Quantum computing

The Bloch sphere is a representation of a qubit, the fundamental building block of quantum computers.

Quantum computing studies theoretical computation systems (quantum computers) that make direct use of quantum-mechanical phenomena, such as superposition and entanglement, to perform operations on data.[1] Quantum computers are different from binary digital electronic computers based on transistors. Whereas common digital computing requires that the data be encoded into binary digits (bits), each of which is always in one of two definite states (0 or 1), quantum computation is analog and uses quantum bits, which can be in an infinite number of superpositions of states. A quantum Turing machine is a theoretical model of such a computer, and is also known as the universal quantum computer. Quantum computers share theoretical similarities with non-deterministic and probabilistic computers. The field of quantum computing was initiated by the work of Paul Benioff[2] and Yuri Manin in 1980,[3] Richard Feynman in 1982,[4] and David Deutsch in 1985.[5] A quantum computer with spins as quantum bits was also formulated for use as a quantum space–time in 1968.[6]

As of 2016, the development of actual quantum computers is still in its infancy, but experiments have been carried out in which quantum computational operations were executed on a very small number of quantum bits.[7] Both practical and theoretical research continues, and many national governments and military agencies are funding quantum computing research in an effort to develop quantum computers for civilian, business, trade, environmental and national security purposes, such as cryptanalysis.[8]

Large-scale quantum computers would theoretically be able to solve certain problems much more quickly than any classical computers that use even the best currently known algorithms, like integer factorization using Shor's algorithm or the simulation of quantum many-body systems. There exist quantum algorithms, such as Simon's algorithm, that run faster than any possible probabilistic classical algorithm.[9] Given sufficient computational resources, a classical computer could in theory simulate any quantum algorithm, as quantum computation does not violate the Church–Turing thesis.[10]:202 On the other hand, quantum computers may be able to efficiently solve problems which are not practically feasible on classical computers.

Basis

A classical computer has a memory made up of bits, where each bit is represented by either a one or a zero. A quantum computer maintains a sequence of qubits. A single qubit can represent a one, a zero, or any quantum superposition of those two qubit states;[10]:13–16 a pair of qubits can be in any quantum superposition of 4 states,[10]:16 and three qubits in any superposition of 8 states. In general, a quantum computer with qubits can be in an arbitrary superposition of up to different states simultaneously[10]:17 (this compares to a normal computer that can only be in one of these states at any one time). A quantum computer operates by setting the qubits in a perfect drift that represents the problem at hand and by manipulating those qubits with a fixed sequence of quantum logic gates. The sequence of gates to be applied is called a quantum algorithm. The calculation ends with a measurement, collapsing the system of qubits into one of the pure states, where each qubit is zero or one, decomposing into a classical state. The outcome can therefore be at most classical bits of information. Quantum algorithms are often probabilistic, in that they provide the correct solution only with a certain known probability.[11] Note that the term non-deterministic computing must not be used in that case to mean probabilistic (computing), because the term non-deterministic has a different meaning in computer science.

An example of an implementation of qubits of a quantum computer could start with the use of particles with two spin states: "down" and "up" (typically written and , or and ). But in fact any system possessing an observable quantity A, which is conserved under time evolution such that A has at least two discrete and sufficiently spaced consecutive eigenvalues, is a suitable candidate for implementing a qubit. This is true because any such system can be mapped onto an effective spin-1/2 system.

Some have expressed reservations about the theoretical soundness of quantum computing given the original Copenhagen interpretation of quantum mechanics and Quantum Bayesianism. Such criticism centers around questioning the mathematical versus physical interpretation of superposition and the possibility that the wave function is merely psi-epistemic rather than psi-ontic.[12][13][14] There has been no empirical test to fundamentally resolve this debate and quantum computing cannot be said to fully rely on established science.

Principles of operation

A quantum computer with a given number of qubits is fundamentally different from a classical computer composed of the same number of classical bits. For example, representing the state of an n-qubit system on a classical computer requires the storage of 2n complex coefficients, while to characterize the state of a classical n-bit system it is sufficient to provide the values of the n bits, that is, only n numbers. Although this fact may seem to indicate that qubits can hold exponentially more information than their classical counterparts, care must be taken not to overlook the fact that the qubits are only in a probabilistic superposition of all of their states. This means that when the final state of the qubits is measured, they will only be found in one of the possible configurations they were in before the measurement. It is in general incorrect to think of a system of qubits as being in one particular state before the measurement, since the fact that they were in a superposition of states before the measurement was made directly affects the possible outcomes of the computation.

Qubits are made up of controlled particles and the means of control (e.g. devices that trap particles and switch them from one state to another).[15]

To better understand this point, consider a classical computer that operates on a three-bit register. If the exact state of the register at a given time is not known, it can be described as a probability distribution over the different three-bit strings 000, 001, 010, 011, 100, 101, 110, and 111. If there is no uncertainty over its state, then it is in exactly one of these states with probability 1. However, if it is a probabilistic computer, then there is a possibility of it being in any one of a number of different states.

The state of a three-qubit quantum computer is similarly described by an eight-dimensional vector . Here, however, the coefficients are complex numbers, and it is the sum of the squares of the coefficients' absolute values, , that must equal 1. For each , the absolute value squared gives the probability of the system being found after a measurement in the -th state. However, because a complex number encodes not just a magnitude but also a direction in the complex plane, the phase difference between any two coefficients (states) represents a meaningful parameter. This is a fundamental difference between quantum computing and probabilistic classical computing.[16]

If you measure the three qubits, you will observe a three-bit string. The probability of measuring a given string is the squared magnitude of that string's coefficient (i.e., the probability of measuring 000 = , the probability of measuring 001 = , etc.). Thus, measuring a quantum state described by complex coefficients gives the classical probability distribution and we say that the quantum state "collapses" to a classical state as a result of making the measurement.

An eight-dimensional vector can be specified in many different ways depending on what basis is chosen for the space. The basis of bit strings (e.g., 000, 001, …, 111) is known as the computational basis. Other possible bases are unit-length, orthogonal vectors and the eigenvectors of the Pauli-x operator. Ket notation is often used to make the choice of basis explicit. For example, the state in the computational basis can be written as:

where, e.g.,

The computational basis for a single qubit (two dimensions) is and .

Using the eigenvectors of the Pauli-x operator, a single qubit is and .

Operation

Unsolved problem in physics:

While a classical 3-bit state and a quantum 3-qubit state are each eight-dimensional vectors, they are manipulated quite differently for classical or quantum computation. For computing in either case, the system must be initialized, for example into the all-zeros string, , corresponding to the vector (1,0,0,0,0,0,0,0). In classical randomized computation, the system evolves according to the application of stochastic matrices, which preserve that the probabilities add up to one (i.e., preserve the L1 norm). In quantum computation, on the other hand, allowed operations are unitary matrices, which are effectively rotations (they preserve that the sum of the squares add up to one, the Euclidean or L2 norm). (Exactly what unitaries can be applied depend on the physics of the quantum device.) Consequently, since rotations can be undone by rotating backward, quantum computations are reversible. (Technically, quantum operations can be probabilistic combinations of unitaries, so quantum computation really does generalize classical computation. See quantum circuit for a more precise formulation.)

Finally, upon termination of the algorithm, the result needs to be read off. In the case of a classical computer, we sample from the probability distribution on the three-bit register to obtain one definite three-bit string, say 000. Quantum mechanically, we measure the three-qubit state, which is equivalent to collapsing the quantum state down to a classical distribution (with the coefficients in the classical state being the squared magnitudes of the coefficients for the quantum state, as described above), followed by sampling from that distribution. This destroys the original quantum state. Many algorithms will only give the correct answer with a certain probability. However, by repeatedly initializing, running and measuring the quantum computer's results, the probability of getting the correct answer can be increased. In contrast, counterfactual quantum computation allows the correct answer to be inferred when the quantum computer is not actually running in a technical sense, though earlier initialization and frequent measurements are part of the counterfactual computation protocol.

For more details on the sequences of operations used for various quantum algorithms, see universal quantum computer, Shor's algorithm, Grover's algorithm, Deutsch–Jozsa algorithm, amplitude amplification, quantum Fourier transform, quantum gate, quantum adiabatic algorithm and quantum error correction.

Potential

Integer factorization, which underpins the security of public key cryptographic systems, is believed to be computationally infeasible with an ordinary computer for large integers if they are the product of few prime numbers (e.g., products of two 300-digit primes).[17] By comparison, a quantum computer could efficiently solve this problem using Shor's algorithm to find its factors. This ability would allow a quantum computer to decrypt many of the cryptographic systems in use today, in the sense that there would be a polynomial time (in the number of digits of the integer) algorithm for solving the problem. In particular, most of the popular public key ciphers are based on the difficulty of factoring integers or the discrete logarithm problem, both of which can be solved by Shor's algorithm. In particular the RSA, Diffie-Hellman, and Elliptic curve Diffie-Hellman algorithms could be broken. These are used to protect secure Web pages, encrypted email, and many other types of data. Breaking these would have significant ramifications for electronic privacy and security.

However, other cryptographic algorithms do not appear to be broken by those algorithms.[18][19] Some public-key algorithms are based on problems other than the integer factorization and discrete logarithm problems to which Shor's algorithm applies, like the McEliece cryptosystem based on a problem in coding theory.[18][20] Lattice-based cryptosystems are also not known to be broken by quantum computers, and finding a polynomial time algorithm for solving the dihedral hidden subgroup problem, which would break many lattice based cryptosystems, is a well-studied open problem.[21] It has been proven that applying Grover's algorithm to break a symmetric (secret key) algorithm by brute force requires time equal to roughly 2n/2 invocations of the underlying cryptographic algorithm, compared with roughly 2n in the classical case,[22] meaning that symmetric key lengths are effectively halved: AES-256 would have the same security against an attack using Grover's algorithm that AES-128 has against classical brute-force search (see Key size). Quantum cryptography could potentially fulfill some of the functions of public key cryptography.

Besides factorization and discrete logarithms, quantum algorithms offering a more than polynomial speedup over the best known classical algorithm have been found for several problems,[23] including the simulation of quantum physical processes from chemistry and solid state physics, the approximation of Jones polynomials, and solving Pell's equation. No mathematical proof has been found that shows that an equally fast classical algorithm cannot be discovered, although this is considered unlikely.[24] For some problems, quantum computers offer a polynomial speedup. The most well-known example of this is quantum database search, which can be solved by Grover's algorithm using quadratically fewer queries to the database than are required by classical algorithms. In this case the advantage is provable. Several other examples of provable quantum speedups for query problems have subsequently been discovered, such as for finding collisions in two-to-one functions and evaluating NAND trees.

Consider a problem that has these four properties:

  1. The only way to solve it is to guess answers repeatedly and check them,
  2. The number of possible answers to check is the same as the number of inputs,
  3. Every possible answer takes the same amount of time to check, and
  4. There are no clues about which answers might be better: generating possibilities randomly is just as good as checking them in some special order.

An example of this is a password cracker that attempts to guess the password for an encrypted file (assuming that the password has a maximum possible length).

For problems with all four properties, the time for a quantum computer to solve this will be proportional to the square root of the number of inputs. It can be used to attack symmetric ciphers such as Triple DES and AES by attempting to guess the secret key.[25]

Grover's algorithm can also be used to obtain a quadratic speed-up over a brute-force search for a class of problems known as NP-complete.

Since chemistry and nanotechnology rely on understanding quantum systems, and such systems are impossible to simulate in an efficient manner classically, many believe quantum simulation will be one of the most important applications of quantum computing.[26] Quantum simulation could also be used to simulate the behavior of atoms and particles at unusual conditions such as the reactions inside a collider.[27]

There are a number of technical challenges in building a large-scale quantum computer, and thus far quantum computers have yet to solve a problem faster than a classical computer. David DiVincenzo, of IBM, listed the following requirements for a practical quantum computer:[28]

Quantum decoherence

One of the greatest challenges is controlling or removing quantum decoherence. This usually means isolating the system from its environment as interactions with the external world cause the system to decohere. However, other sources of decoherence also exist. Examples include the quantum gates, and the lattice vibrations and background thermonuclear spin of the physical system used to implement the qubits. Decoherence is irreversible, as it is non-unitary, and is usually something that should be highly controlled, if not avoided. Decoherence times for candidate systems, in particular the transverse relaxation time T2 (for NMR and MRI technology, also called the dephasing time), typically range between nanoseconds and seconds at low temperature.[16] Currently, some quantum computers require their qubits to be cooled to 20 millikelvins in order to prevent significant decoherence.[29]

These issues are more difficult for optical approaches as the timescales are orders of magnitude shorter and an often-cited approach to overcoming them is optical pulse shaping. Error rates are typically proportional to the ratio of operating time to decoherence time, hence any operation must be completed much more quickly than the decoherence time.

If the error rate is small enough, it is thought to be possible to use quantum error correction, which corrects errors due to decoherence, thereby allowing the total calculation time to be longer than the decoherence time. An often cited figure for required error rate in each gate is 10−4. This implies that each gate must be able to perform its task in one 10,000th of the coherence time of the system.

Meeting this scalability condition is possible for a wide range of systems. However, the use of error correction brings with it the cost of a greatly increased number of required bits. The number required to factor integers using Shor's algorithm is still polynomial, and thought to be between L and L2, where L is the number of bits in the number to be factored; error correction algorithms would inflate this figure by an additional factor of L. For a 1000-bit number, this implies a need for about 104 bits without error correction.[30] With error correction, the figure would rise to about 107 bits. Computation time is about L2 or about 107 steps and at 1 MHz, about 10 seconds.

A very different approach to the stability-decoherence problem is to create a topological quantum computer with anyons, quasi-particles used as threads and relying on braid theory to form stable logic gates.[31][32]

Developments

There are a number of quantum computing models, distinguished by the basic elements in which the computation is decomposed. The four main models of practical importance are:

The Quantum Turing machine is theoretically important but direct implementation of this model is not pursued. All four models of computation have been shown to be equivalent; each can simulate the other with no more than polynomial overhead.

For physically implementing a quantum computer, many different candidates are being pursued, among them (distinguished by the physical system used to realize the qubits):

The large number of candidates demonstrates that the topic, in spite of rapid progress, is still in its infancy. There is also a vast amount of flexibility.

Timeline

In 2001, researchers demonstrated Shor's algorithm to factor 15 using a 7-qubit NMR computer.[48]

In 2005, researchers at the University of Michigan built a semiconductor chip ion trap. Such devices from standard lithography, may point the way to scalable quantum computing.[49]

In 2009, researchers at Yale University created the first solid-state quantum processor. The two-qubit superconducting chip had artificial atom qubits made of a billion aluminum atoms that acted like a single atom that could occupy two states.[50][51]

A team at the University of Bristol, also created a silicon chip based on quantum optics, able to run Shor's algorithm.[52] Further developments were made in 2010.[53] Springer publishes a journal (Quantum Information Processing) devoted to the subject.[54]

In February 2010, Digital Combinational Circuits like adder, subtractor etc. are designed with the help of Symmetric Functions organized from different quantum gates.[55][56]

April 2011, a team of scientists from Australia and Japan made a breakthrough in quantum teleportation. They successfully transferred a complex set of quantum data with full transmission integrity, without affecting the qubits' superpositions.[57][58]

Photograph of a chip constructed by D-Wave Systems Inc., mounted and wire-bonded in a sample holder. The D-Wave processor is designed to use 128 superconducting logic elements that exhibit controllable and tunable coupling to perform operations.

In 2011, D-Wave Systems announced the first commercial quantum annealer, the D-Wave One, claiming a 128 qubit processor.[59] On May 25, 2011 Lockheed Martin agreed to purchase a D-Wave One system.[60] Lockheed and the University of Southern California (USC) will house the D-Wave One at the newly formed USC Lockheed Martin Quantum Computing Center.[61] D-Wave's engineers designed the chips with an empirical approach, focusing on solving particular problems. Investors liked this more than academics, who said D-Wave had not demonstrated they really had a quantum computer. Criticism softened after a D-Wave paper in Nature, that proved the chips have some quantum properties.[62][63] Two published papers have suggested that the D-Wave machine's operation can be explained classically, rather than requiring quantum models.[64][65] Later work showed that classical models are insufficient when all available data is considered.[66] Experts remain divided on the ultimate classification of the D-Wave systems though their quantum behavior was established concretely with a demonstration of entanglement.[67]

During the same year, researchers at the University of Bristol created an all-bulk optics system that ran a version of Shor's algorithm to successfully factor 21.[68]

In September 2011 researchers proved quantum computers can be made with a Von Neumann architecture (separation of RAM).[69]

In November 2011 researchers factorized 143 using 4 qubits.[70]

In February 2012 IBM scientists said that they had made several breakthroughs in quantum computing with superconducting integrated circuits.[71]

In April 2012 a multinational team of researchers from the University of Southern California, Delft University of Technology, the Iowa State University of Science and Technology, and the University of California, Santa Barbara, constructed a two-qubit quantum computer on a doped diamond crystal that can easily be scaled up and is functional at room temperature. Two logical qubit directions of electron spin and nitrogen kernels spin were used, with microwave impulses. This computer ran Grover's algorithm generating the right answer from the first try in 95% of cases.[72]

In September 2012, Australian researchers at the University of New South Wales said the world's first quantum computer was just 5 to 10 years away, after announcing a global breakthrough enabling manufacture of its memory building blocks. A research team led by Australian engineers created the first working qubit based on a single atom in silicon, invoking the same technological platform that forms the building blocks of modern-day computers.[73][74]

In October 2012, Nobel Prizes were presented to David J. Wineland and Serge Haroche for their basic work on understanding the quantum world, which may help make quantum computing possible.[75][76]

In November 2012, the first quantum teleportation from one macroscopic object to another was reported.[77][78]

In December 2012, the first dedicated quantum computing software company, 1QBit was founded in Vancouver, BC.[79] 1QBit is the first company to focus exclusively on commercializing software applications for commercially available quantum computers, including the D-Wave Two. 1QBit's research demonstrated the ability of superconducting quantum annealing processors to solve real-world problems.[80]

In February 2013, a new technique, boson sampling, was reported by two groups using photons in an optical lattice that is not a universal quantum computer but may be good enough for practical problems. Science Feb 15, 2013

In May 2013, Google announced that it was launching the Quantum Artificial Intelligence Lab, hosted by NASA's Ames Research Center, with a 512-qubit D-Wave quantum computer. The USRA (Universities Space Research Association) will invite researchers to share time on it with the goal of studying quantum computing for machine learning.[81]

In early 2014 it was reported, based on documents provided by former NSA contractor Edward Snowden, that the U.S. National Security Agency (NSA) is running a $79.7 million research program (titled "Penetrating Hard Targets") to develop a quantum computer capable of breaking vulnerable encryption.[82]

In 2014, a group of researchers from ETH Zürich, USC, Google and Microsoft reported a definition of quantum speedup, and were not able to measure quantum speedup with the D-Wave Two device, but did not explicitly rule it out.[83][84]

In 2014, researchers at University of New South Wales used silicon as a protectant shell around qubits, making them more accurate, increasing the length of time they will hold information and possibly made quantum computers easier to build.[85]

In April 2015 IBM scientists claimed two critical advances towards the realization of a practical quantum computer. They claimed the ability to detect and measure both kinds of quantum errors simultaneously, as well as a new, square quantum bit circuit design that could scale to larger dimensions.[86]

In October 2015 researchers at University of New South Wales built a quantum logic gate in silicon for the first time.[87]

In December 2015 NASA publicly displayed the world's first fully operational $15-million quantum computer made by the Canadian company D-Wave at the Quantum Artificial Intelligence Laboratory at its Ames Research Center in California's Moffett Field. The device was purchased in 2013 via a partnership with Google and Universities Space Research Association. Despite using quantum effects the algorithm run on the quantum computer does not outperform Selby’s algorithm run on a classical computer.[88]

In August 2016, scientists at the University of Maryland successfully built the first reprogrammable quantum computer.[89]

In October 2016 Basel University described a variant of the electron hole based quantum computer, which instead of manipulating electron spins uses electron holes in a semiconductor at low (mK) temperatures which are a lot less vulnerable to decoherence. This has been dubbed the "positronic" quantum computer as the quasi-particle behaves like it has a positive electrical charge.[90]

Relation to computational complexity theory

The suspected relationship of BQP to other problem spaces.[91]

The class of problems that can be efficiently solved by quantum computers is called BQP, for "bounded error, quantum, polynomial time". Quantum computers only run probabilistic algorithms, so BQP on quantum computers is the counterpart of BPP ("bounded error, probabilistic, polynomial time") on classical computers. It is defined as the set of problems solvable with a polynomial-time algorithm, whose probability of error is bounded away from one half.[92] A quantum computer is said to "solve" a problem if, for every instance, its answer will be right with high probability. If that solution runs in polynomial time, then that problem is in BQP.

BQP is contained in the complexity class #P (or more precisely in the associated class of decision problems P#P),[93] which is a subclass of PSPACE.

BQP is suspected to be disjoint from NP-complete and a strict superset of P, but that is not known. Both integer factorization and discrete log are in BQP. Both of these problems are NP problems suspected to be outside BPP, and hence outside P. Both are suspected to not be NP-complete. There is a common misconception that quantum computers can solve NP-complete problems in polynomial time. That is not known to be true, and is generally suspected to be false.[93]

The capacity of a quantum computer to accelerate classical algorithms has rigid limits—upper bounds of quantum computation's complexity. The overwhelming part of classical calculations cannot be accelerated on a quantum computer.[94] A similar fact takes place for particular computational tasks, like the search problem, for which Grover's algorithm is optimal.[95]

Although quantum computers may be faster than classical computers for some problem types, those described above can't solve any problem that classical computers can't already solve. A Turing machine can simulate these quantum computers, so such a quantum computer could never solve an undecidable problem like the halting problem. The existence of "standard" quantum computers does not disprove the Church–Turing thesis.[96] It has been speculated that theories of quantum gravity, such as M-theory or loop quantum gravity, may allow even faster computers to be built. Currently, defining computation in such theories is an open problem due to the problem of time, i.e., there currently exists no obvious way to describe what it means for an observer to submit input to a computer and later receive output.[97]

See also

References

  1. Gershenfeld, Neil; Chuang, Isaac L. (June 1998). "Quantum Computing with Molecules" (PDF). Scientific American.
  2. Benioff, Paul (1980). "The computer as a physical system: A microscopic quantum mechanical Hamiltonian model of computers as represented by Turing machines". Journal of statistical physics. 22 (5): 563–591. Bibcode:1980JSP....22..563B. doi:10.1007/BF01011339.
  3. Manin, Yu. I. (1980). Vychislimoe i nevychislimoe [Computable and Noncomputable] (in Russian). Sov.Radio. pp. 13–15. Retrieved 2013-03-04.
  4. Feynman, R. P.u (1982). "Simulating physics with computers". International Journal of Theoretical Physics. 21 (6): 467–488. Bibcode:1982IJTP...21..467F. doi:10.1007/BF02650179.
  5. Deutsch, David (1985). "Quantum Theory, the Church-Turing Principle and the Universal Quantum Computer". Proceedings of the Royal Society of London A. 400 (1818): 97–117. Bibcode:1985RSPSA.400...97D. doi:10.1098/rspa.1985.0070.
  6. Finkelstein, David (1968). "Space-Time Structure in High Energy Interactions". In Gudehus, T.; Kaiser, G. Fundamental Interactions at High Energy. New York: Gordon & Breach.
  7. Gershon, Eric (2013-01-14). "New qubit control bodes well for future of quantum computing". Phys.org. Retrieved 2014-10-26.
  8. Quantum Information Science and Technology Roadmap for a sense of where the research is heading.
  9. Simon, D.R. (1994). "On the power of quantum computation". Foundations of Computer Science, 1994 Proceedings., 35th Annual Symposium on: 116–123. doi:10.1109/SFCS.1994.365701. ISBN 0-8186-6580-7.
  10. 1 2 3 4 Chuang, Michael A. Nielsen & Isaac L. (2001). Quantum computation and quantum information (Repr. ed.). Cambridge [u.a.]: Cambridge Univ. Press. ISBN 978-0521635035.
  11. Preskill, John (2015). "Lecture Notes for Ph219/CS219: Quantum Information Chapter 5" (PDF). p. 12.
  12. Schlafly, Roger. "Concise Argument Against Quantum Computing". Dark Buzz.
  13. Schlafly, Roger. "Positivist Interpretation". Dark Buzz.
  14. Mermin, David. "Physics: QBism puts the scientist back into science". nature.com. Nature.
  15. Waldner, Jean-Baptiste (2007). Nanocomputers and Swarm Intelligence. London: ISTE. p. 157. ISBN 2-7462-1516-0.
  16. 1 2 DiVincenzo, David P. (1995). "Quantum Computation". Science. 270 (5234): 255–261. Bibcode:1995Sci...270..255D. doi:10.1126/science.270.5234.255. (subscription required)
  17. Lenstra, Arjen K. (2000). "Integer Factoring" (PDF). Designs, Codes and Cryptography. 19 (2/3): 101–128. doi:10.1023/A:1008397921377.
  18. 1 2 Daniel J. Bernstein, Introduction to Post-Quantum Cryptography. Introduction to Daniel J. Bernstein, Johannes Buchmann, Erik Dahmen (editors). Post-quantum cryptography. Springer, Berlin, 2009. ISBN 978-3-540-88701-0
  19. See also pqcrypto.org, a bibliography maintained by Daniel J. Bernstein and Tanja Lange on cryptography not known to be broken by quantum computing.
  20. Robert J. McEliece. "A public-key cryptosystem based on algebraic coding theory." Jet Propulsion Laboratory DSN Progress Report 42–44, 114–116.
  21. Kobayashi, H.; Gall, F.L. (2006). "Dihedral Hidden Subgroup Problem: A Survey". Information and Media Technologies. 1 (1): 178–185.
  22. Bennett C.H., Bernstein E., Brassard G., Vazirani U., "The strengths and weaknesses of quantum computation". SIAM Journal on Computing 26(5): 1510–1523 (1997).
  23. Quantum Algorithm Zoo – Stephen Jordan's Homepage
  24. Jon Schiller, Phd. "Quantum Computers".
  25. Rich, Steven; Gellman, Barton (2014-02-01). "NSA seeks to build quantum computer that could crack most types of encryption". Washington Post.
  26. Norton, Quinn (2007-02-15). "The Father of Quantum Computing". Wired.com.
  27. Ambainis, Andris (Spring 2014). "What Can We Do with a Quantum Computer?". Institute for Advanced Study.
  28. DiVincenzo, David P. (2000-04-13). "The Physical Implementation of Quantum Computation". arXiv:quant-ph/0002077Freely accessible [quant-ph].
  29. Jones, Nicola (19 June 2013). "Computing: The quantum company". Nature. 498 (7454): 286–288. doi:10.1038/498286a. PMID 23783610.
  30. Dyakonov, M. I. (2006-10-14). "Is Fault-Tolerant Quantum Computation Really Possible?". In: Future Trends in Microelectronics. Up the Nano Creek, S. Luryi, J. Xu, and A. Zaslavsky (eds), Wiley, pp.: 4–18. arXiv:quant-ph/0610117Freely accessible.
  31. Freedman, Michael H.; Kitaev, Alexei; Larsen, Michael J.; Wang, Zhenghan (2003). "Topological quantum computation". Bulletin of the American Mathematical Society. 40 (1): 31–38. arXiv:quant-ph/0101025Freely accessible. doi:10.1090/S0273-0979-02-00964-3. MR 1943131.
  32. Monroe, Don (2008-10-01). "Anyons: The breakthrough quantum computing needs?". New Scientist.
  33. Das, A.; Chakrabarti, B. K. (2008). "Quantum Annealing and Analog Quantum Computation". Rev. Mod. Phys. 80 (3): 1061–1081. Bibcode:2008RvMP...80.1061D. doi:10.1103/RevModPhys.80.1061.
  34. Nayak, Chetan; Simon, Steven; Stern, Ady; Das Sarma, Sankar (2008). "Nonabelian Anyons and Quantum Computation". Rev Mod Phys. 80 (3): 1083. arXiv:0707.1889Freely accessible. Bibcode:2008RvMP...80.1083N. doi:10.1103/RevModPhys.80.1083.
  35. Clarke, John; Wilhelm, Frank (June 19, 2008). "Superconducting quantum bits". Nature. 453 (7198): 1031–1042. Bibcode:2008Natur.453.1031C. doi:10.1038/nature07128. PMID 18563154.
  36. Kaminsky, William M (2004). "Scalable Superconducting Architecture for Adiabatic Quantum Computation". arXiv:quant-ph/0403090Freely accessible [quant-ph].
  37. Imamoğlu, Atac; Awschalom, D. D.; Burkard, Guido; DiVincenzo, D. P.; Loss, D.; Sherwin, M.; Small, A. (1999). "Quantum information processing using quantum dot spins and cavity-QED". Physical Review Letters. 83 (20): 4204. Bibcode:1999PhRvL..83.4204I. doi:10.1103/PhysRevLett.83.4204.
  38. Fedichkin, Leonid; Yanchenko, Maxim; Valiev, Kamil (2000). "Novel coherent quantum bit using spatial quantization levels in semiconductor quantum dot". Quantum Computers and Computing. 1: 58–76. arXiv:quant-ph/0006097Freely accessible. Bibcode:2000quant.ph..6097F.
  39. "Quantum computing in molecular magnets.". Nature. 410 (6830): 789–93. Apr 12, 2001. doi:10.1038/35071024. PMID 11298441.
  40. Knill, G. J.; Laflamme, R.; Milburn, G. J. (2001). "A scheme for efficient quantum computation with linear optics". Nature. 409 (6816): 46–52. Bibcode:2001Natur.409...46K. doi:10.1038/35051009. PMID 11343107.
  41. Nizovtsev, A. P. (August 2005). "A quantum computer based on NV centers in diamond: Optically detected nutations of single electron and nuclear spins". Optics and Spectroscopy. 99 (2): 248–260. Bibcode:2005OptSp..99..233N. doi:10.1134/1.2034610.
  42. Gruener, Wolfgang (2007-06-01). "Research indicates diamonds could be key to quantum storage". Retrieved 2007-06-04.
  43. Neumann, P.; et al. (June 6, 2008). "Multipartite Entanglement Among Single Spins in Diamond". Science. 320 (5881): 1326–1329. Bibcode:2008Sci...320.1326N. doi:10.1126/science.1157233. PMID 18535240.
  44. Millman, Rene (2007-08-03). "Trapped atoms could advance quantum computing". ITPro. Archived from the original on 2007-09-27. Retrieved 2007-07-26.
  45. Ohlsson, N.; Mohan, R. K.; Kröll, S. (January 1, 2002). "Quantum computer hardware based on rare-earth-ion-doped inorganic crystals". Opt. Commun. 201 (1–3): 71–77. Bibcode:2002OptCo.201...71O. doi:10.1016/S0030-4018(01)01666-2.
  46. Longdell, J. J.; Sellars, M. J.; Manson, N. B. (September 23, 2004). "Demonstration of conditional quantum phase shift between ions in a solid". Phys. Rev. Lett. 93 (13): 130503. arXiv:quant-ph/0404083Freely accessible. Bibcode:2004PhRvL..93m0503L. doi:10.1103/PhysRevLett.93.130503. PMID 15524694.
  47. Náfrádi, Bálint; Choucair, Mohammad; Dinse, Klaus-Peter; Forró, László (July 18, 2016). "Room Temperature manipulation of long lifetime spins in metallic-like carbon nanospheres". Nature Communications: 12232. doi:10.1038/ncomms12232.
  48. Vandersypen, Lieven M. K.; Steffen, Matthias; Breyta, Gregory; Yannoni, Costantino S.; Sherwood, Mark H.; Chuang, Isaac L. (2001). "Experimental realization of Shor's quantum factoring algorithm using nuclear magnetic resonance". Nature. 414 (6866): 883–7. doi:10.1038/414883a. PMID 11780055.
  49. "U-M develops scalable and mass-producible quantum computer chip". University of Michigan. 2005-12-12. Retrieved 2006-11-17.
  50. DiCarlo, L.; Chow, J. M.; Gambetta, J. M.; Bishop, Lev S.; Johnson, B. R.; Schuster, D. I.; Majer, J.; Blais, A.; Frunzio, L.; S. M. Girvin; R. J. Schoelkopf (9 July 2009). "Demonstration of two-qubit algorithms with a superconducting quantum processor" (PDF). Nature. 460 (7252): 240–4. Bibcode:2009Natur.460..240D. doi:10.1038/nature08121. PMID 19561592. Retrieved 2009-07-02.
  51. "Scientists Create First Electronic Quantum Processor". Yale University. 2009-07-02. Retrieved 2009-07-02.
  52. "Code-breaking quantum algorithm runs on a silicon chip". New Scientist. 2009-09-04. Retrieved 2009-10-14.
  53. "New Trends in Quantum Computation". Simons Conference on New Trends in Quantum Computation 2010: Program. C.N. Yang Institute for Theoretical Physics.
  54. "Quantum Information Processing". Springer.com. Retrieved on 2011-05-19.
  55. Bhattacharjee, Pijush Kanti (2010). "Digital Combinational Circuits Design by QCA Gates" (PDF). International Journal of Computer and Electrical Engineering (IJCEE), Singapore, vol. 2, no. 1, pp. 67-72, February 2010.
  56. Bhattacharjee, Pijush Kanti (2010). "Digital Combinational Circuits Design with the Help of Symmetric Functions Considering Heat Dissipation by Each QCA Gate" (PDF). International Journal of Computer and Electrical Engineering (IJCEE), Singapore, vol. 2, no. 4, pp. 666-672, August 2010.
  57. "Quantum teleporter breakthrough". University of New South Wales. 2011-04-15. Archived from the original on 2011-04-18.
  58. Lai, Richard (2011-04-18). "First light wave quantum teleportation achieved, opens door to ultra fast data transmission". Engadget.
  59. "Learning to program the D-Wave One". Hack The Multiverse. D-Wave. Retrieved 2011-05-11.
  60. "D-Wave Systems sells its first Quantum Computing System to Lockheed Martin Corporation". D-Wave. 2011-05-25. Retrieved 2011-05-30.
  61. "Operational Quantum Computing Center Established at USC". University of Southern California. 2011-10-29. Retrieved 2011-12-06.
  62. Johnson, M. W.; Amin, M. H. S.; Gildert, S.; Lanting, T.; Hamze, F.; Dickson, N.; Harris, R.; Berkley, A. J.; Johansson, J.; Bunyk, P.; Chapple, E. M.; Enderud, C.; Hilton, J. P.; Karimi, K.; Ladizinsky, E.; Ladizinsky, N.; Oh, T.; Perminov, I.; Rich, C.; Thom, M. C.; Tolkacheva, E.; Truncik, C. J. S.; Uchaikin, S.; Wang, J.; Wilson, B.; Rose, G. (12 May 2011). "Quantum annealing with manufactured spins". Nature. 473 (7346): 194–198. doi:10.1038/nature10012. PMID 21562559.
  63. Simonite, Tom (October 4, 2012). "The CIA and Jeff Bezos Bet on Quantum Computing". Technology Review.
  64. Seung Woo Shin; Smith, Graeme; Smolin, John A.; Vazirani, Umesh (2014-05-02). "How "Quantum" is the D-Wave Machine?". arXiv:1401.7087Freely accessible [quant-ph].
  65. Boixo, Sergio; Rønnow, Troels F.; Isakov, Sergei V.; Wang, Zhihui; Wecker, David; Lidar, Daniel A.; Martinis, John M.; Troyer, Matthias (2013-04-16). "Quantum Annealing With More Than 100 Qbits". Nature Physics. 10 (3): 218. arXiv:1304.4595Freely accessible. doi:10.1038/nphys2900.
  66. Albash, Tameem; Rønnow, Troels F.; Troyer, Matthias; Lidar, Daniel A. (2014-09-12). "Reexamining classical and quantum models for the D-Wave One processor". The European Physical Journal Special Topics. 224 (111): 111. arXiv:1409.3827Freely accessible. doi:10.1140/epjst/e2015-02346-0.
  67. Lanting, T.; Przybysz, A. J.; Smirnov, A. Yu.; Spedalieri, F. M.; Amin, M. H.; Berkley, A. J.; Harris, R.; Altomare, F.; Boixo, S.; Bunyk, P.; Dickson, N.; Enderud, C.; Hilton, J. P.; Hoskinson, E.; Johnson, M. W.; Ladizinsky, E.; Ladizinsky, N.; Neufeld, R.; Oh, T.; Perminov, I.; Rich, C.; Thom, M. C.; Tolkacheva, E.; Uchaikin, S.; Wilson, A. B.; Rose, G. (2014-05-29). "Entanglement in a quantum annealing processor". Physical Review X. prx. 4 (2). doi:10.1103/PhysRevX.4.021041.
  68. Lopez, Enrique Martin; Laing, Anthony; Lawson, Thomas; Alvarez, Roberto; Zhou, Xiao-Qi; O'Brien, Jeremy L. (2011). "Implementation of an iterative quantum order finding algorithm". Nature Photonics. 6 (11): 773–776. arXiv:1111.4147Freely accessible. doi:10.1038/nphoton.2012.259.
  69. Mariantoni, Matteo; Wang, H.; Yamamoto, T.; Neeley, M.; Bialczak, Radoslaw C.; Chen, Y.; Lenander, M.; Lucero, Erik; O'Connell, A. D.; Sank, D.; Weides, M.; Wenner, J.; Yin, Y.; Zhao, J.; Korotkov, A. N.; Cleland, A. N.; Martinis, John M. (2011). "Quantum computer with Von Neumann architecture". Science. 334 (6052): 61–65. arXiv:1109.3743Freely accessible. doi:10.1126/science.1208517.
  70. Xu, Nanyang; Zhu, Jing; Lu, Dawei; Zhou, Xianyi; Peng, Xinhua; Du, Jiangfeng (2011). "Quantum Factorization of 143 on a Dipolar-Coupling NMR system". Physical Review Letters. 109 (26). arXiv:1111.3726Freely accessible. doi:10.1103/PhysRevLett.109.269902.
  71. "IBM Says It's 'On the Cusp' of Building a Quantum Computer". PCMAG. Retrieved 2014-10-26.
  72. "Quantum computer built inside diamond". Futurity. Retrieved 2014-10-26.
  73. "Australian engineers write quantum computer 'qubit' in global breakthrough". The Australian. Retrieved 2012-10-03.
  74. "Breakthrough in bid to create first quantum computer". University of New South Wales. Retrieved 2012-10-03.
  75. Frank, Adam (October 14, 2012). "Cracking the Quantum Safe". New York Times. Retrieved 2012-10-14.
  76. Overbye, Dennis (October 9, 2012). "A Nobel for Teasing Out the Secret Life of Atoms". New York Times. Retrieved 2012-10-14.
  77. "First Teleportation from One Macroscopic Object to Another: The Physics arXiv Blog". MIT Technology Review. November 15, 2012. Retrieved 2012-11-17.
  78. Bao, Xiao-Hui; Xu, Xiao-Fan; Li, Che-Ming; Yuan, Zhen-Sheng; Lu, Chao-Yang; Pan, Jian-wei (November 13, 2012). "Quantum teleportation between remote atomic-ensemble quantum memories". Proceedings of the National Academy of Sciences. 109 (50): 20347. arXiv:1211.2892Freely accessible. doi:10.1073/pnas.1207329109.
  79. "1QBit Founded". 1QBit.com. Retrieved 2014-06-22.
  80. "1QBit Research". 1QBit.com. Retrieved 2014-06-22.
  81. "Launching the Quantum Artificial Intelligence Lab". Research@Google Blog. Retrieved 2013-05-16.
  82. "NSA seeks to build quantum computer that could crack most types of encryption". Washington Post. January 2, 2014.
  83. Defining and detecting quantum speedup, Troels F. Rønnow, Zhihui Wang, Joshua Job, Sergio Boixo, Sergei V. Isakov, David Wecker, John M. Martinis, Daniel A. Lidar, Matthias Troyer, 2014-01-13.
  84. "Quantum Chaos: After a Failed Speed Test, the D-Wave Debate Continues". Scientific American. 2014-06-19.
  85. Gaudin, Sharon (23 October 2014). "Researchers use silicon to push quantum computing toward reality". Computer World.
  86. IBM Scientists Achieve Critical Steps to Building First Practical Quantum Computer
  87. World's First Silicon Quantum Logic Gate Brings Quantum Computing One Step Closer
  88. "3Q: Scott Aaronson on Google's new quantum-computing paper". MIT News. Retrieved 2016-01-05.
  89. MacDonald, Fiona. "Researchers have built the first reprogrammable quantum computer". ScienceAlert. Retrieved 8 August 2016.
  90. https://www.unibas.ch/en/News-Events/News/Uni-Research/A-new-Type-of-Quantum-Bit.html
  91. Nielsen, p. 42
  92. Nielsen, p. 41
  93. 1 2 Bernstein, Ethan; Vazirani, Umesh (1997). "Quantum Complexity Theory". SIAM Journal on Computing. 26 (5): 1411. doi:10.1137/S0097539796300921.
  94. Ozhigov, Yuri (1999). "Quantum Computers Speed Up Classical with Probability Zero". Chaos Solitons Fractals. 10 (10): 1707–1714. arXiv:quant-ph/9803064Freely accessible. Bibcode:1998quant.ph..3064O. doi:10.1016/S0960-0779(98)00226-4.
  95. Ozhigov, Yuri (1999). "Lower Bounds of Quantum Search for Extreme Point". Proceedings of the London Royal Society. A455 (1986): 2165–2172. arXiv:quant-ph/9806001Freely accessible. Bibcode:1999RSPSA.455.2165O. doi:10.1098/rspa.1999.0397.
  96. Nielsen, p. 126
  97. Scott Aaronson, NP-complete Problems and Physical Reality, ACM SIGACT News, Vol. 36, No. 1. (March 2005), pp. 30–52, section 7 "Quantum Gravity": "[…] to anyone who wants a test or benchmark for a favorite quantum gravity theory,[author's footnote: That is, one without all the bother of making numerical predictions and comparing them to observation] let me humbly propose the following: can you define Quantum Gravity Polynomial-Time? […] until we can say what it means for a 'user' to specify an 'input' and ‘later' receive an 'output'—there is no such thing as computation, not even theoretically." (emphasis in original)

Further reading

External links

Wikimedia Commons has media related to Quantum computer.
Lectures
This article is issued from Wikipedia - version of the 12/4/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.