Apr 23, 2009

Ions keep their cool at crossroads


Apr 15, 2009

Ions keep their cool at crossroads


Down at the crossroadsPhysicists in the US have created a junction through which single ultracold ions can pass without having their temperature raised. The junction is contained within a 2D ion trap and could be useful in building large-scale quantum computers.

Quantum computers, like classical computers, work by processing of bits of information. In classical computers such bits can take on only the values 0 and 1, but in quantum computers they can also take on “superpositions” of both 0 and 1. When many of these quantum bits or “qubits” are combined, a quantum computer can process them simultaneously. In principle, this would enable a quantum computer to work exponentially faster than its classical counterpart for certain operations — however many technical challenges must be overcome before practical quantum computers become a reality.

Scientists are working on many devices to take on the role of qubits, but one of the most promising are trapped ions. Inside a trap, the position and ordering of the ions could be changed by running them through a junction, at which point they can be encouraged to go in one direction or the other. However, it is important that this switching does not result in any heating, because that tends to take the ions out of the required electronic ground state.

Off the heat
Now, Brad Blakestad and colleagues at the National Institute of Standards and Technology (NIST) in Boulder, Colorado have created a junction in an ion trap in which there is practically no heating. Constructed from laser-machined alumina, it contains 46 gold-coated electrodes surrounding an X-shaped junction. When the researchers apply a series of voltages to the electrodes, ions are encouraged through the junction a little at a time.

The NIST group managed to get ions through the junction with a 99.99% success rate, and with seven orders of magnitude reduced heating than previous trapped ion systems .

Christopher Monroe, a physicist at the University of Maryland in the US who has worked extensively with trapped ion systems, told physicsworld.com that although it is not surprising that the researchers have shuttled ions without losing coherence of their internal states, it is surprising that the ions have not lost any coherence in their motion.

Exquisite control
“Moreover, to shuttle ions around a multitude of electrodes and around corners requires exquisite control of the applied electrical potentials so that the ions surf smoothly without getting lost (or agitated),” Monroe adds. “The NIST experiment accomplishes all of this, and thus marks an important milestone in one of the only known realistic architectures for a large-scale quantum computer.”


Bernhard Roth, a quantum physicist at the Heinrich Heine University of Düsseldorf, Germany, also thinks the work is important. “In particular the work might be relevant toward efficient large-scale quantum information processing, the main challenge in the field,” he says. “The authors have significantly increased the reliability of the ion transport through an array, have reduced energy gain and preserved coherence. These are all things which are considered essential for the realization large-scale systems.”

The research will be published in Physical Review Letters and a preprint of the paper is available on arXiv.

About the author
Jon Cartwright is a freelance journalist based in Bristol, UK

Apr 12, 2009

News and Views

News and Views

--------------------------------------------------------------------------------

Nature Physics 5, 248 - 249 (2009)
doi:10.1038/nphys1245


Subject Category: Quantum physics

Quantum physics: Schrödinger's cat is still alive
Jörg Wrachtrup1

Jörg Wrachtrup is at the 3rd Institute of Physics, University of Stuttgart, Pfaffenwaldring 57, D-70550 Stuttgart, Germany.
e-mail: wrachtrup@physik.uni-stuttgart.de



--------------------------------------------------------------------------------

AbstractStrong coupling between a mechanical oscillator and the spin of an electron could enable cooling of the oscillator to its quantum ground state and measurement of the zero-point fluctuations.

Top of page
--------------------------------------------------------------------------------

News and Views

Nature 458, 580-581 (2 April 2009) | doi:10.1038/458580a; Published online 1 April 2009


Solid-state physics: Spin's lifetime extended

Jaroslav Fabian1

Top of pageAbstractElectrons in semiconductors are subject to forces that make their spins flip. According to new evidence, if an ensemble of spins curls into a helix, the collective spin lifetime can be greatly enhanced.

Over the past decade, electron spin — the electron's intrinsic rotation, which is commonly described as 'up' and 'down' and which gives rise to its magnetic moment — has come to the forefront of research in solid-state physics. A whole new field, called spintronics1, 2, 3, 4, has emerged as an umbrella for both applied and fundamental research on spin transport and spin control in metals and semiconductors.

Jaroslav Fabian is at the Institute for Theoretical Physics, University of Regensburg, 93040 Regensburg, Germany.
Email: jaroslav.fabian@physik.uni-regensburg.de

The quantum-optical Josephson interferometer

Nature Physics 5, 281 - 284 (2009)
Published online: 22 March 2009 | doi:10.1038/nphys1223

0811.3762v1
Subject Categories: Quantum physics | Optical physics

The quantum-optical Josephson interferometer
Dario Gerace1,2, Hakan E. Türeci1, Atac Imamoglu1, Vittorio Giovannetti3 & Rosario Fazio3,4


Top of pageThe photon-blockade effect, where nonlinearities at the single-photon level alter the quantum statistics of light emitted from a cavity1, has been observed in cavity quantum electrodynamics experiments with atomic2, 3 and solid-state systems4, 5, 6, 7, 8. Motivated by the success of single-cavity quantum electrodynamics experiments, the focus has recently shifted to the exploration of the rich physics promised by strongly correlated quantum-optical systems in multicavity and extended photonic media9, 10, 11, 12, 13, 14. Even though most cavity quantum electrodynamics structures are inherently dissipative, most of the early work on strongly correlated photonic systems has assumed cavity structures where losses are essentially negligible. Here we investigate a dissipative quantum-optical system that consists of two coherently driven linear optical cavities connected through a central cavity with a single-photon nonlinearity (an optical analogue of the Josephson interferometer). The interplay of tunnelling and interactions is analysed in the steady state of the system, when a dynamical equilibrium between driving and losses is established. Strong photonic correlations can be identified through the suppression of Josephson-like oscillations of the light emitted from the central cavity as the nonlinearity is increased. In the limit of a single nonlinear cavity coupled to two linear waveguides, we show that photon-correlation measurements would provide a unique probe of the crossover to the strongly correlated regime.

Top of page
--------------------------------------------------------------------------------

Institute of Quantum Electronics, ETH Zurich, 8093 Zurich, Switzerland
CNISM and Dipartimento di Fisica 'A. Volta', Università di Pavia, 27100 Pavia, Italy
NEST CNR-INFM and Scuola Normale Superiore, Piazza dei Cavalieri 7, 56126 Pisa, Italy
International School for Advanced Studies (SISSA), via Beirut 2–4, 34014 Trieste, Italy
Correspondence to: Dario Gerace1,2 e-mail: gerace@fisicavolta.unipv.it

Apr 4, 2009

Relaxing the requirements for scalable quantum computing

Relaxing the requirements for scalable quantum computing

Fibonacci scheme for fault-tolerant quantum computation

Panos Aliferis and John Preskill

Phys. Rev. A 79, 012332 (Published January 30, 2009)


Quantum Information


To carry out a long calculation on a quantum computer, some form of error correction is necessary. If the error probability of each logical operation is below what is called the “fault-tolerance” threshold, an error correction procedure will actually remove more errors than it introduces, and the overall failure rate can then be made arbitrarily small. The fault-tolerant threshold is typically quoted as 10-4 or 10-5. This is an extremely stringent tolerance, since it says failure must occur in less than 0.01% of the operations.

A few years ago, Emanuel Knill at NIST in Boulder, Colorado, introduced a different approach to error correction that relied primarily on preparing and verifying a (possibly very large) number of auxiliary qubits, called ancillas, in special states that could be used to diagnose the errors in the computer’s qubits, and replace them if necessary. The most attractive feature of these codes was their large error tolerance, which, based on numerical simulations, Knill estimated to be of the order of 1%.

In a paper appearing in Physical Review A, Panos Aliferis, who is at the IBM Watson Research Center, and John Preskill of the California Institute of Technology, rigorously establish a lower bound for the fault-tolerance threshold for one of Knill’s constructions that has relatively small overhead requirements. Their results indicate that fault-tolerant computation should definitely be possible with this scheme, if the error probability per logical operation does not exceed 0.1%. While lower than Knill’s original numerical estimate, this analytical bound is still at least one order of magnitude larger than was thought possible with other codes and it makes the prospect of scalable quantum computing appear that much more feasible. – Julio Gea-Banacloche

Entanglement dies a sudden death

Entanglement dies a sudden death 2007-05-09 18:09

A strange quantum phenomenon that could be a stumbling block to building quantum computers has been observed for the first time by physicists in Brazil. Known as entanglement sudden death (ESD), it involves the rapid decay of the "entangled" pairs of particles that will be central to the operation of quantum computers. Since the particles decay so quickly, the physicists claim that the decay cannot be reversed using the error-correction schemes that have been proposed to increase the lifetimes of entangled particles (Science 316 579).

Decay channel


In the weird world of quantum mechanics, entanglement means that particles can have a much closer relationship than allowed by classical physics. For instance, two photons can be created experimentally such that if one is polarized in the vertical direction, then the other is always polarized horizontally. By measuring the polarization of one of the pair, we immediately know the state of the other, no matter how far apart they are.

Whereas ordinary computers use bits of information that are either 1 or 0, quantum computers use quantum bits of information, or qubits, that can be in a superposition of both 1 and 0 at the same time. A 1 could represent, say, a horizontally polarized photon, while 0 represent a vertically polarized photon. By combining N such qubits, these could entangled to represent 2N values at the same time, which would, in principle, allow a quantum computer to outperform a classical computer for certain tasks.

However, the qubits in any practical quantum computer have to interact with their local environments, which will cause the quantum state of the qubit to change, or decay. A photon reflecting from a mirror, for example, could suffer a change to its polarization, and successive interactions could even lead to the entanglement disappearing altogether. Crucially, the gradual nature of the decay means that it should be possible to restore entanglement during the computation process using error-correction schemes.

However, it had been predicted that interactions that appear to have a small effect on a single qubit can have a devastating effect on an entangled system of two qubits. This effect -- entanglement sudden death, or ESD -- is so rapid and complete that error-correction schemes will not be able to restore entanglement. Now, Luiz Davidovich and colleagues at the Federal University of Rio de Janeiro have observed ESD for the first time.


In their experiment, the researchers prepared entangled pairs of photons, which were then sent along two identical paths that were separated such that there could be no mutual interaction between the photons. Each path contained optical equipment that could be used to cause a deliberate and gradual decay of the vertical polarization component of both photons. The researchers then detected both photons with the aid of interference filters to determine their degree of entanglement – or concurrence.

The researchers studied pairs of photons that were entangled in two different ways: one type had a certain combination of horizontal and vertical polarizations, while the other type had a different combination of these polarizations. Both initial states were created with the same degree of entanglement and both were subjected to the same gradual decay of vertical polarization. It turned out that the entangled pairs that were more vertically than horizontally polarized underwent ESD, whereas the pairs that where the opposite was true decayed relatively slowly as expected. Davidovich reckons that the vertically-rich entanglement suffered ESD because in this experiment, vertical polarization is a higher energy state and is therefore more sensitive to decay via interactions with the environment than is the lower-energy state of horizontal polarization.

Davidovich told Physics Web that ESD should also occur in other systems that have been proposed for use in quantum computers including trapped ions and atoms in cavities. However, he does not believe that ESD precludes the development of quantum computers. “It leads to an upper limit for the duration of the quantum computation”, he said. “Calculations must be made faster than the time for which ESD occurs”.

Davidovich explained that ESD precludes the use of error correction: “Error-correction techniques rely on entanglement. ESD implies that the quantum computer becomes classical at a finite instant of time, after which quantum error correction is no longer possible”.

Single photons make the trek from space

Single photons make the trek from space

A team of Italian and Austrian scientists has shown it is possible to send single photons from a satellite to a receiving station on Earth. The work, carried out using the Matera Laser Ranging Observatory in southern Italy, paves the way for global quantum cryptography and more rigorous tests of quantum mechanics.

Quantum cryptography exploits the laws of quantum mechanics to create keys for encoding and decoding messages. These keys are strings of 1s and 0s, which are represented by the quantum states of individual subatomic particles, such as the polarization of photons. In principle quantum keys are uncrackable — this is because a measurement of a quantum system in general alters the state of that system. In other words, an eavesdropper situated between a sender and a receiver cannot intercept and identify a key without corrupting it.

Quantum cryptographic systems are already available commercially and have been used, for example, to make bank transfers. Indeed, physicists have shown how to transmit quantum keys over distances of more than 100 km by sending single photons either along optical fibres or via telescopes. Extending this range significantly is difficult, however — in optical fibres photon scattering causes unacceptably high losses, and telescopes are subject to atmospheric turbulence, which can distort a photon beam.

High expectations
Now, Paolo Villoresi of the University of Padova, along with other scientists in Italy and a group lead by Anton Zeilinger at the University of Vienna in Austria, have shown how to overcome these limitations by extending quantum cryptography into space. The Matera Laser Ranging Observatory is usually used to measure variations in the Earth’s gravity and motion, by measuring the time it takes for laser pulses to return to the observatory having been reflected off a passing satellite. Villoresi and colleagues employ the same basic technique but make the beam deliberately weak so less than one photon from each pulse returns to Earth. Transmitting individual photons is crucial for realizing quantum cryptography to prevent an eavesdropper siphoning off excess photons without altering the key.

By bouncing the beam off the Japanese Ajisai satellite, which orbits at an altitude of about 1500 km, the researchers calculate that they receive an average of just 0.4 photons per pulse (after taking into account losses such as the inefficiency of their photon detector). Crucially, by precisely calculating when each pulse is to return to the observatory (accounting for the changing position of the satellite), they are able to show that these detected photons are those transmitted by the telescope and not stray photons from background sources. “Not only have we shown that it is possible to detect single photons from a satellite, we have also demonstrated that we can do this using existing technology,” says Villoresi. “We are very happy about that.”

Ultrafast control of donor-bound electron spins with single detuned optical pulses

Ultrafast control of donor-bound electron spins with single detuned optical pulses 2008-10-31 04:29

Kai-Mei C. Fu1, Susan M. Clark2, Charles Santori1, Colin R. Stanley3, M. C. Holland3 & Yoshihisa Yamamoto2,4

Top of page

The ability to control spins in semiconductors is important in a variety of fields, including spintronics and quantum information processing. Due to the potentially fast dephasing times of spins in the solid state1, 2, 3, spin control operating on the picosecond timescale, or faster, may be necessary. Such speeds—which are not possible to reach with standard electron spin resonance techniques based on microwave sources—can be attained with broadband optical pulses. One promising ultrafast technique uses single broadband pulses detuned from resonance in a three-level system4. This technique is robust against optical-pulse imperfections and does not require a fixed optical reference phase. Here we demonstrate, theoretically and experimentally, the principle of coherent manipulation of spins using this approach. Spin rotations with areas exceeding /4 for a single pulse and /2 for two pulses are achieved for donor-bound electrons. This technique might find applications from basic solid-state electron spin resonance spectroscopy to arbitrary single-qubit rotations4, 5 and bang–bang control6 for quantum computation.

Top of page


--------------------------------------------------------------------------------



Information and Quantum Systems Lab, Hewlett-Packard Laboratories, 1501 Page Mill Road, MS1123, Palo Alto, California 94304, USA
Edward L. Ginzton Laboratory, Stanford University, Stanford, California 94305-4088, USA
Department of Electronics and Electrical Engineering, Oakfield Avenue, University of Glasgow, Glasgow, G12 8LT, UK
National Institute of Informatics, 2-1-2 Hitotsubashi, Chiyoda-ku, Tokyo 101-8430, Japan

Correspondence to: Kai-Mei C. Fu1 e-mail: kai-mei.fu@hp.com.

Subtracting photons from arbitrary light fields: experimental test of coherent state invariance by single-photon annihilation

Subtracting photons from arbitrary light fields: experimental test of coherent state invariance by single-photon annihilation
A Zavatta et al 2008 New J. Phys. 10 123006 (10pp) doi: 10.1088/1367-2630/10/12/123006


5 Author to whom any correspondence should be addressed.
E-mail: bellini@inoa.it

Abstract. The operator annihilating a single quantum of excitation in a bosonic field is one of the cornerstones for the interpretation and prediction of the behavior of the microscopic quantum world. Here we present a systematic experimental study of the effects of single-photon annihilation on some paradigmatic light states. In particular, by demonstrating the invariance of coherent states by this operation, we provide the first direct verification of their definition as eigenstates of the photon annihilation operator.

Received 13 October 2008
Published 9 December 2008

Dec 17, 2008

Experiment verifies Nobel-winning theory

Physicists take important step towards “quantum state engineering”

A property of laser light first predicted in 1963 by the future Nobel laureate Roy Glauber has been verified by physicists in Italy.



One photon out

Marco Bellini and colleagues from the University of Florence have shown that if one photon is removed from a beam of coherent laser light, the light remains in the same coherent state. According to Bellini, the ability to remove photons from light in this way could be used to develop quantum information and quantum metrology systems.

Despite being comprised of many photons, the output of a laser can often be described as a single quantum (or coherent) state. What Glauber did in 1963 — five years after the first laser was built — was to use quantum electrodynamics to show that the addition and subtraction of single photons from coherent light does not affect its coherence. Changing the number of photons only changes the amplitude of the beam.

Laboratory tour-de-force

Verifying this prediction in the lab has proved far from easy because it is very difficult to remove just one photon at a time from a beam. Another big problem has been actually measuring the coherence of the beam before and after the photon has been removed.

About five years ago, however, Bellini and colleagues started developing a way of removing single photons from a laser beam. In their experiments, which they report in New Journal of Physics, a relatively intense laser beam is first passed through a highly-reflective beam splitter, which deflects most of the light into a coherent reference beam.

The rest of the light travels straight through the beam splitter and emerges as relatively weak but still coherent beam. This beam is then sent through a second beam splitter, which is extremely inefficient and only occasionally diverts a photon away from the beam and into a very sensitive detector (see “One photon out”). When the detector “clicks”, the team can be fairly certain that just one photon has been removed from the beam.

Hearing a click

In their most recent study, the team then looked for any changes in the coherence of the beam by recombining it with the reference beam in an interferometer. With each successive “click”, the interferometer is used to measure a different aspect of the phase and amplitude of beam. These data are then analysed using a technique called quantum state tomography, which gives the complete quantum state of the light.

The team found that removing a photon from the light did not change its coherent state — verifying Glauber's 1963 prediction.

In similar experiments Bellini and colleagues have worked out a way to add a single photon to a coherent state and have confirmed another pillar of quantum optics called “noncommutivity” — that removing a photon from a coherent state and then adding a photon is not the same as adding a photon and then removing a photon.

As a result, the team has assembled a “toolbox” for quantum optics that includes the “creation” and “annihilation” operators that add and remove photons, as well as establishing the noncommutivity of these operators. They have also shown that a coherent state is an “eigenstate” of the annihilation operator by showing that the state is not altered by the removal of a photon.

Bellini told physicsworld.com that these tools should allow physicists to engineer quantum-optical states that are optimized for a range of applications such as measuring very small changes in distance or the secure transmission of quantum information.

Physicists squeeze light to quantum limit

Physicists squeeze light to quantum limit
A team of University of Toronto physicists has demonstrated a new technique to squeeze light to the fundamental quantum limit, a finding that has potential applications for high-precision measurement, next-generation atomic clocks, novel quantum computing and our most fundamental understanding of the universe.

Krister Shalm, Rob Adamson and Aephraim Steinberg of U of T's Department of Physics and Centre for Quantum Information and Quantum Control, publish their findings in the January 1 issue of the prestigious international journal Nature.

"Precise measurement lies at the heart of all experimental science: the more accurately we can measure something the more information we can obtain. In the quantum world, where things get ever-smaller, accuracy of measurement becomes more and more elusive," explains PhD graduate student Krister Shalm.

Light is one of the most precise measuring tools in physics and has been used to probe fundamental questions in science ranging from special relativity to questions concerning quantum gravity. But light has its limits in the world of modern quantum technology.

The smallest particle of light is a photon and it is so small that an ordinary light bulb emits billions of photons in a trillionth of a second. "Despite the unimaginably effervescent nature of these tiny particles, modern quantum technologies rely on single photons to store and manipulate information. But uncertainty, also known as quantum noise, gets in the way of the information," explains Professor Aephraim Steinberg.

Squeezing is a way to increase certainty in one quantity such as position or speed but it does so at a cost. "If you squeeze the certainty of one property that is of particular interest, the uncertainty of another complementary property inevitably grows," he says.

In the U of T experiment, the physicists combined three separate photons of light together inside an optical fibre, to create a triphoton. "A strange feature of quantum physics is that when several identical photons are combined, as they are in optical fibres such as those used to carry the internet to our homes, they undergo an "identity crisis" and one can no longer tell what an individual photon is doing," Steinberg says. The authors then squeezed the triphotonic state to glean the quantum information that was encoded in the triphoton´s polarization. (Polarization is a property of light which is at the basis of 3D movies, glare-reducing sunglasses, and a coming wave of advanced technologies such as quantum cryptography.)

In all previous work, it was assumed that one could squeeze indefinitely, simply tolerating the growth of uncertainty in the uninteresting direction. "But the world of polarization, like the Earth, is not flat," says Steinberg.

"A state of polarization can be thought of as a small continent floating on a sphere. When we squeezed our triphoton continent, at first all proceeded as in earlier experiments. But when we squeezed sufficiently hard, the continent lengthened so much that it began to "wrap around" the surface of the sphere," he says.

"To take the metaphor further, all previous experiments were confined to such small areas that the sphere, like your home town, looked as though it was flat. This work needed to map the triphoton on a globe, which we represented on a sphere providing an intuitive and easily applicable visualization. In so doing, we showed for the first time that the spherical nature of polarization creates qualitatively different states and places a limit on how much squeezing is possible," says Steinberg.

"Creating this special combined state allows the limits to squeezing to be properly studied," says Rob Adamson. "For the first time, we have demonstrated a technique for generating any desired triphoton state and shown that the spherical nature of polarization states of light has unavoidable consequences. Simply put: to properly visualize quantum states of light, one should draw them on a sphere."

Images are available at www.physics.utoronto.ca/~lshalm/media/

http://www.artsci.utoronto.ca/main/squeeze-light
Submitted by kb on Tue, 2009-01-06 12:26.

Quantum information: Stopping the rot

Nature 453, 167-168 (8 May 2008) | doi:10.1038/453167a; Published online 7 May 2008

Quantum information: Stopping the rot
Philip C. E. Stamp1

Abstract
Uncontrollable outside influences undermine the whole enterprise of quantum computing. Nailing down the sources of this 'decoherence' in a solid-state system is a step towards solving the problem.
In the quest for a quantum computer, no obstacle is more formidable than decoherence — the 'collapse' of an information-encoding quantum wavefunction when it couples to its surroundings. We pressingly need to understand what causes it, how it works and how to get rid of it. Bertaina et al. (page 203 of this issue)1 have passed a milestone on that road. They report the first observation of Rabi oscillations, a signature of coherent spin dynamics, in a magnetic molecule of a kind envisaged as the basic physical carrier of a 'qubit' of quantum information in a quantum computer. Perhaps more importantly, they have also succeeded in pinpointing the sources of decoherence in their system, and so taken the first step towards eliminating them.

Magnetic molecules come in all shapes and sizes, and have spins with values ranging from 1/2, the smallest that quantum theory allows, to more than 30. Their great advantage for making qubits is that all molecules of a species are the same, and have a structure governed purely by quantum mechanics. The authors focus on the vanadium VIV15 molecule, which, at just over a nanometre in diameter, is small-to-middling in size. It has an interesting spin structure, in which 15 vanadium ions, each with a net electronic spin of 1/2, couple strongly into three groups of five.

Because of the way spins add as vector quantities (direction, as well as magnitude, counts), the whole molecule can have an overall spin of 1/2 or 3/2, depending on how the individual electron spins line up. These low-energy spin states are very widely separated from the many higher-energy states. The spin-1/2 state in particular, which has two energy levels corresponding to molecular spin 'up' and molecular spin 'down', is a natural candidate for a two-state qubit.

But it's here that interactions with the environment — decoherence — become a problem. Decoherence was long thought to be a relatively simple process. A popular view was to model the environment as a 'bath' of oscillators that are not localized, but extend throughout space2. Decoherence was the result of transitions in the bath caused by its interactions with the central quantum system of interest. The results of experiments on simple quantum-optical systems3 and on superconductors4 agreed with this picture.

But there were also good reasons to suppose that the oscillator-bath picture should not work in describing low- temperature decoherence in most solid-state systems, in which decoherence is mostly caused by the influence of entities in the local, rather than the extended, environment5. These might be nuclear spins, found almost everywhere, or else one of the many defects (some charged), dislocations and spin impurities found in anything but a perfect crystal. All these objects hop or flip quantum mechanically between a few different states, so that they act as a reservoir of quantum states known as a spin bath. A spin bath often causes little dissipation of energy, but can cause quite devastating decoherence by its interactions with the central quantum system.

Bertaina et al.1 were able to spot Rabi oscillations between the low-energy qubit states of their vanadium molecule — the first time the phenomenon had been seen in a molecular magnet, and clinching proof that a degree of coherence is present in the system. But the authors were also able to work out what was causing decoherence, as manifested in the decay of the Rabi oscillation. They found that the prime source was the 15 vanadium nuclear spins in each molecule, with a rather smaller contribution from hydrogen nuclei (protons) also present in the structure. The experimental decoherence rate differed by only a few per cent from that expected theoretically for spin-bath decoherence in this system5, 6.

This result indicates that the decoherence mechanism is as follows. Each time the spin state of the qubit flips from up to down, it also flips the field on the vanadium nuclear spins. But because both other internal and externally applied fields are present, this nuclear flip is not through fully 180°. The nuclear spins attempt to realign with the field, but because the field is constantly jumping, they end up precessing in a complicated way that depends on the motion of the qubit. Quantum mechanically, this means that the dynamics of the nuclear spin bath are entangled with the qubit dynamics — decoherence has occurred, even though no energy has dissipated from the qubit into the nuclear spin bath2 (Fig. 1). This is a remarkable finding, because the magnetic moments resulting from the nuclear spin are thousands of times smaller than those associated with the electronic spin of the qubit; and yet, like David overcoming Goliath, they prove the stronger party.

Figure 1: Descent into decoherence.

In Bertaina and colleagues' experiment1, a spin qubit flipping between its up and down quantum states (a Rabi oscillation) also flips the field acting on nearby nuclear spins between two orientations. The nuclear spins try to precess in this qubit field, but each time it suddenly changes they must begin anew. Thus, the path they follow is conditional on the specific trajectory of the qubit — the two are quantum-mechanically entangled, which leads to the decoherence of the qubit.

High resolution image and legend (26K)

As far as other possible sources of decoherence are concerned, Bertaina et al. calculated the contribution of lattice vibrations (phonons) and found it to be more than a hundred times weaker than the nuclear-spin contribution.This is because the phonon frequency is much higher than that of the qubit's Rabi oscillation, so that the phonons smoothly follow the qubit dynamics, rather than destroying it. Dipolar interactions between separate vanadium molecules are potentially more dangerous than any other decoherence source, because they are effective over long ranges7, 8; but the authors were able to suppress these effects simply by spacing the vanadium molecules far apart in a solvent.

What are the implications of these results for future work? Certainly, the prospects for using magnetic molecules as qubits are good. If one can get rid of nuclear spins — perhaps using systems with only zero-spin nuclei, prepared by isotopic purification — then the intrinsic decoherence time is about 100 microseconds for a two-level system with an energy separation of around 10 gigahertz. That should be enough to permit a quantum computer to work, given sufficiently weak dipolar interactions7.

The nature of spin-bath decoherence has now been addressed experimentally in both molecular magnets1 and rare-earth metals9, 10, and Rabi oscillations have been seen in both1, 9. Such systems would thus seem to have a clear edge over a rival system posited as a viable basis for a qubit — electron transitions in the semiconductor structures known as quantum dots. In quantum dots, roughly a million nuclear spins can couple to each qubit (although ingenious methods have been proposed to deal with these11). Similarly, the magnetic-molecule qubits are superior to superconducting qubits, which are so large that they inevitably harbour many defects.

But before we get carried away by these latest achievements, two urgent 'architectural' problems must be solved. The first is that, in a real quantum computer, one might not have the option of keeping the qubits very far apart — so a way must be found to arrange the qubits and their interactions to suppress errors arising from dipolar interactions. The second is that the small size of the qubits means that reading out the quantum state of a large number of them, as well as controlling individual qubits externally, has so far defeated our experimental guile. But there is no fundamental reason why these problems cannot be solved. With advances such as that of Bertaina and colleagues1, there would seem to be good grounds for optimism for the future of spin-based quantum computation.


Top of page
References
Bertaina, S. et al. Nature 453, 203–206 (2008). | Article |
Caldeira, A. O. & Leggett, A. J. Ann. Phys. (NY) 149, 374–456 (1983). | Article |
Myatt, C. J. et al. Nature 403, 269–273 (2000). | Article | PubMed | ISI | ChemPort |
Clarke, J. et al. Science 239, 992–997 (1988). | Article | PubMed | ISI |
Prokof'ev, N. V. & Stamp, P. C. E. Rep. Prog. Phys. 63, 669–726 (2000). | Article | ChemPort |
Stamp, P. C. E. & Tupitsyn, I. S. Phys. Rev. B 69, 014401 (2004). | Article | ChemPort |
Aharonov, D., Kitaev, A. & Preskill, J. Phys. Rev. Lett. 96, 050504 (2006). | Article | PubMed | ChemPort |
Morello, A., Stamp, P. C. E. & Tupitsyn, I. S. Phys. Rev. Lett. 97, 207206 (2006). | Article | PubMed | ChemPort |
Bertaina, S. et al. Nature Nanotech. 2, 39–42 (2007). | Article | ChemPort |
Rønnow, H. M. et al. Science 308, 389–392 (2005). | Article | PubMed | ISI | ChemPort |
Yao, W., Lao, R.-B. & Sham, L. J. Phys. Rev. Lett. 98, 077602 (2007). | Article | PubMed | ChemPort |

Multi-particle entanglement in solid is a first

Multi-particle entanglement in solid is a first

An international team of physicists has entangled three diamond nuclei for the first time. The development promotes solid-state systems to a rank of quantum systems including ions and photons that have achieved entanglement for more than two particles.

Entanglement lies at the heart of fields such as quantum computation and quantum teleportation. At its most basic level, if two particles are entangled a measurement of the state of one reveals something about the state of the other, regardless of the distance separating them.

But entanglement is difficult to achieve. It requires quantum states to be manipulated while preventing them from interacting with their environment, which tends to degrade the quantum system into a classical state. Physicists have had some successes, having entangled up to eight calcium ions and up to five photons. So far, however, solid state systems have proved trickier.

Now, a team led by Jöerg Wrachtrup of the University of Stuttgart, Germany, has demonstrated that two or three diamond nuclei can be entangled (Science 320 1326). “If we compare the quality of entanglement in our experiments with those [of ions and photons], our results compare favourably,” says Wrachtrup. His team includes researchers from the University of Tsukuba, the National Institute of Advanced Industrial Science and Technology, and the Nanotechnology Research Institute, Japan, and Texas A&M University, US.

Method ‘not new’
The researchers’ system is a piece of synthetic diamond containing a large proportion of carbon-13 isotopes. At one point in the lattice they place a nitrogen atom, which leaves a defect containing a single electron.

Because this electron interacts with the neighbouring carbon nuclei, Wrachtrup’s team can shine laser light onto it to put some of the nuclei into a certain quantum state. Then, by applying radio-frequency pulses of a magnetic field, they can drive the spin of the nuclei so that they become entangled with one another.

“The method itself is not new,” says Wrachtrup, who adds that equivalent magnetic pulses are also used in nuclear magnetic resonance (NMR) spectroscopy. “It is the system itself which we discovered to be addressable as a single quantum system almost a decade ago. Meanwhile we can engineer this defect to such a degree that we do have excellent control.”
Diamond nuclei are an attractive option as a quantum system for computation. They can be kept in a coherent state for a long time and are easier to control than other systems, which will become vital for minimizing errors. However, because they currently have to be manipulated using the defect electron as an intermediary, it might be difficult to entangle many of them. Wrachtrup says that his team are working on scaling-up their system now, and believes in the future they should be able to control up to five or six nuclei per electron spin.

Solid-state quantum memory using the 31P nuclear spin

Solid-state quantum memory using the 31P nuclear spin

John J. L. Morton1,2, Alexei M. Tyryshkin3, Richard M. Brown1, Shyam Shankar3, Brendon W. Lovett1, Arzhang Ardavan2, Thomas Schenkel4, Eugene E. Haller4,5, Joel W. Ager4 & S. A. Lyon3


Department of Materials, Oxford University, Oxford OX1 3PH, UK
CAESR, Clarendon Laboratory, Department of Physics, Oxford University, Oxford OX1 3PU, UK
Department of Electrical Engineering, Princeton University, Princeton, New Jersey 08544, USA
Lawrence Berkeley National Laboratory, 1 Cyclotron Road, Berkeley, California 94720, USA
Department of Materials Science and Engineering, University of California, Berkeley, California 94720, USA

Correspondence to: John J. L. Morton1,2 Correspondence and requests for materials should be addressed to J.J.L.M. (Email: john.morton@materials.ox.ac.uk).

The transfer of information between different physical forms—for example processing entities and memory—is a central theme in communication and computation. This is crucial in quantum computation1, where great effort2 must be taken to protect the integrity of a fragile quantum bit (qubit). However, transfer of quantum information is particularly challenging, as the process must remain coherent at all times to preserve the quantum nature of the information3. Here we demonstrate the coherent transfer of a superposition state in an electron-spin 'processing' qubit to a nuclear-spin 'memory' qubit, using a combination of microwave and radio-frequency pulses applied to 31P donors in an isotopically pure 28Si crystal4, 5. The state is left in the nuclear spin on a timescale that is long compared with the electron decoherence time, and is then coherently transferred back to the electron spin, thus demonstrating the 31P nuclear spin as a solid-state quantum memory. The overall store–readout fidelity is about 90 per cent, with the loss attributed to imperfect rotations, and can be improved through the use of composite pulses6. The coherence lifetime of the quantum memory element at 5.5 K exceeds 1 s.

《自然》:混合存储方法解决量子计算核心问题



一个国际科学家小组成功完成了对量子计算机存储装置的小型化——将信息存储在原子核内。这一突破成为实现量子计算机的关键一步。相关论文发表于10月23日的《自然》(Nature)杂志上。

在量子理论的世界里,像原子这样的物体可以同时处于多种状态,即它们理论上可以同时位于两个不同位置或者具有多种看起来互斥的属性。量子计算被视为计算领域的“圣杯”,因为在量子计算中,每个单独的信息位,或者叫‘比特’,可以同时存有不止一个数值,而目前的技术中,每个‘比特’只能存一个或0或1的数值。量子计算将带来空前的计算能力,从而将极大地拓展计算机的能力范围。

问题是,如何在嘈杂的环境中隔离一个量子比特来保护这一精密的量子信息,同时还要能够让这个量子比特与外界互动从而可以操作和测量这个量子比特位?

该研究小组由英国牛津大学、美国普林斯顿大学和美国劳伦斯伯克利国家实验室的科学家和工程师组成,研究人员设计了一个混合系统,该系统选用掺有磷原子的硅晶体,利用原子的原子核和电子实现量子信息存储。原子核和电子都将作为一个可以记录量子信息的小型磁性物质,不过在晶体中,电子比原子核大1百万倍,其磁场也要强上1千倍,这使得电子更适合用来进行操作和测量,但是用来存贮信息不是很好,因为存储的信息会迅速消失。这时就显示出了用原子核作为量子比特的意义——当电子上记录的信息可存储时,这些信息将被转移到原子核上从而得以保存更长的时间。

伯克利小组将富含28Si同位素的硅“培育”成为大型晶体,并保持其不受污染、绝对纯净,令实验取得了成功。

论文主要作者、牛津大学圣约翰学院研究员John Morton说:“电子就像原子核和外界的中间人一样,给了我们一个两全其美的办法——利用电子实现对量子信息的快速处理,同时利用原子核实现对量子信息的长期保存。”

关键在于,在原子核存储的信息有1.75秒的生命周期,超过了最近计算出的基于硅的量子计算的时间阀值,只要时间长于计算出的时间值,纠错技术就能够将数据保存任意长的时间。如果不采用这一技术的话,先前研究人员最长只能将基于硅的量子信息保存几十毫秒。

普林斯顿研究小组领导者Steve Lyon说,“先前没有人真正知道在这个系统中原子核到底能保存量子信息多长时间。有了伯克利国家实验室提供的晶体和非常仔细的测量,我们很高兴地看到存储时间超过了门限值。”

科学家正在研究很多不同的构建量子计算机的方法,不过此次研究的方法有一个很大的优势——它是基于硅技术的,这使得它更易与当代的计算机相兼容。(转载于科学网)

Coherent population trapping of an electron spin in a single negatively charged quantum dot

Coherent population trapping of an electron spin in a single negatively charged quantum dot
Xiaodong Xu1, Bo Sun1, Paul R. Berman1, Duncan G. Steel1, Allan S. Bracker2, Dan Gammon2 & L. J. Sham3


Top of page
Coherent population trapping (CPT) refers to the steady-state trapping of population in a coherent superposition of two ground states that are coupled by coherent optical fields to an intermediate state in a three-level atomic system1. Recently, CPT has been observed in an ensemble of donor-bound spins in GaAs (ref. 2) and in single nitrogen-vacancy centres in diamond3 by using a fluorescence technique. Here, we report the demonstration of CPT of an electron spin in a single quantum dot. The observation demonstrates both the CPT of an electron spin and the successful generation of Raman coherence between the two spin ground states of the electron4, 5, 6. This technique can be used to initialize, at about a gigahertz rate, an electron spin state in an arbitrary superposition by varying the ratio of the Rabi frequencies between the driving and probe fields. The results show the potential importance of charged quantum dots for a solid-state approach to the implementation of electromagnetically induced transparency7, 8, slow light9, quantum information storage10 and quantum repeaters11, 12.
Top of page
--------------------------------------------------------------------------------
The H. M. Randall Laboratory of Physics, The University of Michigan, Ann Arbor, Michigan 48109, USA
The Naval Research Laboratory, Washington DC 20375, USA
Department of Physics, The University of California-San Diego, La Jolla, California 92093, USA

Filling a cavity with photons, and watching them leave

Physics 1, 39 (2008) DOI: 10.1103/Physics.1.39

Filling a cavity with photons, and watching them leave
Alexandre BlaisDépartement de Physique et Regroupement Québécois sur les Matériaux de Pointe, Université de Sherbrooke, Sherbrooke, Québec, Canada, J1K 2R1

Jay M. GambettaInstitute for Quantum Computing and Department of Physics and Astronomy, University of Waterloo, Waterloo, Ontario, Canada, N2L 3G1

Published December 8, 2008

Preparing a harmonic oscillator in a state with a well-defined energy is a tricky business. With the new tools provided by cavity and circuit quantum electrodynamics it is now possible to make these pure quantum states and watch how they evolve in time.
A Viewpoint on:
Process Tomography of Field Damping and Measurement of Fock State Lifetimes by Quantum Nondemolition Photon Counting in a Cavity

M. Brune, J. Bernu, C. Guerlin, S. Deléglise, C. Sayrin, S. Gleyzes, S. Kuhr, I. Dotsenko, J. M. Raimond, and S. Haroche

Phys. Rev. Lett. 101, 240402 (2008) – Published December 08, 2008

Download PDF (free)



Measurement of the Decay of Fock States in a Superconducting Quantum Circuit

H. Wang, M. Hofheinz, M. Ansmann, R. C. Bialczak, E. Lucero, M. Neeley, A. D. O’Connell, D. Sank, J. Wenner, A. N. Cleland, and John M. Martinis

Phys. Rev. Lett. 101, 240401 (2008) – Published December 08, 2008

Download PDF (free)

Illustration: Alan Stonebraker/stonebrakerdesignworks.com
Separate Window | Enlarge

Figure 1: By coupling a nonlinear system, such as an atom, to the electromagnetic field, it is possible to create Fock states (eigenstates of the harmonic oscillator). (Top) Brune et al. send atoms (left) into a cavity (center). The atoms are prepared with pulse P1 to be in a superposition of states |e? and |g? before they enter the cavity. The relative phase between these states, which is converted to probability amplitudes for |e? and |g? with pulse P2 when the atoms exit the cavity, depends on the number of photons in the cavity. (Bottom) In place of a cavity, Wang et al. create an electromagnetic field in a microwave resonator (blue). A superconducting qubit, acting as an artificial atom, couples to the center conductor.

The harmonic oscillator is one of the most fundamental systems in quantum mechanics. Equipped with its solution—one of the first that every physics student learns to calculate exactly—it is possible to describe realistic problems, from phonons in a crystal to the interaction of light with an atom. It is perhaps ironic, then, how challenging it is to actually prepare a pure harmonic oscillator state with a well-defined excitation number n, also known as a Fock state. Now, using different methods, two groups—Michel Brune and colleagues at the Laboratoire Kastler Brossel of the CRNS and Collège de France, both in Paris, and Haohua Wang and colleagues at the University of California in Santa Barbara (UCSB)—have created these nonclassical states of the harmonic oscillator and performed a detailed study of how they decay in time. The experiments, reported in Physical Review Letters [1, 2] demonstrate that the lifetime of a Fock state with excitation number n scales as 1/n, as predicted by theory.

Since it is of relevance to both experiments, consider one of the simplest realizations of the harmonic oscillator: the electromagnetic field. Its excited states are photons and a Fock state corresponds to the creation of n photons with the same energy, ?ω. However, when using a classical source with a well-defined frequency (such as a laser) to generate an electromagnetic field, the result is a coherent state: a superposition of Fock states that is nearly indistinguishable from a classical state. The reason is that the energy spectrum of the harmonic oscillator is linear, such that the energy provided by the source will spread over a wide distribution of Fock states. Instead, to directly prepare a purely nonclassical photon state and observe quantum effects, we need to make a sufficiently strong interaction between the electromagnetic field and an additional, nonlinear, component. This is the heart of the experiments from Brune et al. and Wang et al.

In the work from the CNRS group, the nonlinear components are atoms with excited states that are not evenly spaced in energy. In particular, they use circular Rydberg atoms (atoms in highly excited states and with maximum angular momentum: l=|m|=n-1 ). These atoms have large dipole moments that couple strongly to the microwave photons that are used in the experiments. To enhance this coupling, the photons are confined to a cavity made out of two high-quality mirrors that face each other. The CNRS group has studied the interaction between light and matter in a cavity, also known as cavity quantum electrodynamics (QED) [3, 4, 5], with circular Rydberg atoms over the last 20 years. Thus, they have been able to carefully design the experiment so that the coupling strength between the circular Rydberg atoms and the photons in the cavity overwhelms all the decay rates of the combined system: the single-photon decay rate κ out of the cavity, the atomic decay rate Γ1, and the atomic dephasing rate Γ?. It is worth noting that cavities can now be fabricated with extremely large quality factors such that the single photon lifetime 1/κ can be as large as 0.1 seconds [6]. This is long enough for photons to travel 39,000 km back and forth between the two mirrors separated by 2.7 cm!

The experiment by the CNRS group relies on the conditional preparation of Fock states. They first create a coherent state of the cavity field with a microwave source. Then, to prepare a pure but randomly chosen Fock state, they rely upon the magic of quantum measurements: they perform a measurement of photon number with result n, which projects the classical field to the quantum state |n?. The CNRS team’s measurement is special in the sense that it involves no energy exchange. In these so-called quantum nondemolition measurements, atoms that are nonresonant with the photons in the cavity are sent one by one across the cavity. Before entering the cavity, each atom is prepared in a superposition of two of its internal states, labeled |g? and |e? for ground and excited states (see Fig.1, top ). During the time that the atom is in the cavity, this superposition acquires a relative phase proportional to the number of photons in the cavity. After leaving the cavity, a second pulse converts this phase information to probability amplitudes for |g? and |e?, which are then measured by state-selective ionization of the atom. By repeating this process with sufficiently many atoms (roughly 110 in the experiment), the Fock state, which was randomly selected from the initial field distribution, is prepared with high accuracy.

After preparing the Fock state |n?, every atom that is then sent through the cavity reveals information about the subsequent evolution of the cavity field. In this way, Brune et al. follow the time evolution of Fock states n=0 through n=7, and can track how these states decay, something known as quantum process tomography. As expected from theory, they find that the decay rate of a Fock state with n photons is nκ, which is n times faster than for a Fock state with n=1. This enhanced rate simply reflects the fact that each additional photon has its own probability to decay, speeding up the relaxation. Since in this experiment preparing a Fock state is a random process, completely characterizing the state is a costly enterprise, requiring up to a million single atom measurements.

In parallel, researchers have been developing an on-chip version of cavity QED, also known as circuit QED. In this system, the many Rydberg atoms are replaced by a single superconducting qubit and the cavity is a transmission-line resonator, essentially a one-dimensional superconducting coaxial cable (see Fig. 1, bottom). Gaps in the center conductor of the resonator play the role of the mirrors in cavity QED. Moreover, similarly to the cavity used by the CNRS group, excitations of the resonator are microwave photons. These essentially one-dimensional cavities have a small mode volume, resulting in a large electric field per photon. Superconducting qubits are electrical circuits based on Josephson junctions. With their well-defined energy levels, they behave as artificial atoms, providing the essential nonlinearity. In addition, superconducting qubits have a large effective dipole moment. As a result, this system can easily reach the strong coupling regime of cavity QED [7]. Groups at Yale [8] and Delft (in this last case using a different type of on-chip cavity) [9] first demonstrated strong coupling between a superconducting qubit and a microwave resonator in 2004. Because of the very strong coupling, it was predicted [10] and soon confirmed [11] that in circuit QED, Fock states could be resolved by measuring the qubit absorption spectrum. Earlier this year, the UCSB group showed they could prepare Fock states with up to n=6 photons [12].

In the new experiments from the UCSB group [2], the superconducting qubit, playing the role of the atom, is capacitively coupled to the center conductor of the resonator (Fig. 1, bottom). An advantage of this artificial atom is that the energy difference between its |g? and |e? states can be tuned into and out of resonance with the resonator frequency. Starting with the qubit out of resonance and in its ground state |g?, a classical source is used to pump it to |e?. This energy quantum is then transferred to the resonator by tuning the qubit so it is in resonance with the microwave resonator for an appropriate amount of time. By repeating this process, Fock states with n up to 15 have been created. The microwave resonator’s state can in turn be determined by tuning the qubit into resonance with the resonator and measuring the undriven Rabi oscillations of the qubit between |g? and |e?. Since the frequency of these oscillations depends characteristically on the photon number in the microwave resonator, Wang et al. can extract information about the photon distribution of the Fock state. Similarly to the CNRS group, they find that the n-photon Fock state decays at the enhanced rate nκ.

The experiments now reported by Brune et al. and Wang et al. go beyond the groups’ earlier work in that both are able to create Fock states with large n and reach a level of precision with which to probe the decay of these states. By combining two prototypical systems—harmonic oscillators and two-level systems—cavity QED has established itself in the last 20 years as a unique test-bed for fundamental investigations of quantum mechanics. With the recent developments, such as cavities with high-quality factors and circuit QED, new ways to generate, control and measure non-classical states of light are now possible and more surprises are sure to be on their way.

References
M. Brune, J. Bernu, C. Guerlin, S. Deléglise, C. Sayrin, S. Gleyzes, S. Kuhr, I. Dotsenko, J. M. Raimond, and S. Haroche, Phys. Rev. Lett. 101, 240402 (2008).
H. Wang, M. Hofheinz, M. Ansmann, R. C. Bialczak, E. Lucero, M. Neeley, A. D. O’Connell, D. Sank, J. Wenner, A. N. Cleland, and J. M. Martinis, Phys. Rev. Lett. 101, 240401 (2008).
S. Haroche and J.-M. Raimond, Exploring the quantum: atoms, cavities and photons (Oxford University Press, Oxford, 2006).
H. Mabuchi and A. C. Doherty, Science 298, 1372 (2002).
R. Miller, T. E. Northup, K. M. Birnbaum, A. Boca, A. D. Boozer, and H. J. Kimble, J. Phys. B 38, S551 (2005).
S. Gleyzes, S. Kuhr, C. Guerlin, J. Bernu, S. Deleglise, U. Busk Hoff, M. Brune, J.-M. Raimond, and S. Haroche, Nature 446, 297 (2007).
A. Blais, R. S. Huang, A. Wallraff, S. M. Girvin, and R. J. Schoelkopf, Phys. Rev. A 69, 062320 (2004).
A. Wallraff, D. I. Schuster, A. Blais, L. Frunzio, R. S. Huang, J. Majer, S. Kumar, S. M. Girvin, and R. J. Schoelkopf, Nature 431, 162 (2004).
I. Chiorescu, P. Bertet, K. Semba, Y. Nakamura, C. J. P. M. Harmans, and J. E. Mooij, Nature 431, 159 (2004).
J. Gambetta, A. Blais, D. I. Schuster, A. Wallraff, L. Frunzio, J. Majer, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, Phys. Rev. A 74, 042318 (2006).
D. I. Schuster, A. A. Houck, J. A. Schreier, A. Wallraff, J. M. Gambetta, A. Blais, L. Frunzio, J. Majer, B. Johnson, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, Nature 445, 515 (2007).
M. Hofheinz, E. M. Weig, M. Ansmann, R. C. Bialczak, E. Lucero, M. Neeley, A. D. O'Connell, H. Wang, J. M. Martinis, and A. N. Cleland, Nature 454, 310 (2008).

About the Authors

Alexandre Blais
Alexandre Blais received his B.Sc. in 1997 and Ph.D. in 2003, both at the Université de Sherbrooke, Canada. After a postdoctoral fellowship at Yale University, he returned to Sherbrooke as an associate professor in 2006. He is a member of the Canadian Institute for Advanced Research and was awarded a Sloan Research Fellowship in 2008. His main research interests include superconducting qubits, quantum optics, and quantum computation.



Jay M. Gambetta
Jay M. Gambetta received his B.Sc. (Hons.) (2000) and Ph.D. (2004) in theoretical physics at Griffith University, Australia. From 2004 to 2007 he worked as a Postdoctoral Associate at Yale University and is currently a Postdoctoral Fellow at the Institute for Quantum Computing at the University of Waterloo, Canada. His main research interests include superconducting qubits, quantum optics, measurement theory, and quantum computation.

A solid-state light–matter interface at the single-photon level

Nature 456, 773-777 (11 December 2008) | doi:10.1038/nature07607; Received 27 June 2008; Accepted 30 October 2008


A solid-state light–matter interface at the single-photon level
Hugues de Riedmatten1,2, Mikael Afzelius1,2, Matthias U. Staudt1, Christoph Simon1 & Nicolas Gisin1

Group of Applied Physics, University of Geneva, CH-1211 Geneva 4, Switzerland
These authors contributed equally to this work.
Correspondence to: Hugues de Riedmatten1,2 Correspondence and requests for materials should be addressed to H.d.R. (Email: hugues.deriedmatten@unige.ch).

Top of page
Coherent and reversible mapping of quantum information between light and matter is an important experimental challenge in quantum information science. In particular, it is an essential requirement for the implementation of quantum networks and quantum repeaters1, 2, 3. So far, quantum interfaces between light and atoms have been demonstrated with atomic gases4, 5, 6, 7, 8, 9, and with single trapped atoms in cavities10. Here we demonstrate the coherent and reversible mapping of a light field with less than one photon per pulse onto an ensemble of 107 atoms naturally trapped in a solid. This is achieved by coherently absorbing the light field in a suitably prepared solid-state atomic medium11. The state of the light is mapped onto collective atomic excitations at an optical transition and stored for a pre-determined time of up to 1 s before being released in a well-defined spatio-temporal mode as a result of a collective interference. The coherence of the process is verified by performing an interference experiment with two stored weak pulses with a variable phase relation. Visibilities of more than 95 per cent are obtained, demonstrating the high coherence of the mapping process at the single-photon level. In addition, we show experimentally that our interface makes it possible to store and retrieve light fields in multiple temporal modes. Our results open the way to multimode solid-state quantum memories as a promising alternative to atomic gases.

Quantum computing with an electron spin ensemble

Quantum computing with an electron spin ensemble
J.H. Wesenberg,1 A. Ardavan,2 G.A.D. Briggs,1 J.J.L. Morton,1, 2 R.J. Schoelkopf,3 D.I. Schuster,3 and K. Mlmer4
1Department of Materials, University of Oxford, OX1 3PH, United Kingdom
2Clarendon Laboratory, Department of Physics, University of Oxford, OX1 3PH, United Kingdom
3Department of Applied Physics, Yale University, Connecticut 06520
4Lundbeck Foundation Theoretical Center for Quantum System Research,
Department of Physics and Astronomy, University of Aarhus, 8000 Aarhus C, Denmark
(Dated: March 20, 2009)
We propose to encode a register of quantum bits in di erent collective electron spin wave excitations
in a solid medium. Coupling to spins is enabled by locating them in the vicinity of a
superconducting transmission line cavity, and making use of their strong collective coupling to the
quantized radiation eld. The transformation between di erent spin waves is achieved by applying
gradient magnetic elds across the sample, while a Cooper Pair Box, resonant with the cavity eld,
may be used to carry out one- and two-qubit gate operations.
PACS numbers: 03.67.Lx, 33.90.+h, 85.25.Cp, 42.70.Ln
The construction of a large quantum computer is a
challenge for current research. The overarching problem
is to develop physical systems which can reliably store
thousands of qubits and which allow addressability of in-
dividual bits and pairs of bits in gate operations. Propos-
als in which single trapped ions or atoms encode qubits
in their internal state have successfully demonstrated the
building blocks for few-bit devices, while scaling of these
systems to larger register sizes is believed to require in-
terconnects, e.g., with optical transmission. A novel col-
lective encoding scheme for qubits proposes to use many
identical quantum systems to encode each qubit, either
in the collective population of di erent internal states
[1, 2, 3] or in di erent spatial modes of excitation of the
entire system [4, 5].
In this Letter we propose a hybrid approach to quan-
tum computing making use of an ensemble of 1010..1012
electron spins coupled to a superconducting transmission
line cavity. We will describe how a large number of spa-
tial modes can be addressed in the spin ensemble, and
how a transmon Cooper Pair Box (CPB) [6], integrated
in the cavity can provide one- and two-bit gates for quan-
tum computing in the spin ensemble [7, 8]. Our scheme
enables materials for which large spin coherence times
have been demonstrated in ensemble measurements to be
incorporated into a solid state device. In this way, with-
out requiring single spin measurement or strong coupling
to a cavity, full use can be made of the sophisticated tech-
niques which are no w well established for control of large
numbers of spins.



[1] E. Brion, K. Mlmer, and M. Sa man, \Quantum
computing with collective ensembles of multilevel
systems," Phys. Rev. Lett. 99, 260501 (2007).
[2] M. Sa man and K. Mlmer, \Scaling the neutral-atom
rydberg gate quantum computer by collective encoding
in holmium atoms," Phys. Rev. A 78, 012336 (2008).
[3] K. Tordrup and K. Mlmer, \Quantum computing with
a single molecular ensemble and a Cooper-pair box,"
Phys. Rev. A 77, 020301 (2008).
[4] K. Tordrup, A. Negretti, and K. Mlmer, \Holographic
quantum computing," Phys. Rev. Lett. 101, 040501
(2008).
[5] K. Surmacz, J. Nunn, K. Reim, K. C. Lee, V. O.
Lorenz, B. Sussman, I. A. Walmsley, and D. Jaksch,
\Ecient spatially resolved multimode quantum
memory," Phys. Rev. A 78, 033806 (2008).
[6] J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I.
Schuster, J. Majer, A. Blais, M. H. Devoret, S. M.
Girvin, and R. J. Schoelkopf, \Charge-insensitive qubit
design derived from the Cooper pair box," Phys. Rev. A
76, 042319 (2007).

Realization of Tolliffic Gate in Ion Trap

Realization of Tolliffic Gate in Ion Trap

A new important building block for a future quantum computer has been realised by physicists at the Institute for Experimental Physics in Innsbruck and the Institute for Quantum Optics and Quantum Information (IQOQI): a gate acting on three quantum bits, the so-called Toffoli gate, as has been reported in Physical Review Letters.

Quantum mechanical laws allow quantum computers to process information significantly faster and more efficiently than normal computers. Even the most demanding algorithms can be realised in just a few steps. The basic building blocks for quantum computers are gates acting on one or more quantum bits (qubits). Already single qubit operations and one two-qubit gate allow for basic experiments in the world of quantum physics. Innsbruck scientists around Rainer Blatt impressively demonstrated this during the recent years: In 2005, they were able to deterministically teleport quantum information from one atom to another. Last year, they were the first ones to deterministically generate entanglement between noninteracting quantum bits (entanglement swapping).
Gate acting on three qubits
In principle, every quantum algorithm can be realised by using only single- and two-qubit gates. This approach, however, is unfavourable for more complex tasks and would rapidly reach the limits of current implementations. Hence, physicists worldwide strive for the efficient realisation of multi-qubit gates. The scientists in Innsbruck succeeded in implementing a three-qubit gate acting on three trapped calcium ions representing the qubits. The target qubit of the so-called Toffoli gate will only be switched when both control qubits are set to ?1“, while in all other cases the target qubit will not be changed.






Illustration 1: A calcium ion may be in the state S or D. This is equivalent to a quantum bit being in 0 or 1. For a given input state, the figure shows the output state after the Toffoli gate is applied. For almost all input states, nothing happens, while the final two states SSD and SSS (both control qubits are in state S) are being changed.




Important step towards a quantum computer

This novel gate does not only augment the set of available quantum gates, but also raises the achievable efficiency. Thomas Monz, junior scientist at the experiment, explains:?A Toffoli gate based on the conventional approach would require a sequence of six controlled switch operations. In comparison, our approach is three times faster while operating at a reduced error rate.“ Applications of the Toffoli gate lie within quantum error correction or quantum mechanical prime factorization. Thus, it represents an important component of a future quantum computer.




Literature:

"Realization of the quantum To?oli gate with trapped ions"

T. Monz, K. Kim, W. Hänsel, M. Riebe, A. S. Villar, P. Schindler, M. Chwalla, M. Hennrich, and R. Blatt,

Physical Review Letters 102, 040501 (2009), arXiv:0804.0082.

Deterministic entanglement swapping: First successful implementation of a technique for quantum computers

Deterministic entanglement swapping: First successful implementation of a technique for quantum computers 2008-10-30 20:10

Entanglement Swapping with 4 ions. Image: University of Innsbruck

Click here to enlarge image


(PhysOrg.com) -- Scientists led by Rainer Blatt, Markus Hennrich and Mark Riebe of the Institute for Experimental Physics at Innsbruck University recently succeeded for the first time in realizing a deterministic transfer of entanglement in their lab. They reported this important technique for future quantum computing in the online edition of the acclaimed science journal Nature Physics.

“Transfer of entanglement, also known as entanglement swapping is an important technique for quantum information processing and has been demonstrated in labs before. What we managed to achieve here for the first time was a targeted transfer that is called deterministic entanglement swapping,” explained Mark Riebe and Markus Hennrich of the Institute for Experimental Physics at Innsbruck University.

Entanglement is a specific connection between two individual quantum objects. For their experiment the Innsbruck scientists lined up four ions in an electro-magnetic trap and prepared them with laser beams. In a first step the ions were entangled into two pairs. Then the researchers carried out a “Bell measurement” on one ion of each pair which resulted in an entanglement of the previously unentangled ions. Depending on the result of the measurement the ions were manipulated in such a way as to produce a specific entangled state. “The quantum-mechanical entanglement can be transferred in this way by entangling two particles without a joint history,” stated Riebe and Hennrich.

Linking the building blocks of a quantum computer efficiently

This technique would be applied in future quantum computers. Entanglement is the key feature which allows quantum computers to calculate more efficiently than existing computers. The transfer of entanglement also enables the high-quality entanglement of two particles over distances. Rainer Blatt, who leads the group of researchers, explained, “The entangled particles may be separated from each other and are still linked via what Einstein called 'spukhafte Fernwirkung' (spooky action at a distance). With other methods it is very difficult to separate entangled particles without losing the entanglement.”


Entanglement swapping is of particular significance for the next generation of quantum computers. The individual building blocks of a quantum computer would then be put on small microchips and the particles would be shuttled between processing, storage and transfer elements. Rainer Blatt emphasized, “This only works if the individual ions as carriers of the qubits can be deterministically entangled and separated. We have now succeeded for the first time in proving this experimentally. It will be possible to link the different areas of a quantum computer efficiently”.

Financial support for this research project came from the Austrian Science Fund FWF within its special research area Coherent Quantum Systems, from the European Union and the Institut für Quanteninformation GmbH.

Publication: Deterministic entanglement swapping with an ion-trap quantum computer, M. Riebe,T. Monz, K. Kim, A. S. Villar, P. Schindler, M. Chwalla, M. Hennrich, and R. Blatt, Nature Physics, Advance Online Publication 26/10/2008 http://www.nature.com/nphys/

Single spins controlled by an electric field

Single spins controlled by an electric field 2007-11-09 18:30

Researchers in the Netherlands have shown that it is possible to control the spin of a single electron by using an electric field rather than a magnetic field, as is usually the case. The breakthrough could have potential applications for spintronics and quantum computing (Science DOI: 10.1126/science.1148092 ).

Spintronics is a growing area of research that exploits the spin as well as the charge of electrons. It is has already been used to increase the amount of data that can be stored on hard-disks and could someday form the basis of practical quantum computers that perform calculations by manipulating the spins of single electrons.

A key element of spintronics is the ability to flip the spin of an electron from a spin-up to a spin-down state. In the new work, a team led by Lieven Vandersypen at the Kavli Institute of Nanoscience at Delft University of Technology deposited metallic gold gates onto a gallium arsenide substrate, creating a small region where only a single electron can sit. The researchers were then able to use these so-called “quantum dots” to manipulate the spin of the electron in a controlled manner.

Although previously researchers have been able to flip the spins of electrons confined in these dots by applying a magnetic field, it is not easy to generate a magnetic field locally on a chip that is strong enough to rotate the spin. “To then manipulate an array of single spins is almost impossible,” says Vandersypen.

In their new experiments, the team used two quantum dots each separated by 0.2 µm. If the spins in the dots are both parallel, neither electron can hop from one dot to the other because of the Pauli exclusion principle. However, applying an electric field causes one of the spins to rotate.

Indeed, if the field is applied for long enough the electron’s spin can rotate until it is anti-parallel to the other electron, then it can jump across to the other dot and cause a current flow. Eventually, if the field is applied even longer, the spin goes back to being parallel again. Vandersypen’s PhD students Katja Nowack and Frank Koppens, who carried out the experiment, found that the current varies sinusoidally when plotted against the time over which the electric field is applied. Known as Rabi oscillations, this finding proved they were able to control the rotation of the spin.


The driving mechanism for an electric field to control the spin of an electron lies in the spin-orbit interaction. As the electron orbits around a nucleus it produces a magnetic field that changes its own magnetic moment so that, in the electron’s rest frame, an electric field appears as a magnetic field. The team calculated that the coupling from the gallium arsenide electric field to the single electron’s spin in the quantum dot is strong enough to be able to change the direction of its spin when an electric field is applied.

Having shown that it is possible to control single spins in quantum dots via localized electric fields, the researchers at Delft now plan to produce an array of quantum dots where each electron’s spin state can be manipulated. They plan to use these arrays to form controllably coupled spins, which could pave the way for producing entangled states between the electrons.

Phonons fail to explain high-temperature superconductivity

Apr 23, 2008

Phonons fail to explain high-temperature superconductivity

Two independent teams have hammered what could be the final nails in the coffin of a mechanism long thought able to explain high-temperature superconductivity.

Since compounds of copper oxide or “cuprates” were first discovered to exhibit superconductivity at temperatures well above absolute zero in the mid 1980s, physicists have tried to explain how these materials behave by adopting the approach used to understand their low-temperature counterparts. This would mean that the signature of superconductivity — a flow of charge with zero resistance — results from electrons interacting with vibrations in the cuprate’s crystal lattice.

It’s a politically charged problem, a little like trying to find the tomb of Jesus Tom Timusk, McMaster University
In recent years, despite mounting experimental evidence against it, some physicists have clung on to this interpretation. But now teams from Germany and the US have performed calculations to suggest that lattice vibrations in cuprates can at best account for just a small fraction of the materials’ superconducting behaviour.

“It’s a politically charged problem, a little like trying to find the tomb of Jesus,” says Tom Timusk of McMaster University in Canada, who is not involved in the research. “An atheist could not care less, nor would a Muslim or a Buddhist be excited, but for believers it would be a monumental discovery. So it is with the phonon mechanism of [high-temperature] superconductivity.”

From low to high
Low-temperature superconductors, which make a transition to the superconducting phase close to absolute zero, have been well described by Bardeen–Cooper–Schrieffer (BCS) theory for 50 years. In this theory, the natural repulsion between two electrons is overpowered by a lattice vibration, known as a phonon, which binds the electrons into a “Cooper pair”. It turns out that at low temperatures an electron can avoid the effect of any resistance by being with its partner, so the Cooper pair does not experience resistance at all.

BCS theory itself cannot explain how electrons pair up in high-temperature superconductors, but physicists — notably Alex Müller, who shared the 1987 Nobel Prize with Georg Bednorz for discovering the materials — have suggested that it might still be some process based on phonons. Although many techniques including neutron scattering and far-infrared spectroscopy have not agreed, in 2001 separate teams performing photoemission spectroscopy found evidence to back up the phonon interpretation.

They found that, after shining light at different angles onto cuprates, electrons emitted by photoionization had an energy–momentum relationship containing a prominent “kink” between 50 and 80 meV. Such a kink implied that the electrons were interacting in the cuprates with some sort of boson (an integer-spin particle) — though it was not clear whether that boson could be a phonon.

Now Dirk Manske and colleagues at the Institute for Solid-State Physics in Karlsruhe and the Max–Planck Institute for Solid-State Research in Stuttgart have calculated how the energy–momentum relationship should look from first-principles calculations to see to what extent phonons can be the cause of the kink. They used a “local density approximation” to calculate the number of energy states available to electrons in the high-temperature superconductor YBa2Cu3O7, which they combined with a calculation of how the electrons should change in energy due to a phonon interaction.

It’s embarrassing for people to admit they have worked on something for 20 years if it is not true Dirk Manske, Institute for Solid-State Physics, Karlsruhe
Small kink
Manske’s team found that the theoretical energy–momentum relationship produced by these calculations did contain a kink — but about a three to five times smaller than the 2001 observations (Phys. Rev. Lett. 100 137001). This is bad news for physicists who have been hoping phonons can account for all of the behaviour of high-temperature superconductors. “It is embarrassing for people to admit they have worked on something for 20 years if it is not true,” jokes Manske.

Meanwhile, Steven Louie and colleagues at the Univerisity of California in Berkeley have come to a similar conclusion with the cuprate LaSrCuO4. From their calculations, the phonon contribution is almost an order of magnitude too small for the observed kink (Nature 452 975).

Although phonons are now effectively ruled-out as the underlying cause of high-temperature superconductivity in cuprates, researchers still have the task of finding another process that works. One possibility is that the electrons pair up via interactions between their spins. Such interactions occur over much shorter ranges than phonons, and in fact there has already been significant evidence to support this notion.

“There are many experimental techniques being used to interrogate the high-temperature superconductors,” says Louie. “With increasingly better samples and advances in experimental techniques, data from these measurements are revealing more and more detailed, systematic results. The correct theory of superconductivity for the high-temperature cuprates will have to consistently explain all these data.”

Physicists spooked by faster-than-light information transfer

Physicists spooked by faster-than-light information transfer
Quantum weirdness even stranger than previously thought.

Geoff Brumfiel

Two photons can be connected in a way that seems to defy the very nature of space and time, yet still obeys the laws of quantum mechanics.

Physicists at the University of Geneva achieved the weird result by creating a pair of ‘entangled’ photons, separating them, then sending them down a fibre optic cable to the Swiss villages of Satigny and Jussy, some 18 kilometres apart.


PS:

Letter
Nature 454, 861-864 (14 August 2008) | doi:10.1038/nature07121; Received 2 April 2008; Accepted 30 May 2008


Testing the speed of 'spooky action at a distance'
Daniel Salart1, Augustin Baas1, Cyril Branciard1, Nicolas Gisin1 & Hugo Zbinden1

Group of Applied Physics, University of Geneva, 20 Rue de l'Ecole de Médecine, CH-1211 Geneva 4, Switzerland
Correspondence to: Daniel Salart1 Correspondence and requests for materials should be addressed to D.S. (Email: daniel.salart@physics.unige.ch).

Top of page
Correlations are generally described by one of two mechanisms: either a first event influences a second one by sending information encoded in bosons or other physical carriers, or the correlated events have some common causes in their shared history. Quantum physics predicts an entirely different kind of cause for some correlations, named entanglement. This reveals itself in correlations that violate Bell inequalities (implying that they cannot be described by common causes) between space-like separated events (implying that they cannot be described by classical communication). Many Bell tests have been performed1, and loopholes related to locality2, 3, 4 and detection5, 6 have been closed in several independent experiments. It is still possible that a first event could influence a second, but the speed of this hypothetical influence (Einstein's 'spooky action at a distance') would need to be defined in some universal privileged reference frame and be greater than the speed of light. Here we put stringent experimental bounds on the speed of all such hypothetical influences. We performed a Bell test over more than 24 hours between two villages separated by 18 km and approximately east–west oriented, with the source located precisely in the middle. We continuously observed two-photon interferences well above the Bell inequality threshold. Taking advantage of the Earth's rotation, the configuration of our experiment allowed us to determine, for any hypothetically privileged frame, a lower bound for the speed of the influence. For example, if such a privileged reference frame exists and is such that the Earth's speed in this frame is less than 10-3 times that of the speed of light, then the speed of the influence would have to exceed that of light by at least four orders of magnitude.

Experimental Quantum Computing without Entanglement

Experimental Quantum Computing without Entanglement 2008-11-25 18:44

B. P. Lanyon,* M. Barbieri,? M. P. Almeida, and A. G. White
Department of Physics and Centre for Quantum Computer Technology, University of Queensland, Brisbane 4072, Australia
(Received 15 August 2008; published 13 November 2008)
Deterministic quantum computation with one pure qubit (DQC1) is an efficient model of computation
that uses highly mixed states. Unlike pure-state models, its power is not derived from the generation of a
large amount of entanglement. Instead it has been proposed that other nonclassical correlations are
responsible for the computational speedup, and that these can be captured by the quantum discord. In this
Letter we implement DQC1 in an all-optical architecture, and experimentally observe the generated
correlations. We find no entanglement, but large amounts of quantum discord—except in three cases
where an efficient classical simulation is always possible. Our results show that even fully separable,
highly mixed, states can contain intrinsically quantum mechanical correlations and that these could offer a
valuable resource for quantum information technologies.
DOI: 10.1103/PhysRevLett.101.200501 PACS numbers: 03.67.Lx, 03.67.Ac



While a great deal of work has been done on the conventional pure-state models of quantum computing [1,2], relatively little is known about computing with mixed states. Deterministic quantum computation with one pure
qubit (DQC1) is a model of computation that employs only a single qubit in a pure state, alongside a register of qubits in the fully mixed state [3]. While this model is not universal—it cannot implement any arbitrary algorithm—
it can still efficiently solve important problems that are thought to be classically intractable. One of the original
applications identified was the simulation of quantum systems [3]. Since then exponential speedups have been identified in estimating the average fidelity decay under quantum maps [4], quadratically signed weight enumerators[5], and the Jones Polynomial in knot theory [6]. DQC1 also affords efficient parameter estimation at the quantum metrology limit [7]. That such a useful tool could be built with only a single pure quantum bit is particularly appealing given the current state of experimental quantum computing, where decoherence is a significant obstacle in the path to large-scale implementations. Besides its practical applications, DQC1 is also fascinating from a fundamental perspective. Its power is thought to lie somewhere between universal classical and quantum computing—it is strictly less powerful than a universal quantum computer [3] and no efficient classical simulation has been found or thought likely to exist [8,9]. Furthermore its power is thought not to come from the generation of entanglement, which is at most marginally present in DQC1 [9]. This is surprising, as entanglement is widely believed to lie at the heart of the advantages offered by a quantum computer—a belief supported by the discovery that a universal pure-state quantum computer must generate a large amount of entanglement in order to offer any speedup over a classical computer [10,11]. However, no such proof exists for mixed-state models. Instead it has been proposed that DQC1 generates other types of nonclassical correlations and that these are responsible for the computational advantage [8,12–14].
In this Letter we present a small-scale implementation of DQC1 in a linear-optic architecture [15]. We observe and fully characterize the predicted nonclassical correlations. Our results show that while there is no entanglement, other intrinsically quantum mechanical correlations are generated, except in the cases where an efficient classical simulation is always possible. Furthermore, we demonstrate that a small fraction of a single pure quantum bit is enough to implement DQC1 efficiently [9]. This represents the first implementation of DQC1 outside of a liquid-state NMR architecture, in which the question of nonclassical correlations was not addressed [16]. Unlike liquid-state NMR, there are several known paths to scalable linear-optic quantum computing [2,17,18], and there is active development of the necessary technology [19–21].

......


*Corresponding author.
lanyon@physics.uq.edu.au
?Present address: Laboratoire C. Fabry, Institut d’Optique,
France.
[1] M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information (Cambridge University Press, Cambridge, 2000).
[2] R. Raussendorf and H. J. Briegel, Phys. Rev. Lett. 86, 5188 (2001).
[3] E. Knill and R. Laflamme, Phys. Rev. Lett. 81, 5672 (1998).
[4] D. Poulin, R. Blume-Kohout, R. Laflamme, and H. Ollivier, Phys. Rev. Lett. 92, 177906 (2004).
[5] E. Knill and R. Laflamme, Inf. Proc. Lett. 79, 173 (2001).
[6] P.W. Shor and S. P. Jordan, Quantum Inf. Comput. 8, 681 (2008).
[7] S. Boixo and R. D. Somma, Phys. Rev. A 77, 052320 (2008).
[8] A. Datta and G. Vidal, Phys. Rev. A 75, 042310 (2007).
[9] A. Datta, S. T. Flammia, and C. M. Caves, Phys. Rev. A 72, 042316 (2005).
[10] R. Jozsa and N. Linden, Proc. R. Soc. A 459, 2011 (2003).
[11] G. Vidal, Phys. Rev. Lett. 91, 147902 (2003).
[12] A. Datta, A. Shaji, and C. M. Caves, Phys. Rev. Lett. 100, 050502 (2008).
[13] L. Henderson and V. Vedral, J. Phys. A 34, 6899 (2001).
[14] H. Ollivier and W. H. Zurek, Phys. Rev. Lett. 88, 017901 (2001).
[15] B. Lanyon et al., arXiv:0804.0272v1 [Nature Phys. (to be published)].
[16] C. A. Ryan, J. Emerson, D. Poulin, C. Negrevergne, and R. Laflamme, Phys. Rev. Lett. 95, 250502 (2005).
[17] E. Knill, R. Laflamme, and G. J. Milburn, Nature (London) 409, 46 (2001).
[18] J. L. O’Brien, Science 318, 1567 (2007).
[19] Special Issue on Single photons on demand [New J. Phys. 6, 85 (2004)].
[20] Special Issue on Single-photon: detectors, applications, and measurement [J. Mod. Opt. 51, 1265 (2004)].
[21] A. Politi et al., Science 320, 646 (2008).
[22] 2 ¼ ðN MÞ1 PN1 ðai  biÞ2=2i where: N is the number of sample points, M ¼ 3 is the number of fitting parameters in our functions; ai is the ith data point; bi is the ith theoretical point; and i is the uncertainty in ai [23].
[23] J. R. Taylor, An Introduction to Error Analysis (University Science Books, Sausalito, CA, 1997).
[24] V. Coffman, J. Kundu, and W. K. Wootters, Phys. Rev. A 61, 052306 (2000).
[25] A. G. White, D. F.V. James, W. J. Munro, and P. G. Kwiat, Phys. Rev. A 65, 012301 (2001).
[26] See EPAPS Document No. E-PRLTAO-101-019847 for mathematical details of the proofs cited in the text. For more information on EPAPS, see http://www.aip.org/pubservs/epaps.html.
[27] C. H. Bennett et al., Phys. Rev. A 59, 1070 (1999).
[28] G. J. Pryde, J. L. O’Brien, A. G. White, and S. D. Bartlett, Phys. Rev. Lett. 94, 220406 (2005).
[29] N. K. Langford et al., Phys. Rev. Lett. 95, 210504 (2005).

A Quantum Adiabatic Algorithm for Factorization and Its Experimental Implementation

A Quantum Adiabatic Algorithm for Factorization and Its Experimental Implementation

Phys. Rev. Lett. 101, 220405 (2008)

Xinhua Peng1, Zeyang Liao1, Nanyang Xu1, Gan Qin1, Xianyi Zhou1, Dieter Suter2, and Jiangfeng Du1
Hefei National Laboratory for Physical Sciences at Microscale and Department of Modern Physics,
University of Science and Technology of China, Hefei, Anhui 230026, People’s Republic of China and
Fakult¨at Physik, Technische Universit¨at Dortmund, 44221 Dortmund, Germany
(Dated: August 14, 2008)
We propose an adiabatic quantum algorithm capable of factorizing numbers, using fewer qubits
than Shor’s algorithm. We implement the algorithm in an NMR quantum information processor
and experimentally factorize the number 21. Numerical simulations indicate that the running time
grows only quadratically with the number of qubits.
PACS numbers: 87.23.Cc, 05.50.+q, 03.65.Ud




Using quantum mechanical systems as computational devices may be a possible way to build computers that
are qualitatively more powerful than classical computers [1]. The algorithms that are adapted to the special capabilities of these devices are called quantum algorithms.


One of the best known quantum algorithms is Shor’s algorithm for integer factorization [2]. Since no efficient
factorization algorithm is known for classical computers [3], various cryptographic techniques rely on the difficulty
of finding the prime factors of large numbers [4]. However, in 1994, Peter Shor developed a quantum algorithm
that can factorize large numbers in polynomial time [2]. This discovery was one of the main reasons for the subsequent strong interest in quantum computation. An experimental implementation of Shor’s algorithm was
demonstrated by Vandersypen et al. [5], using nuclear spins as qubits to find the prime factors of 15. More recent experiments by Lu et al. [6] and Lanyon et al. [7] used photons as qubits and found the same factors.
While Shor’s algorithm and its experimental implementation are based on the circuit (or network) model
of quantum computing, different models have been proposed later. Here, we consider the adiabatic quantum
computing model proposed by Farhi et al [8], The basis of this model is the quantum adiabatic theorem: A quantum system remains in its instantaneous eigenstate if the system Hamiltonian varies slowly enough and if
there is a gap between this eigenvalue and the rest of the Hamiltonian’s spectrum [9, 10]. It has been proved to be equivalent to the conventional circuit model [11]. Several adiabatic quantum algorithms have been discussed, such as 3SAT and search of unstructured databases [8, 12, 13]. Compared to the network model, the adiabatic scheme appears to offer lower sentivitiy to some perturbations
and thus improved robustness against errors due to dephasing, environmental noise and some unitary control
errors [14, 15].
In this paper, we propose a factorization algorithm that uses the adiabatic approach to quantum information
processing. We also implement this algorithm experimentally, using nuclear spin qubits to factorize the
number 21.



[1] M. A. Nielsen and I . L. Chuang, Quantum Computation
and Quantum Information, Cambridge University Press,
UK, 2000.
[2] P. Shor, in Proc. 35th Annu. Symp. on the Foundations of
Computer Science, editted by S.Goldwasser (IEEE Computer
Society Press, Los Alamitos, California, 1994), p.
124-134.
[3] D.E. Knuth, The Art of Computer Programming Vol.
2, Seminumerical Algorithms (Addison-Wesley Reading,
Massachusetts, 1998).
[4] N. Koblitz, A Course in Number Theory and Cryptography(
Springer, New York, 1994).
[5] L.M.K. Vandersypen et al, Nature 414, 883 (2001).
[6] C. Lu et al, Phys. Rev. Lett 99, 250504 (2007).
[7] B.P. Lanyon et al, Phys. Rev. Lett 99, 250505 (2007).
[8] E. Farhi et al, Science 292, 472 (2001).
[9] A. Messiah, Quantum Mechanics, North-Holland, Amsterdam
1962.
[10] T. Kato, J. Phys. Soc. Jpn. 5, 435 (1950).
[11] A. Mizel, D.A. Lidar and M. Mitchell, Phys. Rev. Lett
99, 070502 (2007).
[12] L.K. Grover, Phys. Rev. Lett 79, 325 (1997).
[13] J. Roland and N.J. Cerf, Phys. Rev. Lett 90, 067903
(2003).
[14] A.M. Childs and E. Farhi, Phys. Rev. A 65, 012322
(2001).
[15] J. Roland and N.J. Cerf, Phys. Rev. A 71, 032330 (2005).
[16] G. Schaller and R. Schutzhold, quant-ph/0708.1882v1
(2007).
[17] E. Farhi and J. Goldstone, “A Numerical Study of the
Performance of a Quantum Adiabatic Evolution Algorithm
for Satisfiability”, quant-ph/0007071v1.
[18] M. Steffen et al, Phys. Rev. A 65, 042308 (2002).
[19] A. Mitra et al, Journal of Magnetic Resonance 177, 285
(2005).
[20] D.G. Cory, A.E Fahmy and T.R Havel, Proc. Natl. Acad.
Sci. 94, 1634 (1997).
[21] N.A. Gershenfeld and I.L. Chuang, Science 275, 350
(1997).
[22] X. Peng, J. Du and D. Suter, Phys. Rev. A 71, 012307

A new phase in quantum computation

Physics 1, 35 (2008) DOI: 10.1103/Physics.1.35

A new phase in quantum computation
Erik SjöqvistDepartment of Quantum Chemistry, Uppsala University, Box 518, Se-751 20 Uppsala, Sweden

Published November 17, 2008

Large-scale quantum computers are hard to construct because quantum systems easily lose their coherence through interaction with the environment. Researchers have tried to avoid this problem by using geometric phase shifts in the design of quantum gates to perform information processing. Experiments and simulations have shown that these gates may be tolerant to certain types of faults, and may therefore be useful for robust quantum computation.

Separate Window | Enlarge

Figure 1: Quantum computers operate on a register of quantum objects that store information (qubits). The upper panel shows the difference between classical bits and qubits. Each classical bit takes a definite value while quantum bits can be prepared in superposition of several values forming entangled states (depicted as tilted bits). The lower panel shows the basic concept of quantum computation. An input register of qubits, initialized in some state, is manipulated by a set of quantum gates (unitary transformations of the qubit states) that act on one (orange) or two (green) qubits at a time. The result of the computation is read out at the end of all these processing steps. [Adapted from [34].]

Illustration: Alan Stonebraker/stonebrakerdesignworks.com
Separate Window | Enlarge

Figure 2: Parallel transport of a vector on a curved surface, in this case a sphere. The transport takes place along geodesic segments (parts of great circles) forming a loop. The angle between the vector and each segment is constant (no local rotation). The final vector has been rotated compared to the initial one, the rotation angle being the solid angle enclosed by the loop. This “global rotation without local rotation” is the holonomy caused by the curvature of the sphere.

Illustration: Alan Stonebraker/stonebrakerdesignworks.com
Separate Window | Enlarge

Figure 3: (a) Tripod structure that can be used for holonomic quantum computation. Three degenerate internal states k=1,2,3 of an atom, say, are coupled by three laser fields fk to an excited state e. (b) Two of the resulting energy levels form a degenerate pair of dark states that encodes the states of a single qubit. (c) One-qubit holonomies can be obtained by slowly varying the strengths and phases of the laser fields so that the initial and final field configurations coincide. A possible cyclic variation A→B→C→A in the special case of real-valued fk is shown. The resulting holonomy is fully determined by the solid angle π/2 which yields a holonomic gate that takes the logical states |0? and |1? into |1? and -|0?, respectively

Illustration: Alan Stonebraker/stonebrakerdesignworks.com
Separate Window | Enlarge

Figure 4: Loop C (and antipodal image C) on the Bloch sphere consisting of geodesic segments that enclose the solid angle. The dynamical phases vanish along each segment. The resulting phases are purely geometric and equal to -1/2 for C and +1/2 for C. The computational states 0 and 1 (such as spin-up and spin-down of a spin?1/2 particle) pick up geometric phases of opposite sign, which yield a phase gate fully determined by the solid angle.

Peter Shor’s demonstration [1] in the mid 1990s of an efficient algorithm for factorizing prime numbers has triggered an immense interest in various aspects of quantum computation. Researchers have proposed several ways to implement quantum computers, ranging from systems that store information in trapped atoms [2] or ions [3] to computers based on condensed matter systems such as Josephson junctions [4] and quantum dots [5]. Such computers would rely on the phenomena of quantum coherence and quantum entanglement among a set of such “qubits” (Fig. 1). Despite these efforts, quantum computers of any useful size still seem far beyond the scope of present day technology, mainly because of the difficulties in maintaining the necessary coherence of all the qubits. Achievable error probabilities for qubit manipulations are still far above the value of ~10-4 required for efficient fault-tolerant quantum computation [6]. A key challenge for quantum computation research is to achieve this precision.

One approach towards this goal is to use quantum geometric phases (that is, the effects of moving a set of quantum parameters around a curved parameter space) [7, 8, 9] to implement quantum gates that manipulate states of physical qubits. Such gates would be the quantum computing equivalent of the logic gates found on today’s microchips. The idea of using geometric phase is known as holonomic or geometric quantum computation, and has become one of the key approaches to achieving quantum computation that is resilient against errors. In 1999, Zanardi and Rasetti [10] laid the theoretical foundations of holonomic quantum computation by showing that any quantum circuit can be generated by using suitable Hamiltonians that depend on experimentally controllable parameters, such as those related to the manipulation of a bosonic mode in a quantum optical system [11]. At the same time, Jones et al. [12] demonstrated experimentally a quantum gate based on geometric phase that was able to entangle a pair of nuclear spins in a nuclear magnetic resonance (NMR) setup. This experiment provided the first explicit example of geometric quantum computation and helped to boost the interest in this field.

Holonomic quantum computation
A quantum holonomy or geometric phase is the quantum analogue of a well-known rotation effect in differential geometry that arises when a vector is parallel transported around a loop on a curved surface (see Fig. 2). This vector may return rotated although there has been no local rotation along the loop. Such a “global rotation without local rotation” is the holonomy caused by the curvature of the underlying space. In quantum mechanics, states are represented by vectors in a Hilbert space and rotations of such vectors are given by applying unitary matrices or phase factors to them. Just as in the differential geometric case, a quantum state vector can be transported without locally rotating it around a loop in some quantum parameter space, and the resulting transformation has the same effect on the state vector as applying a unitary matrix [7] or phase factor [8, 9] that depends only on the global geometric properties of the loop.

In the original form of holonomic quantum computation [10], the states of the quantum bits (the qubits) are encoded in a degenerate energy eigenspace of a suitable parameter-dependent Hamiltonian. When the parameters change adiabatically around a loop, a state that starts in such an energy eigenspace of the initial Hamiltonian will end at another state in the same eigenspace. The final and initial states of the qubits will be related by a unitary rotation that depends only on the properties of the loop in parameter space, but is independent of the energy of the state and the time it takes to traverse the loop. This rotation is the desired holonomy transformation that constitutes the holonomic quantum gate—the basic building block of holonomic quantum computation. This is very different from the more conventional kind of quantum computing where qubits in a register of some kind are acted upon by various logic operations to dynamically evolve the system in time toward a result. Geometric quantum computation, which is discussed in the next section, is holonomic quantum computation, or some nonadiabatic extension thereof, restricted to nondegenerate (one-dimensional) energy eigenspaces of Hamiltonians.

In practice, to perform quantum computation it is sufficient to implement certain elementary one- and two-qubit operations, forming universal sets. These operations are analogous to the fundamental building block OR, AND, NOT operations of conventional microelectronics. Such a set of operations on qubits can be used to simulate any quantum computation with arbitrary precision. A first goal for holonomic quantum computation is to find physical implementations of universal sets of gates that are all-geometric, i.e., based entirely upon quantum holonomies.

Basically what we are looking for are physical qubits (two-level quantum systems) whose evolution can be controlled by means of parameters in a curved space. Unanyan et al. [13] discovered that a four-level atom forming a “tripod” system (see Fig. 3) could represent a curved parameter space. This system consists of three degenerate internal atomic states, each state coupled to an excited state by a laser field. The fields lift the degeneracy of the internal states, but only partially: two of the energy levels remain degenerate for all field configurations. These degenerate levels have zero energy and do not involve the excited state, i.e., they form a pair of “dark” states that can be used to encode the states of a single qubit.

One can perform computation on the qubit if the amplitudes and phases of the laser fields are varied around suitable loops. In this way, one can achieve a pair of noncommuting holonomic gates that are universal for the qubit. Next, if the atoms that store the different qubits are allowed to interact, and the fields are varied around a suitable loop, the result is an entangling two-qubit gate of purely geometric origin, which completes the all-geometric universal set.

Remarkably, the above-described scenario for purely geometric quantum computation can be implemented for other quantum gate architectures, such as ion traps [14], superconducting nanocircuits [15], and semiconductor quantum dots [16]. This has made the tripod energy-level system the paradigm scenario for holonomic quantum computation.

Geometric quantum computation
Geometric quantum computation employs one-dimensional geometric phase factors instead of multidimensional holonomies to achieve universal sets of quantum gates. There are three different ways to achieve quantum computation based on geometric phases.

First, consider the Berry phase [8], which occurs in situations like the Aharonov-Bohm setup where a charged particle confined to a box acquires a geometric phase while slowly taking the box around a magnetic flux. Just as in the more general holonomies, described in the previous section, this phase arises in adiabatic evolution, but now for nondegenerate eigenspaces of Hamiltonians. In fact, the Berry phase can be thought of as an adiabatic quantum holonomy restricted to a one-dimensional energy eigenspace.

Berry phases may be used for quantum computation by encoding the logical states in nondegenerate energy levels, such as in the spin-up and spin-down states of a spin?1/2 particle in a magnetic field. When this field rotates slowly around a loop, the spin states will pick up Berry phases of magnitude given by half the enclosed solid angle and of opposite sign, which defines an adiabatic geometric phase-shift gate acting on the two spin states. But these states also pick up different dynamical phases (due to the Zeeman splitting caused by the magnetic field interacting with the spins), which one needs to compensate for. Jones et al. [12] removed these dynamical phases by a clever sequence of radio-frequency fields interrupted by suitable π-pulses (which swap the spin-up and spin-down states) applied to a pair of coupled nuclear spins in an NMR setup.

The second approach to geometric quantum computation is based on the fact that geometric phases may accumulate also in nonadiabatic processes [9], as in cases where the parameters in the Hamiltonian (for instance the magnetic field in the spin-1/2 system) vary rapidly in time, causing transitions between different energy levels. These phases are determined by geometric properties of loops in the state space, such as the space representing the direction of a quantum spin, rather than loops in a space of slow parameters, such as the direction of the rotating magnetic field in the above Berry phase scenario. Compared to the Berry phase, one can identify two important advantages of nonadiabatic geometric phases: they can be implemented much faster, which means that unwanted decoherence effects have a shorter time to take effect, and they may occur even if the dynamical phases vanish, which circumvents the need to introduce complicated techniques to remove these phases.

The key idea of nonadiabatic geometric quantum computation is to find paths in state space along which the dynamical phase is zero. Let us consider a single qubit whose state space is a two-dimensional sphere, called the Bloch sphere, where the polar angle θ (azimuthal angle ?) describes the relative weight (relative phase) between the two computational basis states 0 and 1. No dynamical phases occur if we move the qubit along a great circle (geodesic) on this sphere. Now, consider evolution around a loop consisting of segments of great circles, forming a geodesic polygon (see Fig. 4). It results in a phase factor of purely geometric origin, since the dynamical phase vanishes along each geodesic segment. The geometric phase becomes half the solid angle enclosed by the loop on the Bloch sphere.

To see how nonadiabatic geometric quantum computation [17, 18] can be implemented in a physical setup, one may think of a spin?1/2 particle in a magnetic field. If this magnetic field is varied so that it is always orthogonal to the evolving spin, the dynamical phase vanishes and the resulting phase becomes purely geometric. The sign of the acquired geometric phase depends on the direction of the spin, i.e., a superposition of spin-up and spin-down states picks up a relative phase that is equal to the enclosed solid angle carved out by the motion of the state vector. This phase rotation corresponds to moving the qubit state in the ?-direction on the Bloch sphere (see Fig. 4) and is a nonadiabatic geometric phase gate, one of the required ingredients for a universal set.

The third way to implement geometric quantum computation is based on the idea that in some cases the dynamical phase may be proportional to the geometric phase. For such evolutions, there is no need to remove the dynamical phase, as it should show the same resilience to errors as the geometric phase. This idea goes under the name unconventional geometric quantum computation [19].

As an example of this, Leibfried et al. [20] performed an experimental realization of a robust two-qubit phase gate based on an unconventional geometric phase. They demonstrated how laser beams can move two beryllium ions in space and how this motion results in a quantum phase that is conditional on the internal states of the ions. Zhu and Wang [19] subsequently proved that this phase was in fact an unconventional geometric phase being, in this case, equal to the conventional geometric phase but with a negative sign. The exceptionally high fidelity of the gate implemented by Leibfried et al. [20] indicates that unconventional geometric phase scenarios, which have also been proposed for superconducting [21] and atomic [22] qubits in cavities, could be useful for quantum computation.

Robustness
Geometric phases and quantum holonomies are global properties of quantum evolutions and are therefore robust to local errors. This is the basic reason for the conjectured resilience of holonomic and geometric quantum computation—a conjecture that has been examined in some detail recently.

To understand how geometric phase ideas can be used to achieve error resilience, consider a spin exposed to a magnetic field that fluctuates around an adiabatic loop. These fluctuations cause errors in the acquired phase of the spin. The errors reside only in the dynamical phase since the solid angle of the loop is preserved on average if the fluctuations are sufficiently random. Thus, if the dynamical phases can be removed, one can expect resilience to parameter fluctuations.

Now, the key point is that the influence of dynamical phases on quantum gates can be removed in a systematic way. There are at least two ways to do this: either use spin-echo, or encode qubit states in dark energy eigenspaces. Spin-echo is achieved by traversing the cyclic evolution twice, with the second application in the reverse direction surrounded by a pair of short π-pulses that flip the qubit(s), such as in the NMR setup demonstrated by Jones et al. [12]. This results in an effective removal of the dynamical phase (including fluctuation-induced corrections) as this phase becomes an overall phase with no influence on the gate. Dark states have zero energy and therefore pick up no dynamical phase. Thus, gates based on dark states are automatically robust for adiabatic evolution—a robustness that has been confirmed in simulations of the tripod system [23]. These results constitute strong evidence for the conjectured robustness of holonomic quantum computation in the case of parameter fluctuations.

Parameter fluctuations, however, are not the only error source in quantum computation. Errors can also be introduced via environment-induced decoherence, i.e., processes where the computational system loses its coherence through entanglement with its environment. But decoherence makes it less obvious how to separate the geometric and dynamical contributions to quantum gates. These two contributions become fundamentally intertwined for decohering evolutions of quantum systems. Therefore, how the geometric phase can protect quantum information from environment-induced errors is indeed a subtle issue.

The behavior of the geometric phase in the presence of decoherence has been analyzed from different perspectives recently. Carollo et al. [24] proposed a concept of geometric phase that is based on the quantum trajectory model. This model starts from an assumption of Markovian-type evolutions, which can be decomposed into sets of trajectories of pure state evolutions. Each such quantum trajectory picks up an error-dependent geometric phase. The point with this approach is that it can be used to understand how the effect of different error sources to the geometric phase or quantum holonomy can be prevented. Using this idea, Cen and Zanardi [25] proposed a method, similar to spin-echo, to prevent dissipative errors associated with the lowest order trajectories, by traversing the trajectories twice but in the opposite direction.

Tong et al. [26] introduced a general theoretical framework to study geometric phases for nonunitary evolutions, based on quantum kinematics in interferometry. This approach catches another feature of nonunitary evolution, namely that the state of a quantum system in contact with an environment must be described by a density matrix, i.e., a statistical mixture of several wave functions. When these wave functions evolve, they acquire different geometric phases resulting in an overall geometric phase being a certain average over the statistical mixture. A key point of this approach is that it is kinematic, which means that it applies to any form of underlying dynamics. It has been used to demonstrate the existence of time [27] and temperature [28] scales on which the geometric phase is practically unaffected by decoherence and therefore useful for quantum computation.

Outlook
Advances in the field of geometric phases and quantum holonomies have led to an interesting merging of ideas in geometry and computation. We have witnessed not only new ways to implement quantum holonomies and geometric phases in real systems, but also how these implementations can be used to do robust quantum computation. The results from several groups have provided novel ideas to address the robustness of geometric phases and holonomic quantum computation.

Experiments on quantum computation using adiabatic holonomies have so far been limited to the one-dimensional case: one- and two-qubit gates implemented by geometric phases. A challenge for the future is to perform also matrix-valued holonomies in the laboratory. These are needed for all-geometric universal quantum computation and have the attractive feature of being intrinsically robust to dynamical errors. Therefore, the experimental realization of matrix-valued holonomies would be a key step towards error resilient quantum computation.

Another experimental issue worth further investigation concerns the geometric phase in the presence of open system effects. Du et al. [29] performed an NMR experiment to measure the geometric phase for mixed quantum states, theoretically proposed in [30]. Leek et al. [31] analyzed experimentally the Berry phase for a superconducting qubit affected by parameter fluctuations. A challenge for the future would be to extend this work to more general forms of open system effects, such as dissipation and quantum jumps. Such experiments would be useful to test the error resilience of holonomic and geometric quantum computation in real systems.

Research has reached a level where it becomes relevant to combine holonomic and geometric quantum computation with other forms of error-avoiding and error-correcting methods to improve error resilience. Some steps in that direction have in fact already been taken. Wu et al. [32] demonstrated how to combine the resilience of holonomic quantum computation to parameter fluctuations with the inherent robustness of decoherence free subspaces. Very recently, Oreshkov et al. [33] demonstrated how to combine holonomic quantum computation with active error correction. It seems highly desirable to develop these ideas further with the aim to find ways to reach the required precision for fault-tolerant quantum computation.

References
P. W. Shor, in Proceedings of the 35th Annual Symposium on the Foundations of Computer Science (IEEE Press, Los Alamitos, CA, 1994)[Amazon][WorldCat].
T. Pellizzari, S. A. Gardiner, J. I. Cirac, and P. Zoller, Phys. Rev. Lett. 75, 3788 (1995).
J. I. Cirac and P. Zoller, Phys. Rev. Lett. 74, 4091 (1995).
A. Shnirman, G. Schön, and Z. Hermon, Phys. Rev. Lett. 79, 2371 (1997).
D. Loss and D. P. DiVincenzo, Phys. Rev. A 57, 120 (1998).
E. Knill, Nature 434, 39 (2005).
F. Wilczek and A. Zee, Phys. Rev. Lett. 52, 2111 (1984).
M. V. Berry, Proc. R. Soc. London A 392, 45 (1984).
Y. Aharonov and J. Anandan, Phys. Rev. Lett. 58, 1593 (1987).
P. Zanardi and M. Rasetti, Phys. Lett. A 264, 94 (1999).
J. Pachos and S. Chountasis, Phys. Rev. A 62, 052318 (2000).
J. A. Jones, V. Vedral, A. Ekert, and G. Castagnoli, Nature 403, 869 (2000).
R. G. Unanyan, B. W. Shore, and K. Bergmann, Phys. Rev. A 59, 2910 (1999).
L.-M. Duan, J. I. Cirac, and P. Zoller, Science 292, 1695 (2001).
L. Faoro, J. Siewert, and R. Fazio, Phys. Rev. Lett. 90, 028301 (2003).
P. Solinas, P. Zanardi, N. Zangh`?, and F. Rossi, Phys. Rev. B 67, 121307 (2003).
W. Xiang-Bin and M. Keiji, Phys. Rev. Lett. 87, 097901 (2001).
S.-L. Zhu and Z. D. Wang, Phys. Rev. Lett. 89, 097902 (2002).
S.-L. Zhu and Z. D. Wang, Phys. Rev. Lett. 91, 187902 (2003).
D. Leibfried, B. DeMarco, V. Meyer, D. Lucas, M. Barrett, J. Britton, W. M. Itano, B. Jelenkovic, C. Langer, T. Rosenband, and D. J. Wineland, Nature 422, 412 (2003).
S.-L. Zhu, Z. D. Wang, and P. Zanardi, Phys. Rev. Lett. 94, 100502 (2005).
C. Y. Chen, X.-L. Zhang, Z. J. Deng, K.-L. Gao, and M. Feng, Phys. Rev. A 74, 032328 (2006).
P. Solinas, P. Zanardi, and N. Zangh`?, Phys. Rev. A 70, 042316 (2004).
A. Carollo, I. Fuentes-Guridi, M. Franca Santos, and V. Vedral, Phys. Rev. Lett. 90, 160402 (2003).
L.-X. Cen and P. Zanardi, Phys. Rev. A 70, 052323 (2004).
D. M. Tong, E. Sjöqvist, L. C. Kwek, and C. H. Oh, Phys. Rev. Lett. 93, 080405 (2004).
F. C. Lombardo and P. I. Villar, Phys. Rev. A 74, 042311 (2006).
A. T. Rezakhani and P. Zanardi, Phys. Rev. A 73, 052117 (2006).
J. F. Du, P. Zou, L. C. Kwek, J.-W. Pan, C. H. Oh, A. Ekert, D. K. L. Oi, and M. Ericsson, Phys. Rev. Lett. 91, 100403 (2003).
E. Sjöqvist, A. K. Pati, A. Ekert, J. S. Anandan, M. Ericsson, D. K. L. Oi, and V. Vedral, Phys. Rev. Lett. 85, 2845 (2000).
P. J. Leek, J. M. Fink, A. Blais, R. Bianchetti, M. G¨oppl, J. M. Gambetta, D. I. Schuster, L. Frunzio, R. J. Schoelkopf, and A. Wallraff, Science 318, 1889 (2008).
L.-A. Wu, P. Zanardi, and D. A. Lidar, Phys. Rev. Lett. 95, 130501 (2005).
O. Oreshkov, T. A. Brun, and D. A. Lidar, arXiv:0806.0875.
J.-S. Tsai and Y. Nakamura, NEC Research, http://www.nec.co.jp/rd/Eng/innovative/E3/top.html.

About the Author

Erik Sjöqvist
Erik Sjöqvist received his B.Sc. in 1988 and Ph.D. in 1995, both at Uppsala University, Sweden. During 1996–1999, he carried out post-doctoral research at University of Durham, UK, and Oxford University, UK, as well as at the Atomic Institute in Vienna. Dr Sjöqvist returned to Uppsala University in 2000 as a junior researcher, financed by the Swedish research council, and was appointed professor in 2005. His main research interests include geometric and topological phases with applications to quantum information theory and quantum interferometry. He has developed and taught a wide range of physics courses at both undergraduate and graduate levels. Dr Sjöqvist has published about 60 refereed articles.