BQP

Last updated
BQP in relation to other probabilistic complexity classes (ZPP, RP, co-RP, BPP, PP), which generalise P within PSPACE. It is unknown if any of these containments are strict. Randomised Complexity Classes 2.svg
BQP in relation to other probabilistic complexity classes (ZPP, RP, co-RP, BPP, PP), which generalise P within PSPACE. It is unknown if any of these containments are strict.

In computational complexity theory, bounded-error quantum polynomial time (BQP) is the class of decision problems solvable by a quantum computer in polynomial time, with an error probability of at most 1/3 for all instances. [1] It is the quantum analogue to the complexity class BPP .

Contents

A decision problem is a member of BQP if there exists a quantum algorithm (an algorithm that runs on a quantum computer) that solves the decision problem with high probability and is guaranteed to run in polynomial time. A run of the algorithm will correctly solve the decision problem with a probability of at least 2/3.

BQP algorithm (1 run)
Answer
produced
Correct
answer
YesNo
Yes≥ 2/3≤ 1/3
No≤ 1/3≥ 2/3
BQP algorithm (k runs)
Answer
produced
Correct
answer
YesNo
Yes> 1 − 2ck< 2ck
No< 2ck> 1 − 2ck
for some constant c > 0

Definition

BQP can be viewed as the languages associated with certain bounded-error uniform families of quantum circuits. [1] A language L is in BQP if and only if there exists a polynomial-time uniform family of quantum circuits , such that

Alternatively, one can define BQP in terms of quantum Turing machines. A language L is in BQP if and only if there exists a polynomial quantum Turing machine that accepts L with an error probability of at most 1/3 for all instances. [2]

Similarly to other "bounded error" probabilistic classes, the choice of 1/3 in the definition is arbitrary. We can run the algorithm a constant number of times and take a majority vote to achieve any desired probability of correctness less than 1, using the Chernoff bound. The complexity class is unchanged by allowing error as high as 1/2 − nc on the one hand, or requiring error as small as 2nc on the other hand, where c is any positive constant, and n is the length of input. [3]

Relationship to other complexity classes

Unsolved problem in computer science:

What is the relationship between and ?

The suspected relationship of BQP to other problem spaces BQP complexity class diagram.svg
The suspected relationship of BQP to other problem spaces

BQP is defined for quantum computers; the corresponding complexity class for classical computers (or more formally for probabilistic Turing machines) is BPP . Just like P and BPP, BQP is low for itself, which means BQPBQP = BQP. [2] Informally, this is true because polynomial time algorithms are closed under composition. If a polynomial time algorithm calls polynomial time algorithms as subroutines, the resulting algorithm is still polynomial time.

BQP contains P and BPP and is contained in AWPP , [4] PP [5] and PSPACE . [2] In fact, BQP is low for PP, meaning that a PP machine achieves no benefit from being able to solve BQP problems instantly, an indication of the possible difference in power between these similar classes. The known relationships with classic complexity classes are:


As the problem of P ≟ PSPACE has not yet been solved, the proof of inequality between BQP and classes mentioned above is supposed to be difficult. [2] The relation between BQP and NP is not known. In May 2018, computer scientists Ran Raz of Princeton University and Avishay Tal of Stanford University published a paper [6] which showed that, relative to an oracle, BQP was not contained in PH. It can be proven that there exists an oracle A such that BQPA PHA. [7] In an extremely informal sense, this can be thought of as giving PH and BQP an identical, but additional, capability and verifying that BQP with the oracle (BQPA) can do things PHA cannot. While an oracle separation has been proven, the fact that BQP is not contained in PH has not been proven. An oracle separation does not prove whether or not complexity classes are the same. The oracle separation gives intuition that BQP may not be contained in PH.

It has been suspected for many years that Fourier Sampling is a problem that exists within BQP, but not within the polynomial hierarchy. Recent conjectures have provided evidence that a similar problem, Fourier Checking, also exists in the class BQP without being contained in the polynomial hierarchy. This conjecture is especially notable because it suggests that problems existing in BQP could be classified as harder than NP-Complete problems. Paired with the fact that many practical BQP problems are suspected to exist outside of P (it is suspected and not verified because there is no proof that P ≠ NP), this illustrates the potential power of quantum computing in relation to classical computing. [7]

Adding postselection to BQP results in the complexity class PostBQP which is equal to PP . [8] [9]

A complete problem for Promise-BQP

Promise-BQP is the class of promise problems that can be solved by a uniform family of quantum circuits (i.e., within BQP). [10] Completeness proofs focus on this version of BQP. Similar to the notion of NP-completeness and other complete problems, we can define a complete problem as a problem that is in Promise-BQP and that every other problem in Promise-BQP reduces to it in polynomial time.

APPROX-QCIRCUIT-PROB

The APPROX-QCIRCUIT-PROB problem is complete for efficient quantum computation, and the version presented below is complete for the Promise-BQP complexity class (and not for the total BQP complexity class, for which no complete problems are known). APPROX-QCIRCUIT-PROB's completeness makes it useful for proofs showing the relationships between other complexity classes and BQP.

Given a description of a quantum circuit acting on qubits with gates, where is a polynomial in and each gate acts on one or two qubits, and two numbers , distinguish between the following two cases:

  • measuring the first qubit of the state yields with probability
  • measuring the first qubit of the state yields with probability

Here, there is a promise on the inputs as the problem does not specify the behavior if an instance is not covered by these two cases.

Claim. Any BQP problem reduces to APPROX-QCIRCUIT-PROB.

Proof. Suppose we have an algorithm that solves APPROX-QCIRCUIT-PROB, i.e., given a quantum circuit acting on qubits, and two numbers , distinguishes between the above two cases. We can solve any problem in BQP with this oracle, by setting .

For any , there exists family of quantum circuits such that for all , a state of qubits, if ; else if . Fix an input of qubits, and the corresponding quantum circuit . We can first construct a circuit such that . This can be done easily by hardwiring and apply a sequence of CNOT gates to flip the qubits. Then we can combine two circuits to get , and now . And finally, necessarily the results of is obtained by measuring several qubits and apply some (classical) logic gates to them. We can always defer the measurement [11] [12] and reroute the circuits so that by measuring the first qubit of , we get the output. This will be our circuit , and we decide the membership of in by running with . By definition of BQP, we will either fall into the first case (acceptance), or the second case (rejection), so reduces to APPROX-QCIRCUIT-PROB.

BQP and EXP

We begin with an easier containment. To show that , it suffices to show that APPROX-QCIRCUIT-PROB is in EXP since APPROX-QCIRCUIT-PROB is BQP-complete.

Claim  

Proof

The idea is simple. Since we have exponential power, given a quantum circuit C, we can use classical computer to stimulate each gate in C to get the final state.

More formally, let C be a polynomial sized quantum circuit on n qubits and m gates, where m is polynomial in n. Let and be the state after the i-th gate in the circuit is applied to . Each state can be represented in a classical computer as a unit vector in . Furthermore, each gate can be represented by a matrix in . Hence, the final state can be computed in time, and therefore all together, we have an time algorithm for calculating the final state, and thus the probability that the first qubit is measured to be one. This implies that .

Note that this algorithm also requires space to store the vectors and the matrices. We will show in the following section that we can improve upon the space complexity.

BQP and PSPACE

Sum of histories is a technique introduced by physicist Richard Feynman for path integral formulation. APPROX-QCIRCUIT-PROB can be formulated in the sum of histories technique to show that . [13]

Sum of Histories Tree Sum of histories tree.png
Sum of Histories Tree

Consider a quantum circuit C, which consists of t gates, , where each comes from a universal gate set and acts on at most two qubits. To understand what the sum of histories is, we visualize the evolution of a quantum state given a quantum circuit as a tree. The root is the input , and each node in the tree has children, each representing a state in . The weight on a tree edge from a node in j-th level representing a state to a node in -th level representing a state is , the amplitude of after applying on . The transition amplitude of a root-to-leaf path is the product of all the weights on the edges along the path. To get the probability of the final state being , we sum up the amplitudes of all root-to-leave paths that ends at a node representing .

More formally, for the quantum circuit C, its sum over histories tree is a tree of depth m, with one level for each gate in addition to the root, and with branching factor .

Define  A history is a path in the sum of histories tree. We will denote a history by a sequence for some final state x.

Define  Let . Let amplitude of the edge in the j-th level of the sum over histories tree be . For any history , the transition amplitude of the history is the product .

Claim  For a history . The transition amplitude of the history is computable in polynomial time.

Proof

Each gate can be decomposed into for some unitary operator acting on two qubits, which without loss of generality can be taken to be the first two. Hence, which can be computed in polynomial time in n. Since m is polynomial in n, the transition amplitude of the history can be computed in polynomial time.

Claim  Let be the final state of the quantum circuit. For some , the amplitude can be computed by .

Proof

We have . The result comes directly by inserting between , and , and so on, and then expand out the equation. Then each term corresponds to a , where

Claim  

Notice in the sum over histories algorithm to compute some amplitude , only one history is stored at any point in the computation. Hence, the sum over histories algorithm uses space to compute for any x since bits are needed to store the histories in addition to some workspace variables.

Therefore, in polynomial space, we may compute over all x with the first qubit being 1, which is the probability that the first qubit is measured to be 1 by the end of the circuit.

Notice that compared with the simulation given for the proof that , our algorithm here takes far less space but far more time instead. In fact it takes time to calculate a single amplitude!

BQP and PP

A similar sum-over-histories argument can be used to show that . [14]

P and BQP

We know , since every classical circuit can be simulated by a quantum circuit. [15]

It is conjectured that BQP solves hard problems outside of P, specifically, problems in NP. The claim is indefinite because we don't know if P=NP, so we don't know if those problems are actually in P. Below are some evidence of the conjecture:

See also

Related Research Articles

<span class="mw-page-title-main">Quantum computing</span> Technology that uses quantum mechanics

A quantum computer is a computer that takes advantage of quantum mechanical phenomena.

Shor's algorithm is a quantum algorithm for finding the prime factors of an integer. It was developed in 1994 by the American mathematician Peter Shor. It is one of the few known quantum algorithms with compelling potential applications and strong evidence of superpolynomial speedup compared to best known classical algorithms. On the other hand, factoring numbers of practical significance requires far more qubits than available in the near future. Another concern is that noise in quantum circuits may undermine results, requiring additional qubits for quantum error correction.

In quantum computing, Grover's algorithm, also known as the quantum search algorithm, is a quantum algorithm for unstructured search that finds with high probability the unique input to a black box function that produces a particular output value, using just evaluations of the function, where is the size of the function's domain. It was devised by Lov Grover in 1996.

<span class="mw-page-title-main">Quantum superposition</span> Principle of quantum mechanics

Quantum superposition is a fundamental principle of quantum mechanics that states that linear combinations of solutions to the Schrödinger equation are also solutions of the Schrödinger equation. This follows from the fact that the Schrödinger equation is a linear differential equation in time and position. More precisely, the state of a system is given by a linear combination of all the eigenfunctions of the Schrödinger equation governing that system.

The Deutsch–Jozsa algorithm is a deterministic quantum algorithm proposed by David Deutsch and Richard Jozsa in 1992 with improvements by Richard Cleve, Artur Ekert, Chiara Macchiavello, and Michele Mosca in 1998. Although of little practical use, it is one of the first examples of a quantum algorithm that is exponentially faster than any possible deterministic classical algorithm.

<span class="mw-page-title-main">PP (complexity)</span> Class of problems in computer science

In complexity theory, PP is the class of decision problems solvable by a probabilistic Turing machine in polynomial time, with an error probability of less than 1/2 for all instances. The abbreviation PP refers to probabilistic polynomial time. The complexity class was defined by Gill in 1977.

In quantum computing and specifically the quantum circuit model of computation, a quantum logic gate is a basic quantum circuit operating on a small number of qubits. Quantum logic gates are the building blocks of quantum circuits, like classical logic gates are for conventional digital circuits.

Quantum error correction (QEC) is used in quantum computing to protect quantum information from errors due to decoherence and other quantum noise. Quantum error correction is theorised as essential to achieve fault tolerant quantum computing that can reduce the effects of noise on stored quantum information, faulty quantum gates, faulty quantum preparation, and faulty measurements. This would allow algorithms of greater circuit depth.

In computational complexity theory, P/poly is a complexity class representing problems that can be solved by small circuits. More precisely, it is the set of formal languages that have polynomial-size circuit families. It can also be defined equivalently in terms of Turing machines with advice, extra information supplied to the Turing machine along with its input, that may depend on the input length but not on the input itself. In this formulation, P/poly is the class of decision problems that can be solved by a polynomial-time Turing machine with advice strings of length polynomial in the input size. These two different definitions make P/poly central to circuit complexity and non-uniform complexity.

In computational complexity theory, PostBQP is a complexity class consisting of all of the computational problems solvable in polynomial time on a quantum Turing machine with postselection and bounded error.

In computational complexity theory, QMA, which stands for Quantum Merlin Arthur, is the set of languages for which, when a string is in the language, there is a polynomial-size quantum proof that convinces a polynomial time quantum verifier of this fact with high probability. Moreover, when the string is not in the language, every polynomial-size quantum state is rejected by the verifier with high probability.

Quantum complexity theory is the subfield of computational complexity theory that deals with complexity classes defined using quantum computers, a computational model based on quantum mechanics. It studies the hardness of computational problems in relation to these complexity classes, as well as the relationship between quantum complexity classes and classical complexity classes.

In discrete mathematics, ideal lattices are a special class of lattices and a generalization of cyclic lattices. Ideal lattices naturally occur in many parts of number theory, but also in other areas. In particular, they have a significant place in cryptography. Micciancio defined a generalization of cyclic lattices as ideal lattices. They can be used in cryptosystems to decrease by a square root the number of parameters necessary to describe a lattice, making them more efficient. Ideal lattices are a new concept, but similar lattice classes have been used for a long time. For example, cyclic lattices, a special case of ideal lattices, are used in NTRUEncrypt and NTRUSign.

Vector logic is an algebraic model of elementary logic based on matrix algebra. Vector logic assumes that the truth values map on vectors, and that the monadic and dyadic operations are executed by matrix operators. "Vector logic" has also been used to refer to the representation of classical propositional logic as a vector space, in which the unit vectors are propositional variables. Predicate logic can be represented as a vector space of the same type in which the axes represent the predicate letters and . In the vector space for propositional logic the origin represents the false, F, and the infinite periphery represents the true, T, whereas in the space for predicate logic the origin represents "nothing" and the periphery represents the flight from nothing, or "something".

Quantum refereed game in quantum information processing is a class of games in the general theory of quantum games. It is played between two players, Alice and Bob, and arbitrated by a referee. The referee outputs the pay-off for the players after interacting with them for a fixed number of rounds, while exchanging quantum information.

Quantum optimization algorithms are quantum algorithms that are used to solve optimization problems. Mathematical optimization deals with finding the best solution to a problem from a set of possible solutions. Mostly, the optimization problem is formulated as a minimization problem, where one tries to minimize an error which depends on the solution: the optimal solution has the minimal error. Different optimization techniques are applied in various fields such as mechanics, economics and engineering, and as the complexity and amount of data involved rise, more efficient ways of solving optimization problems are needed. Quantum computing may allow problems which are not practically feasible on classical computers to be solved, or suggest a considerable speed up with respect to the best known classical algorithm.

Quantum counting algorithm is a quantum algorithm for efficiently counting the number of solutions for a given search problem. The algorithm is based on the quantum phase estimation algorithm and on Grover's search algorithm.

<span class="mw-page-title-main">Bernstein–Vazirani algorithm</span> Quantum algorithm

The Bernstein–Vazirani algorithm, which solves the Bernstein–Vazirani problem, is a quantum algorithm invented by Ethan Bernstein and Umesh Vazirani in 1997. It is a restricted version of the Deutsch–Jozsa algorithm where instead of distinguishing between two different classes of functions, it tries to learn a string encoded in a function. The Bernstein–Vazirani algorithm was designed to prove an oracle separation between complexity classes BQP and BPP.

The One Clean Qubit model of computation is performed an qubit system with one pure state and maximally mixed states. This model was motivated by highly mixed states that are prevalent in Nuclear magnetic resonance quantum computers. It's described by the density matrix , where I is the identity matrix. In computational complexity theory, DQC1; also known as the Deterministic quantum computation with one clean qubit is the class of decision problems solvable by a one clean qubit machine in polynomial time, upon measuring the first qubit, with an error probability of at most 1/poly(n) for all instances.

This glossary of quantum computing is a list of definitions of terms and concepts used in quantum computing, its sub-disciplines, and related fields.

References

  1. 1 2 3 Michael Nielsen and Isaac Chuang (2000). Quantum Computation and Quantum Information. Cambridge: Cambridge University Press. ISBN   0-521-63503-9.
  2. 1 2 3 4 Bernstein, Ethan; Vazirani, Umesh (October 1997). "Quantum Complexity Theory". SIAM Journal on Computing. 26 (5): 1411–1473. CiteSeerX   10.1.1.655.1186 . doi:10.1137/S0097539796300921.
  3. Barak, Sanjeev Arora, Boaz (2009). Computational Complexity: A Modern Approach / Sanjeev Arora and Boaz Barak. Cambridge. p. 122. Retrieved 24 July 2018.{{cite book}}: CS1 maint: location missing publisher (link) CS1 maint: multiple names: authors list (link)
  4. Fortnow, Lance; Rogers, John (1999). "Complexity limitations on Quantum computation" (PDF). J. Comput. Syst. Sci. 59 (2): 240–252. arXiv: cs/9811023 . doi:10.1006/jcss.1999.1651. ISSN   0022-0000. S2CID   42516312. Archived (PDF) from the original on 2022-10-09.
  5. L. Adleman, J. DeMarrais, and M.-D. Huang. Quantum computability. SIAM J. Comput., 26(5):1524–1540, 1997.
  6. George, Michael Goderbauer, Stefan. "ECCC - TR18-107". eccc.weizmann.ac.il. Retrieved 2018-08-03.{{cite web}}: CS1 maint: multiple names: authors list (link)
  7. 1 2 Aaronson, Scott (2010). "BQP and the Polynomial Hierarchy" (PDF). Proceedings of ACM STOC 2010. Archived (PDF) from the original on 2022-10-09.
  8. Aaronson, Scott (2005). "Quantum computing, postselection, and probabilistic polynomial-time". Proceedings of the Royal Society A. 461 (2063): 3473–3482. arXiv: quant-ph/0412187 . Bibcode:2005RSPSA.461.3473A. doi:10.1098/rspa.2005.1546. S2CID   1770389.. Preprint available at
  9. Aaronson, Scott (2004-01-11). "Complexity Class of the Week: PP". Computational Complexity Weblog. Retrieved 2008-05-02.
  10. Janzing, Dominik; Wocjan, Pawel (2007-03-30). "A Simple PromiseBQP-complete Matrix Problem" (PDF). Theory of Computing. 3 (4): 61–79. doi:10.4086/toc.2007.v003a004 . Retrieved 2024-04-18.
  11. Michael A. Nielsen; Isaac L. Chuang (9 December 2010). "4.4 Measurement". Quantum Computation and Quantum Information: 10th Anniversary Edition. Cambridge University Press. p. 186. ISBN   978-1-139-49548-6.
  12. Odel A. Cross (5 November 2012). "5.2.2 Deferred Measurement". Topics in Quantum Computing. O. A. Cross. p. 348. ISBN   978-1-4800-2749-7.
  13. E. Bernstein and U. Vazirani. Quantum complexity theory, SIAM Journal on Computing, 26(5):1411-1473, 1997.
  14. L. Adleman, J. DeMarrais, and M. Huang. Quantum computability, SIAM Journal on Comput- ing 26:1524-1540, 1997.
  15. Nielsen, Michael A.; Chuang, Isaac L. (2000), Quantum Computation and Quantum Information, Cambridge: Cambridge University Press, ISBN 0-521-63235-8, MR 1796805.
  16. 1 2 arXiv:quant-ph/9508027v2 Polynomial-Time Algorithms for Prime Factorization and Discrete Logarithms on a Quantum Computer, Peter W. Shor