Search This Blog

Tuesday, December 24, 2013

Wondering About The Ultimate Beginnings

I hope that chapter eight (my previous blog) has given you a reasonable feel for what is commonly called “deep” time, that is, the geological and biological evolution of our own planet. Given that we do reside here and did evolve here, that was a pretty good place to start. But the universe as a whole did not begin with that of our own world and the rest of our solar system; there is an approximately nine billion year gap between those two events. Besides, as already mentioned, events in the very early universe were quite different than those later on, simply because back then, the cosmos was smaller, denser, and hotter, and the laws of physics needed to understand it were necessarily different as well.
 
Imagine yourself on a journey backward in time, back not just before our Earth and solar system, not just before our Milky Ways galaxy began to form, but much, much further than that, to a time before the first stars and galaxies began to take form. We are at a point in the universe’s evolution where it can be modeled fairly accurately as a gas – a gas composed almost entirely out of hydrogen and some helium although that is not the critical feature determining its behavior. Although it is a somewhat rough analogy, it can essentially be thought of as a gas in a closed flask, characterized by a specific density, pressure, and temperature. As such, it can be modeled reasonably well by the gas laws you learned in first year college chemistry, if you were fortunate enough to have taken them. What’s that? You never took chemistry in college? No matter; it is quite straightforward. The basic law governing the behavior of gasses is the so-called “Ideal Gas” Law, which, placed in equation form is PV = nRT, where P is the pressure of the gas, V its volume, and T its temperature. N is the totals number of moles, or atoms / nolecules of gas in the universe. Never mind the meaning of R here, the proportionality constant, which remains constant in this situation anyway; when the equation is re-written as V T/P ( being the symbol for proportionality), we see that as V, in this case the volume of the universe, decreases, either T must increase or P decrease. A modification to the equation is needed here, however. I am speaking of ideal gasses, which, in reality, don’t actually exist, but serve as models for real gasses. In fact, with real gasses, increasing pressure also raises their temperature. An example of this is the gasoline vapor / air mixture in the cylinder of a car; as the piston presses down on the mixture both its pressure and temperature rise – in a diesel engine, this compression is enough by itself to ignite the mixture, driving the piston upward and turning the crankshaft.
If the temperature of a gas rises high enough, the kinetic energy of the atoms or molecules composing it is sufficient to strip away their electrons, leading to a state of matter known as a plasma. This temperature is fairly high, in the thousands or tens of thousands of degrees. The sun and other stars of equal or greater brightness are both so hot as to be composed of such gaseous plasmas, whose temperature rises well up into the millions of degrees in their centers, enough along with the very high pressures / densities there to allow the thermonuclear fusion reactions which power their enormous energy outputs.
If we continue our backwards time journey, at a point of between three and four hundred thousand years after the start of the Big Bang, we reach the point where the temperature of the universe increases to and above the plasma temperature; after this point, electrons combine with protons and heavier nuclei to form the first atoms. This is a critical time in the cosmos’ evolution: prior to it, the interaction of electromagnetic radiation with the electrically charged electrons and bare atomic nuclei make it opaque; after it, when stable atoms form, the electromagnetic radiation can stream freely through space. This radiation, called the cosmic background radiation, is a measure of the universe’s temperature. Largely in the visible range at first, it cools over the billions of years the universe has been expanding, to the point now where it is almost entirely in the microwave region, a region of much lower energy than visible light, indicating a cosmic temperature of only a few degrees above absolute zero (absolute zero, or 0K on the Kelvin temperature range, is the complete absence of all heat). By the way, it was the (accidental) discovery of this microwave radiation in 1964 by Penzias and Wilson which as much or more than anything else cinched the case for the Big Bang theory.
Back to the universe at 3-400,000 years after the Big Bang. Another important fact of the universe at this time is that, although I have described it as though it were a gas of uniform, or homogenous, density throughout, obviously this could not have been the case. Even then there had to be inhomogenuities present, otherwise there would have been nothing for stars and galaxies and galactic clusters and larger scale structures to gravitationally condense around. These inhomogenuities need only be quite small – you would never notice their equivalent in a pot of mashed potatoes however hard you looked – but they had to be there, or else – well, for else one thing, just as with the discovery of a stable high energy state of the carbon nucleus, we would not be here to make their prediction. In fact, they turn out to be so small that it was not until the 1990s that they were finally unequivocally discovered by a space-based probe called the Cosmic Background Explorer, or COBE for short. The discovery was of such significance that some regard it as the most important scientific discovery of the century, even to the point of making religious analogies (of the Einsteinian nature) to it.
Three or four hundred thousand years is not much on our cosmic timeline, if you recall it from the last chapter, only around four hours. Actually, this is where the line begins to lose its usefulness, because the next set of interesting events occur up to only about 20 real minutes after the Big Bang, and reach back to only a trillionth of a trillionth of a trillionth (10-36) of a second after the beginning. Indeed, it is difficult to come up with any type of line that is intuitively useful; if we make 10-36 equal to one second, then events happening at a trillionth of a second would be 1024 or thirty million billion years later, over two million times longer than the known age of the cosmos! So we are going to have to drop our attempts to make such time intervals intuitively meaningful, and stick with the hard numbers, as difficult as they are to grasp.
As it is also impossible to describe the events that happen during this period of 10-36 second to approximately twenty minutes without some basic knowledge of nuclear physics, a digression is necessary before plunging in. Don’t panic, though; it will only be enough for our purposes here, and besides, I’m not enough of an expert in the subject to make it too abstruse.
As you probably already know from your high schooling at least the atom is composed of a nucleus consisting of at least one (in the case of the simplest element, hydrogen) or more protons, plus zero or more neutrons, along with one or more electrons which (though you will recall from chapter three that this is not really correct) circle it. Neutral atoms have as many electrons as protons, since the negative charge on the former exactly equals the positive charge on the latter. An atom stripped of one or more electrons is called an ion; high enough temperature or energetic enough radiation has the ability to do this, and as already mentioned, matter in this state is called a plasma.
There is something I must take the time to explain here. You probably didn’t learn much about the atomic nuclei in your schooling, but a fairly obvious question should occur to you about it: given that protons are all positively charged, what holds them together in the nucleus? Before answering that question, another thing you should know about both protons and neutrons, which are collectively known as hadrons, is that they themselves are composed of still smaller entities bearing the strange name of quarks (a pun I can’t resist is that there is a type of high-energy quark called the strange quark). You could say, in fact, that instead of describing nuclei as being made of two different types of hadrons, we really should say that they are made of two different kinds, or “flavors”, of quarks, namely up quarks and down quarks.
Quarks have “fractional” electric charges, in that they possess ⅓ of the negative charge of an electron – this is the down quark – or ⅔ of the positive charge of a proton – this is the up quark. Thus, what we call a proton is really a combination of two up quarks with a down quark, and a neutron is composed of one up quark with two down quarks. Add up the charges and you will see they work out, protons having a +1 charge and neutrons having zero charge.
Quarks have other interesting properties as well. Individual quarks cannot be isolated from each other and observed; they always exist in combinations of two or three (or maybe more). In fact, their existence was predicted on purely theoretical grounds in the early 1960s by Murray Gell-Mann and George Zweig, and weren’t indirectly experimentally verified until several years later by particle scattering experiments.
So the question isn’t what holds protons and neutrons together in the nucleus, but what holds the quarks together. Strangely enough, that question was partially answered several decades earlier (although the answer had to be modified to account for the quark structure of hadrons). Again, there is much more to this answer than needs to be covered here, but yet another brief digression, this time on forces, is enough to cover the basics. Also, as I have alluded to this earlier, now is a good time to explain it in more detail.
* * *
There are four “fundamental” forces in the universe – fundamental in that any force you encounter consists of one or a combination of them, working together or against one another. You are actually already familiar with two of these forces: gravity, which pulls all mass objects in the universe toward each, including holding you down on the ground, the moon orbiting Earth, and Earth and the other planets orbiting the sun; and electromagnetism, which you observe every time you use a magnet or electrically charged objects – it is, of course, the force that keeps electrons in their orbits, or orbitals, around the atomic nucleus. Incidentally, the reason you are much more aware of gravity than electromagnetism is that the former is (almost) a universally attractive force, building up as the mass to generate it accumulates. Electromagnetism, on the other hand, is both attractive and repulsive, so you only notice it under the special conditions where an excess of positive or negative charges occurs, and even then the excess is usually quite small, so the effect seems relatively weak compared to gravity. In fact, electromagnetism is some 1039 times more powerful than gravity! Also, the reason you come into direct contact with both forces is that they are infinite in range; both fall off only as the square of the distance between the two attracting (or repelling) objects.
The remaining two forces are called nuclear forces because their intensities fall off so rapidly that they act only on the scale of atomic nuclei; this is the reason we don’t encounter them directly, but only indirectly through their effects. One of these forces, the weak nuclear force, is involved in certain kinds of radioactive decay. I won’t speak more about it here. The other, the strong nuclear force, which I have mentioned before, is what answers our question about what holds the quarks, or the protons and neutrons, together in the atomic nucleus. This force is approximately a hundred times stronger than the electromagnetic force at the ranges typical inside nuclei. Again though, its range is so short that it takes tremendous kinetic energy to overcome the mutual electromagnetic repulsion between two nuclei and allow them to come close enough together to fuse via the strong nuclear force; this is why it takes the incredibly high temperatures in the core of a star, or in a thermonuclear weapon, or in the very early universe, to accomplish this kind of nuclear fusion.
The reason for the digression to discuss these forces is that, according to modern theories of nuclear physics, they are all actually manifestations of a single force, and that at sufficiently high temperatures and pressures, such as what happens as we get closer and closer to time zero, they merge together one by one until there is only a single force. The reason for the digression on quarks is that prior to a certain time, the temperature of the universe is so high that they cannot hold together long enough to make stable protons and neutrons.
* * *
There is one more digression that needs to be made before we talk about the earliest moments of cosmic evolution. It is, or so it seems to me, a non-physicist, to be the Central Problem if we are ever able to fully understand those moments.
The problem is that there are two major edifices of physics twentieth century science has erected to understand matter, energy, space, and time over the last hundred or so years. The first edifice, which we’ve already met, is the physics of the ultra-tiny, the world of the atom and smaller, the physics of quantum mechanics. The other edifice is the physics that describes the universe on the large scale, from approximately planet sized objects on up: Einstein’s General Relativity. And the problem is both simple and deep at the same time: they simply do not look at and model reality in the same way.
A good example of this is how they describe gravity. In quantum mechanics all forces are carried by a type of particle called a virtual boson (bosons are particles which carry forces; the particles which compose mass itself are called fermions). For the electromagnetic force, this boson is the photon; and for the strong nuclear force, the gluon. For gravity it is a hypothetical force dubbed, naturally enough, the graviton. I say hypothetical because gravity is such a weak force that gravitons have yet to be detected, although they are well described theoretically; nevertheless, according to everything we know, they must be there.
According to general relativity, however, gravity is really not a force at all, but the result of the Einsteinian curvature of four dimensional spacetime by massive objects. Another object will fall toward the object because it is only following the path of least resistance through this spacetime. Although this curvature is enough to hold us solidly on Earth, it requires a very massive object to detect it. One way of doing this is by the way it bends light; historically, General Relativity was regarded as proven by the slight deflection of star positions during a solar eclipse in 1919. The bending of light is used to explain a number of other astronomical phenomena as well, such as gravitational lensing, and the splitting of the image of a distance galaxy into two or more images by the presence of an intervening object of sufficient mass.
Another difference between the two theories is how they regard spacetime itself. General relativity requires that spacetime be smooth and relatively flat on all scales. Quantum mechanics however says that that is impossible. The uncertainty principle, which we have already met, means that on small enough scales spacetime must be lumpy and twisted. An analogy to this might be a woolen blanket which from a large distance looks smooth but up close is revealed to be composed of intertwining hair. The uncertainly principle also affects spacetime on small enough scales in another way, by allowing “virtual” particles to come into existence over short enough time periods. This happens because of another way of expressing the uncertainty principle besides the x × s ≤ /m form we encountered in chapter three: t × E ≤ , where t is time and E energy. In this form the equation states that it is possible for particles of any given mass energy (E) to exist as long as they disappear within time t. Despite the term virtual (they are not directly detectable), these particles are not only quite real in their effects, but they are the heart of what explains the four fundamental forces in quantum mechanics.
This conflict between quantum mechanics and general relativity means that neither theory encompasses a complete and fully correct vision of reality. This is not normally a problem for physicists however as generally, they divide reality into two camps, which deal with it on such different scales. In dealing with the very early universe however, they clash like charging elephants at full speed, for we have now delved into a realm of both the extremely small and the extremely massive, a place that no one has gone before and where all our curiosity and imagination and brilliance become less and less able to predict what we will find there. The only thing that is certain is that we are not in Kansas anymore.
* * *
It is time to resume our journey back to the beginning of the universe, or at least as far as our knowledge of physics permits, back towards T = 0, if indeed there was such a time. We had stopped at T + 20 minutes, and for good reason. In the universe today, only the centers of stars are hot enough and dense enough to fuse hydrogen into helium and heavier elements. But there must have been a time, if the Big Bang is true, when the universe as a whole existed in those conditions. There was, and T + 20 minutes marks the end of that time.
Astronomers observing our current cosmos discover that it is, by mass, approximately 75% hydrogen and 24% helium, with only traces of heavier elements. It is impossible to account for more than a tiny fraction of that helium by stellar nucleosynthesis, however. One of the triumphs of Big Bang theory was to account for the remaining helium; the period between T + 3 and T + 20 minutes in our universe had just the right conditions in terms of temperature and density, and lasted just the right amount of time, to create it.
The earliest periods of the Big Bang are referred to by cosmologists as epochs. Despite the name, epochs are mostly extremely short periods of time when the newly born universe was evolving extremely rapidly. Thus, there is the Planck epoch, the grand unification epoch, the inflationary epoch, the quark epoch, and so on. These epochs are defined according to the predominant process(es) or particle(s) which characterize them. The period of nucleosynthesis we are discussing is just a part of the photon epoch, the total length of which is from T + 3 minutes to almost T + three-four hundred thousand years (although the nucleosynthesis fraction of this time, if you’ll recall, only lasts up to T + 20 minutes), a time when most of the energy of the universe is in the form of photons; as mentioned before, this epoch ends when stable atoms finally form and the photons are free to stream through space unhindered as the cosmic background radiation we detect today.
The epoch preceding the photon epoch is the so-called lepton epoch, which takes us back to approximately T + 1 second. Leptons are fermions (a type of mass bearing particle, if you’ll remember) that interact with all forces except the strong nuclear force; the member of this family we are most familiar with is the electron, although there are others, such as the electron neutrino, a very low mass particle involved in certain types of nuclear reactions. There are also high energy, short-lived versions of both these particles, such as the muon and tau high energy analogues of the electron, and their corresponding neutrinos, the muon neutrino and tau neutrino. In the lepton epoch leptons dominate the mass of the universe. Excuse me, I should say leptons and anti-leptons, for we have reached that period of the universe’s evolution where one of its most interesting puzzles needs to be addressed: the cosmic asymmetry between matter and antimatter.
* * *
Antimatter probably sounds like the stuff of science fiction, especially if you are a Star Trek fan (this is admittedly where I first heard of it), but in fact it is very real, and that reality poses a serious problem. The problem is that every mass carrying particle, or fermion, has a corresponding antiparticle, which has the same mass but the opposite electric charge (there are other differences, too). So every electron, say, has an antielectron – also known as a positron – every quark has an antiquark, every neutrino an antineutrino. The real problem is that if a particle and its anti counterpart should encounter each other, say an electron and a positron, the result is cataclysmic: both particles mutually annihilate each other in a burst of high energy photons (photons, like other force carrying particles, are their own anti-particles; there are no such things as anti-photons). No, the real problem is that, in the first few seconds of the cosmos’ existence, both fermions and their anti counterparts ought to be produced in equal numbers, only in the next few seconds to completely annihilate each other, leaving a universe composed of nothing but high energy radiation; no matter, no stars or galaxies, and no us. As the universe today, for good theoretical and observational reasons, appears to be composed almost entirely of matter, with very little if any antimatter, there must have been a certain asymmetry between the number of matter and antimatter particles formed in the early universe. This asymmetry, favoring the creation of matter over antimatter, need only be quite small; once all of the antimatter had mutually annihilated by an equal quantity of matter, the excess of matter would have been left to dominate the cosmos as we see today. But what could have caused this asymmetry, however small?
This is no trivial question because symmetry lies at the heart of much of the laws of physics, especially the laws that govern sub-atomic particles and their behavior. Violations of certain kinds of symmetry, however, are known to occur. Symmetry breaking is, indeed, crucial to the earliest moments of Big Bang cosmology, particularly in the evolution of the four fundamental forces. Recall that these forces merge, one by one, into a single force as we close in on T = 0. So it is not unreasonable to hypothesize that some kind of symmetry breaking is responsible for the matter excess we see in the universe today. This is an area of active research and intense debate among cosmologists.
It is worthwhile to pause here at T + 1 seconds and take stock of where we are and what is happening in our attempt to unravel the earliest moments of the cosmos. I mentioned at the beginning of this chapter that as we went deeper and deeper into the past, we would eventually reach a point where our understanding of the laws of physics begins to get increasingly shaky, shaky to the degree that we are no longer certain of the ground beneath our feet. Like fossil hunters digging into deeper and deeper strata, what we find is less certain, more speculative, and harder to lay out with the same confidence that has carried us this far. My sense and reading and understanding leaves me to believe that we have arrived at this point, or at least are very close to it. The one event before T + 1 which does seem well established, the breaking of electroweak (electromagnetic plus weak forces) symmetry and the ensuing establishment of the weak nuclear force and electromagnetism as two separate forces, occurs at approximately T + 10-12 seconds. At this point all four fundamental forces have achieved their current form (though not current strengths), and the quarks in the quark-gluon plasma that fills the universe acquire their masses via their interaction with a still hypothetical particle (it is currently being actively searched for) called the Higgs boson. The subsequent cooling after this point allows the free quarks to combine into the protons and neutrons and other hadrons we see today.
* * *
I think I can say confidently that what happens before T + 10-12 seconds is entirely the subject of theoretical work. The next symmetry breaking, between the strong nuclear force and the electroweak force, is the subject of so-called Grand Unification Theories, or GUTS, of which there are several varieties. By the way, in a way this name is misleading, as we have still not accounted for gravity yet. But recalling our earlier discussion of general relativity and quantum mechanics, we know that a quantum theory of gravity needs to be formulated and tested before we tread that realm, and that such a theory is still in such a theoretical stage that one of its prime candidates, string theory, has yet to be accepted a real, credible theory by many in the scientific community.
Current estimates of the break between the strong and electroweak forces places it at about T + 10-36 seconds, or a trillionth of a trillionth of a trillionth of a second after the Beginning. And here, at the risk of understatement, is where things begin to get interesting, at least if our theoretical models are correct. For this is where Big Bang cosmology almost fell flat on its face, if I may be pardoned what is about to be another pun.
Besides the matter-antimatter asymmetry, two other features of the current universe need to be explained by events very early in its history: one is that, on very large scale, its shape is very flat; the second is that, on more local scales, it is lumpy and inhomogeneous.
The local inhomogenuity is the easier of the two to understand. We look around ourselves and we see a universe today in which the matter is organized into stars / solar systems, galaxies, clusters of galaxies, clusters of clusters, and so on. This is due to gravity working over billions of years, of course. But there must have been primordial inhomogenuities in the early universe for gravity to work on; if the Big Bang had produced a perfectly homogeneous distribution of mass-energy, then we would not be here to observe a universe composed of non-uniformly distributed hydrogen and helium, bathing in an equally non-uniform sea of background radiation.
Fortunately for us, the universe is inhomogeneous, and has been since the de-coupling of matter and energy around T + 3-400,000 years, as careful studies of the cosmic background radiation (from COBE) have shown. But where did these inhomogenuities arise from? Classical Big Bang theory at the time could not answer this question.
The other problem, that of the flatness of universe on large scales, also stumped classical theory, although it is a little harder to explain. This is an issue raised by general relativity; more precisely, by the so-called “field equations” of general relativity, which have a number of different solutions, under different conditions. These solutions, among other things, describe the cosmic curvature of spacetime due to the presence of mass-energy. There are three possible curvatures, depending on the mass-energy density, measured by a value called omega or Ω: if Ω is greater than one, then the mass-energy density yields a universe characterized by positive spacetime curvature, causing its expansion to eventually stop, then reverse into a contraction phase (which would have already happened by now) which may result in another cosmic singularity and big bang; if Ω is less than one, however, then spacetime is described as hyperbolic and the expansion will continue forever; if Ω is exactly equal to one, than spacetime is flat and the expansion will also continue forever, albeit slower and slower, gradually grinding to a stop it will never quite reach.
An exact measurement of Ω today is difficult, but between the observational data and theoretical considerations, it should be very close to if not exactly equal to one. The problem this creates is that any deviation from Ω = 1 in the early universe would be exponentially magnified by the cosmos’ expansion until today we should see a Ω vastly greater or smaller than one. As Ω appears close to or equal to one today, this must mean that it was even more exquisitely close to one in the early universe as well. Prior to the 1980s, however, nobody had a convincing reason why that should be the case. It simply appeared that Ω was another example of the “fine tuning” problem which we shall return to later.
Human ingenuity is never to be underestimated, however. In the 1980s the work of Alan Guth, Andrei Linde, Andreas Albrecht, and Paul Steinhardt yielded a modified version of Big Bang theory that included a period of exponential expansion very early in the cosmos’ evolution. They called this extra fast expansion Inflation. The idea of an ultra-fast, in fact exponential, expansion meant that during this phase the universe increased in size by many orders of magnitude (by a factor of at least 1026) in a fantastically short period of time, from about T + 10-36 to T + 10-32 seconds. The triggering mechanism for this expansion is not known for sure, but a good candidate appears to be the decoupling of the strong nuclear force from the electroweak force, especially as they appear to happen at the same time. It is also a matter of contention as to what brought inflation to an end, or even whether it ended everywhere at the same time or broke up into “bubbles” of ordinary universes formed at different times, of which ours is one. In fact, inflation could still be going on outside of our own universe, or perhaps “hyperverse” is the better term, still creating new universes with perhaps different laws and constants.
Whatever the physics behind inflation, what initiates it and how it ends, it neatly solves both the problems of local inhomogenuity and cosmic flatness (and a number of other problems as well). The flatness problem is solved because whatever the value of Ω before inflation, the enormous exponential stretching of spacetime brings it essentially so close to one that it will not diverge significantly from this value during the subsequent normal cosmic expansion. The local inhomegenuity problem is also solved, thanks to quantum mechanics: in the pre-inflation epoch the cosmos is so small that random inhomogenuities arise simply due to the uncertainty principle, which says that spacetime and the distribution of mass-energy can never be perfectly uniform; the effect of inflation is to “freeze” and enormously expand these inhomogenuities into the seeds of stars and galaxies and larger structures we see today.
* * *
So. We find ourselves at the decoupling of the strong nuclear force from the electroweak force which, if theory is correct, occurred somewhere between T + 10-36 and T + 10-32 seconds. The next step, going back further, T + 10-43 seconds marks the end of the Planck epoch, named so because according to quantum mechanics, it is approximately the shortest period of time which can be even theoretically measured, the shortest period of time one could say that time can even exist. The Planck epoch is also the time period in which quantum mechanics and general relativity find themselves in full collision. Somehow, some way, somewhere, gravity merges with the strong + electroweak force, although no one knows how with any certainty. We have entered the realm of pure imagination, where some scientists play with entities called cosmic strings and work long hours trying to turn them into the ultimate explanation of matter, energy, space, and time, while other scientists place their time and bets on ideas like quantum loop gravity and other exotic hypotheses. As no one has succeeded to the approval of all, we have also reached the end of our own, personal journey into the past, arriving if possible at where we began in Chapter eight, when we tried to imagine what nothing would really be like and realized that we couldn’t do it no matter how hard we tried. Of course, perhaps what preceded the Big Bang wasn’t nothing at all. Quite possibly our universe is part of a greater reality, in which other universes are also embedded – the multiverse conjecture. There are also a number of cyclic universe models, such as the Steinhardt-Turok model in which the universe oscillates between expansion and contraction, with each Big Bang triggered by a collisions of two “branes” (multi-dimensional strings) in a higher dimensional spacetime. Again, this model could predict many, even an infinite number, of universes.
Although any of these models could be true, there is, I think, a philosophical problem with the whole approach, one ironically not too different from the concept of a supernatural god(s) being responsible for the universe. Just as a god needs a greater god to explain it, ad infinitum, we are potentially postulating an infinite number of greater or higher dimensional cosmoses to explain our own. To me it all seems driven by a pathological inability to accept nothing merely because we are incapable of imagining it. But the limitations of human imagination prove nothing, except our need to accept them, however unpleasant. This is a subject we will return to in the last chapter of the book.

Thailand

From Wikipedia, the free encyclopedia https://en.wikipedia.org/wiki/Thailand Thailand , officially the K...