptive of its physical function than the name spacetime derivative---it is noted that this derivative always has the effect of adding one extra covariant index to the covariantly differentiated tensor. As before, because the covariant derivative comes from the absolute derivative in a straightforward way, the covariant derivative of a tensor of general type follows the same form as that of its absolute derivative. Finally, also the covariant derivative satisfies the linear-combination rule and Leibniz's rule, and the metric tensor is also covariantly constant, $g_{\alpha\beta ;\gamma}=g^{\alpha\beta}_{\ \ \ ;\gamma}=0$.
981: 
982: Once the covariant derivative has been developed, we can recreate with it generalizations of well-known concepts from vector analysis to be useful in the following pages:
983: \begin{itemize}
984: \item[(i)] The \emph{covariant gradient} $T_{;\alpha}$ of an invariant scalar $T$ is by convention (because it has no indices) taken to correspond with its ordinary gradient (ordinary partial derivative), 
985: \begin{equation}
986: T_{;\alpha}=: \frac{\partial T}{\partial x^\alpha} =: T_{,\alpha},	
987: \end{equation}
988: where from now on we also denote partial derivatives with a comma. 
989: 
990: \item[(ii) ]The \emph{covariant divergence} $T^\alpha_{\ \ ;\alpha}$ of a contravariant vector $T^\alpha$ is\footnote{This is a consequence of the simplified form of the appearing Christoffel symbol $\big\{\begin{smallmatrix}\alpha\\\beta\alpha\end{smallmatrix}\big\}$, when this form is made explicit.}
991: \begin{equation}
992: T^\alpha_{\ \ ;\alpha}=T^\alpha_{\ \ ,\alpha}+\frac{1}{2}g^{\alpha\gamma} \frac{\partial g_{\alpha\gamma}}{\partial x^\beta} T^\beta.
993: \end{equation}
994: 
995: \item[(iii)] The \emph{covariant curl} $T_{\alpha ;\beta}-T_{\beta ;\alpha}$ of a covariant vector $T_\gamma$ is its ordinary curl, 
996: \begin{equation}
997: T_{\alpha ;\beta}-T_{\beta ;\alpha}=T_{\alpha ,\beta}-T_{\beta ,\alpha}	
998: \end{equation}
999: on account of the symmetry property of the Christoffel symbol. As a consequence, when $T_\alpha$ is some invariant gradient $U_{,\alpha}$, we recover the spacetime (i.e.\ covariant) version of the well-known result that a gradient is a \emph{conservative vector field}, 
1000: \begin{equation}
1001: U_{,\alpha ,\beta}-U_{,\beta ,\alpha}=0,	
1002: \end{equation}
1003: since ordinary partial derivatives commute. From now on, we denote consecutive ordinary partial derivatives of tensors with a single comma; e.g.\ $U_{,\alpha ,\beta}=: U_{,\alpha\beta}$.
1004: \end{itemize}
1005: 
1006: 
1007: \subsection{The curvature tensor}\label{sec:curvature-tensor}
1008: We saw on page \pageref{flat-space} that the geodesics of a free particle in a spacetime whose quadratic form is everywhere pseudo-Cartesian are straight lines resembling those of flat Euclidean space of classical mechanics. This motivates us to characterize flat spacetime by the condition that it be possible to choose coordinates such that the quadratic form be everywhere pseudo-Cartesian, i.e.\ \mbox{$Q=(\D x^1)^2+(\D x^2)^2+(\D x^3)^2-(\D x^4)^2$}; in other words, that $g_{\alpha\beta}$ can be transformed into $\eta_{\alpha\beta}=\mathrm{diag}(1,1,1,-1)$ at every point. If this were not possible, we take spacetime to be curved.
1009: 
1010: This characterization of flat and curved spacetime is clear and intuitive, but is it practical? Given a quadratic form, how shall we proceed to show that it can or cannot be recast as an everywhere pseudo-Cartesian one? Can we devise an easier way of testing whether spacetime is flat or curved? To do this, we take advantage of the intuitive image given by the spacetime-related concepts inherited from ordinary vector calculus, and think in terms of an analogy. It may be objected that the analogy is far-fetched, but we should not worry about this so long as it naturally guides us to the solution of our problem.
1011: 
1012: We take a covariant vector field $T_\alpha$ and use its second covariant derivative
1013: \begin{equation}
1014: T_{\alpha;\beta\gamma}=: (T_{\alpha;\beta})_{;\gamma} = T_{\alpha;\beta,\gamma}-
1015: \big\{\begin{smallmatrix}\delta\\\alpha\gamma\end{smallmatrix}\big\} T_{\delta;\beta}-
1016: \big\{\begin{smallmatrix}\delta\\\beta\gamma\end{smallmatrix}\big\} T_{\alpha;\delta}
1017: \end{equation}
1018: to construct \emph{a kind of covariant curl} (third-order tensor) $T_{\alpha;\beta\gamma}-T_{\alpha;\gamma\beta}$ of its first covariant derivative $T_{\alpha;\beta}$ (second-order tensor). This curl can be written out explicitly to find
1019: \begin{multline}
1020: T_{\alpha;\beta\gamma}-T_{\alpha;\gamma\beta}=
1021: T_{\alpha;\beta,\gamma}-\big\{\begin{smallmatrix}\delta\\\alpha\gamma\end{smallmatrix}\big\} T_{\delta;\beta}-
1022: \big\{\begin{smallmatrix}\delta\\\beta\gamma\end{smallmatrix}\big\} T_{\alpha;\delta}-
1023: T_{\alpha;\gamma,\beta}+\\
1024: \big\{\begin{smallmatrix}\delta\\\alpha\beta\end{smallmatrix}\big\} T_{\delta;\gamma}+
1025: \big\{\begin{smallmatrix}\delta\\\gamma\beta\end{smallmatrix}\big\} T_{\alpha;\delta}.
1026: \end{multline}
1027: The third and sixth terms on the right-hand side cancel out. Writing out the remaining covariant derivatives, we find
1028: \begin{multline}
1029: T_{\alpha;\beta\gamma}-T_{\alpha;\gamma\beta}=
1030: (T_{\alpha,\beta}-\big\{\begin{smallmatrix}\delta\\\alpha\beta\end{smallmatrix}\big\}T_\delta)_{,\gamma}-
1031: \big\{\begin{smallmatrix}\delta\\\alpha\gamma\end{smallmatrix}\big\}
1032: (T_{\delta,\beta}-\big\{\begin{smallmatrix}\epsilon\\\delta\beta\end{smallmatrix}\big\}T_\epsilon)-\\
1033: (T_{\alpha,\gamma}-\big\{\begin{smallmatrix}\delta\\\alpha\gamma\end{smallmatrix}\big\}T_\delta)_{,\beta}+
1034: \big\{\begin{smallmatrix}\delta\\\alpha\beta\end{smallmatrix}\big\}
1035: (T_{\delta,\gamma}-\big\{\begin{smallmatrix}\epsilon\\\delta\gamma\end{smallmatrix}\big\}T_\epsilon).
1036: \end{multline}
1037: The first and fifth terms, part of the second term and the seventh term, and the third term and part of the sixth term cancel out in pairs. Finally, we find
1038: \begin{equation}\label{covariant-rotor-eq}
1039: T_{\alpha;\beta\gamma}-T_{\alpha;\gamma\beta}= R^\delta_{\ \alpha\beta\gamma}T_\delta,
1040: \end{equation}
1041: where
1042: \begin{equation}\label{curvature-tensor}
1043: R^\delta_{\ \alpha\beta\gamma}=
1044: \big\{\begin{smallmatrix}\delta\\\alpha\gamma\end{smallmatrix}\big\}_{,\beta}-
1045: \big\{\begin{smallmatrix}\delta\\\alpha\beta\end{smallmatrix}\big\}_{,\gamma}+
1046: \big\{\begin{smallmatrix}\epsilon\\\alpha\gamma\end{smallmatrix}\big\}
1047: \big\{\begin{smallmatrix}\delta\\\epsilon\beta\end{smallmatrix}\big\}-
1048: \big\{\begin{smallmatrix}\epsilon\\\alpha\beta\end{smallmatrix}\big\}
1049: \big\{\begin{smallmatrix}\delta\\\epsilon\gamma\end{smallmatrix}\big\}
1050: \end{equation}
1051: must be a fourth-order mixed tensor, since the left-hand side of Eq.~(\ref{covariant-rotor-eq}) is a third-order covariant tensor and $T_\delta$ on the right-hand side is a covariant vector. We call $R^\delta_{\ \alpha\beta\gamma}$ the \emph{mixed curvature tensor} or the \emph{mixed Riemann tensor}. In general, $R^\delta_{\ \alpha\beta\gamma}T_\delta\neq 0$, suggesting that the discrepancy arising from the order of spacetime directions in which a vector is differentiated is the result of a \emph{non-conservative property of spacetime} captured by the tensor field $R^\delta_{\ \alpha\beta\gamma}$. But what is the connection of all this with curvature?
1052: 
1053: If spacetime is flat, there exist coordinates such that the metric field is $\eta_{\alpha\beta}=\mathrm{diag}(1,1,1,-1)$ at every point. In that case, all Christoffel symbols and their ordinary partial derivatives in Eq.~(\ref{curvature-tensor}) are null, and so 
1054: \begin{equation}
1055: R^\delta_{\ \alpha\beta\gamma}T_\delta(x) \equiv 0 	
1056: \end{equation}
1057: at every spacetime point $x$. Since this is a \emph{tensor equation}, its validity is independent of the coordinate system used; if it holds in one coordinate system, it holds in all of them. Now, while this is a necessary condition for the flatness of spacetime, is it also a sufficient condition, such that we may say that there is a one-to-one correspondence between the curvature tensor and curvature? A reading of Einstein's original paper on the general theory of relativity is not auspicious regarding the level of complexity required to answer this question. Einstein \citeyear{Einstein:1952b} prefers not to tackle this problem and, in a footnote, writes: ``The mathematicians have proved that this is also a \emph{sufficient} condition'' (p.~141). Let us nonetheless attempt this demonstration---the physicist's way.
1058: 
1059: We start with an arbitrary covariant vector $T_\alpha(\tilde x)$ at an arbitrary spacetime point, and parallel-transport it first along a worldline $x^\alpha=x^\alpha(u,v_0)$ a separation $\D u$, and subsequently along another worldline $x^\alpha=x^\alpha(\D u,v)$ a separation $\D v$, where $u_0$ and $v_0$ are constants, thus reaching spacetime point $x$ (see Figure \ref{ptgr} on page \pageref{ptgr}). Alternatively, we can reach $x$ from $\tilde x$ by parallel-transporting $T_\alpha(\tilde x)$ first along worldline $x^\alpha=x^\alpha(u_0,v)$ a separation $\D v$, and subsequently along worldline $x^\alpha=x^\alpha(u,\D v)$ a separation $\D u$. The two resulting vectors at $x$ are the same only if the difference
1060: \begin{equation}
1061: \frac{\delta^2T_\alpha}{\delta v \delta u}-
1062: \frac{\delta^2T_\alpha}{\delta u \delta v}=
1063: \left( T_{\alpha;\beta} \frac{\partial x^\beta}{\partial u} \right)_{;\gamma}
1064: \frac{\partial x^\gamma}{\partial v}-
1065: \left( T_{\alpha;\beta} \frac{\partial x^\beta}{\partial v} \right)_{;\gamma}
1066: \frac{\partial x^\gamma}{\partial u}
1067: \end{equation}
1068: is null. Because $\partial x^\beta/\partial u$ and $\partial x^\beta/\partial v$ on the right-hand side are the vectors tangent to the worldlines, they are absolutely constant and can be taken out of the covariant derivative. We find
1069: \begin{align}
1070: \frac{\delta^2T_\alpha}{\delta v \delta u}-
1071: \frac{\delta^2T_\alpha}{\delta u \delta v}&=
1072: T_{\alpha;\beta\gamma}\frac{\partial x^\beta}{\partial u}\frac{\partial x^\gamma}{\partial v}-
1073: T_{\alpha;\beta\gamma}\frac{\partial x^\beta}{\partial v}\frac{\partial x^\gamma}{\partial u}\nonumber\\
1074: &=(T_{\alpha;\beta\gamma}-T_{\alpha;\gamma\beta})\frac{\partial x^\beta}{\partial u}\frac{\partial x^\gamma}{\partial v},
1075: \end{align}
1076: and we now recognize the Riemann tensor in the parenthesized expression:
1077: \begin{equation}
1078: \frac{\delta^2T_\alpha}{\delta v \delta u}-
1079: \frac{\delta^2T_\alpha}{\delta u \delta v}=
1080: R^\delta_{\ \alpha\beta\gamma}T_\delta \frac{\partial x^\beta}{\partial u}
1081: \frac{\partial x^\gamma}{\partial v},
1082: \end{equation}
1083: where the worldlines are arbitrary and so is $T_\alpha$. 
1084: 
1085: Therefore, \emph{if} the Riemann tensor is identically null, then the result of parallel-transporting $T_\alpha(\tilde x)$ to $x$ is a \emph{unequivocal} vector $T_\alpha(x)$ independently of the worldline followed. But then, its covariant derivative must be null, $T_{\alpha;\beta}(x)=0$. Since spacetime point $x$ is arbitrary, given $R^\delta_{\ \alpha\beta\gamma}\equiv 0$ (sufficient condition),\footnote{This is also a necessary condition. The system of equations $T_{\alpha,\beta}=\big\{\begin{smallmatrix}\delta\\\alpha\beta\end{smallmatrix}\big\}T_\delta$ is integrable only if $T_{\alpha,\beta\gamma}=T_{\alpha,\gamma\beta}$. This gives $
1086: \left(\big\{\begin{smallmatrix}\delta\\\alpha\beta\end{smallmatrix}\big\} T_\delta \right)_{,\gamma}=\left(\big\{\begin{smallmatrix}\delta\\\alpha\gamma\end{smallmatrix}\big\} T_\delta\right)_{,\beta}$, implying $R^\delta_{\ \alpha\beta\gamma}\equiv 0$.} we can create a spacetime vector \emph{field} at any spacetime point $x$ via parallel-transport of an original vector. Based on our intuition, we now expect that the unequivocal existence of $T_\alpha(x)$ will imply that spacetime is flat. For example, in the intuitively curved case of a spherical surface, the parallel-transport of a vector from the equator to the north pole produces different vectors at the north pole depending on the path chosen, but the same vector on a flat surface.
1087: 
1088: We now take the covariant vector field $T_\alpha(x)$ to be the gradient $S_{,\alpha}(x)$ of an arbitrary invariant scalar field $S(x)$. We take, moreover, four independent scalars fields, which we choose to be the coordinates $x'^1(x),x'^2(x),x'^3(x)$, and $x'^4(x)$. Now $T_{\alpha;\beta}=0$ translates into $x'^\gamma_{,\alpha;\beta}(x)=0$, which, in turn, gives 
1089: \begin{equation}\label{constant-x}
1090: \frac{\partial^2 x'^\gamma}{\partial x^\alpha\partial x^\beta}=\big\{\begin{smallmatrix}\delta\\ \alpha\beta\end{smallmatrix}\big\}\frac{\partial x'^\gamma}{\partial x^\delta}.
1091: \end{equation}
1092: We now show that in the primed coordinates obtained through parallel transport, the metric tensor $g'_{\alpha\beta}(x')$ is constant.
1093: 
1094: Because 
1095: \begin{equation}
1096: g_{\alpha\beta}=\frac{\partial x'^\gamma}{\partial x^\alpha} \frac{\partial x'^\delta}{\partial x^\beta} g'_{\gamma\delta},	
1097: \end{equation}
1098: we find
1099: \begin{equation}
1100: \frac{\partial g_{\alpha\beta}}{\partial x^\epsilon}=
1101: \frac{\partial^2 x'^\gamma}{\partial x^\alpha\partial x^\epsilon} \frac{\partial x'^\delta}{\partial x^\beta} g'_{\gamma\delta}+
1102: \frac{\partial x'^\gamma}{\partial x^\alpha} \frac{\partial^2 x'^\delta}{\partial x^\beta \partial x^\epsilon} g'_{\gamma\delta}+
1103: \frac{\partial x'^\gamma}{\partial x^\alpha}\frac{\partial x'^\delta}{\partial x^\beta} \frac{\partial g'_{\gamma\delta}}{\partial x^\epsilon},
1104: \end{equation}
1105: which, on account of Eq.~({\ref{constant-x}}), becomes
1106: \begin{equation}
1107: \frac{\partial g_{\alpha\beta}}{\partial x^\epsilon}=
1108: \big\{\begin{smallmatrix}\zeta\\\epsilon\alpha\end{smallmatrix}\big\}
1109: \frac{\partial x'^\gamma}{\partial x^\zeta} \frac{\partial x'^\delta}{\partial x^\beta} g'_{\gamma\delta}+
1110: \big\{\begin{smallmatrix}\zeta\\\epsilon\beta\end{smallmatrix}\big\}
1111: \frac{\partial x'^\gamma}{\partial x^\alpha} \frac{\partial x'^\delta}{\partial x^\zeta} g'_{\gamma\delta}+
1112: \frac{\partial x'^\gamma}{\partial x^\alpha} \frac{\partial x'^\delta}{\partial x^\beta} \frac{\partial g'_{\gamma\delta}}{\partial x^\epsilon}.
1113: \end{equation}
1114: The first and second terms can be simplified to get
1115: \begin{align}
1116: \frac{\partial g_{\alpha\beta}}{\partial x^\epsilon}&=
1117: \big\{\begin{smallmatrix}\zeta\\\epsilon\alpha\end{smallmatrix}\big\}
1118: g_{\zeta\beta}+
1119: \big\{\begin{smallmatrix}\zeta\\\epsilon\beta\end{smallmatrix}\big\}
1120: g_{\alpha\zeta}+
1121: \frac{\partial x'^\gamma}{\partial x^\alpha} \frac{\partial x'^\delta}{\partial x^\beta} \frac{\partial g'_{\gamma\delta}}{\partial x^\epsilon}\nonumber\\
1122: &=[\epsilon\alpha,\beta]+[\epsilon\beta,\alpha]+\frac{\partial x'\gamma}{\partial x^\alpha} \frac{\partial x'^\delta}{\partial x^\beta} \frac{\partial g'_{\gamma\delta}}{\partial x^\epsilon},
1123: \end{align}
1124: finally giving
1125: \begin{equation}
1126: \frac{\partial g_{\alpha\beta}}{\partial x^\epsilon}=
1127: \frac{\partial g_{\alpha\beta}}{\partial x^\epsilon}+ \frac{\partial x'^\gamma}{\partial x^\alpha} \frac{\partial x'^\delta}{\partial x^\beta}\frac{\partial g'_{\gamma\delta}}{\partial x^\epsilon}.
1128: \end{equation}
1129: It follows that $\partial g'_{\gamma\delta}/\partial x^\epsilon=0$, and therefore $\partial g'_{\gamma\delta}/\partial x'^\zeta=0$. In conclusion, if the Riemann curvature tensor is identically null, then coordinates $x'$ exist such that $g'_{\alpha\beta}(x')$ is constant everywhere in spacetime.\footnote{We have not proved, however, that $g'_{\alpha\beta}(x')=\eta_{\alpha\beta}$.}
1130: 
1131: The mixed Riemann tensor $R^\delta_{\ \alpha\beta\gamma}$, its covariant version $R_{\alpha\beta\gamma\delta}$, and its contractions, $R^{\gamma}_{\ \alpha\beta\gamma}$ and $R^\alpha_{\ \alpha}$, satisfy several useful identities. We can see straight from Eq.~(\ref{curvature-tensor}) that the mixed Riemann tensor is antisymmetric on its last two covariant indices, 
1132: \begin{equation}
1133: R^\delta_{\ \alpha\beta\gamma}=-R^\delta_{\ \alpha\gamma\beta},	
1134: \end{equation}
1135: and that the sum of its cyclic permutations is null, 
1136: \begin{equation}
1137: R^\delta_{\ \alpha\beta\gamma}+R^\delta_{\ \gamma\alpha\beta}+R^\delta_{\ \beta\gamma\alpha}=0. 	
1138: \end{equation}
1139: The covariant Riemann tensor can be obtained from the mixed one lowering the contravariant index, $R_{\alpha\beta\gamma\delta}=g_{\alpha\epsilon}R^\epsilon_{\ \beta\gamma\delta}$, which gives\footnote{Express $g_{\alpha\epsilon}\big\{\begin{smallmatrix}\epsilon\\\beta\delta\end{smallmatrix}\big\}_{,\gamma}$ as the derivative of the product minus an extra term, and rewrite both in terms of Christoffel symbols of the first kind and their derivatives.}
1140: \begin{equation}
1141:     R_{\alpha\beta\gamma\delta}=
1142:     [\beta\delta,\alpha]_{,\gamma}-[\beta\gamma,\alpha]_{,\delta}+
1143:     g^{\epsilon\zeta}                         \left(
1144:     [\alpha\delta,\epsilon][\beta\gamma,\zeta]-
1145:     [\alpha\gamma,\epsilon][\beta\delta,\zeta]\right).
1146: \end{equation}
1147: A more useful expression of the covariant Riemann tensor can be obtained by writing the Christoffel symbols of the first kind in terms of those of the second kind, the metric tensor, and its derivatives. We find
1148: \begin{multline}
1149:     R_{\alpha\beta\gamma\delta}=
1150:     \frac{1}{2}(g_{\alpha\delta,\beta\gamma}+g_{\beta\gamma,\alpha\delta}-
1151:       g_{\alpha\gamma,\beta\delta}-g_{\beta\delta,\alpha\gamma})+\\
1152:       g_{\epsilon\zeta}\left(
1153:       \big\{\begin{smallmatrix}\epsilon\\\alpha\delta\end{smallmatrix}\big\}
1154:       \big\{\begin{smallmatrix}\zeta\\\beta\gamma\end{smallmatrix}\big\}-
1155:       \big\{\begin{smallmatrix}\epsilon\\\alpha\gamma\end{smallmatrix}\big\}
1156:       \big\{\begin{smallmatrix}\zeta\\\beta\delta\end{smallmatrix}\big\}
1157:                         \right).
1158: \end{multline}
1159: By direct inspection, we see that the covariant Riemann tensor is antisymmetric on its first two and last two indices, 
1160: \begin{equation}
1161: R_{\alpha\beta\gamma\delta}=-R_{\beta\alpha\gamma\delta} 
1162: \qquad \mathrm{and} \qquad 
1163: R_{\alpha\beta\gamma\delta}=-R_{\alpha\beta\delta\gamma},
1164: \end{equation}
1165: and it is symmetric with respect to a swap of the first and second pairs of indices taken as a unit,
1166: \begin{equation}
1167: R_{\alpha\beta\gamma\delta}=R_{\gamma\delta\alpha\beta}. 		
1168: \end{equation}
1169: Applying all these results together, we also find that the covariant Riemann tensor is symmetric with respect to an opposite reordering of its indices,
1170: \begin{equation}
1171: R_{\alpha\beta\gamma\delta}=R_{\delta\gamma\beta\alpha}.
1172: \end{equation}
1173: 
1174: In addition, the covariant derivatives of the mixed and covariant Riemann tensors each satisfy an identity called the \emph{Bianchi identity}, which says that the sum of the cyclic permutations on the last three covariant indices of the derivative tensor is null. For the mixed Riemann tensor, it reads\footnote{To prove the Bianchi identity, calculate the ``covariant curl'' of the third-order tensor $(A_\alpha B_\beta)_{;\gamma}$ to find $(A_\alpha B_\beta)_{;\gamma\delta}-(A_\alpha B_\beta)_{;\delta\gamma}=A_\alpha R^\epsilon_{\ \beta\gamma\delta}B_\epsilon+B_\beta R^\epsilon_{\ \alpha\gamma\delta}A_\epsilon$. Next identify $A_\alpha B_\beta$ with $T_{\alpha;\beta}$ and form the sum of its three cyclic permutations of $\beta$, $\gamma$, and $\delta$. Express this sum, on the one hand, in terms of the mixed Riemann tensor and first covariant derivatives of $T_\alpha$ and, on the other hand, in terms of covariant derivatives of the product of the mixed Riemann tensor and $T_\alpha$. Use $R^\delta_{\ \alpha\beta\gamma}+R^\delta_{\ \gamma\alpha\beta}+R^\delta_{\ \beta\gamma\alpha}=0$ and simplify.} 
1175: \begin{equation}
1176: R^\epsilon_{\ \alpha\beta\gamma;\delta}+R^\epsilon_{\ \alpha\delta\beta;\gamma}+R^\epsilon_{\ \alpha\gamma\delta;\beta}=0.	
1177: \end{equation}
1178: The corresponding Bianchi identity for the covariant Riemann tensor can be obtained from the former by lowering the contravariant index; it reads 
1179: \begin{equation}\label{Bianchi-II}
1180: R_{\alpha\beta\gamma\delta;\epsilon}+R_{\alpha\beta\epsilon\gamma;\delta}+R_{\alpha\beta\delta\epsilon;\gamma}=0.	
1181: \end{equation}
1182: 
1183: The \emph{Ricci tensor} $R_{\alpha\beta}$ is a useful tensor derived from the covariant Riemann tensor $R_{\delta\alpha\beta\gamma}$ by contracting the first and last indices, 
1184: \begin{equation}
1185: g^{\gamma\delta} R_{\delta\alpha\beta\gamma}=R^\gamma_{\ \alpha\beta\gamma}=: R_{\alpha\beta}.	
1186: \end{equation}
1187: The Ricci tensor is symmetric\footnote{To show this result, write the Ricci tensor in terms of the covariant Riemann tensor and use the symmetry properties of the latter.} and has only 10 independent components, which is the same amount of independent components as that of the metric tensor. This property becomes essential when we form Einstein's field equation linking gravitation with its sources. Another useful quantity, the invariant \emph{curvature scalar} $R$, can be derived from the Ricci tensor by contracting its two indices, 
1188: \begin{equation}
1189: g^{\alpha\gamma}R_{\gamma\alpha}=R^\alpha_{\ \alpha}=: R.	
1190: \end{equation}
1191: 
1192: Do the Ricci tensor and curvature scalar have any noteworthy properties? Taking the second Bianchi identity (\ref{Bianchi-II}) and contracting its indices in a suitable way, we find
1193: \begin{multline}
1194: g^{\beta\gamma}g^{\alpha\delta}R_{\alpha\beta\gamma\delta;\epsilon}+
1195: g^{\alpha\delta}g^{\beta\gamma}R_{\alpha\beta\epsilon\gamma;\delta}+
1196: g^{\beta\gamma}g^{\alpha\delta}R_{\alpha\beta\delta\epsilon;\gamma}=\\
1197: g^{\beta\gamma}R_{\beta\gamma;\epsilon}-g^{\alpha\delta}R_{\alpha\epsilon;\delta}-g^{\beta\gamma}R_{\beta\epsilon;\gamma}=
1198: R_{;\epsilon}-2R^\delta_{\ \epsilon;\delta}=0.
1199: \end{multline}
1200: Multiplying by $g^{\epsilon\zeta}$, we get
1201: \begin{equation}
1202: g^{\epsilon\zeta}R_{;\epsilon}-2R^{\delta\zeta}_{\ \ ;\delta}=
1203: \left( R^{\epsilon\zeta}-\frac{1}{2}g^{\epsilon\zeta}R\right)_{;\epsilon}=0.
1204: \end{equation}
1205: We have found that the symmetric tensor 
1206: \begin{equation}
1207: G^{\alpha\beta} =: R^{\alpha\beta}-\frac{1}{2}g^{\alpha\beta}R, 	
1208: \end{equation}
1209: called the \emph{Einstein tensor}, has the special property that its covariant divergence is null, 
1210: \begin{equation}\label{Einstein-divergence-free}
1211: G^{\alpha\beta}_{\ \ \ ;\beta}=0.	
1212: \end{equation}
1213: 
1214: Finally, noticing that the Einstein tensor $G^{\alpha\beta}$ consists of a second-order tensor $R^{\alpha\beta}$ and a scalar $R$, we consider whether any further, still simpler, term could be added to it such that its covariant divergence continues to be null. Thinking in purely mathematical terms, a term of the form $\Lambda g^{\alpha\beta}$, where $\Lambda$ is some constant unrelated to the Ricci tensor, satisfies $\Lambda g^{\alpha\beta}_{\ \ \ ;\beta}=0$. Adding it to the Einstein tensor for now only with this purely mathematical motivation, we get the \emph{generalized Einstein tensor} 
1215: \begin{equation}
1216: \mathcal{G}^{\alpha\beta}=:R^{\alpha\beta}-\frac{1}{2}g^{\alpha\beta}R+\Lambda g^{\alpha\beta},	
1217: \end{equation}
1218: where 
1219: \begin{equation}
1220: \mathcal{G}^{\alpha\beta}_{\ \ \ ;\beta}=0.	
1221: \end{equation}
1222: Is the inclusion of $\Lambda$ in the Einstein tensor also justified on physical grounds by the display of relevant observable effects? And if so, what is its physical interpretation? Not dark and mysterious, one should hope. We return to this issue later on.
1223: 
1224: \subsection{Geodesic deviation and the rise of rod length}
1225: How do we observe the active effects of the curvature tensor? In a curved spacetime, we expect that two particles that follow at first parallel geodesics will not remain on parallel spacetime tracks but will start to converge or diverge, i.e.\ two test particles released at rest will either move towards or away from each other (discounting their mutual gravitational interaction). This two-geodesic method of probing curvature sounds promising, because the test in question involves only local information.
1226: 
1227: To inspect the behaviour of the separation between two particle geodesics in
1228: relation to curvature, we examine two close-lying geodesics \mbox{$x^\alpha=x^\alpha(u)$} and \mbox{$x'^\alpha=x'^\alpha(u)$} (not its derivative) that are nearly parallel to each other, as shown in Figure \ref{geodesic-deviation} for the case of two massive particles. The small difference \mbox{$\eta^\alpha(u)=x'^\alpha(u)-x^\alpha(u)$} is a differential vector that connects those points in the geodesics that have the same value of parameter $u$. Since the two geodesics are nearly parallel, the rate of change $\D \eta^\alpha/\D u$ of their mutual separation is also very small.
1229: 
1230: \begin{figure}
1231: \centering
1232: \includegraphics[width=60mm]{geodesic-deviation} 
1233: \caption{Deviation of two timelike geodesics as a local active effect of spacetime curvature. When spacetime is flat, two geodesics deviate from each other linearly. When the size $\eta$ of $\eta^\alpha$ does not depend on $s$, \mbox{$\D \eta/\D s=0$}, $\eta^\alpha$ acts like an idealized measuring rod.}
1234: \label{geodesic-deviation}  
1235: \end{figure}
1236: 
1237: In order to obtain an equation for $\eta^\alpha$, we subtract the
1238: geodesic equation of $x^\alpha(u)$ from that of $x'^\alpha(u)$.
1239: Using the first-order approximation to the primed Christoffel symbol,
1240: \begin{equation}
1241: \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}'=
1242: \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}+\big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}_{,\delta}\eta^\delta,
1243: \end{equation}
1244: on account of $x'^\alpha(u)=x^\alpha(u)+\eta^\alpha(u)$, we find
1245: \begin{equation} 
1246: \frac{\D^2 \eta^\alpha}{\D u^2}+
1247: \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}_{,\delta}
1248: \frac{\D x^\beta}{\D u} \frac{\D x^\gamma}{\D u}\eta^\delta+
1249: \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}
1250: \frac{\D x^\beta}{\D u} \frac{\D \eta^\gamma}{\D u}+
1251: \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}
1252: \frac{\D \eta^\beta}{\D u} \frac{\D x^\gamma}{\D u}=0. 
1253: \end{equation}
1254: Introducing the derivative of a sum and subtracting two extra
1255: terms, we get
1256: \begin{multline} 
1257: \frac{\D}{\D u}\left( \frac{\D \eta^\alpha}{\D u} +
1258: \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}\eta^\beta
1259: \frac{\D x^\gamma}{\D u} \right) -
1260: \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}_{,\delta}
1261: \eta^\beta \frac{\D x^\gamma}{\D u} \frac{\D x^\delta}{\D u} -
1262: \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}\eta^\beta
1263: \frac{\D^2 x^\gamma}{\D u^2}+\\
1264: \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}_{,\delta}
1265: \frac{\D x^\beta}{\D u} \frac{\D x^\gamma}{\D u}\eta^\delta+
1266: \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}
1267: \frac{\D x^\beta}{\D u} \frac{\D \eta^\gamma}{\D u}=0. 
1268: \end{multline}
1269: Finally, using the geodesic equation to rewrite $\D^2x^\gamma/\D
1270: u^2$, adding and subtracting a term (fourth and last below) to form an extra $\delta \eta^\delta/\delta u$, and rearranging the indices, we find 
1271: \begin{multline} 
1272: \frac{\D}{\D u}\left( \frac{\D \eta^\alpha}{\D u} +
1273: \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}\eta^\beta
1274: \frac{\D x^\gamma}{\D u} \right) +
1275: \big\{\begin{smallmatrix}\alpha\\\delta\epsilon\end{smallmatrix}\big\}
1276: \left( \frac{\D \eta^\delta}{\D u} +
1277: \big\{\begin{smallmatrix}\delta\\\beta\gamma\end{smallmatrix}\big\}\eta^\beta
1278: \frac{\D x^\gamma}{\D u} \right)\frac{\D x^\epsilon}{\D u}+\\ \left(
1279: \big\{\begin{smallmatrix}\alpha\\\gamma\delta\end{smallmatrix}\big\}_{,\beta}-
1280: \big\{\begin{smallmatrix}\alpha\\\beta\gamma\end{smallmatrix}\big\}_{,\delta}+
1281: \big\{\begin{smallmatrix}\epsilon\\\gamma\delta\end{smallmatrix}\big\}
1282: \big\{\begin{smallmatrix}\alpha\\\beta\epsilon\end{smallmatrix}\big\}-
1283: \big\{\begin{smallmatrix}\epsilon\\\beta\gamma\end{smallmatrix}\big\}
1284: \big\{\begin{smallmatrix}\alpha\\\delta\epsilon\end{smallmatrix}\big\}
1285: \right)\eta^\beta \frac{\D x^\gamma}{\D u}\frac{\D x^\delta}{\D
1286: u}=0. 
1287: \end{multline} 
1288: In the first two parenthesized expressions, we recognize the absolute derivative of the geodesic separation and, in the last one, the Riemann tensor. The equation satisfied by the geodesic separation, which connects spacetime curvature with the way the two geodesics deviate from one another, can then be expressed in simpler terms as
1289: \begin{equation} 
1290: \frac{\delta^2\eta^\alpha}{\delta u^2}+R^\alpha_{\
1291: \gamma\beta\delta}\eta^\beta \frac{\D x^\gamma}{\D u} \frac{\D
1292: x^\delta}{\D u}=0. 
1293: \end{equation}
1294: The solution is completely determined when we know $\eta^\alpha$ and $\D
1295: \eta^\alpha/\D u$ for some value of the parameter $u$. In flat
1296: spacetime, where the curvature tensor is identically null, the
1297: equation reduces to 
1298: \begin{equation}
1299: \frac{\D^2 \eta^\alpha}{\D u^2}=0.	
1300: \end{equation}
1301: Its solutions are $\eta^\alpha(u)= A^\alpha u +B^\alpha$, which means that in flat spacetime the separation vector between two geodesics changes linearly with $u$---the two geodesics are straight lines in spacetime and, therefore, converge or diverge linearly.
1302: 
1303: The separation between timelike spacetime geodesics gives us now the
1304: possibility to recover the concept of \emph{rod distance} $\eta$ from
1305: chronometric foundations. In relativity theory, the concept of
1306: length between two particles is not  straightforward because, in 
1307: Riemann-Einstein spacetime, the present---that is, when the length measurement should be made---is tied to each observer's worldline; in classical physics, on the other hand, the present is an absolute notion. This difficulty
1308: notwithstanding, the rod length between two particles can be determined
1309: for special cases in relativity theory too, as follows.
1310: 
1311: The geodesic separation $\eta^\alpha(s)$ is a spacelike vector,
1312: whose size 
1313: \begin{equation}
1314: \eta(s) =: \D s_{\eta(s)}=\sqrt{g_{\alpha\beta}[x(s)]\eta^\alpha(s)\eta^\beta(s)}	
1315: \end{equation}
1316: depends on the value of the geodesic parameter $s$. This size, indicating the separation between two neighbouring free massive particles, is measured with a clock and a photon; showing the dependence on $s$ explicitly, it is 
1317: \begin{equation}
1318: \eta(s)=\D s_{s_P s_A} \D s_{s_A s_Q},	
1319: \end{equation}
1320: where $P$ and $Q$ are the events of photon emission and reception and $A$ marks the tail of $\eta^\alpha(s)$ (cf.\ Figure \ref{spacelike-separation} on page \pageref{spacelike-separation}). If $\eta(s)$ does not depend on $s$, the spacelike separation must be constant, 
1321: \begin{equation}
1322: \frac{\D \eta}{\D s}=0.	
1323: \end{equation}
1324: In that case, we can identify the two particles joined by $\eta^\alpha$ with the end-points of an idealized rigid measuring rod and $\eta$ with its length. In practice, this condition is obtained when all the photons sent out from the first particle and bouncing back from the second particle arrive at constant clock intervals. In this roundabout manner, the concept of rod length arises as a particular case of measurements of spacetime intervals with clocks and photons.
1325: 
1326: \section{The sources of gravitation}\label{Sources-of-gravitation}
1327: 
1328: \subsection{The vacuum}
1329: We start the search for the different instances of Einstein's field equation, which connects spacetime curvature with sources in spacetime, in each case looking for guidance in Newton's classical theory of gravitation. At the outset with the case of the vacuum, Newtonian theory helps us realize that there is no \emph{straightforward} general-relativistic counterpart of the classical gravitational potential $\Phi(\vec r)$, but it also provides a useful analogical picture on which to base the search for the different instances of the general-relativistic field equation.
1330: 
1331: In the vacuum outside a spherically symmetric mass distribution, the Newtonian gravitational potential $\Phi(\vec r)$, connected with the universal force of gravitation according to $\vec \nabla \Phi(\vec r)=-\vec g(\vec r)$, satisfies Laplace's equation 
1332: \begin{equation}
1333: \nabla^2\Phi(\vec r)=0.	
1334: \end{equation}
1335: This follows from the fact that $\vec g(\vec r)$ is divergence-free, $\vec\nabla\cdot \vec g(\vec r)=0$, where $\vec g(\vec r)=-GM\hat r/r^2$. In Cartesian coordinates, Laplace's equation becomes 
1336: \begin{equation}
1337: \sum_{i=1}^3 \frac{\partial^2 \Phi(\vec r)}{(\partial x^i)^2}=0.	
1338: \end{equation}
1339: 
1340: In special relativity, an equation corresponding to the classical Laplace equation is 
1341: \begin{equation}
1342: \square \phi(x)=:\eta^{\alpha\beta}\phi_{,\alpha\beta}(x)=0.	
1343: \end{equation}
1344: This equation has the form of the Klein-Gordon equation for a scalar field associated with a massless particle; expressed in the form of a wave equation, it is 
1345: \begin{equation}
1346: \sum_{i=1}^3 \frac{\partial^2 \phi(x)}{(\partial x^i)^2}-\frac{\partial^2 \phi(x)}{(\partial x^4)^2}=0.
1347: \end{equation}
1348: 
1349: Could we now generalize Laplace's equation for the general-relativistic case to find the equation of gravitation in the vacuum? Should we perhaps, in analogy, take our equation to be $g^{\alpha\beta}\phi_{;\alpha\beta}(x)=0$? But what is now the meaning of field $\phi(x)$? Having come this far in the formulation of general relativity, we know it for a fact that its foundation is given by the quadratic form $\D s^2=\epsilon g_{\alpha\beta}(x)\D x^\alpha \D x^\beta$, so why should all of a sudden an extraneous field like $\phi(x)$ be needed? How about $g^{\alpha\beta}g_{\gamma\delta;\alpha\beta}(x)=0$ then? Not really, because the covariant derivative of the metric tensor is identically null, $g_{\gamma\delta;\alpha}(x)\equiv 0$, and so no real information can be extracted from such an equation. Once again, it seems that simply $g_{\alpha\beta}(x)$ is not enough to serve as the ``new gravitational field.''
1350: 
1351: As we suspected earlier, gravitation should instead be related to spacetime curvature, and we know that curvature is characterized by the Riemann tensor. But how should one take advantage of this knowledge to build the field equation in vacuum? If we say that ``spacetime curvature is null in vacuum,'' would this mean that spacetime is flat, and thus gravitation-free, in vacuum even when there is matter elsewhere and not necessarily very far away? Such an interpretation goes against our intuition, because gravitational effects  exist in the vacuum near massive objects; spacetime cannot therefore be flat there.
1352: 
1353: When we say, as we will, that spacetime curvature is null in vacuum, we take the meaning of this assertion in direct analogy with Newtonian theory, by identifying a null curvature tensor with the null Laplacian of the Newtonian potential. And so, just like in classical physics the null Laplacian can give the solution to the non-null gravitational potential \emph{outside} a material distribution, so can a null curvature tensor give the non-null curvature solution \emph{outside} a material distribution. Only when the vacuum extends thoroughly everywhere are gravitational effects nonexistent and, in relativistic physics, we can say that spacetime is flat.
1354: 
1355: So far we have developed several versions of the curvature tensor. Which of the following five alternative field equations: (i) $R_{\alpha\beta\gamma\delta}(x)=0$, (ii) $R^\delta_{\ \alpha\beta\gamma}(x)=0$, (iii) $R_{\alpha\beta}(x)=0$, 
1356: (iv) $R(x)=0$, and (v) $G_{\alpha\beta}(x)=0$ should we now choose as the analogue of $\nabla^2\Phi(\vec r)=0$? The field equation we search should be capable of unequivocally
1357: determining the 10 independent components $g_{\alpha\beta}$ of the
1358: metric tensor, and so we can narrow down the possibilities to
1359: expressions (iii) and (v)---a null Ricci or Einstein tensor,
1360: respectively---the only ones suitable from this perspective, because only they have no more nor less than 10 independent components. But which of the two should we choose?
1361: 
1362: For the purposes of this and the next two sections, that is, for the
1363: study of the vacuum, which of the two tensors we choose is irrelevant, because in the vacuum, as we see below, the Einstein tensor reduces to the Ricci tensor. However, when we later on consider solutions inside a distribution of matter (i.e.\ matter as an explicit source), only the choice of the Einstein tensor will be suitable, because only it is capable of leading us to a law for the local conservation of the momentum and energy of the matter source, i.e.\ to the generalization of the classical law of
1364: continuity of matter. This is made possible by the extra mathematical property of Eq.~(\ref{Einstein-divergence-free}) seen earlier, namely, that the Einstein tensor is divergence-free (in the covariant sense). When, after establishing a connection between curvature and matter, we demand that the matter source be divergence-free for conservation purposes, this mathematical property of $G^{\alpha\beta}$ becomes physically relevant.
1365: 
1366: In general, we should then choose the field equation in vacuum to be
1367: \begin{equation} 
1368: G_{\alpha\beta}(x)=0. 
1369: \end{equation} 
1370: In practice, however, the curvature scalar $R$ is null in vacuum,\footnote{Express $G_{\alpha\beta}(x)=0$ in explicit form and contract its indices with the metric tensor to find $-R=0$.} and in that case we can use $R_{\alpha\beta}(x)=0$ for simplicity.
1371: 
1372: The simplest solution to the field equation in vacuum is obtained when the vacuum extends everywhere. Then the curvature tensor is identically null, $R_{\alpha\beta}(x)\equiv 0$, and coordinates can be found such that the solution to the field equation is $g_{\alpha\beta}(x)=\eta_{\alpha\beta}$, that is, flat spacetime.
1373: 
1374: Another exact solution to the spacetime metric can be found outside a spherically symmetric matter distribution of mass $M$. This solution was found by Karl Schwarzschild in 1916. Considerations of spherical symmetry lead to constraints in the possible form that the metric field can take. Using spherical coordinates, the metric can only be of the form 
1375: \begin{equation}
1376: g_{\alpha\beta}(x)=\mathrm{diag}\left( e^{\lambda(x^1)},(x^1)^2, (x^1)^2\sin^2(x^2),-e^{\nu(x^1)}\right),	
1377: \end{equation}
1378: with asymptotic flatness at infinity, i.e.\ with boundary conditions $\lambda(x^1)\rightarrow 0$ and $\nu(x^1)\rightarrow 0$ when $x^1\rightarrow \infty$.
1379: 
1380: The calculation of the 13 non-null Christoffel symbols gives
1381: \begin{align}
1382: \big\{\begin{smallmatrix}1\\11\end{smallmatrix}\big\}&=\frac{1}{2}\lambda', &\qquad \big\{\begin{smallmatrix}1\\22\end{smallmatrix}\big\}&=-x^1e^{-\lambda},\nonumber\\
1383: \big\{\begin{smallmatrix}1\\33\end{smallmatrix}\big\}&=-x^1\sin^2(x^2)e^{-\lambda}, &\qquad \big\{\begin{smallmatrix}1\\44\end{smallmatrix}\big\}&=\frac{1}{2}e^{\nu-\lambda}\nu', \nonumber\\
1384: \big\{\begin{smallmatrix}2\\12\end{smallmatrix}\big\}&=\big\{\begin{smallmatrix}2\\21\end{smallmatrix}\big\}=\frac{1}{x^1}, & \qquad 
1385: \big\{\begin{smallmatrix}2\\33\end{smallmatrix}\big\}&=-\sin(x^2)\cos(x^2),\\
1386: \big\{\begin{smallmatrix}3\\13\end{smallmatrix}\big\}&=\big\{\begin{smallmatrix}3\\31\end{smallmatrix}\big\}=\frac{1}{x^1}, \qquad &
1387: \big\{\begin{smallmatrix}3\\23\end{smallmatrix}\big\}&=\big\{\begin{smallmatrix}3\\32\end{smallmatrix}\big\}=\cot(x^2),\nonumber\\
1388: \big\{\begin{smallmatrix}4\\14\end{smallmatrix}\big\}& =\big\{\begin{smallmatrix}4\\41\end{smallmatrix}\big\}=\frac{1}{2}\nu'. && \nonumber
1389: \end{align}
1390: As a result, the Ricci tensor is diagonal, with components given by
1391: \begin{align}
1392: R_{11}&=\frac{1}{2}\left( \nu''-\frac{1}{2}\lambda'\nu'+\frac{1}{2}\nu'^2-\frac{2}{x^1}\lambda' \right),\\
1393: R_{22}&=-1+e^{-\lambda}\left[1+\frac{1}{2}x^1(\nu'-\lambda') \right],\\
1394: R_{33}&=\sin^2(x^2)R_{22},\\
1395: R_{44}&=-\frac{1}{2}e^{\nu-\lambda}\left( \nu''-\frac{1}{2}\lambda'\nu'+\frac{1}{2}\nu'^2+\frac{2}{x^1}\nu' \right).
1396: \end{align}
1397: To solve the system of four equations $R_{\alpha\alpha}=0$, we consider first
1398: $R_{11}=0$ and $R_{44}=0$ together with the condition of asymptotic flatness. We find straightforwardly that $\nu(x^1)=-\lambda(x^1)$. Next we use this result to simplify and solve $R_{22}=0$. Grouping the variables $x^1$ and $\lambda$ with their respective differentials and integrating once, we find
1399: \begin{equation}
1400: e^\lambda=\left(1-\frac{C}{x^1} \right)^{-1}
1401: \end{equation}
1402: and, therefore,
1403: \begin{equation}
1404: e^\nu=1-\frac{C}{x^1}.
1405: \end{equation}
1406: 
1407: The value of the integration constant $C$ is found by comparing with the Newtonian limit (expressed in natural units). Because\footnote{See next section, Eq.~(\ref{g44}).} 
1408: \begin{equation}
1409: -\left(1-\frac{C}{x^1}\right)=g_{44}\overset{\mathrm{c}}{=}-(1+2\Phi)=-\left(1-\frac{2GM}{c^2r}\right), 
1410: \end{equation}
1411: then $C=2GM/c^2$, where $x^1$ can be identified with the Euclidean distance $r$ only when $x^1 \rightarrow \infty$. The quadratic form in the vacuum outside a spherical matter distribution then is 
1412: \begin{multline}
1413: \D s^2=\epsilon\Bigg\{ \left(1-\frac{2GM}{c^2x^1}\right)^{-1}(\D x^1)^2+
1414: (x^1)^2[(\D x^2)^2+\sin^2(x^2)(\D x^3)^2]-\\\left(1-\frac{2GM}{c^2x^1}\right)(\D x^4)^2 \Bigg\},
1415: \end{multline}
1416: where $0\leq x^2\leq \pi$, $0\leq x^3\leq 2\pi$, and $-\infty< x^4< \infty$. This solution is valid either for (i) $x^1_B<x^1<\infty$, where $x^1_B$ is the radius of the material edge of the matter distribution or (ii) $x^1_S<x^1<\infty$, where $x^1_S=2GM/c^2$ is the Schwarzschild radius of the matter distribution, depending on which of the two values is reached first as we approach from infinity. If $x^1_S=2GM/c^2$ is reached first, the spacetime region in question is called a (non-rotating) \emph{black hole}.
1417: 
1418: \subsection{Tremors in the vacuum}
1419: If not in the general way described (and rejected) at the beginning of the previous section, do the classical gravitational potential $\Phi(\vec r)$ and the Laplace equation $\nabla^2\Phi(\vec r)=0$ have some kind of relativistic counterpart in some special case? 
1420: 
1421: To find out, we consider physically weak tremors (from the spacetime perspective) induced by a small deviation from metric flatness, 
1422: \begin{equation}\label{tremors}
1423: g_{\alpha\beta}(x)=\eta_{\alpha\beta}+\phi_{\alpha\beta},	
1424: \end{equation}
1425: where all $\phi_{\alpha\beta}$ and their derivatives are very small. In this case, it is accurate enough to solve the equation $R_{\alpha\beta}(x)=0$ to first order in $\phi_{\alpha\beta}$. What kind of weak ``gravitational field'' $\phi_{\alpha\beta}(x)$ results from this approximation?
1426: 
1427: To first order ($\sim$) in $\phi_{\alpha\beta}$, we have that the inverse metric tensor is 
1428: \begin{equation}\label{tremors-inverse}
1429: g^{\alpha\beta}(x)=\eta^{\alpha\beta}-\phi^{\alpha\beta}	
1430: \end{equation}
1431: because 
1432: \begin{equation}
1433: g^{\alpha\beta}g_{\beta\gamma} \sim (\eta^{\alpha\beta}-\phi^{\alpha\beta})(\eta_{\beta\gamma}-\phi_{\beta\gamma})=\delta^\alpha_{\ \gamma}+\phi^\alpha_{\ \gamma}-\phi^\alpha_{\ \gamma}=\delta^\alpha_{\ \gamma}.	
1434: \end{equation}
1435: In the following, the metric tensor $g_{\alpha\beta}$ is the only tensor whose indices will be raised and lowered with itself; for any other field, we use the flat-spacetime metric $\eta_{\alpha\beta}$ or $\eta^{\alpha\beta}$.
1436: 
1437: The next step is to find the first-order approximation to the Ricci tensor. From expression (\ref{curvature-tensor}) of the mixed Riemann tensor, we know that the exact Ricci tensor is
1438: \begin{equation}\label{Ricci-tensor}
1439: R_{\alpha\beta}=
1440: \big\{\begin{smallmatrix}\gamma\\\alpha\gamma\end{smallmatrix}\big\}_{,\beta}-
1441: \big\{\begin{smallmatrix}\gamma\\\alpha\beta\end{smallmatrix}\big\}_{,\gamma}+
1442: \big\{\begin{smallmatrix}\epsilon\\\alpha\gamma\end{smallmatrix}\big\}
1443: \big\{\begin{smallmatrix}\gamma\\\epsilon\beta\end{smallmatrix}\big\}-
1444: \big\{\begin{smallmatrix}\epsilon\\\alpha\beta\end{smallmatrix}\big\}
1445: \big\{\begin{smallmatrix}\gamma\\\epsilon\gamma\end{smallmatrix}\big\}.
1446: \end{equation}
1447: Because to first order the Christoffel symbol is
1448: \begin{equation}
1449: \big\{\begin{smallmatrix}\gamma\\\alpha\beta\end{smallmatrix}\big\}\sim
1450: \frac{1}{2}\eta^{\gamma\delta}(\phi_{\alpha\delta,\beta}+\phi_{\beta\delta,\alpha}-\phi_{\alpha\beta,\delta}),
1451: \end{equation}
1452: we then have 
1453: \begin{equation}
1454: \big\{\begin{smallmatrix}\gamma\\\alpha\beta\end{smallmatrix}\big\}_{,\gamma}\sim
1455: \frac{1}{2}(\phi_{\alpha\ ,\beta\gamma}^{\ \gamma}+\phi_{\beta\ ,\alpha\gamma}^{\ \gamma}-\square \phi_{\alpha\beta}) 
1456: \end{equation}
1457: and
1458: \begin{equation}
1459: \big\{\begin{smallmatrix}\gamma\\\alpha\gamma\end{smallmatrix}\big\}_{,\beta} \sim \frac{1}{2} \phi_{,\alpha\beta}.
1460: \end{equation}
1461: To first order, disregarding products of $\phi_{\alpha\beta}$ or its derivatives, the Ricci tensor is
1462: \begin{equation}
1463: R_{\alpha\beta}\sim
1464: \big\{\begin{smallmatrix}\gamma\\\alpha\gamma\end{smallmatrix}\big\}_{,\beta}-
1465: \big\{\begin{smallmatrix}\gamma\\\alpha\beta\end{smallmatrix}\big\}_{,\gamma}\sim
1466: \frac{1}{2}(\phi_{,\alpha\beta}-\phi_{\alpha\ ,\beta\gamma}^{\ \gamma}-\phi_{\beta\ ,\alpha\gamma}^{\ \gamma}+\square \phi_{\alpha\beta}).
1467: \end{equation}
1468: Because we find ourselves outside any material distribution, we set \mbox{$R_{\alpha\beta}=0$} to get
1469: \begin{equation}
1470: \square \phi_{\alpha\beta}+\phi_{,\alpha\beta}-\phi_{\alpha\ ,\beta\gamma}^{\ \gamma}-\phi_{\beta\ ,\alpha\gamma}^{\ \gamma}=0.\footnote{Taking advantage of the gauge freedom of this equation, it can be recast simply as $\square \phi_{\alpha\beta}=0$ using the gauge $\chi_{\alpha\ ,\gamma}^{\ \gamma}=0$, where $\chi_{\alpha\beta}=\phi_{\alpha\beta}-\eta_{\alpha\beta}\phi/2$. In some more detail, the linearized equation is invariant with respect to the gauge transformation $\phi_{\alpha\beta} \rightarrow \tilde\phi_{\alpha\beta}=\phi_{\alpha\beta}+(\lambda_{\alpha,\beta}+\lambda_{\beta,\alpha})/2$, where $\lambda_\alpha(x)$ is an arbitrary vector field. When $\square \lambda_{\alpha}=2\chi_{\alpha\ \beta}^{\ \beta}$, the gauge condition is satisfied (i.e.\ $\tilde\chi_{\alpha\ ,\gamma}^{\ \gamma}=0$) and the linearized equation simplified with respect to the new field $\tilde\phi_{\alpha\beta}$. This is the starting point of the study of gravitational waves.}
1471: \end{equation}
1472: 
1473: This is the linearized general-relativistic equation for a perturbation $\phi_{\alpha\beta}(x)$ of the metric field. Its form coincides with the Bargmann-Wigner equation for a field in flat spacetime associated with a massless spin-2 particle. On account of its formal correspondence with the linearized general-relativistic equation for the metric perturbation $\phi_{\alpha\beta}(x)$, this particle is sometimes called a ``graviton.'' However, the correspondence appears to be only formal, because the Bargmann-Wigner field equation was derived for a field in Minkowski spacetime, which is flat and has therefore nothing to do with gravitation. As hard as one might try to establish the resemblance between a ``graviton'' and other quantum particles, a ``graviton'' does not stand on the same conceptual footing as quantum particles, because the field associated with it is not an \emph{ordinary field on} flat spacetime but is itself (part of) the \emph{metric field of} spacetime.
1474: 
1475: To see what $\phi_{\alpha\beta}(x)$ may have to do with the Newtonian potential $\Phi(\vec r)$, we start by considering a ``slow particle'' such that the components $V^a$ ($a=1,2,3$) of its four-velocity $V^\alpha=\D x^\alpha/\D s$ are very small and, therefore, $g_{\alpha\beta}V^\alpha V^\beta=-1$ becomes 
1476: \begin{equation}
1477: g_{44}(V_4)^2 \sim -1,	
1478: \end{equation}
1479: from where 
1480: \begin{equation}\label{V4}
1481: V^4\sim \frac{1}{\sqrt{-g_{44}}}.	
1482: \end{equation}
1483: When the slow-particle condition is applied to the geodesic equation for $\alpha=a$, we find 
1484: \begin{equation}
1485: \frac{\D V^a}{\D s}\sim -\big\{\begin{smallmatrix}a\\44\end{smallmatrix}\big\}(V^4)^2. 	
1486: \end{equation}
1487: But because 
1488: \begin{equation}
1489: \frac{\D V^a}{\D s}=\frac{\D V^a}{\D x^\alpha} V^\alpha \sim \frac{\D V^a}{\D x^4}V^4, 	
1490: \end{equation}
1491: we get 
1492: \begin{equation}
1493: \frac{\D V^a}{\D x^4}\sim -\big\{\begin{smallmatrix}a\\44\end{smallmatrix}\big\}V^4. 
1494: \end{equation}
1495: The left-hand side of this equation is comparable with the Newtonian acceleration $\D v_a/\D t$.
1496: 
1497: Assuming next that the full spacetime metric is ``quasistatic,'' i.e.\ that $g_{\alpha\beta,4}$ is much smaller than $g_{\alpha\beta,a}$, we get
1498: \begin{equation}
1499: \big\{\begin{smallmatrix}a\\44\end{smallmatrix}\big\} \sim -\frac{1}{2}g^{a\alpha}\frac{\partial g_{44}}{\partial x^\alpha} \sim -\frac{1}{2}g^{ab} \frac{\partial g_{44}}{\partial x^b}.	
1500: \end{equation}
1501: Using this result, we find 
1502: \begin{equation}
1503: \frac{\D V^a}{\D x^4}\sim  \frac{1}{2}g^{ab} \frac{\partial g_{44}}{\partial  x^b} V^4.	
1504: \end{equation}
1505: Using Eq.~(\ref{V4}), we come to the expression 
1506: \begin{equation}
1507: \frac{\D V^a}{\D x^4} \sim -g^{ab} \frac{\partial \sqrt{-g_{44}}}{\partial x^b}.	
1508: \end{equation}
1509: 
1510: Finally, because we are dealing only with tremors in the vacuum, 
1511: \begin{equation}
1512: g^{44}=1-\phi^{44}	
1513: \end{equation}
1514: on account of Eq.~(\ref{tremors-inverse}). The counterpart of the Newtonian acceleration thus becomes 
1515: \begin{equation}
1516: \frac{\D V^a}{\D x^4} \sim -(\eta^{ab}-\phi^{ab}) \frac{\partial (1-\phi_{44}/2)}{\partial x^b} \sim \frac{1}{2} \eta^{ab} \frac{\partial\phi_{44}}{\partial x^b};	
1517: \end{equation}
1518: that is
1519: \begin{equation}
1520: \frac{\D V_a}{\D x^4}\sim \frac{1}{2}\frac{\partial \phi_{44}}{\partial x^a}.
1521: \end{equation}
1522: Comparing this expression with the Newtonian acceleration \begin{equation}
1523: 	\frac{\D v_a}{\D t}=g_a(\vec r)=-\frac{\partial \Phi(\vec r)}{\partial x^a},
1524: \end{equation} 
1525: we discover the correspondences (valid only in the context of the approximations made)
1526: \begin{equation}
1527: \Phi \overset{\mathrm{c}}{=} -\frac{1}{2}\phi_{44} 
1528: \end{equation}
1529: and
1530: \begin{equation}\label{g44}
1531: 	g_{44} \overset{\mathrm{c}}{=} -(1+2\Phi).
1532: \end{equation}
1533: 
1534: A rough estimate of how $\Phi$ compares with unity in Eq.~(\ref{g44}) at the surface of the earth, where $\Phi(r)=-GM/c^2r$,\footnote{In natural units, as used so far, $\Phi$ is a dimensionless number. On the other hand, $-GM/r$ has SI units $L^2T^{-2}$. To convert a quantity from SI units to natural units ($c=1$), we must divide it by $c^2$ as we just did; to convert a quantity from natural units to SI units, we must multiply it by $c^2$.} gives 
1535: \begin{equation}
1536: 	\Phi \sim -10^{-9} 
1537: 	\qquad \mathrm{and} \qquad
1538: 	\frac{\partial \Phi}{\partial x^a} \sim 10^{-16}\ \mathrm{m}^{-1}.
1539: \end{equation} 
1540: From this, the acceleration at the surface of the earth turns out to be the familiar result
1541: \begin{equation}
1542: c^2 \frac{\partial \Phi}{\partial x^a} = 10\ \mathrm{ms}^{-2}.
1543: \end{equation}
1544: The relative size of the gravitational tremors with respect to the flat-spacetime background is comparable with ripples rising 1 centimetre above the surface of an (unearthly) lake 10,000 kilometres in depth.
1545: 
1546: We find, moreover, that in this correspondence the Laplacian $\nabla^2 \Phi$ of the Newtonian potential is the 44-component of the Ricci tensor,
1547: \begin{multline}
1548: R_{44}\sim \frac{1}{2}(\phi_{,44}-2\phi_{4\ ,4\gamma}^{\ \gamma}+\square \phi_{44})\sim
1549: \frac{1}{2}\eta^{\alpha\beta}\phi_{44,\alpha\beta}\sim \frac{1}{2}\eta^{ab}\phi_{44,ab}=\\
1550: \frac{1}{2}\sum_{a=1}^3 \phi_{44,aa}=\frac{1}{2}\nabla^2\phi_{44}\overset{\mathrm{c}}{=}-\nabla^2\Phi.
1551: \end{multline}
1552: Thus, the Laplace equation $\nabla^2\Phi=0$ for the Newtonian potential in vacuum corresponds to $R_{44}=0$.
1553: 
1554: The special role of coordinate $x^4$ in the way the correspondence with classical physics was established deserves special mention. Early on, $x^4$ played a crucial part in the local characterization of the past, the present, and the future in relativity theory. Now it seems that, in this approximation, $x^4$ has become the absolute time $t$ of classical physics. The likelihood of this conclusion is strengthened by a look at Einstein's \citeyear{Einstein:1952b} original paper, where, on making this approximation, he writes that ``we have set $\D s=\D x_4=\D t$'' (p.~158). But is this true \emph{to first order in} $\Phi$? It is not, because we have \mbox{$\Delta s=\Delta t$} in the Newtonian case, but
1555: \begin{equation}
1556: 	\frac{\D x^4}{\D s}=V^4 \sim \frac{1}{\sqrt{1-\phi_{44}}} \sim 1+\frac{1}{2}\phi_{44} \overset{\mathrm{c}}{=} 1-\Phi \neq 1.
1557: \end{equation} 
1558: We conclude that the identification of the fourth coordinate $x^4$ with absolute time $t$ is not possible even in the relativistic approximation to Newtonian physics. If, in spite of this, the association $x^4=ct$ still wants to be made, one must remember that in this case $t$ is just a notation---the misleading name of a spacetime coordinate, not absolute time.
1559: 
1560: \subsection{The cosmic vacuum}
1561: So far we have studied the field equations $G_{\alpha\beta}=0$ or $R_{\alpha\beta}=0$, where the constant $\Lambda$ is null. In neglecting this constant, we have assumed that $|\Lambda|$ is small enough, but now we would like to make sure that this is correct by setting an upper bound for $|\Lambda|$. 
1562: 
1563: We start by noting that, because
1564: \begin{equation}
1565: 	g^{\alpha\beta}\mathcal{G}_{\alpha\beta}=R-2R+4\Lambda=0, 
1566: \end{equation}
1567: then $R=4\Lambda$, and the field equation reduces to 
1568: \begin{equation}\label{Einstein-lambda}
1569: 	R_{\alpha\beta}-\Lambda g_{\alpha\beta}=0. 
1570: \end{equation}
1571: Using again the correspondence $R_{44}\overset{\mathrm{c}}{=}-\nabla^2\Phi$ in the context of the Newtonian approximation, to first order in $\Lambda$ we have 
1572: \begin{equation}
1573: 	\Lambda g_{44}\overset{\mathrm{c}}{=}-\Lambda(1+2\Phi)\sim -\Lambda.
1574: \end{equation} 
1575: If we now take $\Lambda g_{44}$ \emph{as if} it was a source of curvature (instead of being a curvature term itself), we can write the Poisson equation 
1576: \begin{equation}\label{Poisson-lambda}
1577: 	\nabla^2\Phi=\frac{4\pi G\rho}{c^2}+\Lambda 
1578: \end{equation}
1579: for the Newtonian potential \emph{inside} a medium of density $\rho+\Lambda c^2/4\pi G$ (expressed in SI units). Here $\rho$ stands for the ordinary density of matter, while $\Lambda c^2/4\pi G$ stands for the mass density of a new source of gravitation pervading all space in the manner of an ubiquitous ether, present also inside matter. 
1580: 
1581: We see from Eq.~(\ref{Poisson-lambda}) that, if the classical dynamics inside a system of density $\rho$ is well described by Newtonian theory, then $|\Lambda|\ll 4\pi G \rho/c^2$, obtaining an upper bound for this new constant. The more dilute a system inside of which Newtonian theory holds, the more stringent the upper bound set on $|\Lambda|$ is and, consequently, the smaller it must be. 
1582: 
1583: For a conservative estimate, we take our galaxy as a system inside which Newtonian mechanics can be applied successfully. Given that the average density of the Milky Way is approximately $\rho= 1.7 \times 10^{-21}\ \mathrm{kg/m}^3$,\footnote{Expressed differently, $\rho= 10^6 \times 1.7 \times 10^{-27}\ \mathrm{kg/m}^3$, that is, $10^6$ hydrogen-atom masses per cubic meter. This is one million times the density of the intergalactic ``vacuum.''} we find, $|\Lambda|\ll 1.4 \times 10^{-30}\ \mathrm{s}^{-2}$. To express this bound in a more familiar way in terms of natural units, we divide by $c^2$ to get $|\Lambda|\ll 1.6 \times 10^{-47} \mathrm{m}^{-2}$.  
1584: 
1585: It has been conjectured by other methods that a more accurate but less conservative estimate is $|\Lambda|\leq 10^{-35}\ \mathrm{s}^{-2}$ or, more commonly in natural units, $|\Lambda|\leq 10^{-52}\ \mathrm{m}^{-2}$. On account of Eq.~(\ref{Poisson-lambda}), the gravitational effect of a constant $\Lambda$ of this magnitude would begin to be felt inside a material medium of density $\rho\leq 1.7 \times 10^{-26}\ \mathrm{kg/m}^3$, i.e.\ 10 hydrogen-atom masses per cubic metre. This value corresponds to the density of the \emph{intergalactic medium}. Thus, in order for $\Lambda$ to have non-negligible effects, we must consider cosmological distances---hence its name, the \emph{cosmological constant}.
1586: 
1587: To see what kind of gravitational effects are caused by the cosmological constant---whose so far enigmatic source, we recall, pervades all space---in the absence of any other material medium (i.e.\ $\rho=0$), we consider a point-like\footnote{Or spherical, in which case we are only interested to find the gravitational potential in the vacuum outside it.} mass $M$ surrounded by a perfect vacuum in the ordinary sense of the word. In the Newtonian approximation under consideration, we should then solve Eq.~(\ref{Poisson-lambda}).
1588: 
1589: Since $\Lambda$ is constant and the central mass point-like (or spherical), we solve for $\Phi(r)$ in a spherically symmetric context:
1590: \begin{equation}
1591: \frac{1}{r^2}\frac{\D}{\D r}\left( r^2\frac{\D \Phi}{\D r} \right)=\Lambda.
1592: \end{equation}
1593: Integrating twice and absorbing one constant into the potential, we find
1594: \begin{equation}
1595: \Phi(r)=-\frac{C'}{r}+\frac{1}{6}\Lambda r^2,
1596: \end{equation}
1597: which leads to a gravitational force on a particle of mass $m$ of the type
1598: \begin{equation}
1599: \vec F_g=-m\frac{\D \Phi}{\D r}\hat r=-m\left( \frac{C'}{r^2}+\frac{1}{3}\Lambda r\right) \hat r,
1600: \end{equation}
1601: where $C'=GM/c^2$. If $\Lambda$ is positive, it generates an extra attractive gravitational force between the central mass and a test mass. More interestingly, if it is negative, it generates an extra repulsive gravitational force between the two masses. In both cases, the extra force increases linearly as we recede from the central mass. For negative $\Lambda$, there exists then a distance at which the attraction from the centre is balanced by the repulsion caused by the so far enigmatic source of the cosmological constant. Given the upper bound $|\Lambda|\leq 10^{-52}\ \mathrm{m}^{-2}$, we can get a feel for the distance involved when the central mass is the sun ($M\approx 2\times 10^{30}\ \mathrm{kg}$). We find that, in an otherwise perfect universal vacuum, the cosmological constant is so small that its repulsive force cancels out the attraction of the sun at a distance 
1602: \begin{equation}
1603: r=\sqrt[3]{\frac{3GM}{\Lambda c^2}}\approx 3.5 \times 10^{18}\ \mathrm{m} \approx 375\ \mathrm{ly}. 	
1604: \end{equation}
1605: The value of this result is only pedagogical. Nothing near the perfect universal vacuum here assumed exists either in the solar system or in the galaxy. We should go as far out as intergalactic space to find a medium of low enough density that the cosmological constant can exert its influence at all. This is a distance of millions or tens of millions of light-years---four to five orders of magnitude greater than the mere 375 light-years just obtained.
1606: 
1607: Leaving classical theory behind, we can now consider $\Lambda$ and find an exact solution to the generalized Einstein field equation $\mathcal{G}_{\alpha\beta}=0$, which, outside an ordinary material distribution, we saw reduces to Eq.~(\ref{Einstein-lambda}). Following essentially the same method as for the Schwarzschild solution with the added extra source terms,\footnote{To solve the resulting differential equation by integration, rewrite $(1+x^1 \nu')e^\nu$ as the derivative of a product.} we get
1608: \begin{equation}
1609: e^\nu=e^{-\lambda}=\left[1-\frac{C}{x^1}+\frac{1}{3}\Lambda (x^1)^2\right].
1610: \end{equation}
1611: To consider a cosmos everywhere void of matter, we set $C=0$, since the origin of this constant is a localized matter distribution. As a result, we get
1612: \begin{multline}\label{de-Sitter}
1613: \D s^2=\epsilon \Bigg\{ \left[1+\frac{1}{3}\Lambda (x^1)^2\right]^{-1}(\D x^1)^2+
1614: (x^1)^2[(\D x^2)^2+\sin^2(x^2)(\D x^3)^2]-\\\left[ 1+\frac{1}{3}\Lambda (x^1)^2\right](\D x^4)^2 \Bigg\}.
1615: \end{multline}
1616: This solution was found by Willem de Sitter in 1917. For $\Lambda < 0$, the value $x^1_H=\sqrt{-3/\Lambda}$ characterizes the radius of the \emph{event horizon} of the spacetime in question, beyond which no information can reach an observer.
1617: 
1618: When all is said and done, what are we to make of the cosmological constant? What physical meaning are we to attach to it? When the generalized field equation is written in the form $R_{\alpha\beta}=\Lambda g_{\alpha\beta}$, one gets the impression that the cosmological constant is a material source of spacetime curvature. But since it certainly is not related to matter, we appear to be confronted by a mysterious kind of stuff that is all-pervading and whose origin we know nothing of, but which has active physical effects---a demiether. Evoking this sense of wonder, the presumed physical stuff behind the cosmological constant has been named ``dark energy.'' But is in question here a physical mystery or only a mysterious physical interpretation? The power of psychological inducement perhaps deserves more attention that is usually paid to it. If the generalized field equation is instead simply written in the form $R_{\alpha\beta}-\Lambda g_{\alpha\beta}=0$, the impression vanishes, and one is free to see the cosmological constant simply as an \emph{extra term of spacetime curvature} possibly relevant in a cosmic near-vacuum. \label{lambda-solution}
1619: 
1620: In the context of present theories, consideration of a negative cosmological constant is deemed justified by the observed cosmological redshift of distant galaxies. If this interpretation of the cosmological redshift stood the test of time in the light of hypothetical better cosmological theories to come, and the need for a cosmological constant became thus better secured, we could then, in the light of the previous paragraph, regard it as nothing more than a \emph{new fundamental constant} related to the curvature of spacetime. If we do not worry  and try to explain the origin of the gravitational constant $G$, the speed of light $c$, and Planck's constant $h$, why should we any more worry about the origin of the \emph{curvature constant} $\Lambda$?
1621: 
1622: \subsection{The dust-filled cosmos}
1623: 
1624: The next and final step is to generalize the field equation for the
1625: case of a non-empty cosmos. Now, instead of the homogeneous field
1626: equation $G_{\alpha\beta}(x)=0$ in vacuum, we expect an
1627: inhomogeneous equation with a source term characterizing the matter
1628: distribution in analogy with the classical Poisson equation
1629: \begin{equation}
1630:       \nabla^2\Phi=4\pi G\rho.
1631: \end{equation} 
1632: The source term is usually expressed as
1633: $\kappa  T_{\alpha\beta}$, where $\kappa $ is a constant. Further,
1634: because $G_{\alpha\beta}$ is symmetric and divergence-free (a choice based on a mathematical property so far without physical relevance), $T_{\alpha\beta}$ should be so too.
1635: 
1636: A simple case is that of a cosmos filled with matter resembling an
1637: ideal fluid whose pressure is null (or approximately null) and its constituent
1638: particles are free, interacting with each other only gravitationally. This kind of fluid is usually called dust. When we identify the fluid particles with galaxies, we obtain a useful application of this idea for the study of the cosmos as a whole. Since we are dealing with a continuous material distribution, instead of considering its four-momentum \mbox{$P^\alpha(x)=mV^\alpha(x)$}, it is more appropriate to characterize this fluid by means of its invariant \emph{matter density} $\rho(x)$ and four-velocity $V^\alpha(x)$. Together they form a field $\rho(x)V^\alpha(x)$, called the \emph{four-momentum-density field}, the natural counterpart of the four-momentum field $mV^\alpha(x)$.
1639: 
1640: However, $\rho V^\alpha$ is not yet a viable choice for the source
1641: term $T^{\alpha\beta}$ we seek, because this should be a second-order tensor. Since we are looking for a symmetric second-order tensor and the only fields
1642: characterizing the ideal pressureless fluid are its density and its
1643: four-velocity, we do not seem to have any other options but to
1644: propose that 
1645: \begin{equation}
1646: 	T^{\alpha\beta}=\rho V^\alpha V^\beta
1647: \end{equation}
1648: is the source term we are looking for. If it is a bad choice, it should
1649: next lead us to physically unpleasant results.
1650: 
1651: Because $T^{\alpha\beta}$ should be divergence-free, we have
1652: \begin{equation}\label{divergence-free-matter} 
1653: T^{\alpha\beta}_{\ \ \ ;\beta}= (\rho V^\beta)_{;\beta}V^\alpha+ \rho V^\beta
1654: V^\alpha_{\ \ ;\beta}=0. 
1655: \end{equation} 
1656: Multiplying by $V_\alpha$ on both sides, we get 
1657: \begin{equation} 
1658: (\rho V^\beta)_{;\beta}V_\alpha V^\alpha+\rho V^\beta V_\alpha V^\alpha_{\ \ ;\beta}=0. \end{equation}
1659: The second term is null\footnote{Writing
1660: $V_\alpha V^\alpha_{\ \ ;\beta}=(V_\alpha V^\alpha)_{;\beta}-V^\alpha
1661: V_{\alpha;\beta}$, we find that the first term is null because
1662: $V^\alpha V_\alpha=-1$ and that the second is
1663: $-g^{\alpha\gamma}V_\gamma V_{\alpha;\beta}=-V_\gamma V^\gamma_{\ \ ;\beta}$. Therefore, $V_\alpha V^\alpha_{\ \ ;\beta}=0$.} and, because
1664: $V^\alpha V_\alpha=-1$, the equation reduces to 
1665: \begin{equation}
1666: (\rho V^\beta)_{;\beta}=0. 
1667: \end{equation} 
1668: This is a conservation equation for the four-momentum density, and it is the
1669: general-relativistic generalization of the classical continuity equation 
1670: \begin{equation}
1671: 	\vec\nabla\cdot (\rho\vec v)+\frac{\partial \rho}{\partial t}=0, 
1672: \end{equation}
1673: expressing the local conservation of matter. This is a physically pleasant result. Furthermore, it justifies the choice of $G_{\alpha\beta}$ over $R_{\alpha\beta}$ from a physical point of view.
1674: 
1675: Returning to Eq.~(\ref{divergence-free-matter}), since the first term is null, it follows that so must be the second term. Because neither $\rho$ nor $V^\beta$ are in general null, then 
1676: \begin{equation}
1677: 	V^\alpha_{\ \ ;\beta}=0. 
1678: \end{equation}
1679: In words, the \emph{free particles} composing the ideal pressureless fluid follow spacetime geodesics. And this is as it should be---another physically pleasant result. Our presumption about the form of $T^{\alpha\beta}$ is now confirmed.
1680: 
1681: In this ideal-fluid picture, the field equation is then
1682: \begin{equation}
1683: 	R^{\alpha\beta}-\frac{1}{2}g^{\alpha\beta}R=\kappa \rho V^\alpha V^\beta.
1684: \end{equation}
1685: Contracting the indices with $g_{\alpha\beta}$, we find that
1686: $R=\kappa \rho$. The field equation is thus simplified to
1687: \begin{equation}
1688: 	R^{\alpha\beta}=\kappa \rho\left[V^\alpha V^\beta+\frac{1}{2}g^{\alpha\beta}\right].
1689: \end{equation}
1690: To determine $\kappa $ we resort again to the correspondence
1691: with Newtonian theory. The 44-component of the Ricci
1692: tensor gives 
1693: \begin{equation}
1694: 	-\nabla^2\Phi \overset{\mathrm{c}}{=}R^{44}=\kappa \rho \left[(V^4)^2+\frac{1}{2}g^{44}\right]. 
1695: \end{equation}
1696: To the lowest non-trivial order in $\kappa \rho$, $(V^4)^2\sim
1697: 1$ and $g^{44}\sim -1$, and so 
1698: \begin{equation}
1699: 	\nabla^2\Phi=-\frac{1}{2}\kappa \rho.
1700: \end{equation}
1701: Comparison with Poisson's equation $\nabla^2\Phi=4\pi G\rho$ leads
1702: finally to the result\footnote{In SI units, the constant of proportionality is $\kappa =-8\pi G/c^4$.} 
1703: \begin{equation}
1704: 	\kappa =-8\pi G.
1705: \end{equation} 
1706: In the pressureless-fluid picture here considered, the field equation finally becomes
1707: \begin{equation} 
1708: G^{\alpha\beta}=-8\pi G\rho V^\alpha V^\beta.
1709: \end{equation}
1710: 
1711: In his original paper, Einstein \citeyear{Einstein:1952b} holds that the general expression of this equation, namely,
1712: \begin{equation}
1713: G^{\alpha\beta}=-8\pi G T^{\alpha\beta}
1714: \end{equation}
1715: is indeed the field equation describing ``the influence of the gravitational field on all processes, without our having to introduce any new hypothesis whatever'' (pp.~151--152). This is to say that the equation describes the gravitational interaction of \emph{all} fields related to matter and radiation (velocity field, mass-density field, pressure field, electromagnetic field, etc.), so long as they are described by a matter-radiation tensor $T_{\alpha\beta}$ that is divergence-free. Finally, when we consider the cosmological constant as well, the most general field equation for gravitational interactions is
1716: \begin{equation}
1717: \mathcal{G}^{\alpha\beta}=-8\pi G T^{\alpha\beta}.
1718: \end{equation}
1719: The curvature of spacetime becomes in this way connected to the distribution of matter and radiation in it.
1720: 
1721: \subsection{The problem of cosmology}\label{Problem-cosmology}
1722: ``The problem of cosmology,'' said Einstein \citeyear{Einstein:1982} in his Kyoto lecture, ``is related to the geometry of the universe and to time'' (p.~47). We finish this chapter with a brief analysis of the insight contained in this deceptively simple statement.
1723: 
1724: In order for the study of the cosmos as a whole to be practically possible, standard cosmology is forced to make several assumptions. Firstly, the assumption is made that galaxies, conceived as material particles, are in free fall, interacting with each other only gravitationally (cf.\ previous section). Next follows a much less innocuous supposition, according to which cosmic space is locally isotropic, i.e.\ looks the same in all directions from anywhere at any one time. Against this supposition, it might be argued that an approximation in which the cosmos is materially smooth is too unrealistic, and that, moreover, the cosmos may only look approximately smooth from our own vantage point, which is both severely restricted spatially to the Milky Way and temporally to a few thousand years of astronomical observations. Martin \citeyear{Martin:1996}, for example, readily admits that ``we have no way of telling that the Universe is at all homogeneous and isotropic when seen from other places or at other times'' (p.~143). 
1725: 
1726: Indeed, all this and more could be argued against the cosmological premise of local isotropy. But it is the \emph{unspoken} presumption lying beneath this cosmological premise that must be regarded with a much higher degree of suspicion. For what is ``cosmic space at one time'' supposed to mean in general relativity? How shall we separate the whole universal four-dimensional spacetime into universal spaces at a given time, when this cannot even be done for smaller spacetime regions? What kind of time is it that the existence of cosmic space implies?  
1727: 
1728: In cosmology, the separation of spacetime into a spatial part and a temporal part is translated into geometric language by expressing the line element in the form
1729: \begin{equation}\label{space-time}
1730: 	\D s^2=\epsilon\left[g_{ab}(x)\D x^a\D x^b-(\D x^4)^2\right]. 
1731: \end{equation}
1732: Once this split is made, the local isotropy of cosmic space is expressed geometrically by recasting the (non-invariant) spatial part $Q_s=:g_{ab}(x)\D x^a\D x^b$ of this equation in spherical coordinates as 
1733: \begin{equation}
1734: 	Q_s=e^{A(x^4,x^1)} \left[ (\D x^1)^2+(x^1)^2 (\D x^2)^2+
1735: 	(x^1)^2\sin^2(x^2)(\D x^3)^2 \right].
1736: \end{equation}
1737: 
1738: How are the coordinates interpreted in this space-time setting? Coordinate $x^4$ is taken to represent ``cosmic time'' and denoted $x^4=c\tau$. ``Cosmic time'' $\tau$ works as a ``universal time,'' equal for all ``cosmic observers'' travelling along galactic worldlines. This means that ``cosmic observers'' throughout the cosmos measure the same \emph{absolute} ``cosmic separation'' 
1739: \begin{equation}\label{cosmic-time}
1740: 	\D s^2=c^2 \D\tau^2. 
1741: \end{equation}
1742: Cosmic space is then nothing but the three-dimensional hypersurfaces orthogonal to the galactic worldlines at $\tau=\tau_{\mathrm{now}}$; i.e.\ cosmic space is an instantaneous slice of spacetime at absolute ``cosmic time'' $\tau_{\mathrm{now}}$. But is the assumption of ``cosmic time'' realistic in a general-relativistic world? 
1743: 
1744: As for $x^1$, it is interpreted as the radial distance to an event from a cosmic observer and denoted $x^1=\rho$. The supposition is then made that the function $A(\tau,\rho)$ can be separated into a part $B(\tau)$ connected only with ``cosmic time'' and to be related to the expansion of cosmic space, and a part $C(\rho)$ connected only with matter and to be related to the curvature of cosmic space, i.e.\ 
1745: \begin{equation}
1746: 	A(\tau,\rho)=B(\tau)+C(\rho).
1747: \end{equation}
1748: Since the cosmos is spatially homogeneous at a given ``cosmic instant,'' the spatial part of the tensor describing its matter and radiation content must be of the form \mbox{$T^a_{\ b}=\mathrm{constant}\ \delta^a_{\ b}$}.
1749: 
1750: On all these assumptions and with suitable integration constants, the field equation leads to the solution
1751: \begin{equation}
1752: \D s^2=\epsilon \left[ e^{B(\tau)}\left( 1+\frac{k\rho^2}{4\rho_0^2}\right)^{-2}(\D\rho^2+\rho^2\D\theta^2+\rho^2\sin^2(\theta)\D\phi^2)-\\
1753: c^2\D\tau^2\right],
1754: \end{equation}
1755: where $k=0,1$ or $-1$. An equivalent solution was found by Howard Robertson in 1935 and by Arthur Walker in 1936, and it is the geometric foundation of relativistic cosmology. It is interpreted to picture a flat, positively, or negatively curved locally isotropic space---pictured by $\exp[C(\rho)]$---``expanding'' in ``cosmic time''---pictured by $\exp[B(\tau)]$.
1756: 
1757: A similar interpretation is usually made of the de Sitter solution (\ref{de-Sitter}) for the cosmic vacuum we saw earlier. By means of a change of coordinates, 
1758: \begin{equation}\left\{
1759: \begin{array}{lll}
1760: 	x^1&=&\rho e^{c\tau\sqrt{-\Lambda/3}} \\
1761:       x^4&=&c\tau+\frac{1}{2}\sqrt{-3/\Lambda}\ln\left[1+(\Lambda/3)\rho^2 e^{2c\tau\sqrt{-\Lambda/3}}\right]
1762: \end{array} \right.,
1763: \end{equation}
1764: the de Sitter line element can be expressed as a spatially isotropic part that, for $\Lambda<0$, ``expands'' in ``cosmic time'' $\tau$:
1765: \begin{equation}
1766: \D s^2=\epsilon \left\{ e^{2c\tau\sqrt{-\Lambda/3}}[\D\rho^2+\rho^2\D\theta^2+\rho^2\sin^2(\theta)\D\phi^2]-c^2\D\tau^2\right\}.
1767: \end{equation}
1768: This pictures, as it were, an expanding empty cosmos or ``motion without matter.'' It stands in contrast to Einstein's proposed solution of a matter-filled static universe or ``matter without motion.''
1769: 
1770: We can now see that when Einstein deemed the problem of cosmology to be tied to the geometry of the universe (i.e.\ of space) and to time, he was dead right. Moreover, realizing now that the former problem depends on the latter, we recognize that the problem of cosmology really boils down to its presumptions about the \emph{nature of time}.
1771: 
1772: The problem of space refers to the question of the spatial curvature of the locally isotropic cosmos. But, as we just saw, in order to be able to even talk about spatial properties, spacetime should be separable into a spatial part $g_{ab}(x)\D x^a\D x^b$ and a temporal part $g_{44}(x^1,x^2,x^3)(\D x^4)^2$. This is feasible, for example, for flat spacetime or for the spacetime around an isolated spherical mass distribution, but it is not a general property of spacetime. In fact, it seems \emph{least} circumspect to assume the validity of the space-time split in the cosmological setting, since we are here dealing with the \emph{totality} of spatiotemporal events, which is therefore the \emph{most general} case there is. 
1773: 
1774: This unwarranted space-time split of cosmology signals a regress to a Newtonian worldview: an absolute cosmic space whose existence requires that of absolute time. Through ``cosmic time,'' the universe becomes endowed with an absolute present (absolute space) and a history. The validity of the space-time split in cosmology, however, runs counter to the basic lessons of general relativity, and it is a presumption whose truth can only be hoped for and guessed at. The psychologically suggestive notation $x^4=c\tau$ does not by itself turn $\tau$ into a physically meaningful absolute time.
1775: 
1776: What motivates this cosmological regress to the classical view of time? Ultimately, only our human quest for understanding. The problem of cosmology lies essentially in the suppositions it must make about the nature of time in order to create a theoretical picture of something as vast, complex, and remote as the cosmos simple enough to be humanly manageable and understandable. It lies in the reversed logic of a situation in which we fit the great wild cosmos to our existing, familiar tools of thought instead of trying to develop new tools of thought that suit and do justice to the cosmos. 
1777: 
1778: It seems ironic---although on second thoughts only natural---that the more we have learnt, the deeper and farther we have probed the natural world in pursuit of the disclosure of one more of its secrets, the closer we have come to experience the deficiencies of our own cognitive tools. As if the deeper we sink in an oceanic expedition, the more the dimming light and building pressure interfere with our intended objective study of the ocean's fauna---until, in the utmost darkness of the inhospitable depths, we become oblivious to all fauna, only remaining aware of the feeble light our torch dispenses and the low oxygen supply in our tank. Dimming light leads to virtual blindness and oxygen deficiency to brain damage. The first symptoms of these conditions are easy for a diver to self-diagnose. The same, regrettably, cannot be said of their counterparts in the human expedition to the intellectually inhospitable depths of the cosmos. 
1779: 
1780: In an often-quoted passage, Eddington captured, with rare insight, this human dilemma. Concluding his book on spacetime and gravitation, he wrote:
1781: \begin{quote}
1782: [W]e have found that where science has progressed the farthest, the mind has but regained from nature that which the mind has put into nature.
1783: 
1784: We have found a strange foot-print on the shores of the unknown. We have devised profound theories, one after another, to account for its origin. At last, we have succeeded in reconstructing the creature  that made the foot-print. And Lo! it is our own. \mbox{\cite[pp.~200--201]{Eddington:1920}}
1785: \end{quote}
1786: 
1787: The geometric analysis of clock time, leading us to the classical and relativistic worldviews, is now complete.
1788: 
1789: What next?
1790: 
1791: