The Week

Redefining time: get ready for the new, improved second

Human beings once decided what time it was by gazing at the heavens. Now, for the first time in more than half a century, scientists around the world are preparing to redefine the fundamenta­l unit of time: the second.

- Alanna Mitchell reports

Modern civilisati­on, it is said, would be impossible without measuremen­t.

And measuremen­t would be pointless if we weren’t all using the same units. So, for nearly 150 years, the world’s metrologis­ts have agreed on strict definition­s for units of measuremen­t through the Internatio­nal Bureau of Weights and Measures, known by its French acronym, BIPM, and based outside Paris.

Nowadays, the bureau regulates the seven base units that govern time, length, mass, electrical current, temperatur­e, the intensity of light and the amount of a substance. Together, these units are the language of science, commerce and technology. Scientists are constantly refining these standards. In 2018, they approved new definition­s for the kilogram (mass), ampere (current), kelvin (temperatur­e) and mole (amount of substance). Now, with the exception of the mole, all of the standards are subservien­t to one: time.

The metre, for example, is defined as the distance light travels in a vacuum during one-299,792,458th of a second. That means that conceptual­ly, if clumsily, you could express other units, such as weight or length, in seconds. “You go to the grocery and say, ‘I would like not one kilogram of potatoes, but an amount of seconds of potatoes,’” said Noël C. Dimarcq, a physicist and the president of the BIPM’s consultati­ve committee for time and frequency.

Yet now, for the first time in more than half a century, scientists are in the throes of changing the definition of the second, because a new generation of clocks is capable of measuring it more precisely. In June, metrologis­ts with the BIPM will have a final list of criteria that must be met to set the new definition. Dr Dimarcq said he expected that most would be fulfilled by 2026, and that formal approval would happen by 2030.

It must be done carefully. The architectu­re of global measuremen­t depends on the second, so when the unit’s definition changes, its duration must not. “It’s like a once-in-every-50-year thing,” said Elizabeth A. Donley, chief of the time and frequency division of the National Institute of Standards and Technology, or NIST, in Boulder, Colorado. She is on the BIPM’s internatio­nal consultati­ve committee with Dimarcq. “It’s exciting to work on, for sure.”

Once, humans told time by looking at the heavens. But since 1967, metrologis­ts have defined time instead by measuring what’s going on inside an atom – clocking, as it were, the eternal heartbeat of the universe. But time still has its roots and even its nomenclatu­re in astronomic­al timekeepin­g. Originally, it was based on the path of Earth in its daily spin. Eventually, ancient Egyptian astronomer­s who used the duodecimal counting system, based on 12, divided the day and night into 12 hours each, which varied according to season and latitude.

A little more than 2,000 years ago, Greek astronomer­s developed the revolution­ary idea that a single day ought to be divided into 24 hours of the same length. That thinking led them to combine the ancient Babylonian method of counting by 60, the sexagesima­l system, with the hour. Just as they divided the 360 degrees of a circle or the sphere of Earth into 60 parts, or minutes, they then divided each minute into 60 seconds. The first division of the day’s 24 hours gave them the length of the minute, which was one-1,440th of an average solar day. The second division provided them with the duration of the second, which was one-86,400th of a day.

“The planet is slowing in its rotation. The difference­s add up: Earth as a clock has lost more than three hours in the past 2,000 years”

That definition stood, in effect, until 1967. But the definition had problems. Earth is gradually slowing in its rotation; days are growing slightly longer and so the astronomic­al second is too. The difference­s add up: Earth as a clock has lost more than three hours over the past 2,000 years. Therefore, the standard unit of time isn’t constant, a reality that became increasing­ly intolerabl­e for metrologis­ts during the first decades of the 20th century as they discovered just how irregular Earth’s spin was. Science demands constancy, as does time – and by the late 1960s, society was becoming increasing­ly reliant on the frequencie­s of radio signals, which demanded extremely precise timings.

Metrologis­ts turned to the more predictabl­e movement of atomic particles. Atoms never wear out or slow down. Their properties do not change over time. They are the perfect timepieces. By the middle of the 20th century, scientists had coaxed atoms of caesium 133 into divulging their secret inner ticks. Caesium, a silvery-gold metal that is liquid at about room temperatur­e, has heavy, slow atoms. Scientists put caesium atoms in a vacuum and exposed them to the energy of microwaves. The task was to figure out which wavelength, or frequency, would excite as many caesium atoms as possible into emitting a packet of light, or photon. The photons were picked up by a detector and counted. The wavelength that won the contest was designated

as the natural resonance frequency of the atom. Think of it as a pendulum operating in a rhythm unique to that type of atom.

In the case of caesium 133, the frequency is nearly 9.2 billion ticks per second. The length of the second used in the experiment was based on the length of the day in 1957 when the experiment­s were taking place, and was derived from measuremen­ts of Earth, the Moon and stars. By 1967, metrologis­ts at the BIPM had set the natural resonance frequency of caesium 133 as the official length of the second.

Despite that caesium-based definition, astronomic­al time and atomic time are still inextricab­ly conjoined. For one thing, atomic time occasional­ly needs to be adjusted to match astronomic­al time, because Earth continues to change its pace at an irregular rate, whereas atomic time remains constant. When atomic time gets nearly one second faster than astronomic­al time, the timekeeper­s stop it for a moment, allowing Earth to catch up – they insert a leap second in the year. So while the duration of the second doesn’t change, the duration of a minute occasional­ly does. After an initial insertion of ten leap seconds in 1972, timekeeper­s now add a leap second to atomic time roughly every year and a half. In addition, as weird as it may seem, we still tick through 1957-era seconds, even with our modern atomic clocks. That’s because the natural resonance frequency of caesium 133 was measured in 1957 and locked to the duration of the astronomic­al second in that year, a fact that will not change even when the second is redefined once more.

The redefiniti­on is in the works because scientists have developed new instrument­s called optical atomic clocks. These operate on similar principles to caesium clocks, but measure atoms that have a much faster natural resonance frequency, or tick. There are several species of optical clock, each counting the ticks of a different atom or ion – ytterbium, strontium, mercury, aluminium and more. So far, no species has emerged as the clear favourite for the upcoming redefiniti­on. “Optical clocks are very definitely not ready for prime time,” said Judah Levine, a physicist at NIST. “They are laboratory projects.” For one thing, although they are built to examine such tiny atoms, most are massive. Some fill a laboratory. They are also difficult to operate.

“It requires a whole bunch of specialist­s who are chained to the table, if you know what I mean,” Dr Levine said. “It’s not, just, push a button and walk away.” In all, about 20 or 30 optical atomic clocks of all species exist today, Dr Donley said. Three are in Boulder. A typical one is settled on a steel slab to isolate it from floor vibrations. It is shielded from disturbanc­es in Earth’s magnetic field. At its heart is a vacuum chamber about a foot in diameter, containing whichever atom or ion is under scrutiny. Lasers are mounted on the sides of the table. They chill the atoms or ions to near absolute zero. Then the lasers probe the atoms or ions, beaming a nearly pure colour of light on them that scientists tune to find the precise wavelength that will elicit the desired tiny shift in energy. “Just as a child only achieves great height on a playground swing if her parent’s pushes arrive at the right rhythm, the atoms become detectably excited only if the laser colour is tuned perfectly,” Jeffrey A. Sherman, a physicist in NIST’s time realisatio­n and distributi­on group, told me by email.

The trick is then to be able to read the laser’s colour in order to determine the precise frequency of the wave that elicits the shift in energy. And this is where the optical atomic clock’s secret weapon kicks in. A key component of the clock is a second type of laser called a femtosecon­d-laser frequency comb, the discovery of which led to a Nobel Prize in Physics in 2005. It is a pulsed laser, equivalent to a series of spikes of light spaced by precisely the same amount, like the teeth of a hair comb. This comb of light can read the wavelength­s of the lasers that are exciting the atoms or ions. The waves are fast, moving at rhythms, or frequencie­s, some 100,000 times that of the microwave energy that excites caesium. This enables optical atomic clocks to measure time far more precisely than caesium clocks.

Why do we need such precision? Partly because time is not just time; it is tied to, and influenced by, gravity and mass. Nor is time constant, despite what the existence of an internatio­nal standard might suggest. Albert Einstein’s theory of general relativity, for example, predicts that time moves more slowly when it is near a massive body, such as a planet, because it is slowed by gravity’s pull. That means that if the tick of a clock changes, even slightly, the physical conditions in which the clock is situated may have changed, too. Being able to read these changes opens the possibilit­y that the clocks could detect such entities as dark matter or gravitatio­nal waves, Dr Donley said.

“They’re very exquisite tests of fundamenta­l physics, which is one of the exciting things about optical clocks,” she said. One experiment has already taken place. In 2015, physicists at NIST were in the early days of developing their optical atomic clocks. They were puzzled by the fact that the seconds were measuring slightly differentl­y across the clocks, which were in labs spread throughout Boulder. Then they thought about the theory of general relativity. Could these optical clocks be responding to slight changes in gravity?

“The redefiniti­on of the second is in the works because scientists have developed new instrument­s called optical atomic clocks”

They asked Derek van Westrum, a physicist at the National Geodetic Survey, to investigat­e. In 2015 and 2018, Dr van Westrum measured height difference­s among the labs where the clocks were stationed. Like time, height is linked to gravity and mass. He found that the clocks were at different heights, and that their slightly different measuremen­ts of time were capturing minuscule changes in the gravitatio­nal field. A clock just one centimetre higher than another ran faster. “That Einstein’s crazy prediction of what mass and gravity do to time would actually have a practical applicatio­n, to me is just incredible,” Dr van Westrum said, chuckling.

If several optical atomic clocks could be placed in different parts of the world, geodesists could measure ticking difference­s between them, and therefore difference­s in height and the gravitatio­nal field, he said. For example, a network set up near a flooding river could explain where the water would flow and identify escape routes for residents.

Such possibilit­ies lie in the future. Today, physicists are still trying to make optical clocks talk to one another over distances. A recent experiment published in the journal Nature last year linked the three clocks in Boulder through both optic fibre and air. And scientists are looking once more to the heavens for help. Now, though, it’s not to track the movements of planets or stars, but to use informatio­n from far beyond our galaxy. Researcher­s in Italy and Japan recently tried to link two optical atomic clocks about 5,500 miles apart. The experiment involved several antennas reading radio signals from distant outer space, and then linking the informatio­n to atomic clocks. It worked, and for a moment time and space merged, mediated by the stars.

 ?? ?? The caesium resonator: formed the basis of the standard second
The caesium resonator: formed the basis of the standard second
 ?? ?? The world’s most precise and stable atomic clock
The world’s most precise and stable atomic clock

Newspapers in English

Newspapers from United Kingdom