To change the world, change the stories humans believe.
Home2023

October 2023

So, via the might of science, I am now able to answer the recurring question(s) that is: “Why don’t you have a girlfriend?”, “Are you gay?”, “You must have a tiny schlong”.

The answer – The Drake Equation! The Drake equation is used to estimate the number of highly evolved civilizations that might exist in our galaxy. And with a little bit of tweaking where necessary, I can use it to find out the number of potential girlfriends for me.

The equation is generally specified as:

G = R ⋅ fP ⋅ne ⋅ fl ⋅ fi ⋅ fc ⋅ L

Where:

  • G = The number of civilizations capable of interstellar communication
  • R = The rate of formation of stars capable of supporting life (stars like our Sun)
  • fP = The fraction of these stars that have planets
  • ne = The average number of planets similar to Earth per planetary system
  • fl = The fraction of the Earth-like planets supporting life of any kind
  • fi = The fraction of life-supporting planets where intelligent life develops
  • fc = The fraction of planets with intelligent life that are capable of interstellar communication (those which have electromagnetic technology like radio or TV)
  • L = The length of time such communicating civilizations survive

Using this equation Prof. Drake estimated that 10,000 communicative civilizations probabilistically exist in the Milky Way alone. Astronomers estimate that there are between 200 and 400 billion stars in the Milky Way. Let’s call it 300 billion. This makes the probability of a star chosen at random supporting life capable of interstellar communication 0.00000003%.

Another way to think about this is that this is the probability of the conditions necessary for us to communicate with an alien civilization being satisfied. These seem like slim odds at best, but the probability is positive (There is a chance!) and this approach is widely accepted by astronomers (This isn’t science fiction!). The idea that there could be 10,000 civilizations that we are capable of communicating with is very exciting indeed.

While extraterrestrial civilizations may be rare, there is something that is seemingly rarer still: A girlfriend. For me. What might the approach employed in the estimation of the number of alien civilizations tell us about the number of potential girlfriends for me? A somewhat less scientific question, I admit, but one of substantial personal importance.

The parameters are re-defined as follows with the values in brackets:

  • G = The number of potential girlfriends:

One can easily substitute boyfriends in here but as I am a heterosexual male I will focus on the search for a girlfriend.

  • R = The rate of formation of people in Kenya (i.e. population growth):

This is about 1,000,000 people per year over the last 60 years.

  • fW = The fraction of people in Kenya who are women. (0.51)

The Kenya National Bureau of Statistics puts it at just over half of the population.

  • fL = The fraction of women in Kenya who live in Nairobi. (0.09)

I would like my girlfriend to be nearby so that we can see each other. This makes it easier to get to know each other, avoids the difficulties of a long-distance relationship and saves me the bus fare.

  • fA = The fraction of the women in Nairobi who are age-appropriate. (0.19)

I am 27 years old (Thank you, I know I don’t look it). I would like my girlfriend to be near my age. I don’t want to feel older than I am by not being able to keep up with a spritely eighteen-year-old, or because I haven’t watched Purple Hearts and I don’t know who Olivia Rodrigo is. Nor do I want to fall prey to a voracious cougar or to be regaled with stories of the fight for multiparty democracy. Let’s say I am looking for a woman between 23 and 29 years of age.

  • fU = The fraction of age-appropriate women in Nairobi with a university education. (0.01)

I am not trying to be an elitist or anything, but I would like my girlfriend to have a university education. I think we would have more in common and I would like someone I could discuss my work with sometimes. I know that there are many intelligent people who don’t go to university, so don’t get all righteously indignant. Everyone has preferences. How many women out there have dated men shorter than themselves? I rest my case.

  • fB = The fraction of university-educated, age-appropriate women in Nairobi who I find physically attractive. (0.05)

Physical attractiveness is important. It is often the first thing people notice about each other and it makes sex easier. Not that my potential girlfriend need be considered attractive by anyone else, but I must find her attractive. This is a tough parameter to estimate. Let’s be generous and say I find 1 in 20, or 5% of age-appropriate women in Nairobi with a university education physically attractive.

  • L = The length of time in years that I have been alive thus making an encounter with a potential girlfriend possible.

27? Good lord, I am old.

We can simplify the above specification by recognizing that the number of people who have ever lived in Kenya is related to the population growth rate by:

N = ∫0T R(t).dt

where T is the age of Kenya. If we assume that R is constant over the period T then N = R ⋅ T. While this simplification is often used for the Drake Equation’s intended purpose, it is not a good assumption when adapting the equation for our purposes here. Instead, we use N*, the population of Kenya as of mid-2023, where:

N* = 55,100,586

With this simplification, we can re-specify the Drake equation as:

G = N* ⋅ fW ⋅ fL ⋅ fA ⋅ fU ⋅ fB

If we plug in the above values we get:

G = 55,100,586 ⋅ 0.51⋅ 0.09 ⋅ 0.19 ⋅ 0.01 ⋅ 0.05

or:

G = 240

So, what this means is that 240 people in Kenya satisfy these most basic criteria for being my girlfriend. That is 0.00044% of Kenyans and 0.0047% of Nairobians, which does not seem so good. On a given night in Nairobi, there is greater than a 1 in 10,000 chance that I will meet an attractive woman between the ages of 23 and 29 with a university degree. Of course, this does not take into account the fraction of these women who will find me attractive (depressingly low), the fraction of these women who will be single (falling with age) and, perhaps most importantly, the fraction of these women who I will get along with. Including such factors would greatly reduce the above figure of 240. A rough estimate puts the number of potential girlfriends accounting for these three additional criteria (1 in 10 of the women find me attractive, half are single and I get along with 1 in 10) at 2. That’s correct. There are only 2 women in Kenya with whom I might have a wonderful relationship with. So, on a given night out in Nairobi, there is a 0.000045% chance of meeting one of these special people, about 1000 times better than finding an alien civilization we can communicate with. That’s a 1 in 27,550,293 chance. Not great. At all!

Make of this what you will. It might cheer you up, it might depress you. I guess it depends on what you thought your chances were before reading this. But how do you think I feel? I spent aeons perfecting this formula, only to prove that I am the highest note in a sad song.

For us in the tropics, the lack of a chiller wasn’t much of a problem. Smoking, drying, salting, fermentation, or a combination of these methods were applied.

But for our siblings in the north, it used to be hard to keep food from spoiling in the summertime. There was a person called “the iceman”. He would go to their house and sell them a big block of ice. They’d keep it in something called an “icebox” to preserve the kinds of food that spoiled quickly. But that was a drag because the ice kept melting. It would drip all over the floor.

So, somebody thought up another way to keep food cold. It was a gas-powered system that used ammonia or sulphur dioxide as a coolant. No more lugging blocks of ice. What could be bad about that? Well, the chemicals were not only poisonous, they smelled terrible and there were leaks.

A substitute coolant was badly needed. One that would circulate inside the refrigerator, but would not poison anyone if the refrigerator leaked, or pose a danger if it was sent to the junkyard. Something that wouldn’t make you sick, wouldn’t burn your eyes, or attract bugs, or even bother the cat. But in all of nature, no such material seemed to exist.

So, chemists invented a class of molecules. Little collections of even tinier things called atoms, that had never existed on Earth before. They called them chlorofluorocarbons, or CFCs, because they were made up of one or more carbon atoms and some chlorine and/or fluorine atoms. These new molecules were wildly successful, far exceeding the expectations of their inventors. Not only did CFCs become the chief coolant in refrigerators, but also in air conditioners. There were so many things you could do with CFCs:

  • People used them to propel great fluffy mounds of shaving cream.
  • And to protect your hair from wind and rain.
  • It was also the propellant that made fire extinguishers and spray paint cans so much fun.
  • It was good for foam insulation, industrial solvents and cleansing agents.

The most famous brand name of these chemicals was Freon, a trademark of DuPont. It was used for decades and no harm ever seemed to come from it. Safe as safe could be, everyone figured. Until, in the early 1970s, two atmospheric chemists at the University of California, Irvine were studying Earth’s atmosphere.

Mario Molina (right) was a Mexican immigrant and a young laser chemist. Sherwood Rowland was a chemical kineticist, someone who studied the motions of molecules and gases under varying conditions.

Molina wanted to grow as a scientist. He was looking for a project that would take him as far from his previous research experience as possible. He wondered. What happens to those Freon molecules when they leak out of the air conditioner? This was a time when the Apollo astronauts were still making regularly scheduled trips to the Moon. And NASA was contemplating weekly launches of a space shuttle. Would all that burning rocket fuel pose a danger to the stratosphere, that place where Earth’s atmosphere meets the blackness of space? And this is how science works a lot of the time. You set out to solve one problem, and you happen on a completely different, unexpected phenomenon.

Those wonderfully inert, “harmless” CFCs, the magic molecules of shaving cream and hair spray, didn’t simply vanish when we were done with them. They had an afterlife at the edge of space, where they accumulated in the trillions. They were silently congregating high above the Earth, and they were up to no good. Molina and Rowland were alarmed to discover that the CFCs had thinned the protective layer that shielded us from the Sun’s harmful ultraviolet radiation. And it was getting worse all the time. When UV light hits a CFC molecule, it strips away the chlorine atoms. Once that happens, the chlorine atoms start devouring the precious ozone molecules.

A single chlorine atom can destroy 100,000 ozone molecules.

It wasn’t until our planet developed an ozone layer, about two and a half billion years ago, that it was safe for life to leave the ocean for the land.

CFCs were in everything, and the manufacturers couldn’t imagine a world without them. The corporate response to this danger was that the science hadn’t been settled. People had a hard time believing that we had become powerful enough as a species to endanger life on the planet. They looked for non-human causes for the loss of the ozone in the sky. One Reagan administration official suggested that everyone just wear more sunblock and put on a hat and sunglasses. But the scientists pointed out that the plankton, those tiny plants at the base of the global food chain, and the larger plants, were unlikely to do so.

Molina and Rowland tirelessly worked to warn the world.

What’s the use of having developed a science well enough to make predictions if, in the end, all we’re willing to do is stand around and wait for them to come true?

Sherwood Rowland

But then something amazing happened. There was a global outcry. People all over the world got involved. In the 1960s, the women of the world demanded an end to atmospheric nuclear testing because they didn’t want to nurse their babies with poisoned milk. Then, in the ’80s, consumers demanded that the corporations stop manufacturing CFCs. And you know what? The governments listened. The Montreal Protocol – the international treaty designed to protect the ozone layer by phasing out the production of numerous substances responsible for ozone depletion – was signed 36 years ago today. CFCs were banned in 197 countries. That’s just about as many countries as there are on this planet. The ozone layer has been getting thicker ever since.

But what would’ve happened if Rowland and Molina hadn’t been curious about the stratosphere, or if their warnings had been ignored? By 2060, the ozone would have been all but gone from the entire planet. You would never have been able to take your children out to bask in the sunshine. The food crops would have completely failed. The herbivores, those who live off them, would have died out. The carnivores would subsist on their corpses for a while, but ultimately, they, too, would be doomed.

If we continue to safeguard the ozone layer, it will be completely mended by 2050. And that’s why this is one danger you can cross off your worry list.

And so, as we reflect on the remarkable journey of discovery, responsibility, and global cooperation sparked by the lessons of the Montreal Protocol, we find ourselves standing at the crossroads of another monumental challenge — the looming threat of global warming and climate change.

Embers of Heritage in the Digital Age

Tales formed a crucial part of my childhood. Growing up, I used to love going upcountry to visit my grandparents all cause of the interesting stories they had about their many years of existence on this speck of dust we call home. This is only one instance of how stories and legends shared over the flames of bonfires have always linked generations. But, in the age of screens and clicks, that connection is fraying. Don’t worry though: within the world of technology exists a modern-day tapestry to preserve our stories, ensuring they echo even on the far horizons of the Moon and Mars – someday when we settle on these cosmic neighbours.

Tales used to be shared over bonefires

Weaving Tales into the Cosmic Fabric

Stay with me… let’s consider a world in which the substance of our ancestral stories does not fade but rather evolves. Consider a virtual bonfire, a meeting place for the global (interplanetary) village to share stories not only across generations but throughout the cosmos. Picture a digital loom where history’s threads intertwine, weaving a tapestry for posterity’s gaze.

This is not a new concept to us. We started thinking about sharing our stories throughout the cosmos as early as 1977 when Voyager 1 and Voyager 2 were launched from the NASA Kennedy Space Center at Cape Canaveral, Florida aboard the Titan-Centaur expendable rockets.

These two space crafts that are now approximately 24 billion kilometres from Earth (Voyager 1) and 20 billion kilometres (Voyager 2) carry within them The Golden Record: a phonographic record that contains a curated selection of sounds and images that represent the diversity of life and culture on Earth.

The records are a time capsule, intended to communicate a story of our world to other civilizations that may exist out there.

Chronicles in the Cloud: More Than Just Data

Voyager chronicles aside, let’s get technical, but not too technical. Consider your narrative to be a file. It might have been passed down in your generation or one that you have collected as a hobby. Instead of this file being tucked away, it’s floating in a digital cloud, where it can scatter stories like cosmic confetti. And it’s not just words; there are images, sounds, and animations as well, making our stories the life of this virtual party.

Cultural Constellations: A Galaxy of Narratives

Every culture is represented by a star in the narrative galaxy. These stars are shining even brighter as a result of technological advancements. Imagine exchanging bedtime stories from all around the world with your new Moon friends, creating a constellation of stories that would make even the Milky Way jealous! Do you want to know what the best part is? You’d be able to interact with tales from different cultures around the world. Whether it’s the Māori or the Native Americans or even Aborigines and Early man.

AR & VR Voyages: From Campfires to Cosmic Camps

Now with all these stories available at the click of a button, put on your AR/VR goggles and enter a world where campfires are more than simply logs—they’re not even real! But what about the stories? They’re just as real as they’ve always been. Even if the fire isn’t hot enough for roasting marshmallows, you’re right there, feeling the crackling energy of stories.

10 points for Gryffindor if you can be able to talk to someone in New Zealand while you are in your sustainable city apartment in Nairobi, and ask them what emotions they are feeling at that very moment. A virtual world, given a personal touch.

Passing the Digital Torch: Where Bytes Become Legends

We need to keep our tales alive like a torch in a relay race. However, in this cosmic race, we’re not simply passing a torch, but also a legacy. Technology is more than simply a gadget; it is the link between generations. And, like a phoenix emerging from the ashes, our stories rise from the old to embrace the new, illuminating the path for future generations.

Buckle up, fellow Voyager, for we’re travelling to new galaxies with technology as our spaceship and stories as our fuel. It’s not only about storytelling; it’s about preserving our humanity in the vastness of the universe.

“Alfred, it’s spinning.” Roy Kerr, a New Zealand-born physicist in his late 20s, had, for half an hour, been chain-smoking his way through some fiendish mathematics. Alfred Schild, his boss at the newly built Centre for Relativity at the University of Texas, had sat and watched. Now, having broken the silence, Kerr put down his pencil. He had been searching for a new solution to Albert Einstein’s equations of general relativity, and at last, he could see in his numbers and symbols a precise description of how space-time—the four-dimensional universal fabric those equations describe—could be wrapped into a spinning ball. He had found what he was looking for.

When this happened, in 1962, the general theory of relativity had been around for almost half a century. It was customarily held up as one of the highest intellectual achievements of humanity. And it was also something of an intellectual backwater. It was mathematically taxing and mostly applied to simple models with little resemblance to the real world, and thus not widely worked on. Kerr’s spinning solution changed that. Given that pretty much everything in the universe is part of a system that spins at some rate or other, the new solution had relevance to real-world possibilities—or, rather, out-of-this-world ones—that previous work in the field had lacked. It provided science with a theoretical basis for understanding a bizarre object that would soon bewitch the public imagination: the black hole.

General relativity was presented to the Prussian Academy of Sciences over the course of four lectures in November 1915; it was published on December 2nd that year. The theory explained, to begin with, remarkably little, and unlike quantum theory, the only comparable revolution in 20th-century physics, it offered no insights into the issues that physicists of the time cared about most. Yet it was quickly and widely accepted, not least thanks to the sheer beauty of its mathematical expression; a hundred years on, no discussion of the role of aesthetics in scientific theory seems complete without its inclusion.

When gravity fails

Today its appeal goes beyond its elegance. It provides a theoretical underpinning to the wonders of modern cosmology, from black holes to the Big Bang itself. Its equations have recently turned out to be useful in describing the physics of earthly stuff too. And it may still have secrets to give up: enormous experiments are underway to see how the theory holds in the most extreme physical environments that the universe has to offer.

The theory was built on the insights of Einstein’s first theory of relativity, the “special theory”, one of a trio of breakthroughs that made his reputation in 1905. That theory dramatically abandoned the time-honoured description of the world in terms of absolute space and time in favour of a four-dimensional space-time (three spatial dimensions, one temporal one). In this new space-time, observers moving at different speeds got different answers when measuring lengths and durations; for example, a clock moving quickly with respect to a stationary observer would tell the time more slowly than one sitting still. The only thing that remained fixed was the speed of light, c, which all observers had to agree on (and which also got a starring role in the signature equation with which the theory related matter to energy, E=mc2).

Special relativity applied only to special cases: those of observers moving at constant speeds in a straight line. Einstein knew that a general theory would need to deal with accelerations. It would also have to be reconciled with Isaac Newton’s theory of gravity, which relied on absolute space, made no explicit mention of time at all, and was believed to act not at the speed of light but instantaneously.

Einstein developed all his ideas about relativity with “thought experiments”: careful imaginary assessments of highly stylised states of affairs. In 1907 one of these provided him with what he would later refer to as his “happiest thought”: that someone falling off a roof would not feel his own weight. Objects in free-fall, he realised, do not experience gravity. But the curved trajectories produced by gravity—be they the courses of golf balls or planets—seemed to imply some sort of pushing or pulling. If golf balls and planets, like people falling off roofs, felt no sort of push or pull, why then did they not fall in straight lines?

The central brilliance of general relativity lay in Einstein’s subsequent assertion that they did. Objects falling free, like rays of light, follow straight lines through space-time. But that space-time itself is curved. And the thing that made it curve was mass. Gravity is not a force; it is a distortion of space-time. As John Wheeler, a physicist given to pithy dictums about tricky physics, put it decades later: “Space-time tells matter how to move; matter tells space-time how to curve.”

The problem was that, in order to build a theory on this insight, Einstein needed to be able to create those descriptions in warped four-dimensional space-time. The Euclidean geometry used by Newton and everyone else was not up to this job; fundamentally different and much more challenging mathematics was required. Max Planck, the physicist who set off the revolution in quantum mechanics, thought this presented Einstein with an insurmountable problem. “I must advise you against it,” he wrote to Einstein in 1913, “for in the first place you will not succeed, and even if you succeed no one will believe you.”

Handily for Einstein, though, an old university chum, Marcel Grossmann, was an expert in Riemannian geometry, a piece of previously pure mathematics created to describe curved multi-dimensional surfaces. By the time of his lectures in 1915 Einstein had, by making use of this unorthodox geometry, boiled his grand idea down to the elegant but taxing equations through which it would become known.

Just before the fourth lecture was to be delivered on November 25th, he realised he might have a bit more to offer than thought experiments and equations. Astronomers had long known that the point in Mercury’s orbit closest to the sun changed over time in a way Newton’s gravity could not explain. In the 1840s oddities in the orbit of Uranus had been explained in terms of the gravity of a more distant planet; the subsequent discovery of that planet, Neptune, had been hailed as a great confirmation of Newton’s law. Attempts to explain Mercury’s misbehaviour in terms of an undiscovered planet, though, had come to nought.

Famous long ago

Einstein found that the curvature of space-time near the sun explained Mercury’s behaviour very nicely. At the time of the lectures, it was the only thing he could point to that general relativity explained and previous science did not. Martin Rees, Britain’s Astronomer Royal, is one of those who sees the nugatory role played by evidence in the development of the theory as one of the things “that makes Einstein seem even more remarkable: he wasn’t motivated by any mysterious phenomena he couldn’t explain.” He depended simply on his insight into what sort of thing gravity must be and the beauty of the mathematics required to describe it.

After the theory was published, Einstein started to look for ways to test it through observation. One of them was to compare the apparent positions of stars that were in the same part of the sky as the sun during a solar eclipse with their apparent positions at other times. Rays of light, like free-falling objects, trace straight lines in space-time. Because the sun’s mass warps that space-time, the positions of the stars would seem to change when the rays skirted the sun (see diagram).

In 1919 Arthur Eddington, a famed British astronomer, announced that observations of an eclipse made on the West African island of Principe showed just the distortion Einstein had predicted (one of his images is pictured). “LIGHTS ALL ASKEW IN THE HEAVENS”, read the New York Times headline, adding helpfully that “Nobody Need Worry”. Einstein, while pleased, had faith enough in his idea not to have been on tenterhooks. When asked what he would have done had Eddington found a different result, he replied, “Then I would feel sorry for the good Lord. The theory is correct.”

As far as the rest of the world was concerned, Eddington’s result put general relativity more or less beyond doubt. But that did not make it mainstream. For one thing, it was hard to grasp. At a public event, Eddington was momentarily stumped by the suggestion that he “must be one of the three persons in the world who understand general relativity”. When the silence was taken for modesty, he replied “On the contrary, I am trying to think who the third person is!”

General relativity also seemed somewhat beside the point. The quantum revolution that Planck had begun, and that Einstein had contributed to in one of his other great papers of 1905, was bearing fascinating fruit. Together with a blossoming understanding of the atomic nucleus, it was at the centre of physicists’ attention. Special relativity had a role in the excitement; its most famous expression, E=mc2, gave a measure of the energy stored in those fascinating nuclei. General relativity had none.

What it offered instead was a way to ask questions not about what was in the universe, but about the structure of the universe as a whole. There were solutions to the equations in which the universe was expanding; there were others in which it was contracting. This became a topic of impassioned debate between Einstein and Willem de Sitter, a Dutch physicist who had found one of the expanding-universe solutions. Einstein wanted a static universe. In 1917 he added to his equations a “cosmological constant” which could be used to fix the universe at a given size.

That became an embarrassment when, in 1929, an American astronomer put forward strong evidence that the universe was, indeed, getting bigger. Edwin Hubble had measured the colour of the light from distant galaxies as a way of studying their motion; light from objects approaching the Earth looks bluer than it would otherwise, and light from objects receding looks redder. Hubble found that, on average, the more distant the galaxy, the more its light was shifted towards the red; things receded faster the farther away they were. The evidence for an expanding universe these redshifts provided led Einstein to reject the cosmological constant as the “greatest blunder of my life”.

The theory had other implications at which its architect initially baulked. In the 1930s nuclear physicists worked out that stars were powered by nuclear reactions, and that when those reactions ran out of fuel the stars would collapse. Something like the sun would collapse into a “white dwarf” about the size of the Earth. Bigger stars would collapse yet further into “neutron stars” as dense as an atomic nucleus and just 20 kilometres or so across. And the biggest stars would collapse into something with no length, breadth or depth but infinite density: a singularity.

Finding singularities in a theory is highly distasteful to the mathematically minded; they are normally signs of a mistake. Einstein did not want any of them in his universe, and in 1939 he published a paper attempting to show that the collapse of giant stars would be halted before a singularity could be formed. Robert Oppenheimer, a brilliant young physicist at Berkeley, used the same relativistic physics to contradict the great man and suggest that such extreme collapses were possible, warping space-time so much that they would create regions from which neither light nor anything else could ever escape: black holes.

Oppenheimer boi

Oppenheimer’s paper, though, was published on the day Germany invaded Poland, which rather put the debate on hold. Just a month before, Einstein had written to Franklin Roosevelt highlighting the military implications of E=mc2; it would be for realising those implications, rather than for black holes, that Oppenheimer would be remembered.

In part because of Oppenheimer’s government-bewitching success, new sorts of physical research flourished in the post-war years. One such field, radio astronomy, revealed cosmic dramas that observations using light had never hinted at. Among its discoveries were sources of radio waves that seemed at the same time small, spectacularly powerful and, judging by their redshifts, phenomenally distant. The astronomers dubbed them quasars and wondered what could possibly produce radio signals with the power of hundreds of billions of stars from a volume little bigger than a solar system.

Roy Kerr’s solution to the equations of general relativity provided the answer: a supermassive spinning black hole. Its rotation would create a region just outside the hole’s “event horizon”—the point of no return for light and everything else—in which matter falling inward would be spun up to enormous speeds. Some of that matter would be squirted out along the axis of rotation, forming the jets seen in radio observations of quasars.

Disappear like smoke

For the first time, general relativity was explaining new phenomena in the world. Bright young minds rushed into the field; wild ideas that had been speculated on in the fallow decades were buffed up and taken further. There was talk of “wormholes” in space-time that could connect seemingly distant parts of the universe. There were “closed time-like curves” that seemed as though they might make possible travel into the past. Less speculatively, but with more profound impact, Stephen Hawking, a physicist (pictured, with a quasar), and Roger Penrose, a mathematician, showed that relativistic descriptions of the singularities in black holes could be used to describe the Big Bang in which the expansion of the universe began—that they were, in fact, the only way to make sense of it. General relativity gave humans their first physical account of creation.

Hawking boi

Dr Hawking went on to bring elements of quantum theory into science’s understanding of the black hole. Quantum mechanics says that if you look at space on the tiniest of scales you will see a constant ferment in which pairs of particles pop into existence and then recombine into nothingness. Dr Hawking argued that when this happens at the event horizon of a black hole, some of the particles will be swallowed up, while some will escape. These escaping particles mean, in Dr Hawking’s words, that “black holes ain’t so black”—they give off what is now called “Hawking radiation”. The energy lost this way comes ultimately from the black hole itself, which gives up mass in the process. Thus, it seems, a black hole must eventually evaporate away to nothingness.

Adding quantum mechanics to the description of black holes was a step towards what has become perhaps the greatest challenge in theoretical physics: reconciling the theory used to describe all the fields and particles within the universe with the one that explains its overall shape. The two theories view reality in very different ways. In quantum theory, everything is, at some scale, bitty. The equations of relativity are fundamentally smooth. Quantum mechanics deals exclusively in probabilities—not because of a lack of information, but because that is the way the world actually is. In relativity all is certain. And quantum mechanics is “non-local”; an object’s behaviour in one place can be “entangled” with that of an object kilometres or light-years away. Relativity is proudly local; Einstein was sure that the “spooky action at a distance” implied by quantum mechanics would disappear when a better understanding was reached.

It hasn’t. Experiment after experiment confirms the non-local nature of the physical world. Quantum theory has been stunningly successful in other ways, too. Quantum theories give richly interlinked accounts of electromagnetism and of the strong and weak nuclear forces—the processes that hold most atoms together and split some apart. This unified “standard model” now covers all observable forms of matter and all their interactions—except those due to gravity.

Some people might be satisfied just to let each theory be used for what it is good for and to worry no further. But people like that do not become theoretical physicists. Nor will they ever explain the intricacies of the Big Bang—a crucible to which grandiose theory-unifiers are ceaselessly drawn. In the very early universe, space-time itself seems to have been subject to the sort of fluctuations fundamental to the quantum world (like those responsible for Hawking radiation). Getting to the heart of such shenanigans requires a theory that combines the two approaches.

There have been many rich and subtle attempts at this. Dr Penrose has spent decades elaborating an elegant way of looking at all fields and particles as new mathematical entities called “twistors”. Others have pursued a way of adding quantum bittiness to the fabric of space-time under the rubric of “loop quantum gravity”. Then there is the “Exceptionally Simple Theory of Everything”—which isn’t. As Steven Weinberg, one of the unifiers whose work built the standard model, puts it, “There are so many theories and so few observations that we’re not getting very far.”

Dr Weinberg, like many of his colleagues, fancies an approach called superstring theory. It is an outgrowth of the standard model with various added features that seem as though they would help in the understanding of space-time and which its proponents find mathematically beguiling. Ed Witten of the Institute for Advanced Study (IAS) in Princeton, Einstein’s institutional home for the last 22 years of his life, is one of those who has raised it to its current favoured status. But he warns that much of the theory remains to be discovered and that no one knows how much. “We only understand bits and pieces—but the bits and pieces are staggeringly beautiful.”

This piecemeal progress, as Dr Witten tells it, offers a nice counterpoint to the process which led up to November 1915. “Einstein had the conception behind general relativity before he had the theory. That’s in part why it has stood: it was complete when it was formulated,” he says. “String theory is the opposite, with many manifestations discovered by happy accident decades ago.”

Entangled up in the blue

And the happy accidents continue. In 1997 Juan Maldacena, an Argentine theoretician who now also works at the IAS, showed that there is a deep connection between formulations of quantum mechanics known as conformal field theories and solutions to the Einstein equations called anti-de Sitter spaces (similar to the expanding-universe solution derived by Willem de Sitter, but static and much favoured by string theorists). Neither provides an account of the real world, but the connection between them lets physicists recast intractable problems in quantum mechanics into the sort of equations found in general relativity, making them easier to crack.

This approach is being gainfully employed in solving problems in materials science, superconductivity and quantum computing. It is also “influencing the field in a totally unexpected way,” says Leonard Susskind, of Stanford University. “It’s a shift in our tools and our methodology and our way of thinking about how phenomena are connected.” One possibility Dr Maldacena and Dr Susskind have developed by looking at things this way is that the “wormholes” relativity allows (which can be found in the anti-de Sitter space) may be the same thing as the entanglement between distant particles in quantum mechanics (which is part of the conformal field theory). The irony of Einstein’s spooky quantum bête noire playing such a crucial role has not gone unremarked.

There is more to the future of relativity, though than its eventual subsumption into some still unforeseeable follow-up theory. As well as offering new ways of understanding the universe, it is also providing new ways of observing it.

This is helpful because there are bits of the universe that are hard to observe in other ways. Much of the universe consists of “dark matter” which emits no radiation. But it has mass, and so it warps space, distorting the picture of more distant objects just as the eclipse-darkened sun distorted the positions of Eddington’s stars. Studying distortions created by such “gravitational lenses”—both luminous (pictured, with Einstein) and dark—allows astronomers with the precise images of the deep sky today’s best telescopes provide to measure the distribution of mass around the universe in a new way.

A century ago Albert Einstein changed the way humans saw the universe. His work is still offering new insights today.
Einstein boi

Another form of relativity-assisted astronomy uses gravitation directly. Einstein’s equations predict that when masses accelerate around each other they will create ripples in space-time: gravitational waves. As with black holes and the expanding universe, Einstein was not keen on this idea. Again, later work has shown it to be true. A pair of neutron stars discovered spinning around each other in the 1970s are exactly the sort of system that should produce such waves. Because producing gravitational waves requires energy, it was realised that these neutron stars should be losing some. And so they proved to be—at exactly the rate that relativity predicts. This indirect but convincing discovery garnered a Nobel prize in 1993.

As yet, though, no one has seen a wave in action by catching the expansion and contraction of space that should be seen as one goes by, because the effects involved are ludicrously small. But researchers at America’s recently upgraded Laser Interferometer Gravitational-wave Observatory (LIGO) now think they can do it. At LIGO’s two facilities, one in Louisiana and one in Washington state, laser beams bounce up and down 4km-long tubes dozens of times before being combined in a detector to make a pattern. A passing gravitational wave that squashes space-time by a tiny fraction of the radius of an atomic nucleus in one arm but not the other will make a discernible change to that pattern. Comparing measurements at the two sites could give a sense of the wave’s direction.

Step into the light

The aim is not just to detect gravitational waves—though that would be a spectacular achievement—but to learn about the processes that produce them, such as mergers of neutron stars and black holes. The strengths of the warping effects in such cataclysms are unlike anything seen to date; their observation would provide a whole new type of test for the theory.

And history suggests there should be completely unanticipated discoveries, too. Kip Thorne, a specialist in relativity at the California Institute of Technology and co-founder of LIGO, says that “every time we’ve opened a new window on the cosmos with new radiation, there have been unexpected surprises”. For example, the pioneers of radio astronomy had no inkling that they would discover a universe full of quasars—and thus black holes. A future global array of gravitational-wave observatories could open a whole new branch of observational astronomy.

A century ago general relativity answered no one’s questions except its creator’s. Many theories are hit upon by two or more people at almost the same time; but if Einstein had not devoted years to it, the curvature of space-time which is the essence of gravity might not have been discovered for decades. Now it has changed the way astronomers think about the universe, has challenged them to try and build theories to explain its origin, and even offered them new ways to inspect its contents. And still, it retains what most commended it to Einstein: its singular beauty revealed first to his eyes alone but appreciated today by all who have followed. “The Einstein equations of general relativity are his best epitaph and memorial,” Stephen Hawking has written. “They should last as long as the universe.”

Should we feel excited or frightened by the idea of an AI model directing a robot?

I am a data science student and more than once I’ve wondered if I’m doing all this studying for nought, because GPT, the AI, can do what I do; well except stand in front of my bosses and do a presentation, until they can hire some robots. But for now, I remain relevant.

“If the AI can replace my work, then I don’t think I’m not doing a good job,” uttered one Alex Konrad, Forbes Senior Editor in an interview. Safe to say I live by these words now.

It is the age of AI. Albeit, AI has been there for a while, the likes of Siri, Cortana, Google Assistant, and Alexa, but not as much as what this year has brought us. We have OpenAI’s ChatGPT, Midjourney, Google’s Bard, Jasper, Stability, Bing AI, etc. Image generation from a prompt is incredibly brilliant, don’t you think? Coding capabilities, are also very impressive (but does it work? Sometimes, yeah).

Artificial Intelligence has many a definition, the crux of it being, “the theory and development of computer systems able to perform tasks normally requiring human intelligence”. How does the computer learn? Through Machine Learning – it detects the patterns from training data and predicts and performs tasks without being manually or explicitly programmed. We have one more, Deep learning (done via neural networks) – a method in AI that teaches computers to process data in a way that is inspired by the human brain.

A frequently asked and debated question is, “Does the AI know what it’s doing? “.

ChatGPT is an artificial intelligence chatbot developed by OpenAI and launched on November 30, 2022. In full it is Chat Generative Pre-Training Transformer. When you feed it a prompt, it gives a very coherent output, mostly true, might be false, might be completely made up. Why is this? Why does a supposedly intelligent machine lie? Does it know it’s lying? Most of the ‘experts’ (or insert equally semantic word here) say it doesn’t. This is because the output is merely a prediction of text based on the prompt you have given. ChatGPT was trained on a massive corpus of text data, around 570GB of datasets, including web pages, books, and other sources. It was born after running trillions of words for the equivalent of 300 years, through supercomputers processing in parallel for months. After all this, the computer made about 170 billion connections between all these words. Incomprehensible, isn’t it? Math is beautiful. So anytime you enter a prompt, ChatGPT calculates through all these connections to give you the most appropriate prediction of words that had the highest probability after all the back-end math was done. (Look into neural networks, it’s interesting to watch a machine be taught how to recognize handwritten numbers.) So yes, it is very possible that ChatGPT doesn’t know what you’re asking it, or what it’s replying to. It’s all math and predictions. However, this will change as we keep teaching AI to not just predict data but to understand and learn any intellectual task.

Is there a law to protect us against AI?

In April 2021, the European Commission proposed the first EU regulatory framework for AI. It says that AI systems that can be used in different applications are analyzed and classified according to the risks they pose to users. The different risk levels will mean more or less regulation. The first ever AI Act was created on the 8th of June, 2023 by the European Commission for the use of artificial intelligence in the European Union as part of its digital strategy. It contains different rules for different risk levels and Generative AI transparency requirements. This is a good start because there are several concerns about the data used to train the AI. For example, in February, Getty Images sued Stability AI, a smaller AI start-up, alleging it illegally used its photos to train its image-generating bot. ChatGPT maker OpenAI is facing a class action over how it used people’s data. We have people creating content via AI and passing it off as theirs, and this information might be false. There are many concerns over who is responsible if a human was to use AI for ‘wrong’.

Is AI sentient?

Can AI perceive or feel things? Current applications of AI, like language models (i.e., GPT and Google’s LaMDA), are not sentient. They are only trained to sound like they know what they are talking about – ‘they’ being AI collectively. Will we know if it ever became sentient? There is no consensus on accurately determining if an AI is conscious, given our current understanding of consciousness. Scary sounding, isn’t it? That’s because it is, but it is also very exciting.

The AI revolution is here, and a lot is about to change. Are we ready?

A particular type of enthusiasm fills the air as the sun sets over the Nairobi skyline. For many years, astronomy was thought to be a subject only for experts and astrophysics students. However, a rising number of Kenyans are becoming interested in the mysteries of the night sky.

The societal preconception that astronomy is a complex and foreign notion is being broken down, thanks to the efforts of various individuals and organizations. Amateur astronomy activities and exercises have been introduced, making it available to the public at large. These programs, which range from webinars to skywatching events, astronomy lecture nights to planetarium displays, are affordable, engaging, and open to anybody who is curious about the expanse of space.

But why this recent enthusiasm in astronomy? The answer is found in the fact that astronomy is one of the oldest natural sciences. Throughout history, several cultures employed astronomy for navigating, telling time, predicting the weather, farming, and even celebrating various holidays.

Organizations in Kenya have participated in efforts such as the NameExoWorlds campaign, which allows nations to name exoplanets in their local language. The Kenya Space Agency has collaborated with several groups to give information to amateur astronomers, institutions, and the general public, with the goal of increasing student and citizen interest in STEM fields.

The Agency’s space club program teaches the next generation on the necessity of GEO-STEAM (Geography, Science, Technology, Engineering, Arts, and Mathematics) in supporting the space sector. And the efforts are bearing fruit. Kenyans are becoming innovative in their quest of astronomy, creating water-powered rockets and hosting public astronomy presentations on the capital’s streets.

TikTokers are creating content about the planetarium at the Traveling Telescope, where people go to learn about the universe with family, friends, or significant others. It is clear that astronomy is no longer a novel notion in Nairobi, and we can anticipate to see more people dressed warmly and eagerly attending these activities in the future.

These efforts are not only aimed at unraveling the mysteries of the universe, but also at connecting our Kenyan culture to space and astronomy. Astronomy is an excellent opportunity to bring people from all walks of life together and foster a greater respect for the universe. So, as the stars twinkle above, let us continue to gaze up and marvel at the cosmos’ beauty.

In a remarkable feat, Kenya continues to make waves in astronomy as it achieves yet another milestone in naming celestial bodies. Following last year’s successful participation in the NameExoWorlds campaign, Kenya has now secured its second exoplanet and host star with Kenyan names. The significance of this achievement lies not only in the recognition of Kenya’s cultural heritage but also in the country’s growing influence in the field of astronomy.

The team behind this accomplishment consisted of prominent individuals and institutions, including Penda Kujua, Leo Sky Africa, the Technical University of Kenya, the Kenya Space Agency, and the University of Nairobi. Together, they collaborated to come up with the approved name for the newly named exoplanet and its host star.

Astronomy has been embraced in Kenya

Previously, Kenya participated in the NameExoWorlds campaign, resulting in the naming of a star “Kalausi” and its accompanying exoplanet. The name “Kalausi,” meaning a strong whirlwind in the Dholuo language, carries the essence of Kenyan culture and has become a testament to the country’s rich linguistic diversity.

Building upon this success, the recent NameExoWorlds 2022 contest has now bestowed a Kenyan name upon a second exoplanet and its host star. Enaiposha, named after a large body of water like a lake or sea in the Maa language spoken in Kenya and Tanzania, is a fitting tribute to the captivating exoplanet GJ 1214 b. GJ 1214 b, known as a “sub-Neptune” due to its intermediate size between Earth and Neptune, has been extensively studied by scientists and continues to fascinate researchers around the globe.

Accompanying Enaiposha is the star GJ 1214, which hosts the planetary system. This celestial body, now bearing the name Orkaria, further highlights the cultural significance of the naming initiative. Orkaria adds depth and meaning to the exoplanetary system, connecting Kenya’s rich heritage to the wonders of the cosmos. The selection of Enaiposha and Orkaria as the second exoplanet and star to be given Kenyan names emphasizes the country’s growing prominence in the field of astronomy. It underscores Kenya’s commitment to expanding scientific knowledge and promoting public engagement in astronomical exploration.

Through the NameExoWorlds campaigns, the International Astronomical Union recognizes and celebrates the importance of cultural connections to the celestial realm. By incorporating indigenous languages and cultural themes, the initiative fosters a sense of inclusivity and encourages a global appreciation for the wonders of the universe.

With each named exoplanet and host star, Kenya’s presence on the international astronomical stage becomes more pronounced. These achievements inspire, fueling the curiosity and passion of current and future generations of astronomers and space enthusiasts in Kenya and beyond.

I want to make you look up into the sky and comprehend, maybe for the first time, the darkness that lies beyond the evanescent wisp of the atmosphere, the endless depths of the cosmos, a desolation by degrees

I want the Earth to turn beneath you and knock your balance off, carry you eastward at a thousand miles an hour, into the light, and the dark, and the light again. I want you to watch the Earth rising you up to meet the rays of the morning sun

I want the sky to stop you dead in your tracks on your walk home tonight because you happened to glance up and among all the shining pinpricks you recognized one as the light of an alien world

I want you to taste the iron in your blood and see its likeness in the rust-red sands on the long dry dunes of Mars, born of the same nebular dust that coalesced random flotsam of stellar debris into rocks, oceans, your own beating heart

I want to reach into your consciousness and cast it outward, beyond the light of other suns, to expand it like the universe, not encroaching on some envelope of emptiness, but growing larger, unfolding inside itself

I want you to see your world from four billion miles away, a tiny glint of blue in the sharp white light of an ordinary star in the darkness. I want you to try to make out the boundaries of your nation from that vantage point and fail

I want you to feel it, in your bones, in your breath, when two black holes colliding a billion light years away send a tremor through spacetime that makes every cell in your body stretch, and strain

I want to make you nurse nostalgia for the stars long dead, the ones that fused your carbon nuclei and the ones whose last thermonuclear death throes outshined the entire galaxy to send a single photon into your eye

I want you to live forward but see backwards, farther and deeper into the past, because in a relativistic universe, you don’t have any other choice. I want the stale billion-year-old starlight of a distant galaxy to be your reward

I want to utterly disorient you and let you navigate back by the stars. I want you to lose yourself, and find it again, not just here, but everywhere, in everything

I want you to believe that the universe is a vast, random, uncaring place, in which our species, our world, has absolutely no significance. And I want you to believe that the only response is to make our own beauty and meaning and to share it while we can

I want to make you wonder what is out there. What dreams may come in waves of radiation across the breadth of an endless expanse? What we may know, given time, and what splendours might never, ever reach us

I want to make it mean something to you. That you are in the cosmos. That you are of the cosmos. That you are born from stardust and to stardust you will return. That you are a way for the universe to be in awe of itself.

The first millennium BC witnessed the appearance of three potentially universal orders, whose devotees could for the first time imagine the entire world and the entire human race as a single unit governed by a single set of laws. Everyone was ‘us’, at least potentially. There was no longer ‘them’. The first universal order to appear was economic: the monetary order. The second universal order was political: the imperial order. The third universal order was religious: the order of universal religions such as Buddhism, Christianity and Islam.

Merchants, conquerors and prophets were the first people who managed to transcend the binary evolutionary division, ‘us vs them’, and to foresee the potential unity of humankind. For the merchants, the entire world was a single market and all humans were potential customers. They tried to establish an economic order that would apply to all, everywhere. For the conquerors, the entire world was a single empire and all humans were potential subjects, and for the prophets, the entire world held a single truth and all humans were potential believers. They too tried to establish an order that would be applicable to everyone everywhere.

During the last three millennia, people made more and more ambitious attempts to realize that global vision. Let us begin with the story of the greatest conqueror in history, a conqueror possessed of extreme tolerance and adaptability, thereby turning people into ardent disciples.

This conqueror is money.

People who do not believe in the same God or obey the same king are more than willing to use the same money. Osama Bin Laden, for all his hatred of American culture, American religion and American politics, was very fond of American dollars.

How did money succeed where gods and kings failed?

In order to understand why, consider a hypothetical case. Assume that when regular trade opened between Mali and the Mediterranean, Malians were uninterested in gold, so it was almost worthless. But in the Mediterranean, gold was a coveted status symbol, hence its value was high. What would happen next? Merchants travelling between Mali and the Mediterranean would notice the difference in the value of gold. In order to make a profit, they would buy gold cheaply in Mali and sell it dearly in the Mediterranean. Consequently, the demand for gold in Mali would skyrocket, as would its value. At the same time, the Mediterranean would experience an influx of gold, whose value would consequently drop. Within a short time, the value of gold in Mali and the Mediterranean would be quite similar. The mere fact that Mediterranean people believed in gold would cause Malians to start believing in it as well. Even if Malians still had no real use for gold, the fact that Mediterranean people wanted it would be enough to make the Malians value it.

Similarly, the fact that another person believes in cowry shells, dollars, or electronic data, is enough to strengthen our own belief in them, even if that person is otherwise hated, despised or ridiculed by us. Christians and Muslims who could not agree on religious beliefs could nevertheless agree on a monetary belief because, whereas religion asks us to believe in something, money asks us to believe that other people believe in something.

That’s how money works. Cowry shells and banknotes have value only in our common imagination. Their worth is not inherent in the chemical structure of the shells and paper, their colour, or their shape. In other words, money isn’t a material reality – it is a psychological construct. It works by converting matter into mind. But why does it succeed? Why should anyone be willing to exchange a fertile rice paddy for a handful of useless cowry shells? Why are you willing to flip hamburgers, sell health insurance or babysit three obnoxious brats when all you get for your exertions is a few pieces of coloured paper?

People are willing to do such things when they trust the figments of their collective imagination. Trust is the raw material from which all types of money are minted. When a wealthy farmer sold his possessions for a sack of cowry shells and travelled with them to another province, he trusted that upon reaching his destination other people would be willing to sell him rice, houses and fields in exchange for the shells. Money is accordingly a system of mutual trust, and not just any system of mutual trust: money is the most universal and most efficient system of mutual trust ever devised.

Why do I believe in the cowry shell or gold coin or dollar bill? Because my neighbours believe in them. And my neighbours believe in them because I believe in them. And we all believe in them because our king believes in them and demands them in taxes and because our priest believes in them and demands them in tithes. We accept the dollar in payment because we trust in the US secretary of the treasury. The crucial role of trust explains why our financial systems are so tightly bound up with our political, social and ideological systems, why financial crises are often triggered by political developments, and why the stock market can rise or fall depending on the way traders feel on a particular morning.

For thousands of years, philosophers, thinkers and prophets have besmirched money and called it the root of all evil. Be that as it may, money is also the apogee of human tolerance. Money is more open-minded than language, state laws, cultural codes, religious beliefs and social habits. Money is the only trust system created by humans that can bridge almost any cultural gap, and that does not discriminate on the basis of religion, gender, race, age or sexual orientation. Thanks to money, even people who don’t know each other and don’t trust each other can nevertheless cooperate effectively.

Money is based on two universal principles:

  • Universal convertibility: with money as an alchemist, you can turn land into loyalty, justice into health, and violence into knowledge.
  • Universal trust: with money as a go-between, any two people can cooperate on any project.

These principles have enabled millions of strangers to cooperate effectively in trade and industry. But these seemingly benign principles have a dark side. When everything is convertible, and when trust depends on anonymous coins and cowry shells, it corrodes local traditions, intimate relations and human values, replacing them with the cold laws of supply and demand.

Human communities and families have always been based on the belief in ‘priceless’ things, such as honour, loyalty, morality and love. These things lie outside the domain of the market, and they shouldn’t be bought or sold for money. Even if the market offers a good price, certain things just aren’t done. Parents mustn’t sell their children into slavery; a devout Christian must not commit a mortal sin; a loyal knight must never betray his lord, and ancestral tribal lands shall never be sold to foreigners.

Money has always tried to break through these barriers, like water seeping through cracks in a dam. Parents have been reduced to selling some of their children into slavery in order to buy food for others. Devout Christians have murdered, stolen and cheated – and later used their spoils to buy forgiveness from the Church. Ambitious knights auctioned their allegiance to the highest bidder while securing the loyalty of their own followers by cash payments. Tribal lands were sold to foreigners from the other side of the world in order to purchase an entry ticket into the global economy.

Money has an even darker side. Although money builds universal trust between strangers, this trust is invested not in humans, communities or sacred values, but in money itself and in the impersonal systems that back it. We do not trust the stranger or the next-door neighbour – we trust the coin they hold. If they run out of coins, we run out of trust. As money brings down the dams of community, religion and state, the world is in danger of becoming one big and rather heartless marketplace. Hence the economic history of humankind is a delicate dance. People rely on money to facilitate cooperation with strangers, but they’re afraid it will corrupt human values and intimate relations. On one hand, people willingly destroy the communal dams that held at bay the movement of money and commerce for so long. Yet, on the other hand, they build new dams to protect society, religion and the environment from enslavement by market forces.

It is common nowadays to believe that the market always prevails and that the dams erected by kings, priests and communities cannot long hold back the tides of money. This is naïve. Brutal warriors, religious fanatics and concerned citizens have repeatedly managed to trounce calculating merchants, and even reshape the economy. It is therefore impossible to understand the unification of humankind as a purely economic process. In order to understand how thousands of isolated cultures coalesced over time to form the global village of today, we must take into account the role of gold and silver, but we cannot disregard the equally crucial role of steel. Enter politics . . .

Seventy thousand years ago, Homo sapiens was still an insignificant animal minding its own business right here in East Africa.

In the following millennia, it transformed itself into the master of the entire planet and the terror of the ecosystem. Today it stands on the verge of becoming a god, poised to acquire not only eternal youth but also the divine abilities of creation and destruction.

Unfortunately, the Sapiens regime on Earth has so far produced little that we can be proud of. We have mastered our surroundings, increased food production, built cities, established empires and created far-flung trade networks. But did we decrease the amount of suffering in the world? Time and again, massive increases in human power did not necessarily improve the well-being of individual Sapiens and usually caused immense misery to other animals.

In the last few decades we have at last made some real progress as far as the human condition is concerned, with the reduction of famine, plague and war. Yet the situation of other animals is deteriorating more rapidly than ever before, and the improvement in a lot of humanity is too recent and fragile to be certain of.

Moreover, despite the astonishing things that humans are capable of doing, we remain unsure of our goals and we seem to be as discontented as ever. We have advanced from canoes to galleys to steamships to space shuttles – but nobody knows where we’re going. We are more powerful than ever before but have very little idea of what to do with all that power. Worse still, humans seem to be more irresponsible than ever. Self-made gods with only the laws of physics to keep us company, we are accountable to no one. We are consequently wreaking havoc on our fellow animals and the surrounding ecosystem, seeking little more than our own comfort and amusement, yet never finding satisfaction.

Is there anything more dangerous than dissatisfied and irresponsible gods who don’t know what they want?