Now Reading
The Storied History of the Spooky Physics Behind China’s Teleportation Success

The Storied History of the Spooky Physics Behind China’s Teleportation Success

If quantum mechanics is a true theory of nature, does it mean god plays dice with the universe? Credit: johan;/Flickr, CC BY 2.0

At the heart of the history of quantum entanglement lies a famous debate between two groups of physicists, a clever paradox and an iconoclastic way out of it.

If quantum mechanics is a true theory of nature, does it mean god plays dice with the universe? Credit: johan;/Flickr, CC BY 2.0
If quantum mechanics is a true theory of nature, does it mean god plays dice with the universe? Credit: johanl/Flickr, CC BY 2.0

God does not play dice with the universe. He plays an ineffable game of his own devising, which might be compared, from the perspective of any of the other players, to being involved in an obscure and complex version of poker in a pitch dark room, with blank cards, for infinite stakes, with a dealer who won’t tell you the rules, and who smiles all the time.

– Terry Pratchett

In June 2017, a group of scientists in China announced that they had used the country’s Micius satellite, launched a year earlier, to teleport information from Earth to space in an instant. In other words, they had moved it across over 500 km in literally no time. To achieve this, they had relied on a natural phenomenon called quantum entanglement. The name itself correctly suggests that it belongs in the realm of quantum mechanics, the realm of subatomic particles. The Chinese scientists’ experiment had bested a previous record, when in 2012 their leader himself had lead a team that had teleported information across 97 km.

Very few ideas in science enjoy the popularity that teleportation does: it has been equally awe-inspiring among scientists and laypeople. To the more inspired, what is fascinating is not how an object “leaves” one point in space and “arrives” at another but that it traverses the intervening distance in an instant. The implications of such travel are significant at first sight. The day when we will be able to “beam” a person up and down across space – a la Star Trek – might still be very far away but in the meantime we could use quantum entanglement to, for example, teleport digital security keys between two computers and prevent most forms of eavesdropping by hackers.

In the earlier experiment, Jian-Wei Pan, a physics professor at the University of Science and Technology of China, Hefei, and his colleagues used quantum entanglement to teleport information across Qinghai Lake in the country’s west. Using an ultraviolet laser pointed at a barium crystal, Pan’s team generated pairs of entangled photons. Each photon of a pair was transmitted using a telescope to two parties on either sides of the lake.

The nature of quantum entanglement

Let’s call the parties A and B.

Making a measurement on the photons yields a good description of the state the photons are in. It refers to the values of a few fixed variables. If the variables have a particular combination of values, then the system is said to be in a particular state. States are usually independent of extrinsic properties like mass. So, A’s and B’s goals are to see if a third party interacting with these photons ends up in a state similar to the control group even when separated by 97 km of free-space.

To measure this, the researchers at A let photons generated locally – i.e. at A itself – to interact with the incoming modified photons in a fixed, predictable way. This changed state is then measured and compared with the state of the photons at B. Pan & co. found that the states of the modified photons at A and those of the unmodified photons at B were the same 80% of the time.

What is wonderful is that the particles didn’t have to end up with the same state. Eighty per cent is a value large enough to rule out any coincidence. This long-distance “communication” between minuscule, fragile particles is proof that their pre-travel entanglement was durable and resulted in a predictability of state that let the particles behave similarly in two very different measurement experiments.

The precise nature of this entanglement, which Albert Einstein called “spooky action at a distance”, that baffles most scientists. When two groups of photons are said to be quantum-entangled, it means that the states that the groups are in are related to each other by means of a variable. If the variable changes, then the properties of the photons change, too. However, how the groups themselves are related to each other does not change.

The existence of this variable is not as much disputed as it is hoped into existence. We haven’t found it yet – assuming it exists. And because it remains outside the realm of human control, experiments with teleportation tend to leave this variable alone and instead focus on how much the measurement sites can be separated by, how efficiently large molecules can be entangled, etc. That is, they stick to testing its limits.

To do this, the photons are subjected to a simplified treatment, one conceived with the fewest assumptions as well as the fewest sources of error. Instead of groups of photons, physicists address them two at a time. The state that each half of this pair can exist is in is defined thus. Let’s say the two particles are ‘a’ and ‘b’ and the states are ‘0’ and ‘1’. The four possible combinations of states then are:

{0, 0}
{0, 1}
{1, 0}
{1, 1}

Entanglement is said to have occurred when b is in a particular state when a is in a particular state. That is, if b is 1 every time a is 0, then a and b could be entangled. Since this property is commutative, a will be 0 every time b is 1 as well. Further, the change occurs instantaneously irrespective of the distance between the two particles, giving the impression that they’re “communicating” at a speed faster than that of light. The presence of such an order, together with the four possible outcomes, makes each outcome a particular state of the system. These states are called Bell states, named for the Scottish physicist John Stuart Bell.

To find out what the current Bell state of a particle is, a Bell measurement is made. However, Heisenberg’s uncertainty principle, however, messes this up: the principles dictates that the act of making the measurement will change the state of the system. This is how, for example, the principle prohibits us from knowing an electron’s position and momentum at the same time. However, this alteration does not matter as long as the pre-measurement state is observed and recorded. In Pan’s experiment, with six initial possible states, the Bell measurement was made not by a direct observation per se but by observing how the local and incoming photons interacted.

Earlier, another experiment had been conducted that demonstrated the teleportation of quantum information across 16 km. The principal shortcoming of that experiment was that the photons to be teleported had been specially generated within the lab under careful conditions. In practice, this is a highly ideal condition that could make it difficult to be used as ‘everyday technology’. Pan and his colleagues had eliminated this necessity in their 97-km experiment by generating local photons with random states.

Between the classical and the quantum

The history of quantum entanglement is as entertaining as teleportation itself is. At its heart lies a furious debate between two groups of physicists, a clever paradox and an iconoclastic way out of it.

To ease into it, consider an experiment. Imagine two devices separated by a large distance. These are devices that receive inputs and spit out results. There are two kinds of inputs: classical inputs, which are governed by classical physics, and quantum-mechanical inputs, defined by the rules of quantum mechanics. An input is generated by a common source and is delivered to the devices in an instant.

Now, a pair of inputs is generated at the source such that each input may instruct the device to yield a result ‘x’ or ‘y’. The device called A reads the instructions and yields a result, A*. The device called B reads the instructions and yields a result, B*. If A* and B* are in the same state, then they may be said to be entangled. To have achieved this, A and B – the devices that yielded them – must have communicated in some way to, if nothing else, come to an ‘agreement’. Alternatively, they could have been in possession of some information since before the observation phase.

If it so happened that A and B communicated instantaneously – i.e., exchanged information at faster than the speed of light – then they may be said to be entangled. Let’s remember that, in a quantum mechanical context, the results are found to be identical only after they are observed. Thus, ‘the act of observing the result’ also participates in the measurement process.

This is because Heisenberg’s uncertainty principle kicks in when the particles are observed. When we make the measurement, we are changing the value of some state variable of the particle, so it is the final state that we end up observing. Bell was the first to make this observation and added that the act of observation was somehow tied in with quantum entanglement. In fact, he concluded that the results were entangled in some way because of the act of observing.

Now, the act of observing is a classical phenomenon because the devices A and B that enable the measurement are classical devices. That said, Bell argued that this is where the line between classical mechanics and quantum mechanics blurred. He wrote in 1971:

Theoretical physicists live in a classical world, looking out into a quantum-mechanical world. The latter we describe only subjectively, in terms of procedures and results in our classical domain. … Now nobody knows just where the boundary between the classical and the quantum domain is situated. … More plausible to me is that we will find that there is no boundary. The wave functions would prove to be a provisional or incomplete description of the quantum-mechanical part. It is this possibility, of a homogeneous account of the world, which is for me the chief motivation of the study of the so-called “hidden variable” possibility.

That we often call quantum mechanics ‘quirky’ is because it allows things like entanglement to occur. However, the people who first noticed that this was possible were also hoping to use it to make the point that quantum mechanics could not be a true theory of nature. They were Einstein, Boris Podolsky and Nathan Rosen, commonly referred to as EPR.

The principal target of their ire was the wave function, a mathematical function that adherents of quantum mechanics thought could describe the properties of a quantum mechanical entity, like a particle. For example, by ‘solving’ a wave function, physicists could elicit some of a particle’s states. While a wave function could ‘encode’ a particle, the particle itself could not influence its own wave function. Physicists also believed that each wave function depended on the whole configuration of the universe. According to EPR, these properties, among others, meant that any interpretation of quantum mechanics that included the wave function would allow Heisenberg’s uncertainty principle to be violated.

The EPR paradox

In 1935, the trio published a paper describing a paradox – a phenomenon – that has since been called quantum entanglement. EPR tried to refute quantum mechanics by showing up the flaws of quantum entanglement (objects that are entangled share the same wave function). In their paper, they argued that, since entanglement occurred only on conjugate entities – particles that are somehow, but surely, paired – then the measurement of one of the A*-state variables should have rendered the corresponding state variable in B* indeterminate (because of the uncertainty principle). However, entanglement has already been observed. This means that either the two particles should have communicated or that they should have had the information necessary to generate the same outcome.

EPR preferred the latter explanation, asserting that some “hidden local variable” was responsible for controlling the outcome of the ‘act of observing’. They had made two assumptions to come to this conclusion: locality and realism. The principle of locality states that an object is affected directly only by its immediate surroundings, not by an event that is occurring a large distance away and at the same time. Realism is the ability to assume the existence of objects and parameters even when they have not been observed. Together, they made for a classical way to explain a quantum mechanical effect, and so remove one of the features that made quantum mechanics weird and make it more palatable to Einstein. After all, it was he who had asserted “god does not play dice with the universe” in response to quantum mechanics’ whimsy. (E.g., we can’t know the state of a particle before observing it, so it could be in any state, including in both states at once).

In 1964, Bell proposed a now-famous theorem that refuted the EPR paradox’s preferred explanation. He observed that any local realist theories are incompatible with quantum mechanics. Essentially, this means that since a great number of experiments agree with the predictions of quantum mechanics, and since many of the results are stronger than to be explicable by just local hidden variables, either locality or realism is in conflict with quantum mechanics. Specifically, in his theorem, Bell had posited that locality had been violated and that faster-than-light communication was happening.

Bell’s hypothesis was based on the de Broglie-Bohm theory (initially rejected because of Bohm’s support for communism), which interpreted quantum mechanical effects as being caused by the wave function. This, we now understand, immediately requires that the principle of locality be violated (because a wave function was influenced by the entire universe). We also see that teleportation (of quantum information) is an instance of non-locality because it implies instantaneous communication. If two particles can communicate faster than at the speed of light to replicate quantum mechanical effects, then perhaps complex objects can someday be replicated instantaneously across large distances by simultaneously reproducing the quantum states of the particles associated with the object.

Of course, such a possibility is hinged on Bell’s theorem being true and on the EPR paradox’s implied existence of locality being false. To date, numerous experiments have been conducted that have neither conclusively validated nor invalidated Bell’s theorem. Reactions to the theorem itself have ranged from apathetic to celebratory, with one physicist stating, “Anybody who’s not bothered by Bell’s theorem has to have rocks in his head.” The difficulty lies in what it implied for the real world: it made quantum mechanics and local realism mutually exclusive. Either quantum mechanics was falling short of explaining some physical parameters or superluminal information transfer was happening. (Bell told BBC in 1985 that if the latter is to be disallowed, then we should assume the more-disconcerting notion that there is no such thing as free-will in the universe.)

If looking behind the curtain kills some of the fantasy, that is not the case with teleportation at least. Entanglement continues to elude understanding, and simplifying something so enigmatic to problems in linear algebra – as we have seen – is simply not enough to make sense of whatever is allowing it. With their paper in 2012, Pan and his team were sitting pretty at the forefront of quantum mechanical teleportation – as they are today in 2017. Even if we still have a long way go, the knowledge of Pan’s experiments have given us the best shot at ultimately achieving the teleportation of more sophisticated information systems. But as Bell and EPR have helped elucidate, what they have achieved may be awesome but it brings with it an implication that many of us continue to find difficult to accept.

This article was originally published on the author’s blog in 2012, and has been reproduced here with edits.

Scroll To Top