20.07.2013 Views

quantum entanglement - Univerza v Ljubljani

quantum entanglement - Univerza v Ljubljani

quantum entanglement - Univerza v Ljubljani

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

University of Ljubljana<br />

Faculty of Mathematics and Physics<br />

QUANTUM ENTANGLEMENT<br />

Seminar<br />

Jure Koncilija<br />

Mentor: doc. dr. Irena Drevenšek - Olenik<br />

Abstract<br />

Quantum <strong>entanglement</strong> is a physical resource, like energy, associated with the peculiar<br />

nonclassical correlations that are possible between separated <strong>quantum</strong> systems.<br />

Entanglement can be measured, transformed, and purified. A pair of <strong>quantum</strong> systems in<br />

an entangled state can be used as a <strong>quantum</strong> information channel to perform<br />

computational and cryptographic tasks that are impossible in classical systems. In this<br />

seminar I will describe two and multi particle entangled states and the possibilities that<br />

this states bring to <strong>quantum</strong> information technologies.<br />

September 2005


Contents<br />

1. Introduction 3<br />

2. Quantum information 4<br />

2.1. Quantum superposition 4<br />

2.2. Qubit 4<br />

2.3. Single-Qubit transformations 5<br />

2.4. Entanglement 6<br />

2.5. The EPR argument and Bell's inequality 7<br />

3. Quantum <strong>entanglement</strong> 9<br />

3.1. Entangled pair 9<br />

3.2. Producing <strong>quantum</strong> entangled states 9<br />

3.3. Generation of correlated photon pairs in type-II parametric down-conversion 10<br />

3.4. Multi photon <strong>entanglement</strong> 12<br />

4. Quantum dense coding 13<br />

4.1. Quantum dense coding protocol 13<br />

4.2. Experimental dense coding with qubits 14<br />

5. Conclusion 16<br />

6. References 17<br />

2


1. Introduction<br />

One hundred years ago Albert Einstein introduced the concept of the photon.<br />

Although in the early years after 1905 the evidence for <strong>quantum</strong> nature of light was not<br />

compelling, modern experiments, especially those using photon pairs, have beautifully<br />

confirmed its corpuscular character. Research on the <strong>quantum</strong> properties of light<br />

(<strong>quantum</strong> optics) triggered the evolution of the whole field of <strong>quantum</strong> information<br />

processing, which now promises new technology, such as <strong>quantum</strong> cryptography and<br />

even <strong>quantum</strong> computers [1].<br />

Although Einstein was one of the creators of <strong>quantum</strong> theory, he never liked all of<br />

its implications. In particular, he simply could not accept the idea that randomness should<br />

be an inherent principle of nature. He felt that the theory did not and could not explain<br />

why <strong>quantum</strong> effects should appear random to us. He revealed all his doubts in a paper<br />

written together with B. Podolsky and N. Rosen entitled "Can <strong>quantum</strong>-mechanical<br />

description of physics reality be considered complete?" In their paper, EPR argued that<br />

description of nature should obey the following two properties. First, anything that<br />

happens here and now can influence the result of a measurement elsewhere, but only if<br />

enough time has elapsed for a signal to get there without traveling faster than the speed of<br />

light. Second, the result of a measurement is predetermined, particularly if one can<br />

predict it with complete certainty (in other words a result is fixed even if we do not carry<br />

out the measurement itself). Afterwards they examined what impact would these two<br />

conditions have on observations of <strong>quantum</strong> particles that had previously interacted with<br />

one another. They concluded that such particles would exhibit correlations that lead to<br />

contradictions with Heisenberg's uncertainty principle.<br />

In the same year of 1935 Erwin Schrödinger published a response to the EPR<br />

paper, in which he introduced the notion of "<strong>entanglement</strong>" to describe such <strong>quantum</strong><br />

correlations. He said that <strong>entanglement</strong> was the essence of <strong>quantum</strong> mechanics and that it<br />

illustrated the difference between the <strong>quantum</strong> and classical worlds in the most<br />

pronounced way. Schrödinger realized that two entangled particles have to be seen as a<br />

whole, rather than as two separate entities. [2]<br />

After that the studies of <strong>entanglement</strong> paused for thirty years until John Bell's<br />

reconsideration and extension of the EPR argument [1]. Bell looked at <strong>entanglement</strong> in<br />

simpler systems, matching correlations between two-valued dynamical quantities, such as<br />

polarization or spin, of two separated systems in an entangled states. His investigation<br />

generated an ongoing debate on the foundations of <strong>quantum</strong> mechanics. One important<br />

feature of this debate was confirmation that <strong>entanglement</strong> can persist over long distances.<br />

[2] But it was yet in the 1980s that physicists, computer scientists and cryptographers<br />

began to regard the non-local correlations of entangled <strong>quantum</strong> states, which have<br />

become an important part of <strong>quantum</strong> information processing and <strong>quantum</strong> computations.<br />

Physicists nowadays are making great efforts to entangle as many particles as possible<br />

and in 2004 a group of scientist lead by Jian Wei Pan succeeded in entangling five<br />

photons [5]. Although four photons had been entangled before, this was a great<br />

breakthrough because five is the minimum number needed for universal error correction<br />

in <strong>quantum</strong> computation [4].<br />

3


2. Quantum information<br />

2.1. Quantum superposition<br />

The superposition principle plays the most central role in all considerations of<br />

<strong>quantum</strong> information. The essential experiment to describe the superposition is the<br />

double-slit experiment. It is composed of a source, a double-slit assembly, and an<br />

observation screen on which we observe the interference fringes. The interference fringes<br />

may easily be understood on the basis of assuming a wave property of the particles<br />

emerging from the source.<br />

Figure 1: The double slit experiment [3]<br />

Quantum mechanically, the state is the coherent superposition:<br />

1<br />

ψ = ( ψa + ψb<br />

) (1)<br />

2<br />

where ψ a and ψ b describe the <strong>quantum</strong> state with only slit a or slit b open. The<br />

interesting feature in <strong>quantum</strong> double-slit experiment is the observation that the<br />

interference pattern can be collected one by one (one particle interferes with itself), and<br />

that it is impossible of knowing which of the two slits a particle really takes in the<br />

experiment. Any kind of experiment determining this information (we would need to<br />

interact with the particle) would lead to decoherence - loss of interference.<br />

2.2. Qubit<br />

The most fundamental entity in information science is the bit. This is a system<br />

that carries two possible values: "0" and "1", which determine two distinguished states.<br />

No spontaneous transition can occur between the two states.<br />

The <strong>quantum</strong> analog of a bit, the qubit, therefore also has to be a two-state system<br />

where the two states are simply called 0 and 1 . Basically any <strong>quantum</strong> system that<br />

has at least two states can serve as a qubit 1 . The most essential property of <strong>quantum</strong><br />

1 A two-state <strong>quantum</strong> system, that can be used as a qubit is for example spin. We distinguish to opposite<br />

spin states: up ↑ and down ↓ .<br />

4


states, when used to encode bits, is the possibility of coherence and superposition, the<br />

general state being:<br />

Q = α 0 + β 1<br />

(2)<br />

with<br />

probability<br />

2 2<br />

α + β = 1.<br />

This means that if we measure the qubit, we will find it with<br />

2<br />

2<br />

α to carry the value "0" and with probability β to carry the value "1". It<br />

is important to know that eq. (2) describes a coherent superposition rather than incoherent<br />

mixture between "0" and "1". For a coherent superposition there is always a basis in<br />

which the value of the qubit is well defined (as shown in further example), while for an<br />

incoherent mixture it is a mixture whatever way we choose to describe it. Let us consider<br />

a specific state<br />

1<br />

Q′ = ( 0 + 1 ) (3)<br />

2<br />

This clearly means that with 50% probability the qubit will be found to be rather "0" or<br />

"1". Interestingly in basis rotated for 45 in Hilbert space the value of the qubit is well<br />

defined. One of the most basic transformations in <strong>quantum</strong> information science is<br />

Hadamard transformation ( ):<br />

°<br />

ˆH<br />

ˆ 1 ˆ 1<br />

H 0 0 1 , H 1 0<br />

2 2<br />

→ ( + ) → ( − )<br />

1 (4)<br />

Applying this to the qubit Q′ above, results in ˆ HQ′ = 0 , a well defined state, which<br />

is never possible with an incoherent mixture.<br />

2.3. Single-Qubit transformations<br />

Insight of some basic experimental procedures in <strong>quantum</strong> information physics<br />

can be gained by investigating the action of a simple 50/50 beamsplitter. Let us<br />

investigate the case of just two incoming modes and two outgoing modes.<br />

Figure 2: The 50/50 beamsplitter [3]<br />

The effect of such a BS (beamsplitter) on an ingoing - incident state Q = 0 is the<br />

in in<br />

final state Q = 1<br />

2 ( 0 + 1 ), and on an ingoing state Q = 1 the final state is<br />

out<br />

in in<br />

5


(<br />

Q = 1 0 − 1<br />

2 ). For a 50/50 BS, a particle incident either from above or from<br />

out<br />

below has the same probability of 50% of emerging in either output beam, above or<br />

below. The beamsplitter is non-absorbing. A very simple way to describe the action of a<br />

beamsplitter is by the Hadamard transformation eq. (4). The incident state is in general<br />

Q = α 0 + β 1 . Then the action of the beamsplitter results in the final state:<br />

in in in<br />

1<br />

Q = Hˆ Q = ( ( α + β) 0 + ( α − β)<br />

1<br />

out in out out ) (5)<br />

2<br />

where ( α + β ) is now the probability amplitude for finding the particle in the outgoing<br />

upper beam and (α − β ) is the probability amplitude for finding it in the lower beam. It is<br />

interesting and instructive to consider sequences of such beamsplitters because they<br />

realize sequences of Hadamard transformations. The double application of Hadamard<br />

Q = Hˆ Hˆ Q = Q .<br />

transformation is the identity operation ( )<br />

out in in<br />

Another important <strong>quantum</strong> gate besides the Hadamard gate is the phase shifter,<br />

which is introduced into the double beamslitter sequence:<br />

Figure 3: Double beamslitter sequence including a phase shifter in one of the two beams [3]<br />

Its operation is simply to introduce a phase change ϕ to the amplitude of one of the two<br />

beams. The action can be described as: Φ ˆ i<br />

0 = e 0 , Φ ˆ 1 = 1 . The output qubit can be<br />

ϕ<br />

calculated as: Q = HˆΦ ˆ Hˆ Q . For example let us take the input qubit ( α = 1, β = 0 )<br />

out in<br />

Q = 0 . The final state becomes:<br />

in<br />

ˆ ˆ ˆ 1 iϕ iϕ<br />

HΦ H 0 = ( ( e + 1) 0 + ( e − 1) 1 ) (6)<br />

2<br />

This means that the phase shift is able to switch the output qubit between "0" and "1".<br />

For instatnce, if one arranges ϕ = π then the state 0 is transferred to in ( − 1 out ) .<br />

2.4. Entanglement<br />

Let us consider a source which emits a pair of particles such that one particle<br />

emerges to the left and the other to the right . The source is such that the particles are<br />

emitted with opposite momenta.<br />

6


Figure 4: A two particle interferometer verification:<br />

source S emits two qubits in an entangled state [3]<br />

If the particle emerging to the left, which we call particle 1, is found in the upper beam,<br />

then the particle 2 traveling to the right is always found in the lower beam and vice versa.<br />

In qubit language we would say that the two particles carry different bit values. Quantum<br />

1 i<br />

0 1 e 1 0<br />

2 χ<br />

+ .<br />

mechanically this is a two-particle superposition state of the form ( )<br />

The phase χ is determined by the internal properties of the source and we assume for<br />

simplicity 0<br />

χ = . The entangled state is therefore described as:<br />

1<br />

Φ = ( 0 1 + 1 0 ) (7)<br />

2<br />

Interesting property is that neither of the two qubits carries a definite value, but what is<br />

known from <strong>quantum</strong> state is that as soon as one of the two qubits is subject to a<br />

measurement, the result of this measurement being completely random, the other one will<br />

immediately be found to carry the opposite value. This phenomenon represents the<br />

<strong>quantum</strong> non-locality, since the two qubits could be separated by arbitrary distances at<br />

the time of the measurement.<br />

2.5. The EPR argument and Bell's inequality<br />

Immediately after the discovery of modern <strong>quantum</strong> mechanics, it was realized<br />

that it contains novel, counterintuitive features. While Einstein initially tried to argue that<br />

<strong>quantum</strong> mechanics is inconsistent, he later reformulated his argument towards<br />

demonstrating that <strong>quantum</strong> mechanics is incomplete. In the seminal paper, EPR consider<br />

<strong>quantum</strong> systems consisting of two particles such that, while neither position nor<br />

momentum of either particle is well defined, the sum of their positions (that is the centre<br />

of mass) and the difference of their momenta (that is their individual momenta in the<br />

centre of mass system) are both precisely defined. It then follows that a measurement of<br />

either position or momentum performed on, say, particle 1 immediately implies a precise<br />

position or momentum, respectively, for particle 2, without interacting with that particle.<br />

Assuming that the two particles can be separated by arbitrary distances, EPR suggests<br />

that a measurement on particle 1 cannot have any actual influence on particle 2 (locality<br />

condition); thus the property of particle 2 must be independent of the measurement<br />

performed on particle 1. It follows that both position and momentum can simultaneously<br />

be well defined properties of a <strong>quantum</strong> system [3].<br />

7


In 1951 David Bohm introduced spin-entangled systems and in 1964 John Bell<br />

showed that, for such entangled systems, measurements of correlated quantities should<br />

yield different results in the <strong>quantum</strong> mechanical case to those expected if one assumes<br />

that the properties of the system measured prior to, and independent of, the observation.<br />

Let us briefly present the line of reasoning that leads to an inequality equivalent to<br />

the original Bell inequality. Consider a source emitting two qubits in the polarisation<br />

entangled state eq. (7). Such a state can be obtained from a decaying atom of cesium,<br />

which emits two photons in opposite directions. The photons are circularly polarised one<br />

right and the other left. The emitting process therefore conserves both the momentum and<br />

the angular momentum of the cesium atom.<br />

Figure 5: Correlation measurement experiment [3]<br />

One qubit is sent to Alice (to the left in Fig. 5), the other is sent to Bob (to the right). On<br />

both sides the polarisation measurements are made using polarising beamsplitter with two<br />

single-photon detectors in the output ports. Alice will obtain the measurement result "0"<br />

or "1", corresponding to the detection of a qubit by detector D1 or D2 respectively, each<br />

with equal probability. This statement is valid in whatever polarisation basis she decides<br />

to perform the measurement, the actual results being completely random. Yet, if Bob<br />

chooses the same basis as Alice, he will always obtain the same result as Alice.<br />

Following the first step of EPR reasoning, Alice can predict with certainty, what Bob's<br />

result will be. The second step employs the locality hypothesis - the assumption that no<br />

physical influence can instantly go from Alice's apparatus to Bob's, therefore his<br />

measured result should only depend on the properties of his qubit and on the apparatus he<br />

chose. Combining the two steps, J. Bell investigated possible correlations for the case that<br />

Alice and Bob choose detection bases which are at oblique angles. For three arbitrary<br />

angular orientations α, βγ , (used by Alice and Bob in different combinations) the<br />

following inequality must be fulfilled:<br />

N(1 ,1 ) ≤ N(1 ,1 ) + N(1<br />

, 0 )<br />

(8)<br />

α β α γ β γ<br />

N0<br />

2<br />

where N(1 α,1 β) = cos ( α − β )<br />

2<br />

2 is the <strong>quantum</strong>-mechanical prediction for the number<br />

of cases where Alice obtains "1" with her apparatus at orientation α and Bob achieves<br />

"1" with orientation β and N0<br />

is the number of pairs emitted by the source. The<br />

inequality is violated by the <strong>quantum</strong>-mechanical prediction if we choose, for example<br />

2 The first value in the brackets always belongs to Alice's measurement and the second to Bob's<br />

measurement.<br />

8


( α − β) = ( β − γ)<br />

= 30<br />

. The violation implies that at least one of the assumptions<br />

entering Bell's inequality must be in conflict with Quantum mechanics. This is usually<br />

viewed as evidence for non-locality [3].<br />

3. Quantum <strong>entanglement</strong><br />

3.1. Entangled pair<br />

Quantum <strong>entanglement</strong> is a physical resource, like energy, associated with the<br />

peculiar nonclassical correlations that are possible between separated <strong>quantum</strong> systems.<br />

Entanglement can be defined on any two-valued dynamical quantity that a par of particles<br />

posses. The most common is the <strong>entanglement</strong> on the level of spin or polarization. There<br />

are many natural sources of entangled pairs. A very simple example is a pion which has<br />

no spin. When the pion decays it splits into two photons that shoot away in opposite<br />

directions. These two photons have spin. If we measure the spin value of one photon we<br />

instantaneously know the spin value of the other because their spin has to add up into no<br />

spin - that pion has in the first place. Such pairs of photons or other particles are called<br />

entangled pairs.<br />

3.2. Producing <strong>quantum</strong> entangled states<br />

In order to understand both the nature of <strong>entanglement</strong> and ways of producing it,<br />

one has to realize that in states of form eq. (7) we have a superposition between product<br />

states. As we know, superposition means that there is no way to tell which of the two<br />

possibilities forming the superposition actually pertains. This rule must also be applied to<br />

1<br />

Ψ = 0 1 + 1 0<br />

the understanding of <strong>quantum</strong> <strong>entanglement</strong>. In the state ( )<br />

2<br />

12 1 2 1 2<br />

there is no way of telling whether qubit 1 carries the value "0" or "1" and likewise<br />

whether qubit 2 carries value "0" or "1". Yet, if one qubit is measured the other one<br />

immediately assumes a well-defined <strong>quantum</strong> state.<br />

To produce entangled <strong>quantum</strong> states, one has various possibilities. First, one can<br />

create a source which, through its physical construction, is such that the <strong>quantum</strong> states<br />

emerging already have the indistinguishibility feature discussed above. This is realized,<br />

for example, by the decay of a spin-0 particle into two spin- 1 2 particles under<br />

conservation of the internal angular momentum. In this case, thee two spins of the<br />

emerging particles have to be opposite and, if no further mechanism exist which permits<br />

us to distinguish the possibilities right at the source, the emerging <strong>quantum</strong> state is:<br />

where<br />

( )<br />

1<br />

Ψ = ↑ ↓ − ↓ ↑ (9)<br />

12 1 2 1 2<br />

2<br />

↑ means particle 1 with spin up. This state has the remarkable property that it is<br />

1<br />

rotationally invariant: the two spins are anti-parallel along whichever direction we choose<br />

to measure.<br />

9


A second possibility is that a source might actually produce <strong>quantum</strong> states of the<br />

1<br />

form of the individual components in the superposition Ψ = 12 ( 0 1 + 1 0<br />

1 2 1 2 ) .<br />

2<br />

But the states might still be distinguished in some way. This happens, for example, in<br />

type-II parametric down-conversion 3 (see reference [9] for detailed description), where<br />

along a certain chosen direction the two emerging photon states are H V and<br />

1 2<br />

V H . That means that either photon 1 is horizontally polarized and photon 2 is<br />

1 2<br />

vertically polarized, or vice versa. Because of the different speeds of light for the H and<br />

V polarized photon inside down-conversion crystal, the time correlation between the two<br />

photons is different in the two cases. However the entaglement can be produced by<br />

shifting the two photon-wave packets after their production relative to each other such<br />

that they become indistinguishable on the basis of their position in time. The entangled<br />

state for this case can be written as:<br />

1<br />

iϕ<br />

Ψ = 12 ( H V + e V H<br />

1 2 1 2 ) (10)<br />

2<br />

A third means of producing entangled states is to project a non-entangled state<br />

onto an entangled one. Consider a source producing the non-entangled state 0 1 . 1 2<br />

Suppose this state is now sent through filter described by the projection operator<br />

P =Ψ 12 Ψ , where 12 Ψ is in the state of eq. (7):<br />

12<br />

( 0 1 )<br />

P =<br />

1 2<br />

1<br />

= ( 0<br />

2<br />

1 + 1 0 )( 0 1 + 1 0 ) 0 1<br />

1<br />

= ( 0 1 2<br />

1 + 1 2 1 0 2)<br />

1 2 1 2 1 2 1 2 1 2<br />

The state is no longer normalized to unity because the projection procedure implies a loss<br />

of qubits.<br />

3.3. Generation of correlated photon pairs in type-II parametric downconversion<br />

Most commonly used way to produce entangled pairs is parametric down<br />

conversion. The basic optical configuration used for this process is shown in Fig. 6.<br />

3<br />

type-II parametric down-conversion: a nonlinear crystal splits incoming photons into pairs of photons<br />

of lower energy whose combined energy<br />

and momentum is equal to the energy and momentum of the<br />

original photon: k = k + k<br />

<br />

ω = ω + ω . "Parametric" refers to the fact that the state of the<br />

ω ω<br />

in<br />

1<br />

ω and<br />

2 in 1 2<br />

crystal is left unchanged in the process, which is why energy and momentum are conserved. Type-II means<br />

that the phase-matching condition is noncolinear.<br />

=<br />

(11)<br />

10


Figure 6: left: geometrical arrangement for type-II parametric down-conversion [9]<br />

right: the two rings of type-II down-conversion as seen through a narrow-band filter[3]<br />

A pump beam enters a nonlinear optical crystal as an extraordinary beam, and is partially<br />

converted into pairs of photons, obeying energy and momentum conservation. The two<br />

photons in a pair created in this process using the type-II phase-matching condition are<br />

always orthogonally polarised [9]. At certain angles between the pump-beam and the<br />

optical axis of the conversion crystal the phase-matching conditions will be such that the<br />

photons are emitted along cones, which do not have common axis (as illustrated in fig. 6)<br />

One of the cones is ordinarily polarised (idler) and the other extraordinarily (signal). This<br />

cones will in generally intersect along two directions where the emitted light is<br />

unpolarised, because we cannot distinguish whether a certain photon belongs to one or<br />

the other cone. Yet this is not exactly true, because in the crystal the e-polarised and opolarised<br />

photons will propagate with different velocities, so in principle we can<br />

distinguish the two cases by the order of their detection times. It is however possible to<br />

compensate for that "walkoff", by inserting identical crystals of half the thickness rotated<br />

by 90 in each of the two beams. This procedure completely erases any velocity and<br />

arrival time differences and we have a true polarisation-entangled state which can be<br />

described by eq.(10). Further more we can use these compensator crystals to change the<br />

phase<br />

<br />

χ between the two components of the entangled state. If we use an additional halfwave<br />

plate in one of the two beams we can also produce states:<br />

± 1<br />

Φ = ( V V ± H H<br />

12<br />

1 2 1 2)<br />

(12)<br />

2<br />

11


Now we have become acquainted with all four maximally entangled two-qubit<br />

states or Bell states:<br />

± 1<br />

Φ = ( x x ± y y<br />

1 2 1 2)<br />

12 2<br />

(13)<br />

± 1<br />

Ψ = ( x y ± y x<br />

1 2 1 2)<br />

12 2<br />

This is a general expression, where x and y represent qubit values. For spin <strong>entanglement</strong><br />

the values of x and y are written as ↑ and ↓ and for polarisation <strong>entanglement</strong> as H and<br />

V.<br />

3.4. Multi photon <strong>entanglement</strong><br />

The method to entangle more than two photons was suggested by a group of<br />

scientists Greenberger-Horne-Zeilinger and is therefore called GHZ <strong>entanglement</strong>. We<br />

start by two polarization-entangled photon pairs (1-2 and 3-4) which are in the first Bell<br />

+<br />

state Φ :<br />

+<br />

Φ =<br />

12<br />

1<br />

2<br />

+<br />

+<br />

Φ =<br />

34<br />

1<br />

2<br />

+<br />

( H H V V )<br />

1 2 1 2<br />

( H H V V )<br />

3 4 3 4<br />

where H denotes horizontal polarization and V vertical. One photon out of each pair<br />

(photon 2 and photon 3) is then steered to a polarizing beam-splitter (PBS) 4 as shown on<br />

figure 7:<br />

Figure 7: The scheme of the GHZ <strong>entanglement</strong> method for four photons<br />

The path lengths of each photon have to be adjusted such that they arrive simultaneously<br />

on the splitter. Because the PBS transmits H and reflects V polarization, coincidence<br />

detection between the two outputs of PBS, implies that either both photons, 2 and 3, are<br />

H polarized or both V polarized. After PBS the renormalized state corresponding to a<br />

fourfold coincidence is:<br />

4 A polarising beam-splitter is an optical element composed of two prisms, that reflect horizontal<br />

polarization and transmit vertical polarization of the incoming light beam.<br />

(14)<br />

12


1<br />

Φ = 1234 ( H H H H + V V V V<br />

1 2 3 4 1 2 3 4 )<br />

(15)<br />

2<br />

which exhibits four-photon GHZ <strong>entanglement</strong>.<br />

To generate five-photon <strong>entanglement</strong> (five is the highest number of entangled<br />

photons achieved to this day) we further prepare one more photon in the state<br />

1<br />

( H V 1<br />

2 + 1 ) from a single photon source. The scheme of the experiment is shown on<br />

figure 8:<br />

Figure 8: The scheme of the GHZ <strong>entanglement</strong> method for five photons [5]<br />

As shown on figure 2, two PBS are needed: PBS12 to entangle photon 1 from the single<br />

photon source with photon 2 of the entangled pair 23, and PBS34 to entangle photon 3<br />

from the first pair and photon 4 from the second pair of entagled photons. The path<br />

lengths of the photons 1,2,3 and 4 must be carefully adjusted such that they arrive<br />

simultaneously at the designated PBS. After the photons pass through the two PBS, the<br />

state corresponding to a fivefold coincidence is given by:<br />

1<br />

Φ = 12345 ( H H H H H + V V V V V<br />

1 2 3 4 5 1 2 3 4 5 ) (16)<br />

2<br />

This is exactly a five-photon GHZ state.<br />

Multiphoton <strong>entanglement</strong> is needed to realizing <strong>quantum</strong> computations, <strong>quantum</strong><br />

teleportation and building <strong>quantum</strong> computers.<br />

4. Quantum dense coding<br />

4.1. Quantum dense coding protocol<br />

The scheme for <strong>quantum</strong> dense coding utilises <strong>entanglement</strong> between two qubits,<br />

each of which individually has two orthogonal states 0 and 1 . Quantum mechanics<br />

allows one to encode information superpositions of states of two or more entangled<br />

particles. The most convenient basis in which to represent such states is formed by<br />

maximally entangled Bell states - eq.13. Identifying each Bell state with different<br />

13


information we can encode two bits of information by manipulating only one of the two<br />

particles.<br />

This is achieved in the following <strong>quantum</strong> communication scheme. Initially, Alice<br />

+<br />

and Bob each obtain one particle of an entangled pair, let us say, in the state of Ψ<br />

given in eq. 13. Bob then performs one out of four possible unitary transformations on his<br />

particle alone. The four such transformations are:<br />

1. Identity operation (not changing the original two-particle state<br />

+<br />

Ψ )<br />

12<br />

2. State change ( 0 → 1 and 1 → 0 ; changing the two-particle state to<br />

2 2<br />

2 2<br />

3. State-dependent phase shift (differing by π for 0 and 1 ; transforming to<br />

2<br />

2<br />

4. State change and phase shift together (giving the state<br />

−<br />

Φ )<br />

12<br />

+<br />

Φ )<br />

12<br />

12<br />

−<br />

Ψ )<br />

12<br />

Since the four manipulations result in the four orthogonal Bell states, four distinguishable<br />

messages, 2 bits of information, can be sent via Bob's two state particle to Alice, who<br />

finally reads the encoded information (see fig.9) by determining the Bell state of the two<br />

particle system.<br />

4.2. Experimental dense coding with qubits<br />

A <strong>quantum</strong>-optical demonstration [3] of the <strong>quantum</strong> dense coding scheme<br />

described in previous section requires three distinct parts; EPR source, Bob's station of<br />

encoding the messages by a unitary transformation of his particle and Alice's Bell-state<br />

analyser to read the signal sent to Bob (fig.9). The polarisation entangled photons (EPR<br />

source) can be produced by type-II parametric down-conversion.<br />

Figure 9: Experimental setup for <strong>quantum</strong> dense coding [3].<br />

14


One particle of entangled state<br />

+<br />

Ψ is directed to Bob's encoding station, the<br />

other directly to Alice's analyser.<br />

For polarisation encoding, the necessary transformation of Bob's particle is<br />

performed using a half-wave retardation plate for changing the polarisation and a quarterwave<br />

plate to generate the polarisation dependent phase shift. The beam manipulated in<br />

this way then combined with the other beam at Alice's Bell-state analyser. It consisted of<br />

a single 50/50 beamsplitter followed by two-channel PBS polarisators in each of its<br />

outputs and a proper coincidence analysis between four single photon detectors.<br />

−<br />

Since only the state Ψ has an antisymmetric spatial part, only this state will be<br />

registered by coincidence detection between the different outputs of the beamsplitter<br />

(coincidence between H D and V D ′ , or between H D ′ and DV<br />

). For the remaining three<br />

states both photons exit into the same output port of the beamsplitter. The state<br />

+<br />

Ψ can<br />

easily be distinguished from the other two due to the different polarisations of the two<br />

photons, giving behind the two channel polariser a coincidence between detectors H D<br />

and DV or H D ′ and D V ′ . The two states<br />

−<br />

Φ and<br />

+<br />

Φ cannot be distinguished. The<br />

following table gives an overview of the different manipulations and detection<br />

probabilities of Bob's encoder and Alice's receiver.<br />

Bob's setting<br />

λ λ<br />

2 4<br />

0 0 <br />

0 90 <br />

45 0 <br />

45 90 <br />

State sent Alice's registration events<br />

+<br />

Ψ<br />

−<br />

Ψ<br />

+<br />

Φ<br />

−<br />

Φ<br />

coinc. between D H and DV or H D ′ and V D ′<br />

coinc. between H D and V D ′ or H D ′ and<br />

2 photons in either D H , DV , H D ′ or<br />

2 photons in either D H , DV , H D ′ or<br />

To characterize the interference observable at Alice's Bell-state analyser, the path length<br />

difference ∆ of the two beams is varied with the optical trombone. For ∆ c no<br />

interference occurs and one obtains classical statistics for coincidence count rates at the<br />

detectors. For optimal path length tunning (<br />

l <br />

∆ = 0 ), interference enables one to read the<br />

encoded information.<br />

D V ′<br />

D V ′<br />

DV<br />

15


(a) (b)<br />

Figure 10: coincidental rates C HV ( • ) and C HV ′ ( ) as function of the path length difference ∆ [3] for<br />

two different Bell states: (a)<br />

+<br />

Ψ and (b)<br />

5. Conclusion<br />

Entanglement breaks new record in June 2004, when physicists succeeded in<br />

entangling five photons for the first time. This was a key step, which makes it possible to<br />

check computations for errors and teleport <strong>quantum</strong> information within and between<br />

computers. With the GHZ method of multiparticle <strong>entanglement</strong>, the doors are open to<br />

entangle as many particles wanted, which announces the progress of <strong>quantum</strong> computers.<br />

By taking advantage of <strong>quantum</strong> phenomena such as <strong>entanglement</strong>, teleportation (a<br />

process where a certain <strong>quantum</strong> state of a single particle is transferred from Alice to<br />

Bob, without actually delivering the particle itself) and superposition, a <strong>quantum</strong><br />

computer could, in principle, outperform a classical computer in certain computational<br />

tasks (For example, a system of 500 qubits, which is impossible to simulate classically,<br />

represents a <strong>quantum</strong> superposition of as many as 2 500 states. Each state would be<br />

classically equivalent to a single list of 500 1's and 0's. Any <strong>quantum</strong> operation on that<br />

system - a particular pulse of radio waves, for instance, whose action might be to execute<br />

a controlled NOT operation on the 100th and 101st qubits - would simultaneously operate<br />

on all 2 500 states. Hence with one fell swoop, one tick of the computer clock, a <strong>quantum</strong><br />

operation could compute not just on one machine state, as serial computers do, but on<br />

2 500 machine states at once!) [4], [8].<br />

New discoveries regarding <strong>quantum</strong> <strong>entanglement</strong> could revolutionize long<br />

distance communication.<br />

−<br />

Ψ<br />

16


6. References<br />

[1] A. Zeilinger, G. Weihs, T. Jennewein, M. Aspelmeyer, "Happy centenary, photon",<br />

Nature, Vol. 433, pp. 230-238, 2005<br />

[2] H. Weinfurther, "The power of <strong>entanglement</strong>", Physics World, January 2005<br />

[3] D. Boukwmeester, A. Ekert, A. Zeilinger, The Physics of Quantum Information,<br />

Springer-Verlag Berlin Heidelberg, 2000<br />

[4] http://physicsweb.org/articles/news/8/6/18/1<br />

[5] Z. Zhao et al., "Experimental demonstration of five-photon <strong>entanglement</strong> and<br />

open-destination teleportation", Nature, Vol. 430, pp. 54-58, 2004<br />

[6] http://beige.ovpit.indiana.edu/B679/node70.html<br />

[7] C. Kurtsiefer, M. Oberparleiter, H. Weinfurter," Generation of correlated photon<br />

pairs in type-II parametric down-conversion", Journal of modern optics, February 7,<br />

2001<br />

[8] http://www.cs.caltech.edu/~westside/<strong>quantum</strong>-intro.html<br />

[9] R.W. Boyd, Nonlinear optics, Academic press, 1992, pp. 85 -90<br />

17

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!