September 24, 2002  The New York Times

Here They Are, Science's 10 Most Beautiful Experiments

By GEORGE JOHNSON

Whether they are blasting apart subatomic particles in accelerators, sequencing the genome or analyzing the wobble of a distant star, the experiments that grab the world's attention often cost millions of dollars to execute and produce torrents of data to be processed over months by supercomputers. Some research groups have grown to the size of small companies.

But ultimately science comes down to the individual mind grappling with something mysterious. When Robert P. Crease, a member of the philosophy department at the State University of New York at Stony Brook and the historian at Brookhaven National Laboratory, recently asked physicists to nominate the most beautiful experiment of all time, the 10 winners were largely solo performances, involving at most a few assistants. Most of the experiments — which are listed in this month's Physics World — took place on tabletops and none required more computational power than that of a slide rule or calculator.

What they have in common is that they epitomize the elusive quality scientists call beauty. This is beauty in the classical sense: the logical simplicity of the apparatus, like the logical simplicity of the analysis, seems as inevitable and pure as the lines of a Greek monument. Confusion and ambiguity are momentarily swept aside, and something new about nature becomes clear.

The list in Physics World was ranked according to popularity, first place going to an experiment that vividly demonstrated the quantum nature of the physical world. But science is a cumulative enterprise — that is part of its beauty. Rearranged chronologically and annotated below, the winners provide a bird's-eye view of more than 2,000 years of discovery.

Eratosthenes' measurement of the Earth's circumference

At noon on the summer solstice in the Egyptian town now called Aswan, the sun hovers straight overhead: objects cast no shadow and sunlight falls directly down a deep well. When he read this fact, Eratosthenes, the librarian at Alexandria in the third century B.C., realized he had the information he needed to estimate the circumference of the planet. On the same day and time, he measured shadows in Alexandria, finding that the solar rays there had a bit of a slant, deviating from the vertical by about seven degrees.

The rest was just geometry. Assuming the earth is spherical, its circumference spans 360 degrees. So if the two cities are seven degrees apart, that would constitute seven-360ths of the full circle — about one-fiftieth. Estimating from travel time that the towns were 5,000 "stadia" apart, Eratosthenes concluded that the earth must be 50 times that size — 250,000 stadia in girth. Scholars differ over the length of a Greek stadium, so it is impossible to know just how accurate he was. But by some reckonings, he was off by only about 5 percent. (Ranking: 7)

Clockwise, from top left:
 
Eratosthenes,
Galileo Galilei,
Henry Cavendish
and Isaac Newton.


Galileo's experiment on falling objects

In the late 1500's, everyone knew that heavy objects fall faster than lighter ones. After all, Aristotle had said so. That an ancient Greek scholar still held such sway was a sign of how far science had declined during the dark ages.

Galileo Galilei, who held a chair in mathematics at the University of Pisa, was impudent enough to question the common knowledge. The story has become part of the folklore of science: he is reputed to have dropped two different weights from the town's Leaning Tower showing that they landed at the same time. His challenges to Aristotle may have cost Galileo his job, but he had demonstrated the importance of taking nature, not human authority, as the final arbiter in matters of science. (Ranking: 2)

Galileo's experiments with rolling balls down inclined planes

Galileo continued to refine his ideas about objects in motion. He took a board 12 cubits long and half a cubit wide (about 20 feet by 10 inches) and cut a groove, as straight and smooth as possible, down the center. He inclined the plane and rolled brass balls down it, timing their descent with a water clock — a large vessel that emptied through a thin tube into a glass. After each run he would weigh the water that had flowed out — his measurement of elapsed time — and compare it with the distance the ball had traveled.

Aristotle would have predicted that the velocity of a rolling ball was constant: double its time in transit and you would double the distance it traversed. Galileo was able to show that the distance is actually proportional to the square of the time: Double it and the ball would go four times as far. The reason is that it is being constantly accelerated by gravity. (Ranking: 8)

Newton's decomposition of sunlight with a prism

Isaac Newton was born the year Galileo died. He graduated from Trinity College, Cambridge, in 1665, then holed up at home for a couple of years waiting out the plague. He had no trouble keeping himself occupied.

The common wisdom held that white light is the purest form (Aristotle again) and that colored light must therefore have been altered somehow. To test this hypothesis, Newton shined a beam of sunlight through a glass prism and showed that it decomposed into a spectrum cast on the wall. People already knew about rainbows, of course, but they were considered to be little more than pretty aberrations. Actually, Newton concluded, it was these colors — red, orange, yellow, green, blue, indigo, violet and the gradations in between — that were fundamental. What seemed simple on the surface, a beam of white light, was, if one looked deeper, beautifully complex. (Ranking: 4)

 


Cavendish's torsion-bar experiment

Another of Newton's contributions was his theory of gravity, which holds that the strength of attraction between two objects increases with the square of their masses and decreases with the square of the distance between them. But how strong is gravity in the first place?

In the late 1700's an English scientist, Henry Cavendish, decided to find out. He took a six-foot wooden rod and attached small metal spheres to each end, like a dumbbell, then suspended it from a wire. Two 350-pound lead spheres placed nearby exerted just enough gravitational force to tug at the smaller balls, causing the dumbbell to move and the wire to twist. By mounting finely etched pieces of ivory on the end of each arm and in the sides of the case, he could measure the subtle displacement. To guard against the influence of air currents, the apparatus (called a torsion balance) was enclosed in a room and observed with telescopes mounted on each side.

The result was a remarkably accurate estimate of a parameter called the gravitational constant, and from that Cavendish was able to calculate the density and mass of the earth. Erastothenes had measured how far around the planet was. Cavendish had weighed it: 6.0 x 1024 kilograms, or about 13 trillion trillion pounds. (Ranking: 6)

Young's light-interference experiment

Newton wasn't always right. Through various arguments, he had moved the scientific mainstream toward the conviction that light consists exclusively of particles rather than waves. In 1803, Thomas Young, an English physician and physicist, put the idea to a test. He cut a hole in a window shutter, covered it with a thick piece of paper punctured with a tiny pinhole and used a mirror to divert the thin beam that came shining through. Then he took "a slip of a card, about one-thirtieth of an inch in breadth" and held it edgewise in the path of the beam, dividing it in two. The result was a shadow of alternating light and dark bands — a phenomenon that could be explained if the two beams were interacting like waves.

Bright bands appeared where two crests overlapped, reinforcing each other; dark bands marked where a crest lined up with a trough, neutralizing each other.

The demonstration was often repeated over the years using a card with two holes to divide the beam. These so-called double-slit experiments became the standard for determining wavelike motion — a fact that was to become especially important a century later when quantum theory began. (Ranking: 5)

Clockwise, from top left:

Thomas Young,
Jean-Bernard-Leon Foucault,
 Ernest Rutherford
and Robert Millikan.

Foucault's pendulum

Last year when scientists mounted a pendulum above the South Pole and watched it swing, they were replicating a celebrated demonstration performed in Paris in 1851. Using a steel wire 220 feet long, the French scientist Jean-Bernard-Léon Foucault suspended a 62-pound iron ball from the dome of the Panthéon and set it in motion, rocking back and forth. To mark its progress he attached a stylus to the ball and placed a ring of damp sand on the floor below.

The audience watched in awe as the pendulum inexplicably appeared to rotate, leaving a slightly different trace with each swing. Actually it was the floor of the Panthéon that was slowly moving, and Foucault had shown, more convincingly than ever, that the earth revolves on its axis. At the latitude of Paris, the pendulum's path would complete a full clockwise rotation every 30 hours; on the Southern Hemisphere it would rotate counterclockwise, and on the Equator it wouldn't revolve at all. At the South Pole, as the modern-day scientists confirmed, the period of rotation is 24 hours. (Ranking: 10)

Millikan's oil-drop experiment

Since ancient times, scientists had studied electricity — an intangible essence that came from the sky as lightning or could be produced simply by running a brush through your hair. In 1897 (in an experiment that could easily have made this list) the British physicist J. J. Thomson had established that electricity consisted of negatively charged particles — electrons. It was left to the American scientist Robert Millikan, in 1909, to measure their charge.

Using a perfume atomizer, he sprayed tiny drops of oil into a transparent chamber. At the top and bottom were metal plates hooked to a battery, making one positive and the other negative. Since each droplet picked up a slight charge of static electricity as it traveled through the air, the speed of its descent could be controlled by altering the voltage on the plates. (When this electrical force matched the force of gravity, a droplet — "like a brilliant star on a black background" — would hover in midair.)

Millikan observed one drop after another, varying the voltage and noting the effect. After many repetitions he concluded that charge could only assume certain fixed values. The smallest of these portions was none other than the charge of a single electron. (Ranking: 3)

Rutherford's discovery of the nucleus

When Ernest Rutherford was experimenting with radioactivity at the University of Manchester in 1911, atoms were generally believed to consist of large mushy blobs of positive electrical charge with electrons embedded inside — the "plum pudding" model. But when he and his assistants fired tiny positively charged projectiles, called alpha particles, at a thin foil of gold, they were surprised that a tiny percentage of them came bouncing back. It was as though bullets had ricocheted off Jell-O.

Rutherford calculated that actually atoms were not so mushy after all. Most of the mass must be concentrated in a tiny core, now called the nucleus, with the electrons hovering around it. With amendments from quantum theory, this image of the atom persists today. (Ranking: 9)

Young's double-slit experiment applied to the interference of single electrons

Neither Newton nor Young was quite right about the nature of light. Though it is not simply made of particles, neither can it be described purely as a wave. In the first five years of the 20th century, Max Planck and then Albert Einstein showed, respectively, that light is emitted and absorbed in packets — called photons. But other experiments continued to verify that light is also wavelike.

It took quantum theory, developed over the next few decades, to reconcile how both ideas could be true: photons and other subatomic particles — electrons, protons, and so forth — exhibit two complementary qualities; they are, as one physicist put it, "wavicles."

To explain the idea, to others and themselves, physicists often used a thought experiment, in which Young's double-slit demonstration is repeated with a beam of electrons instead of light. Obeying the laws of quantum mechanics, the stream of particles would split in two, and the smaller streams would interfere with each other, leaving the same kind of light- and dark-striped pattern as was cast by light. Particles would act like waves.

According to an accompanying article in Physics World, by the magazine's editor, Peter Rodgers, it wasn't until 1961 that someone (Claus Jönsson of Tübingen) carried out the experiment in the real world.

By that time no one was really surprised by the outcome, and the report, like most, was absorbed anonymously into science. (Ranking: 1)

 

 

 

 

 

 

 

 

 



October 20, 1999 SCIENTIFIC AMERICAN
Modeling the Atomic Universe
By Shawn Carlson
Grant that the universe is filled with atomic-size billiard balls. Then, with a few insightful definitions and some mathematical gymnastics, physicists can provide you with a near-perfect explanation of our everyday world. The theory is called statistical mechanics. Many people know that it limits the amount of work a machine can deliver. But it actually goes much further. Statistical mechanics describes the engines that drive the earth's weather. It governs the temperatures and pressures inside stars and constrains the evolution of the cosmos. It even sheds light on the arrow of time--why we remember the past and not the future. Indeed, Albert Einstein and Richard Feynman saw the theory as the highest achievement of classical physics.

Sadly, many amateurs have avoided this important subject because, in this case, the highest plateau is also the hardest to reach. One cubic centimeter of air at atmospheric pressure contains more than 10 billion billion atoms of various sizes, all smashing into one another at different speeds. No computer can project the exact trajectories of all these particles, and even if one could, no human mind could make sense of it. Therefore, physicists have devised clever but devilishly difficult mathematical methods to extract comprehension from the chaos.

Molecular Dynamics
Image: George Musser

ATOMS of helium (gold) and krypton (red) clump when the temperature is low....

Molecular Dynamics
Daniels & Daniels; Source: Molecular Modeling
 
but as the gas heats up, the lighter heliums are torn asunder....
Daniels & Daniels; Source: Molecular Modeling

...and at still higher temperatures, the heavier kryptons fly apart, too.
 

But the subject is not as abstruse as it seems. The trick is to find models that let you visualize how these random collisions average out to yield the familiar properties of matter, such as temperature, pressure and entropy. The right mental pictures can elucidate the behavior of materials and in turn can help advance amateur projects involving chemistry, sound, heat transfer, crystals and vacuum techniques. That's why I'm pleased to let you know about Molecular Dynamics, an innovative piece of educational software. It doesn't cover every topic within classical statistical mechanics, and it ignores quantum-mechanical effects completely. But it is still the most accessible modeling software I've seen. What is more, the authors of the program at Stark Design in Morristown, N.J., have made it available to Scientific American readers for free until October 2000.

This kind of simulation is nothing new. Many amateur scientists fondly remember writing such programs back in the days of hobbyist computing [see Computer Recreations, by A. K. Dewdney; Scientific American, March 1988], and several limited versions are available on the World Wide Web (such as a Maxwell's demon game).

But Molecular Dynamics takes this all to a new level. It allows you to conduct an impressive array of virtual experiments to see how different atoms interact under all kinds of conditions. The program consists of numerous modules that demonstrate diffusion, osmotic pressure, the relation between temperature and pressure, the distribution of molecular speeds in a gas and many other topics. And you can use the software to discover things that even the most mathematically gifted physicist would be hard-pressed to wrestle from the basic the

The simulation runs so fast that when I first saw it at a conference I was certain it was a trick. The presenter put about 50 each of four different kinds of electrically neutral atoms inside a three-dimensional volume. The particle positions updated so quickly that I thought it had to be a computer movie, not a real-time simulation. So I decided to challenge the fellow.

In nature even neutral atoms can bond together. The mutual repulsion of the orbital electrons polarizes the atoms, and it turns out there is a range of distances over which these polarized atoms are attracted. So I asked the presenter to add these electrostatic interactions and then slowly decrease the temperature. He did. The heavier atoms began clumping together while the lighter ones kept speeding about, just as they should. He then rapidly brought the temperature to zero. The free atoms settled into small isolated clumps, again just as they should. That made me a believer.

 
Molecular Dynamics
Daniels & Daniels; Source: Molecular Modeling
 
3-D VIEW shows a cool crystal of krypton with a few helium atoms on its surface.
Molecular Dynamics
Daniels & Daniels; Source: Molecular Modeling
 
At higher temperatures, the heliums meander about the surface....
Maxwell-Boltzmann Speed Distribution
Daniels & Daniels; Source: Molecular Modeling
 
...and at still higher temperatures, the whole thing disintegrates.

Geologists see this clumping effect because a volcanic rock that cools slowly possesses larger mineral grains than one that cools quickly. Molecular Dynamics makes it possible to study the underlying principles of this process (called annealing) by varying the number and kind of atoms as well as the rate of cooling. By pausing the simulation at each temperature and rotating the virtual container, one can count the clumps and see how many atoms of which type are in each. That suggests an interesting study. Try repeating the experiment a few times and plotting the average size of the clumps versus the cooling rate. You may discover some fundamental facts about annealing that are quite difficult to derive mathematically.

One delightful demo starts with a cubic crystal of 63 krypton atoms. A few added helium atoms quickly bond to the surface. Tweaking the temperature upward causes the helium atoms to walk randomly on the crystal's face. At a little higher temperature the heliums leave the crystal, and if you raise the temperature still further, the crystal will fly apart. These kinds of effects are observed in real crystals. You can do other experiments here as well. Try lowering the temperature and see whether you can get the crystal to re-form. Then plot the time required for the krypton crystal to form versus the number of hydrogen atoms bouncing about. Does the hydrogen interfere with the crystal formation and, if so, why?

You can also explore gas behavior, such as how a gas adjusts to changes in temperature, volume, or number and types of its atoms. The simulation can approximately reproduce the proportionalities that are combined into the well-known ideal gas law. But only approximately. That is because the ideal gas law itself is just an approximation. It holds only if the gas atoms occupy a negligible fraction of the container's volume and if the atoms' kinetic energies are much larger than the interatomic potential energies that tend to make them clump together. As a result, any real gas departs from the ideal gas law at high densities or low temperatures. Molecular Dynamics includes these effects automatically.

My favorite module, "Maxwell-Boltzmann Speed Distribution," lets you discover how few atoms you need before the physicists' mathematical tricks start working. One of the early triumphs of statistical mechanics in the 19th century was its ability to predict the fraction of atoms moving with a particular range of speeds within a gas at a given temperature. The curve of the fraction versus speed has a sharp rise--meaning there are fewer atoms at lower speeds--and a long tail, indicating that some atoms have speeds that are much higher than the average. I placed 100 atoms of helium and argon into the box and watched the distribution of speeds in real time. After just a few collisions, the two curves took on the expected shape. The heavier atoms peaked at a slower speed, as the theory predicts. You might enjoy removing atoms and observing how the distributions deterior

Unfortunately, the software does have some glaring omissions. For instance, it does not allow treatment of heat flow, work or entropy. You cannot, for example, simulate a piston. Also, the support materials were clearly developed by educators with different views of the target audience; some sections are aimed at beginners, whereas others are perhaps more appropriate for graduate students. The software designer has set up a special Web page for Scientific American readers to submit suggestions for a future version. Despite its limitations, Molecular Dynamics is a wonderful aid for understanding how atoms build up our universe. And for free, how can you possibly go wrong?


Note from Dr H:   The software is now called "Atomic Microscope"   
To learn more, visit the Stark Design website

 


January 28, 2003

Lab Coat Chic: The Arts Embrace Science

By DENNIS OVERBYE

We get tremory   

In this world with no memory

Life makes only partial sense

Knowing only the present tense.

You might have thought that "On the Electrodynamics of Moving Bodies," the 1905 paper in which Albert Einstein proposed the theory of relativity, would be an unlikely subject for song and dance. After all, haven't art and science been at war since the British scientist and novelist C. P. Snow said so in a famous 1959 lecture, "The Two Cultures"?

The four lines quoted above are from the libretto of a musical under development and sung last week in the Kaufmann Theater at the American Museum of Natural History. Called "Einstein's Dreams," produced by Brian Schwartz and written by Joanne Sydney Lessner and Joshua Rosenblum, it is based on the best-selling novel of the same name, a tone poem of ruminations on time and mortality in early 20-century Switzerland by Dr. Alan Lightman. The producers have their own dreams of a Broadway run.


Pamela Traynor
 
A whiff of science in the cultural breezes: Jared Coseglia portrays Einstein in one of two stage versions of "Einstein's Dreams" to play in Manhattan last week.

 

Remarkably, that was not the sole relativity show in town last week. A completely different adaptation of "Einstein's Dreams" is playing until Feb. 1 at the Culture Project Theater in Greenwich Village.

If you squint hard enough, you can imagine a cultural moment occurring.

On Saturday at the Sundance Film Festival in Utah — the mecca of independent filmmakers, epitome of gritty cool — the Alfred P. Sloan Foundation handed out the first of its $20,000 awards, to be given yearly, for films about science or technology. The winner, "Dopamine," directed by Mark Decena and written by him and Timothy Breitbach, is a romantic comedy about a computer engineer who believes that love is merely a chemical reaction and a schoolteacher who has a more romantic bent.

All this follows a string of plays that deal with science or scientists that have lighted up Broadway and beyond, led by "Copenhagen," about two physicists arguing about the atomic bomb, and "Proof," about mathematicians, each winning a Tony Award for best play. A year ago, "A Beautiful Mind," based on Sylvia Nasar's biography of the troubled mathematician John Nash, won the Oscar for best picture.

Meanwhile, a glossy new science magazine, Seed, dedicated to documenting "the global science scene" and promising never to put a dinosaur on the cover, has begun putting out issues with fashion models on the cover and articles on biowarfare inside.

Once, a decade or so ago, there were Science Wars. Scientists were being pilloried by religious conservatives for undermining spiritual values and by postmodernists for their pretensions of objectivity. If the scientists were in the movies, they were the bad guys, chasing E. T.

Are we now on the verge of Science Chic?

Science will, of course, always be hip to scientists. Should they care whether anyone else thinks they are hip? Like the rest of us, they would like a date for the prom or a table at Nobu. And the perception of uncool presents a recruiting problem for the next generation.

But the hippest thing, any scientist will tell you, is the thrill of the intellectual chase and the discovery itself, the moment you alone know something new about the genome or the 11th dimension.

Translating that thrill into the popular culture has always been a hard sell. "If we judge by movies and TV, science is a nonstarter," said Dr. Leon Lederman, a Nobel laureate and a former director of the Fermi National Accelerator Laboratory who has been involved in an effort to develop and sell an "ER"-style television series about scientists.

When this writer sent e-mail requests to astronomers and physicists for memorable portrayals of science, there was nothing but complaints. Movie producers, they said, often spend a lot of money to get small details right — Post-it notes on the computers in "Contact," snippets of authentic jargon or real equations in "Good Will Hunting." Yet they get the big picture wrong, often portraying scientists as madmen or geeks.

"On the other hand," said Dr. Sean Carroll, an astrophysicist at the University of Chicago, "we don't necessarily do so bad compared to doctors or lawyers, I would guess."

The Sundance Award is the latest step in a multipronged effort by the Sloan Foundation to redress the situation by encouraging filmmakers, writers and playwrights to explore scientific and technological themes. The effort includes collaborations with the Ensemble Studio Theater and the Manhattan Theater Club to commission and produce new plays.

The film project began six years ago with awards and workshops in film schools, said Doron Weber, program director of the foundation, and it has been working its way up the food chain since then.

The point of such films, which scientists do not always get, is not to "teach" science — an effort that is invariably fatal to a novel or a movie. "You can treat the drama of scientific thinking," said Dr. Lightman, who was a Sundance judge.

A trim soft-spoken man with a slight Southern accent, Dr. Lightman was an astrophysicist with degrees from Princeton and Caltech before he gradually shifted his energies to writing.

He said the idea for his novel came to him first as a title. He had been thinking, he explained, about the tension between science and art, the rational and the intuitive. Einstein's dreams, he realized, were the bridge.

Dr. Lightman said he did not buy into Snow's "two cultures" separated by a gulf, which he characterized as a "negative perception." Rather, he said, science and art are complementary to each other, "two different ways of being in the world."

Science is about questions that have answers. Art is about questions that do not. It is the lack of answers and the sense of being haunted by them that gives art its power, Dr. Lightman said.

"Einstein's Dreams," published in 1993, is a series of poetic parables in which different notions of time apply. Dr. Lightman said that the film rights had been optioned several times, but that no one had been able to figure out what to do with it.

On the other hand, he estimated that 15 to 20 versions of his novel had been produced on stage. He has not seen all of them. "Sometimes we don't even know about them until afterward," he said.

Now, in terms of audience and money, the lowliest martial arts movie is a bigger deal than the biggest Broadway splash. So if physics seems to be doing better on stage than on screen lately, maybe that is because plays require fewer resources. So they can afford to be experimental. Or maybe it's just that words, and thus ideas, are more important in theater than in the movies, a visual medium.

But a recent immersion in "Einstein's Dreams," the plays, made one wonder whether the Janus faces of science and art might melt together more easily in the shadowy half-light of the stage, where a little greasepaint and our own conspiring imaginations help create the scene, than in the blinding information-rich literalness of celluloid.

Halfway through a Holderness Theater Company production of "Einstein's Dreams," written by Kipp Erante Cheng and directed by Rebecca Holderness, there occurs a moment that can only happen in theater. In a passage from the book, Einstein plays the violin and wonders whether he should leave his wife.

In the play a baby-faced, frizzy-haired Einstein, played with mute intensity by Jared Coseglia, picks up his wife, Mileva (Kate Kohler Amory). She wraps her legs around his waist, and he walks around playing her, sliding the violin bow across her backside.

Einstein looks as if he is thinking, but it is our own thoughts that seem to matter. A man making love to a woman? A scientist coaxing secrets out of nature? Who is this man? What is this alchemy called science?

It reminded me of a scene near the end of "Copenhagen," when the German nuclear physicist Werner Heisenberg thinks out loud as he begins to run a calculation crucial to the atomic bomb through his head. Suddenly a flash and an apocalyptic roar — Einstein's darkest dream — flood the theater. If Heisenberg does the calculation, the Germans build their bomb and history is horribly altered.

At moments like this, theater returns to its mythic roots as a place where the actual and the symbolic, the sacred and profane, pity and awe, meet. Once it was the gods giving us grief. But the lesson of the 20th century has been that they have left the house.

Now it is just us on the stage we have conjured, with the godlike powers and responsibilities that science has given us — and with it the possibility of truly godlike failures. Science is the new mythology.

"There is a hope that science will open like a flower and reveal to us how we came to consciousness," Ms. Holderness said.

Her company had been looking for a break from doing Shakespeare, she said, and she was attracted by the language and "big ideas" in Einstein.

"There is a strong quest for God in Einstein," Ms. Holderness said. "If you're going to spend a lot of time in a dark room, you might as well have something to think about."