Blockbuster Science Read online

Page 2


  By the way, the term “science fiction” made its public debut in 1851 in William Wilson's book of essays, A Little Earnest Book upon a Great Old Subject. He writes,

  We hope it will not be long before we may have other works of Science-Fiction [like Richard Henry Horne's “The Poor Artist”], as we believe such books likely to fulfil a good purpose, and create an interest, where, unhappily, science alone might fail.1

  I like Wilson's thinking. I like it so much that I wrote the book you're holding now.

  Anyway, before the end of the nineteenth century, Mary Shelley's science fiction spark grew into a blazing fire.2 Jules Verne teased people with his extrapolations on technology. He was all about plausibility. Get this: he believed that someday people would zip across the seas in electric submarines. Crazy idea. Right? That was in Twenty Thousand Leagues under the Sea, written back in 1870. How about his idea of using solar sails (described in chapter 17) for space travel in his 1865 novel From the Earth to the Moon?

  Then H. G. Wells arrived on the scene with science fiction that contained a more sociological bent. Like Verne, he wrote about changes in technology, but he was less interested in the plausibility of the science than in how the changes might affect people. These two styles have served as the foundations of science fiction ever since.

  Following in the style of Verne's hard (plausible) science, the “Big Three” authors of the early twentieth century emerged. Robert Heinlein took technology that existed, such as the telephone, and made it into something that resembles today's devices, like a mobile phone (Space Cadet, 1948). Isaac Asimov went robot crazy starting in 1939 (you will read about robots rising up in chapter 14).

  The third author, Arthur C. Clarke, wrote about artificial satellites in a stationary orbit twenty years before a real one orbited Earth. He also had ideas about space elevators (learn more about them in chapter 17). All three were friends, but Asimov and Clarke found it necessary to create the Clarke-Asimov treaty.

  Under the terms of the treaty, Asimov was required to insist that Clarke was the best science fiction writer in the world (reserving second-best for himself). Clarke, meanwhile, was required to insist that Asimov was the best science writer in the world (reserving second-best for himself).3 Which led to the following dedication in Clarke's 1972 novel Report on Planet Three and Other Speculations:

  In accordance with the terms of the Clarke-Asimov treaty, the second-best science writer dedicates this book to the second-best science-fiction writer.

  So, I'm not the first author writing about science and science fiction who also likes humor. As I mentioned before, Wells demonstrated how not all the science in science fiction needs to be plausible (it just needs to not be mystical). Science concepts in fiction can be a tool with which to explore topics that other types of early fiction often avoided.

  Although Clarke's stories are generally known for their hard science, he wasn't shy about tackling social issues. In Imperial Earth, released in 1975, the main character is of African descent and his sexuality is flexible. Recall how radical it was back then to provide a starring role for a character who happens to have an ethnic heritage…let alone to address the fluidity in human sexuality.

  Ursula K. Le Guin is the queen of humanistic science fiction. Her stories explore the cultural and social structures of alien cultures. In her, the heart of an anthropologist blends with the writing chops of a Heinlein. She gives alien cultures consistent motivations, albeit different from the ones most readers have experienced before, and she skillfully makes readers sympathetic to the aliens’ cause.

  And then there is science fiction that borders on philosophy. Philip K. Dick used fiction as a tool to question identity and what constitutes reality…sometimes his own. You might harbor your own doubts about what is real after you read chapter 20, where we question existence. Don't worry. You are real. Maybe.

  Okay, I need to take a deep breath. I could continue gushing about great authors, but I'll stop for the sake of keeping this introduction under one thousand pages. I think you get the idea. Science fiction has exploded onto the scene with books and films and television shows based on real science. And it all began with Mary Shelley, a bet, and a cold summer.

  I'm thinking that a lot of you might be interested in fantasy. That's cool. Fantasy is a living ancestor of science fiction and graciously stops by several of the upcoming chapters for an occasional visit. Keep in mind that science fiction, unlike fantasy, is about rationality. This book sticks to what can be proven using the scientific method and how scientific concepts and theories have been, or can be, extrapolated for fiction.

  Nonetheless, a little magical thinking (ghosts, unicorns, etc.) mixed in with plausible science can be fun. Think about steampunk fiction, where modern technology is realized in Victorian England or the American Wild West. I believe that as long as fantasy follows the internal rules of the imaginary world in which it is set, its family resemblance to science is quite striking. After all, science is a consistent and replicable explanation of the natural world.

  In fantasy, the author creates a world with its own set of laws. If these laws are consistently applied, and things that happened to one character must happen to others under the same conditions, then you have the fantasy world's equivalent of science. Let's look at this for an example: On a world called Meryton, a young sorceress (like all the magicians in this world) relies on nonrenewable magic. She is running low because of her wasteful father and must marry well to replenish the family's magical treasure chest. Mayhem ensues. Title: Pride and Prejudice in Space. Sorry, I couldn't resist.

  Why do I and so many other humanoids love science fiction? And yes, for me, both reading and writing science fiction are a love affair. Here is my answer to the question: it is so beloved because of how inclusive it is. Any bookstore (brick-and-mortar or online) will stock authors and characters who represent a full array of racial, ethnic, and gender groups. And because fictional worlds can support a variety of societies and cultures, a multitude of worldviews, norms, and sexual orientations have appeared.

  Because of this inclusiveness, the typical science fiction fan isn't the clichéd man-child who lives in his parents’ basement. I think the typical science fiction fan is all of us: men, women, engineers, lawyers, scientists, actors, sports stars, and so on. If you are a creator (or aspiring creator), learn who your potential readers might be. Guess what? They can be anyone from anywhere in the world.

  Before you plunge into writing a book or reading (or viewing) your next science fiction story, let's separate the science from the fiction. Science asks, “What if?” Fiction speculates on what will happen to people or societies when that thorny question is answered. But what actually constitutes science? Broadly speaking, it's a way of knowing. Scientific methodologies seek the truth of the physical world in which we live. Of course, there are other ways of knowing, such as art (personal truth), religion (revealed truth), and so on, but this book isn't about those. This book is evidence based.

  Science has two distinct aspects. First, it is a method, a set of steps used to question phenomena in the natural world through reproducible observations and controlled experiments. The steps are fairly concrete: conduct observation, devise a hypothesis to attempt to explain the observations, experiment in ways that test the hypothesis, formulate a theory, confirm the theory through more experiments, and use the theory for prediction.

  But wait, that isn't all. The second aspect of science is that it is also a collective term for the accumulation of what was learned. There is the act of doing chemistry, and there is chemistry itself.

  Don't worry. This book avoids getting too technical on any particular subject. But, for the strong of heart and the endlessly curious geeks, bonus sections provide extra details and/or further extrapolations on some of the topics. Each chapter also includes examples of the good, the bad, and the ugly use of the chapters’ scientific topics in novels, television, and movies. A diverse selection of authors and characters wi
ll be heard from through these pages. There is also a glossary at the end of the book listing a few of the noteworthy terms used in the chapters, and a reading/movie/song list of sci-fi-themed books, films, music, etc.

  Also, don't worry about the math that occurs here and there, because these references are very limited. Never fear math. It is the language of science. In fact, as with spoken languages, it is fraught with tongue twisters that scientists sometimes take too seriously. On occasion, that has led them astray from the methods and demands of science.

  For example, consider string theory (the topic of chapter 3). Intuitively it may seem hard to follow, but mathematically it is internally consistent. It might theoretically explain phenomena that can't otherwise be explained using more conventional models. However, because it isn't observable and because no direct experiments can test the theory, for now it dwells more in the realm of philosophy than science.

  For creators of science fiction, though, who cares? Whatever else can be said of string theory, it's a great story generator. I hope you will see from this and other examples how math can nudge science toward the world of science fiction. So remember this mantra: math is good.

  Blockbuster Science can be read out of sequence if you want to hone right in on a specific topic, but the chapters do have some tendency to build upon each other. I also encourage speculations—your own, not just those that deal with fiction—throughout the book. I want this to be collaborative.

  The first two chapters describe the twin pillars of twentieth-century physics: quantum mechanics and Albert Einstein's theories of relativity. Special and general relativity show us that time is woven into our universe and that our perception of time is correlated with gravity.

  Before the theories of relativity were developed, studying the universe was like studying a three-dimensional cube as if it were a square—not too helpful. Einstein united space with time, matter with energy, and everything to gravity. From his theories, scientists have deduced the existence of wormholes, black holes, and the big bang.

  Another major topic of relativity is the energy-mass equivalency (E=mc2), which is really amazing, and in that chapter 1, you find out why. Are you interested in time travel and space travel? (Who isn't?) This is the chapter for you (along with a few others that dip into the same topic).

  In chapter 2, we learn that size really does matter…at least for quantum mechanics. We will consider really small stuff, subatomic small, to discover why uncertainty will always be a part of our universe. This is the area of physics where determinism is dumped. At the tiny scale, everything is fuzzy because tangible particles are also waves.

  Wait, it gets weirder. The tangible particle can be anywhere in its own wave, where every occurrence of the particle in its own wave is merely a probability. Until an outcome is observed, all the occurrences exist simultaneously. This fuzziness has challenged the way scientists understand reality (and it might do the same for you). Chapter 2 also considers a couple of the more popular interpretations of what they believe this fuzziness and lack of determinism could mean.

  Don't worry about the weirdness. The trickier concepts of quantum mechanics are broken down into bite-sized chunks. By consuming these earlier chapters so daintily, in later chapters you will understand the theory behind topics like zero point energy, virtual particles, quantum entanglement, quantum computing, quantum teleportation, loop quantum gravity, quantum suicide and immortality, and time-traveling text messages.

  Topics in later chapters include string theory, parallel worlds, antimatter, neutrinos, tachyons, invisibility, holograms, extraterrestrial life, interstellar communication, bioengineering, terraforming, global warming, cosmology, evolution, the origin of life (carbon-based life, at any rate), rocketry, genetic modification, thermodynamics, the “arrow of time,” what might be next for computers (artificial intelligence), ranking civilizations, plus much more.

  Geez, that's a lot. We are going to have fun! At least until the last chapter, which covers the end of everything—the earth, the sun, the universe…everything.

  I hope this book will give you a sense of wonder and the desire to explore these topics even more.

  The moral of this introduction: Science tells us what is, not what we want. Science fiction has no such restriction.

  The moral of this introduction (originally printed on a parallel Earth): Science fiction is driven by fear or hope, while science is driven by necessity or curiosity. The overlap between their motivations is huge.

  I am going to conclude this introduction by listing Arthur C. Clarke's three laws of prediction.4 Popular works of science fiction from Doctor Who to Star Trek have cited his third law ad nauseam. I encourage you to embrace all of them, but for fun, try living the second:

  When a distinguished but elderly scientist states that something is possible, he is almost certainly right. When he states that something is impossible, he is very probably wrong.

  The only way of discovering the limits of the possible is to venture a little way past them into the impossible.

  Any sufficiently advanced technology is indistinguishable from magic.

  Buckle up, because here we go.

  Logic will get you from A to Z; imagination will get you everywhere.

  —Albert Einstein

  Science fiction books, movies, and television shows get a lot of mileage out of driving their characters through space and through time. In some cases, the hero manages to travel through both with the benefit of a Time and Relative Dimensions in Space (TARDIS) machine. This is the space and time vehicle of choice in the television series Doctor Who.

  They might also journey via the time dilation that affects rapid acceleration ships as seen in Orson Scott Card's novel (and 2013 movie) Ender's Game. This type of travel is possible, but the truth behind traveling through time and space is much simpler. You don't actually need fancy technology because you are always traveling through time and space. If you jog a distance of five miles at an average speed of five miles per hour, then you have moved five miles through space and one hour through time.

  I know, I know, this is obvious. But I want to point out that space and time are so intertwined in the physics of the universe that they must always be considered together. Physicists came up with a creative name for this unification: spacetime. The unification is also sometimes referred to as the space-time continuum.

  Because of time's connection to space, scientists no longer believe, as Isaac Newton did, that time is absolute and flows uniformly. Albert Einstein's theory of special relativity proved that time is woven into the fabric of space and, oddly, is connected to the speed of light. This was a wow concept, but what did he mean by relativity?

  If you are sitting at home reading this book you might feel stationary. This is an illusion. The earth spins as it circles the sun. The sun is traveling within the Milky Way galaxy, which itself is rotating around a supermassive black hole called Sagittarius A* (pronounced Sagittarius A-star). The Milky Way also moves within the local cluster of galaxies.

  Ever since the big bang, space itself has continued expanding in all directions. So, every point (location) you can think of can be considered the center of our growing universe. There is no absolute position anywhere. No place within the universe is stationary; all movement is relative to something. Everything depends on point of view.

  Think, for example, of a baseball game. The pitcher throws a fastball, and the batter takes a swing. From the bat's point of reference, the ball is moving toward it. But from the ball's reference, the bat is approaching.

  An untrue tale: A physicist stood on one bank of a river, and Einstein stood on the other. “Hey,” the physicist called. “How do I get to the other side?” Einstein thought for a few moments as he puffed on his pipe. Finally he called back, “You are on the other side!”

  Einstein understood that almost everything is really a matter of perspective, the exception being the speed of light. That is absolute.

  I wonder what sci
ence fiction author Theodore Sturgeon might have thought about this. According to the law named for him, nothing is always absolutely so. Then again, he wasn't a scientist (or, dare I say, an Einstein). Sturgeon is also credited with having said that ninety percent of everything is crap.1 I promise you, special relativity is 100 percent crap-free. It has been proven experimentally and the results replicated. That's science.

  With this in mind, the easiest way to reduce special relativity into something easy to learn (although granted not necessarily the most intuitive way) is to understand that everything moves at the speed of light. Yes, that's a weird idea, but our velocity through time added to our velocity through space always equals the speed of light.2

  Some cool math hangs around behind this, but rather than blinding you with flashy equations, I'll tell you their conclusion: the faster we travel through space, the slower our journey becomes through time. In other words, as our speed increases, time bends to conserve the speed of light. This time bending is called time dilation. Don't worry, at our earthly speeds the time effects are minimal. Only after we have accelerated our starships to high speeds do the relativistic effects on time become more severe.

  According to special relativity, as we travel through spacetime, we take our frame of reference—clocks—with us. On Earth, our personal clock is mostly in sync with everyone else's because of our slow speeds relative to each other. Now, if we board a starship and accelerate at great speeds away from our home planet, our clocks begin to differ from those we left behind. As our speed through space increases, we must travel more slowly into the future to conserve the speed of light.

  In the vacuum of space, the speed of light is 299,792,458 meters/second (670,616,629 miles/hour).3 If anyone ever asks, just say, “About 300,000,000 meters/second.” It's easier to remember, and you'll look really, really smart.