HOME

TheInfoList



OR:

In
computational complexity theory In theoretical computer science and mathematics, computational complexity theory focuses on classifying computational problems according to their resource usage, and relating these classes to each other. A computational problem is a task solved ...
, PostBQP is a
complexity class In computational complexity theory, a complexity class is a set (mathematics), set of computational problems of related resource-based computational complexity, complexity. The two most commonly analyzed resources are time complexity, time and spa ...
consisting of all of the
computational problem In theoretical computer science, a computational problem is a problem that may be solved by an algorithm. For example, the problem of factoring :"Given a positive integer ''n'', find a nontrivial prime factor of ''n''." is a computational probl ...
s solvable in polynomial time on a quantum Turing machine with
postselection In probability theory, to postselect is to condition a probability space upon the occurrence of a given event. In symbols, once we postselect for an event E, the probability of some other event F changes from \operatorname /math> to the conditional ...
and bounded error (in the sense that the algorithm is correct at least 2/3 of the time on all inputs). Postselection is not considered to be a feature that a realistic computer (even a quantum one) would possess, but nevertheless postselecting machines are interesting from a theoretical perspective. Removing either one of the two main features (quantumness, postselection) from PostBQP gives the following two complexity classes, both of which are subsets of PostBQP: * BQP is the same as PostBQP except without postselection * BPPpath is the same as PostBQP except that instead of quantum, the algorithm is a classical randomized algorithm (with postselection) The addition of
postselection In probability theory, to postselect is to condition a probability space upon the occurrence of a given event. In symbols, once we postselect for an event E, the probability of some other event F changes from \operatorname /math> to the conditional ...
seems to make quantum Turing machines much more powerful:
Scott Aaronson Scott Joel Aaronson (born May 21, 1981) is an American theoretical computer scientist and David J. Bruton Jr. Centennial Professor of Computer Science at the University of Texas at Austin. His primary areas of research are quantum computing a ...
proved. Preprint available a

/ref> PostBQP is equal to PP (complexity), PP, a class which is believed to be relatively powerful, whereas BQP is not known even to contain the seemingly smaller class NP. Using similar techniques, Aaronson also proved that small changes to the laws of quantum computing would have significant effects. As specific examples, under either of the two following changes, the "new" version of BQP would equal PP (complexity), PP: * if we broadened the definition of 'quantum gate' to include not just unitary operations but linear operations, or * if the probability of measuring a basis state , x\rangle was proportional to , \alpha_x, ^p instead of , \alpha_x, ^2 for any even integer ''p > 2''.


Basic properties

In order to describe some of the properties of PostBQP we fix a formal way of describing quantum postselection. Define a quantum algorithm to be a family of
quantum circuit In quantum information theory, a quantum circuit is a model for quantum computation, similar to classical circuits, in which a computation is a sequence of quantum gates, measurements, initializations of qubits to known values, and possibly o ...
s (specifically, a uniform circuit family). We designate one qubit as the ''postselection qubit P'' and another as the ''output qubit Q''. Then PostBQP is defined by postselecting upon the event that the postselection qubit is , 1\rangle. Explicitly, a language ''L'' is in PostBQP if there is a quantum algorithm ''A'' so that after running ''A'' on input ''x'' and measuring the two qubits ''P'' and ''Q'', * ''P'' = 1 with nonzero probability * if the input ''x'' is in ''L'' then Pr ''P'' = 1≥ 2/3 * if the input ''x'' is not in ''L'' then Pr ''P'' = 1≥ 2/3. One can show that allowing a single postselection step at the end of the algorithm (as described above) or allowing intermediate postselection steps during the algorithm are equivalent. Here are three basic properties of PostBQP (which also hold for BQP via similar proofs): # PostBQP is ''closed under complement''. Given a language ''L'' in PostBQP and a corresponding deciding circuit family, create a new circuit family by flipping the output qubit after measurement, then the new circuit family proves the complement of ''L'' is in PostBQP. # You can do ''probability amplification'' in PostBQP. The definition of PostBQP is not changed if we replace the 2/3 value in its definition by any other constant strictly between 1/2 and 1. As an example, given a PostBQP algorithm ''A'' with success probability 2/3, we can construct another algorithm which runs three independent copies of ''A'', outputs a postselection bit equal to the
conjunction Conjunction may refer to: * Conjunction (grammar), a part of speech * Logical conjunction, a mathematical operator ** Conjunction introduction, a rule of inference of propositional logic * Conjunction (astronomy), in which two astronomical bodies ...
of the three "inner" ones, and outputs an output bit equal to the majority of the three "inner" ones; the new algorithm will be correct with conditional probability (2/3)^3 + 3(1/3)(2/3)^2 = 20/27, greater than the original 2/3. # PostBQP is ''closed under intersection''. Suppose we have PostBQP circuit families for two languages and , with respective postselection qubits and output qubits . We may assume by probability amplification that both circuit families have success probability at least 5/6. Then we create a composite algorithm where the circuits for and are run independently, and we set ''P'' to the conjunction of and , and ''Q'' to the conjunction of and . It is not hard to see by a union bound that this composite algorithm correctly decides membership in L_1 \cap L_2 with (conditional) probability at least 2/3. More generally, combinations of these ideas show that PostBQP is closed under union and BQP truth-table reductions.


=

Scott Aaronson Scott Joel Aaronson (born May 21, 1981) is an American theoretical computer scientist and David J. Bruton Jr. Centennial Professor of Computer Science at the University of Texas at Austin. His primary areas of research are quantum computing a ...
showed that the complexity classes (postselected bounded error quantum polynomial time) and PP (complexity), PP (probabilistic polynomial time) are equal. The result was significant because this quantum computation reformulation of gave new insight and simpler proofs of properties of . The usual definition of a circuit family is one with two outbit qubits ''P'' (postselection) and ''Q'' (output) with a single measurement of ''P'' and ''Q'' at the end such that the probability of measuring has nonzero probability, the conditional probability Pr ''P'' = 1≥ 2/3 if the input ''x'' is in the language, and Pr ''P'' = 1≥ 2/3 if the input ''x'' is not in the language. For technical reasons we tweak the definition of as follows: we require that for some constant ''c'' depending on the circuit family. Note this choice does not affect the basic properties of , and also it can be shown that any computation consisting of typical gates (e.g. Hadamard, Toffoli) has this property whenever .


Proving ⊆

Suppose we are given a family of circuits to decide a language ''L''. We assume without loss of generality (e.g. see the inessential properties of quantum computers) that all gates have transition matrices that are represented with real numbers, at the expense of adding one more qubit. Let denote the final quantum state of the circuit before the postselecting measurement is made. The overall goal of the proof is to construct a algorithm to decide ''L''. More specifically it suffices to have ''L'' correctly compare the squared amplitude of in the states with to the squared amplitude of in the states with to determine which is bigger. The key insight is that the comparison of these amplitudes can be transformed into comparing the acceptance probability of a machine with 1/2.


Matrix view of PostBQP algorithms

Let ''n'' denote the input size, denote the total number of qubits in the circuit (inputs, ancillary, output and postselection qubits), and denote the total number of gates. Represent the ''i''th gate by its transition matrix ''A''''i'' (a real unitary 2^B \times 2^B matrix) and let the initial state be , x\rangle (padded with zeroes). Then \Psi = A^G A^\dotsb A^2 A^1 , x\rangle. Define ''S1'' (resp. ''S''0) to be the set of basis states corresponding to (resp. ) and define the probabilities : \pi_1 := \text =1,Q=1\sum_ \Psi^2_\omega : \pi_0 := \text =1,Q=0\sum_ \Psi^2_\omega. The definition of ensures that either \pi_ \ge 2\pi_0 or \pi_0 \ge 2\pi_1 according to whether ''x'' is in ''L'' or not. Our machine will compare \pi_ and \pi_. In order to do this, we expand the definition of matrix multiplication: :: \Psi_\omega = \sum_A^_A^_\dotsb A^2_ A^1_ x_ where the sum is taken over all lists of ''G'' basis vectors \alpha_i. Now \pi_1 and \pi_0 can be expressed as a sum of pairwise products of these terms. Intuitively, we want to design a machine whose acceptance probability is something like \tfrac(1+\pi_1-\pi_0), since then x \in L would imply that the acceptance probability is \tfrac(1+\pi_-\pi_)>\tfrac 1 2, while x \not\in L would imply that the acceptance probability is \tfrac(1+\pi_1-\pi_0)<\tfrac 1 2.


Technicality: we may assume entries of the transition matrices are rationals with denominator for some polynomial ''f''(''n'').

The definition of tells us that \pi_ \ge \tfrac(\pi_0+\pi_1) if ''x'' is in ''L'', and that otherwise \pi_ \ge \tfrac(\pi_0+\pi_1). Let us replace all entries of ''A'' by the nearest fraction with denominator 2^ for a large polynomial that we presently describe. What will be used later is that the ''new'' values satisfy \pi_1 > \tfrac(\pi_0+\pi_1) if ''x'' is in ''L'', and \pi_0 > \tfrac(\pi_0+\pi_1) if ''x'' is not in ''L''. Using the earlier technical assumption and by analyzing how the 1-norm of the computational state changes, this is seen to be satisfied if (1+2^2^)^-1 < \tfrac2^, thus clearly there is a large enough ''f'' that is polynomial in ''n''.


Constructing the PP machine

Now we provide the detailed implementation of our machine. Let denote the sequence \_^G and define the shorthand notation :: \Pi(A, \omega, \alpha, x) := A^_A^_\dotsb A^2_ A^1_ x_, then :: \pi_1 - \pi_0 = \sum_ \sum_ \Pi(A, \omega, \alpha, x)\Pi(A, \omega, \alpha', x) - \sum_ \sum_ \Pi(A, \omega, \alpha, x)\Pi(A, \omega, \alpha', x). We define our machine to * pick a basis state uniformly at random * if \omega \not\in S_0 \cup S_1 then STOP and accept with probability 1/2, reject with probability 1/2 * pick two sequences \alpha,\alpha' of ''G'' basis states uniformly at random * compute X = \Pi(A, \omega, \alpha, x)\Pi(A, \omega, \alpha', x) (which is a fraction with denominator 2^ such that -1 \le X \le 1) * if \omega \in S_1 then accept with probability \tfrac and reject with probability \tfrac (which takes at most coin flips) * otherwise (then \omega \in S_0) accept with probability \tfrac and reject with probability \tfrac (which again takes at most coin flips) Then it is straightforward to compute that this machine accepts with probability \frac+\frac, so this is a machine for the language ''L'', as needed.


Proving ⊆

Suppose we have a machine with time complexity on input ''x'' of length n := , x, . Thus the machine flips a coin at most ''T'' times during the computation. We can thus view the machine as a deterministic function ''f'' (implemented, e.g. by a classical circuit) which takes two inputs (''x, r'') where ''r'', a binary string of length ''T'', represents the results of the random coin flips that are performed by the computation, and the output of ''f'' is 1 (accept) or 0 (reject). The definition of tells us that :: x \in L \Leftrightarrow \#\ \ge 2^ Thus, we want a algorithm that can determine whether the above statement is true. Define ''s'' to be the number of random strings which lead to acceptance, :: s := \#\ and so 2^T-s is the number of rejected strings. It is straightforward to argue that without loss of generality, s \not\in \; for details, see a similar
without loss of generality ''Without loss of generality'' (often abbreviated to WOLOG, WLOG or w.l.o.g.; less commonly stated as ''without any loss of generality'' or ''with no loss of generality'') is a frequently used expression in mathematics. The term is used to indicat ...
assumption in the proof that is closed under complementation.


Aaronson's algorithm

Aaronson's algorithm for solving this problem is as follows. For simplicity, we will write all quantum states as unnormalized. First, we prepare the state , x \rangle \otimes \sum_ , r \rangle , f(x,r) \rangle. Second, we apply
Hadamard gate The Hadamard transform (also known as the Walsh–Hadamard transform, Hadamard–Rademacher–Walsh transform, Walsh transform, or Walsh–Fourier transform) is an example of a generalized class of Fourier transforms. It performs an orthogon ...
s to the second register (each of the first ''T'' qubits), measure the second register and postselect on it being the all-zero string. It is easy to verify that this leaves the last register (the last qubit) in the residual state :: , \psi \rangle := (2^T-s), 0 \rangle + s, 1 \rangle. Where ''H'' denotes the Hadamard gate, we compute the state :: H , \psi\rangle = (2^T , 0\rangle + (2^T - 2s), 1 \rangle)/\sqrt . Where \alpha, \beta are positive real numbers to be chosen later with \alpha^2+\beta^2=1, we compute the state \alpha , 0\rangle, \psi\rangle+\beta , 1\rangle, H\psi\rangle and measure the second qubit, postselecting on its value being equal to 1, which leaves the first qubit in a residual state depending on \beta/\alpha which we denote :: , \phi_ \rangle := \alpha s , 0\rangle + \frac(2^T-2s), 1\rangle. Visualizing the possible states of a qubit as a circle, we see that if s > 2^, (i.e. if x \in L) then \phi_ lies in the open quadrant Q_ := (-, 1\rangle, , 0\rangle) while if s < 2^, (i.e. if x \not\in L) then \phi_ lies in the open quadrant Q_ := (, 0\rangle,, 1\rangle). In fact for any fixed ''x'' (and its corresponding ''s''), as we vary the ratio \beta/\alpha in (0, \infty), note that the image of , \phi_\rangle is precisely the corresponding open quadrant. In the rest of the proof, we make precise the idea that we can distinguish between these two quadrants.


Analysis

Let , +\rangle = (, 1\rangle+, 0\rangle)/\sqrt, which is the center of Q_, and let , -\rangle be orthogonal to , +\rangle. Any qubit in Q_, when measured in the basis \, gives the value , +\rangle less than 1/2 of the time. On the other hand, if x \not\in L and we had picked \beta/\alpha = r^* := \sqrts / (2^T-2s) then measuring , \phi_ \rangle in the basis \ would give the value , +\rangle all of the time. Since we don't know ''s'' we also don't know the precise value of ''r*'', but we can try several (polynomially many) different values for \beta/\alpha in hopes of getting one that is "near" ''r*''. Specifically, note 2^ < r* < 2^T and let us successively set \beta/\alpha to every value of the form 2^i for -T \leq i \leq T. Then elementary calculations show that for one of these values of ''i'', the probability that the measurement of , \phi_ \rangle in the basis \ yields , +\rangle is at least (3+2\sqrt)/6 \approx 0.971. Overall, the algorithm is as follows. Let ''k'' be any constant strictly between 1/2 and (3+2\sqrt)/6. We do the following experiment for each -T \leq i \leq T: construct and measure , \phi_ \rangle in the basis \ a total of C \log T times where ''C'' is a constant. If the proportion of , +\rangle measurements is greater than ''k'', then reject. If we don't reject for any ''i'', accept.
Chernoff bound In probability theory, the Chernoff bound gives exponentially decreasing bounds on tail distributions of sums of independent random variables. Despite being named after Herman Chernoff, the author of the paper it first appeared in, the result is d ...
s then show that for a sufficiently large universal constant ''C'', we correctly classify ''x'' with probability at least 2/3. Note that this algorithm satisfies the technical assumption that the overall postselection probability is not too small: each individual measurement of , \phi_ \rangle has postselection probability 1/2^ and so the overall probability is 1/2^.


Implications

* See Quantum computation reformulation of PP


References

{{DEFAULTSORT:Postbqp Articles containing proofs Quantum complexity theory Probabilistic complexity classes