Quantum tomography

Last updated

Quantum tomography or quantum state tomography is the process by which a quantum state is reconstructed using measurements on an ensemble of identical quantum states. [1] The source of these states may be any device or system which prepares quantum states either consistently into quantum pure states or otherwise into general mixed states. To be able to uniquely identify the state, the measurements must be tomographically complete. That is, the measured operators must form an operator basis on the Hilbert space of the system, providing all the information about the state. Such a set of observations is sometimes called a quorum. The term tomography was first used in the quantum physics literature in a 1993 paper introducing experimental optical homodyne tomography. [2]

Contents

Figure 1: One harmonic oscillator represented in the phase space by its momentum and position Figure 1.PNG
Figure 1: One harmonic oscillator represented in the phase space by its momentum and position
Figure 2: Many identical oscillators represented in the phase space by their momentum and position ManyHarmonicOscillatorsPhaseSpaceRepresentation.PNG
Figure 2: Many identical oscillators represented in the phase space by their momentum and position

In quantum process tomography on the other hand, known quantum states are used to probe a quantum process to find out how the process can be described. Similarly, quantum measurement tomography works to find out what measurement is being performed. Whereas, randomized benchmarking scalably obtains a figure of merit of the overlap between the error prone physical quantum process and its ideal counterpart.

The general principle behind quantum state tomography is that by repeatedly performing many different measurements on quantum systems described by identical density matrices, frequency counts can be used to infer probabilities, and these probabilities are combined with Born's rule to determine a density matrix which fits the best with the observations.

This can be easily understood by making a classical analogy. Consider a harmonic oscillator (e.g. a pendulum). The position and momentum of the oscillator at any given point can be measured and therefore the motion can be completely described by the phase space. This is shown in figure 1. By performing this measurement for a large number of identical oscillators we get a probability distribution in the phase space (figure 2). This distribution can be normalized (the oscillator at a given time has to be somewhere) and the distribution must be non-negative. So we have retrieved a function W(x,p) which gives a description of the chance of finding the particle at a given point with a given momentum.

For quantum mechanical particles the same can be done. The only difference is that the Heisenberg's uncertainty principle mustn't be violated, meaning that we cannot measure the particle's momentum and position at the same time. The particle's momentum and its position are called quadratures (see Optical phase space for more information) in quantum related states. By measuring one of the quadratures of a large number of identical quantum states will give us a probability density corresponding to that particular quadrature. This is called the marginal distribution, pr(X) or pr(P) (see figure 3). In the following text we will see that this probability density is needed to characterize the particle's quantum state, which is the whole point of quantum tomography.

Figure 3: Marginal Distribution MarginalDistribution.PNG
Figure 3: Marginal Distribution

What quantum state tomography is used for

Quantum tomography is applied on a source of systems, to determine the quantum state of the output of that source. Unlike a measurement on a single system, which determines the system's current state after the measurement (in general, the act of making a measurement alters the quantum state), quantum tomography works to determine the state(s) prior to the measurements.

Quantum tomography can be used for characterizing optical signals, including measuring the signal gain and loss of optical devices, [3] as well as in quantum computing and quantum information theory to reliably determine the actual states of the qubits. [4] [5] One can imagine a situation in which a person Bob prepares many identical objects (particles or fields) in the same quantum states and then gives them to Alice to measure. Not confident with Bob's description of the state, Alice may wish to do quantum tomography to classify the state herself.

Methods of quantum state tomography

Linear inversion

Using Born's rule, one can derive the simplest form of quantum tomography. Generally, being in a pure state is not known in advance, and a state may be mixed. In this case, many different types of measurements will have to be performed, many times each. To fully reconstruct the density matrix for a mixed state in a finite-dimensional Hilbert space, the following technique may be used.

Born's rule states , where is a particular measurement outcome projector and is the density matrix of the system. Given a histogram of observations for each measurement, one has an approximation to for each .

Given linear operators and , define the inner product

where is representation of the operator as a column vector and a row vector such that is the inner product in of the two.

Define the matrix as

.

Here Ei is some fixed list of individual measurements (with binary outcomes), and A does all the measurements at once.

Then applying this to yields the probabilities:

.

Linear inversion corresponds to inverting this system using the observed relative frequencies to derive (which is isomorphic to ).

This system is not going to be square in general, as for each measurement being made there will generally be multiple measurement outcome projectors . For example, in a 2-D Hilbert space with 3 measurements , each measurement has 2 outcomes, each of which has a projector Ei, for 6 projectors, whereas the real dimension of the space of density matrices is (2⋅22)/2=4, leaving to be 6 x 4. To solve the system, multiply on the left by :

.

Now solving for yields the pseudoinverse:

.

This works in general only if the measurement list Ei is tomographically complete. Otherwise, the matrix will not be invertible.

Continuous variables and quantum homodyne tomography

In infinite dimensional Hilbert spaces, e.g. in measurements of continuous variables such as position, the methodology is somewhat more complex. One notable example is in the tomography of light, known as optical homodyne tomography. Using balanced homodyne measurements, one can derive the Wigner function and a density matrix for the state of the light. [6]

One approach involves measurements along different rotated directions in phase space. For each direction , one can find a probability distribution for the probability density of measurements in the direction of phase space yielding the value . Using an inverse Radon transformation (the filtered back projection) on leads to the Wigner function, , [7] which can be converted by an inverse Fourier transform into the density matrix for the state in any basis. [5] A similar technique is often used in medical tomography.

Example: single-qubit state tomography

The density matrix of a single qubit can be expressed in terms of its Bloch vector and the Pauli vector :

.

The single-qubit state tomography can be performed by means of single-qubit Pauli measurements: [8]

  1. First, create a list of three quantum circuits, with the first one measuring the qubit in the computational basis (Z-basis), the second one performing a Hadamard gate before measurement (which makes the measurement in X-basis), and the third one performing the appropriate phase shift gate (that is ) followed by a Hadamard gate before measurement (which makes the measurement in Y-basis);
  2. Then, run these circuits (typically thousands of times), and the counts in the measurement results of the first circuit produces , the second circuit , and the third circuit ;
  3. Finally, if , then a measured Bloch vector is produced as , and the measured density matrix is ; If , it'll be necessary to renormalize the measured Bloch vector as before using it to calculate the measured density matrix.

This algorithm is the foundation for qubit tomography and is used in some quantum programming routines, like that of Qiskit. [9] [10]

Example: homodyne tomography.

Electromagnetic field amplitudes (quadratures) can be measured with high efficiency using photodetectors together with temporal mode selectivity. Balanced homodyne tomography is a reliable technique of reconstructing quantum states in the optical domain. This technique combines the advantages of the high efficiencies of photodiodes in measuring the intensity or photon number of light, together with measuring the quantum features of light by a clever set-up called the homodyne tomography detector.

Quantum homodyne tomography is understood by the following example. A laser is directed onto a 50-50% beamsplitter, splitting the laser beam into two beams. One is used as a local oscillator (LO) and the other is used to generate photons with a particular quantum state. The generation of quantum states can be realized, e.g. by directing the laser beam through a frequency doubling crystal [11] and then onto a parametric down-conversion crystal. This crystal generates two photons in a certain quantum state. One of the photons is used as a trigger signal used to trigger (start) the readout event of the homodyne tomography detector. The other photon is directed into the homodyne tomography detector, in order to reconstruct its quantum state. Since the trigger and signal photons are entangled (this is explained by the spontaneous parametric down-conversion article), it is important to realize that the optical mode of the signal state is created nonlocal only when the trigger photon impinges the photodetector (of the trigger event readout module) and is actually measured. More simply said, it is only when the trigger photon is measured, that the signal photon can be measured by the homodyne detector.

Now consider the homodyne tomography detector as depicted in figure 4 (figure missing). The signal photon (this is the quantum state we want to reconstruct) interferes with the local oscillator, when they are directed onto a 50-50% beamsplitter. Since the two beams originate from the same so called master laser, they have the same fixed phase relation. The local oscillator must be intense, compared to the signal so it provides a precise phase reference. The local oscillator is so intense, that we can treat it classically (a = α) and neglect the quantum fluctuations. The signal field is spatially and temporally controlled by the local oscillator, which has a controlled shape. Where the local oscillator is zero, the signal is rejected. Therefore, we have temporal-spatial mode selectivity of the signal. The beamsplitter redirects the two beams to two photodetectors. The photodetectors generate an electric current proportional to the photon number. The two detector currents are subtracted and the resulting current is proportional to the electric field operator in the signal mode, depended on relative optical phase of signal and local oscillator.

Since the electric field amplitude of the local oscillator is much higher than that of the signal the intensity or fluctuations in the signal field can be seen. The homodyne tomography system functions as an amplifier. The system can be seen as an interferometer with such a high intensity reference beam (the local oscillator) that unbalancing the interference by a single photon in the signal is measurable. This amplification is well above the photodetectors noise floor.

The measurement is reproduced a large number of times. Then the phase difference between the signal and local oscillator is changed in order to ‘scan’ a different angle in the phase space. This can be seen from figure 4. The measurement is repeated again a large number of times and a marginal distribution is retrieved from the current difference. The marginal distribution can be transformed into the density matrix and/or the Wigner function. Since the density matrix and the Wigner function give information about the quantum state of the photon, we have reconstructed the quantum state of the photon.

The advantage of this balanced detection method is that this arrangement is insensitive to fluctuations in the intensity of the laser.

The quantum computations for retrieving the quadrature component from the current difference are performed as follows.

The photon number operator for the beams striking the photodetectors after the beamsplitter is given by:

,

where i is 1 and 2, for respectively beam one and two. The mode operators of the field emerging the beamsplitters are given by:

The denotes the annihilation operator of the signal and alpha the complex amplitude of the local oscillator. The number of photon difference is eventually proportional to the quadrature and given by:

,

Rewriting this with the relation:

Results in the following relation:

,

where we see clear relation between the photon number difference and the quadrature component . By keeping track of the sum current, one can recover information about the local oscillator's intensity, since this is usually an unknown quantity, but an important quantity for calculating the quadrature component .

Problems with linear inversion

One of the primary problems with using linear inversion to solve for the density matrix is that in general the computed solution will not be a valid density matrix. For example, it could give negative probabilities or probabilities greater than 1 to certain measurement outcomes. This is particularly an issue when fewer measurements are made.

Another issue is that in infinite dimensional Hilbert spaces, an infinite number of measurement outcomes would be required. Making assumptions about the structure and using a finite measurement basis leads to artifacts in the phase space density. [5]

Maximum likelihood estimation

Maximum likelihood estimation (also known as MLE or MaxLik) is a popular technique for dealing with the problems of linear inversion. By restricting the domain of density matrices to the proper space, and searching for the density matrix which maximizes the likelihood of giving the experimental results, it guarantees the state to be theoretically valid while giving a close fit to the data. The likelihood of a state is the probability that would be assigned to the observed results had the system been in that state.

Suppose the measurements have been observed with frequencies . Then the likelihood associated with a state is

where is the probability of outcome for the state .

Finding the maximum of this function is non-trivial and generally involves iterative methods. [12] [13] The methods are an active topic of research.

Problems with maximum likelihood estimation

Maximum likelihood estimation suffers from some less obvious problems than linear inversion. One problem is that it makes predictions about probabilities that cannot be justified by the data. This is seen most easily by looking at the problem of zero eigenvalues. The computed solution using MLE often contains eigenvalues which are 0, i.e. it is rank deficient. In these cases, the solution then lies on the boundary of the n-dimensional Bloch sphere. This can be seen as related to linear inversion giving states which lie outside the valid space (the Bloch sphere). MLE in these cases picks a nearby point that is valid, and the nearest points are generally on the boundary. [4]

This is not physically a problem, the real state might have zero eigenvalues. However, since no value may be less than 0, an estimate of an eigenvalue being 0 implies that the estimator is certain the value is 0, otherwise they would have estimated some greater than 0 with a small degree of uncertainty as the best estimate. This is where the problem arises, in that it is not logical to conclude with absolute certainty after a finite number of measurements that any eigenvalue (that is, the probability of a particular outcome) is 0. For example, if a coin is flipped 5 times and each time heads was observed, it does not mean there is 0 probability of getting tails, despite that being the most likely description of the coin. [4]

Bayesian methods

Bayesian mean estimation (BME) is a relatively new approach which addresses the problems of maximum likelihood estimation. It focuses on finding optimal solutions which are also honest in that they include error bars in the estimate. The general idea is to start with a likelihood function and a function describing the experimenter's prior knowledge (which might be a constant function), then integrate over all density matrices using the product of the likelihood function and prior knowledge function as a weight.

Given a reasonable prior knowledge function, BME will yield a state strictly within the n-dimensional Bloch sphere. In the case of a coin flipped N times to get N heads described above, with a constant prior knowledge function, BME would assign as the probability for tails. [4]

BME provides a high degree of accuracy in that it minimizes the operational divergences of the estimate from the actual state. [4]

Methods for incomplete data

The number of measurements needed for a full quantum state tomography for a multi-particle system scales exponentially with the number of particles, which makes such a procedure impossible even for modest system sizes. Hence, several methods have been developed to realize quantum tomography with fewer measurements.

The concept of matrix completion and compressed sensing have been applied to reconstruct density matrices from an incomplete set of measurements (that is, a set of measurements which is not a quorum). In general, this is impossible, but under assumptions (for example, if the density matrix is a pure state, or a combination of just a few pure states) then the density matrix has fewer degrees of freedom, and it may be possible to reconstruct the state from the incomplete measurements. [14]

Permutationally Invariant Quantum Tomography [15] is a procedure that has been developed mostly for states that are close to being permutationally symmetric, which is typical in nowadays experiments. For two-state particles, the number of measurements needed scales only quadratically with the number of particles. [16] Besides the modest measurement effort, the processing of the measured data can also be done efficiently: It is possible to carry out the fitting of a physical density matrix on the measured data even for large systems. [17] Permutationally Invariant Quantum Tomography has been combined with compressed sensing in a six-qubit photonic experiment. [18]

Quantum measurement tomography

One can imagine a situation in which an apparatus performs some measurement on quantum systems, and determining what particular measurement is desired. The strategy is to send in systems of various known states, and use these states to estimate the outcomes of the unknown measurement. Also known as "quantum estimation", tomography techniques are increasingly important including those for quantum measurement tomography and the very similar quantum state tomography. Since a measurement can always be characterized by a set of POVM's, the goal is to reconstruct the characterizing POVM's . The simplest approach is linear inversion. As in quantum state observation, use

.

Exploiting linearity as above, this can be inverted to solve for the .

Not surprisingly, this suffers from the same pitfalls as in quantum state tomography: namely, non-physical results, in particular negative probabilities. Here the will not be valid POVM's, as they will not be positive. Bayesian methods as well as Maximum likelihood estimation of the density matrix can be used to restrict the operators to valid physical results. [19]

Quantum process tomography

Quantum process tomography (QPT) deals with identifying an unknown quantum dynamical process. The first approach, introduced in 1996 and sometimes known as standard quantum process tomography (SQPT) involves preparing an ensemble of quantum states and sending them through the process, then using quantum state tomography to identify the resultant states. [20] Other techniques include ancilla-assisted process tomography (AAPT) and entanglement-assisted process tomography (EAPT) which require an extra copy of the system. [21]

Each of the techniques listed above are known as indirect methods for characterization of quantum dynamics, since they require the use of quantum state tomography to reconstruct the process. In contrast, there are direct methods such as direct characterization of quantum dynamics (DCQD) which provide a full characterization of quantum systems without any state tomography. [22]

The number of experimental configurations (state preparations and measurements) required for full quantum process tomography grows exponentially with the number of constituent particles of a system. Consequently, in general, QPT is an impossible task for large-scale quantum systems. However, under weak decoherence assumption, a quantum dynamical map can find a sparse representation. The method of compressed quantum process tomography (CQPT) uses the compressed sensing technique and applies the sparsity assumption to reconstruct a quantum dynamical map from an incomplete set of measurements or test state preparations. [23]

Quantum dynamical maps

A quantum process, also known as a quantum dynamical map, , can be described by a completely positive map

,

where , the bounded operators on Hilbert space; with operation elements satisfying so that .

Let be an orthogonal basis for . Write the operators in this basis

.

This leads to

,

where .

The goal is then to solve for , which is a positive superoperator and completely characterizes with respect to the basis. [21] [22]

Standard quantum process tomography

SQPT approaches this using linearly independent inputs , where is the dimension of the Hilbert space . For each of these input states , sending it through the process gives an output state which can be written as a linear combination of the , i.e. . By sending each through many times, quantum state tomography can be used to determine the coefficients experimentally.

Write

,

where is a matrix of coefficients. Then

.

Since form a linearly independent basis,

.

Inverting gives :

.

Related Research Articles

<span class="mw-page-title-main">Quantum entanglement</span> Correlation between quantum systems

Quantum entanglement is the phenomenon that occurs when a group of particles are generated, interact, or share spatial proximity in a way such that the quantum state of each particle of the group cannot be described independently of the state of the others, including when the particles are separated by a large distance. The topic of quantum entanglement is at the heart of the disparity between classical and quantum physics: entanglement is a primary feature of quantum mechanics not present in classical mechanics.

<span class="mw-page-title-main">Schrödinger equation</span> Description of a quantum-mechanical system

The Schrödinger equation is a linear partial differential equation that governs the wave function of a quantum-mechanical system. Its discovery was a significant landmark in the development of quantum mechanics. It is named after Erwin Schrödinger, who postulated the equation in 1925 and published it in 1926, forming the basis for the work that resulted in his Nobel Prize in Physics in 1933.

In quantum mechanics, a density matrix is a matrix that describes the quantum state of a physical system. It allows for the calculation of the probabilities of the outcomes of any measurement performed upon this system, using the Born rule. It is a generalization of the more usual state vectors or wavefunctions: while those can only represent pure states, density matrices can also represent mixed states. Mixed states arise in quantum mechanics in two different situations:

  1. when the preparation of the system is not fully known, and thus one must deal with a statistical ensemble of possible preparations, and
  2. when one wants to describe a physical system which is entangled with another, without describing their combined state.
<span class="mw-page-title-main">Quantum superposition</span> Principle of quantum mechanics

Quantum superposition is a fundamental principle of quantum mechanics. In classical mechanics, things like position or momentum are always well-defined. We may not know what they are at any given time, but that is an issue of our understanding and not the physical system. In quantum mechanics, a particle can be in a superposition of different states. However, a measurement always finds it in one state, but before and after the measurement, it interacts in ways that can only be explained by having a superposition of different states.

<span class="mw-page-title-main">Quantum decoherence</span> Loss of quantum coherence

Quantum decoherence is the loss of quantum coherence, the process in which a system's behaviour changes from that which can be explained by quantum mechanics to that which can be explained by classical mechanics. In quantum mechanics, particles such as electrons are described by a wave function, a mathematical representation of the quantum state of a system; a probabilistic interpretation of the wave function is used to explain various quantum effects. As long as there exists a definite phase relation between different states, the system is said to be coherent. A definite phase relationship is necessary to perform quantum computing on quantum information encoded in quantum states. Coherence is preserved under the laws of quantum physics.

In physics, specifically in quantum mechanics, a coherent state is the specific quantum state of the quantum harmonic oscillator, often described as a state which has dynamics most closely resembling the oscillatory behavior of a classical harmonic oscillator. It was the first example of quantum dynamics when Erwin Schrödinger derived it in 1926, while searching for solutions of the Schrödinger equation that satisfy the correspondence principle. The quantum harmonic oscillator arise in the quantum theory of a wide range of physical systems. For instance, a coherent state describes the oscillating motion of a particle confined in a quadratic potential well. The coherent state describes a state in a system for which the ground-state wavepacket is displaced from the origin of the system. This state can be related to classical solutions by a particle oscillating with an amplitude equivalent to the displacement.

In quantum physics, a measurement is the testing or manipulation of a physical system to yield a numerical result. A fundamental feature of quantum theory is that the predictions it makes are probabilistic. The procedure for finding a probability involves combining a quantum state, which mathematically describes a quantum system, with a mathematical representation of the measurement to be performed on that system. The formula for this calculation is known as the Born rule. For example, a quantum particle like an electron can be described by a quantum state that associates to each point in space a complex number called a probability amplitude. Applying the Born rule to these amplitudes gives the probabilities that the electron will be found in one region or another when an experiment is performed to locate it. This is the best the theory can do; it cannot say for certain where the electron will be found. The same quantum state can also be used to make a prediction of how the electron will be moving, if an experiment is performed to measure its momentum instead of its position. The uncertainty principle implies that, whatever the quantum state, the range of predictions for the electron's position and the range of predictions for its momentum cannot both be narrow. Some quantum states imply a near-certain prediction of the result of a position measurement, but the result of a momentum measurement will be highly unpredictable, and vice versa. Furthermore, the fact that nature violates the statistical conditions known as Bell inequalities indicates that the unpredictability of quantum measurement results cannot be explained away as due to ignorance about "hidden variables" within quantum systems.

<span class="mw-page-title-main">Bloch sphere</span> Geometrical representation of the pure state space of a two-level quantum mechanical system

In quantum mechanics and computing, the Bloch sphere is a geometrical representation of the pure state space of a two-level quantum mechanical system (qubit), named after the physicist Felix Bloch.

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 functional analysis and quantum information science, a positive operator-valued measure (POVM) is a measure whose values are positive semi-definite operators on a Hilbert space. POVMs are a generalization of projection-valued measures (PVM) and, correspondingly, quantum measurements described by POVMs are a generalization of quantum measurement described by PVMs.

A quasiprobability distribution is a mathematical object similar to a probability distribution but which relaxes some of Kolmogorov's axioms of probability theory. Quasiprobabilities share several of general features with ordinary probabilities, such as, crucially, the ability to yield expectation values with respect to the weights of the distribution. However, they can violate the σ-additivity axiom: integrating over them does not necessarily yield probabilities of mutually exclusive states. Indeed, quasiprobability distributions also have regions of negative probability density, counterintuitively, contradicting the first axiom. Quasiprobability distributions arise naturally in the study of quantum mechanics when treated in phase space formulation, commonly used in quantum optics, time-frequency analysis, and elsewhere.

Photon polarization is the quantum mechanical description of the classical polarized sinusoidal plane electromagnetic wave. An individual photon can be described as having right or left circular polarization, or a superposition of the two. Equivalently, a photon can be described as having horizontal or vertical linear polarization, or a superposition of the two.

The theoretical and experimental justification for the Schrödinger equation motivates the discovery of the Schrödinger equation, the equation that describes the dynamics of nonrelativistic particles. The motivation uses photons, which are relativistic particles with dynamics described by Maxwell's equations, as an analogue for all types of particles.

Resonance fluorescence is the process in which a two-level atom system interacts with the quantum electromagnetic field if the field is driven at a frequency near to the natural frequency of the atom.

The Hong–Ou–Mandel effect is a two-photon interference effect in quantum optics that was demonstrated in 1987 by three physicists from the University of Rochester: Chung Ki Hong (홍정기), Zheyu Ou (区泽宇), and Leonard Mandel. The effect occurs when two identical single-photons enter a 1:1 beam splitter, one in each input port. When the temporal overlap of the photons on the beam splitter is perfect, the two photons will always exit the beam splitter together in the same output mode, meaning that there is zero chance that they will exit separately with one photon in each of the two outputs giving a coincidence event. The photons have a 50:50 chance of exiting (together) in either output mode. If they become more distinguishable, the probability of them each going to a different detector will increase. In this way, the interferometer coincidence signal can accurately measure bandwidth, path lengths, and timing. Since this effect relies on the existence of photons and the second quantization it can not be fully explained by classical optics.

This is a glossary for the terminology often encountered in undergraduate quantum mechanics courses.

The Maxwell–Bloch equations, also called the optical Bloch equations describe the dynamics of a two-state quantum system interacting with the electromagnetic mode of an optical resonator. They are analogous to the Bloch equations which describe the motion of the nuclear magnetic moment in an electromagnetic field. The equations can be derived either semiclassically or with the field fully quantized when certain approximations are made.

<span class="mw-page-title-main">Symmetry in quantum mechanics</span> Properties underlying modern physics

Symmetries in quantum mechanics describe features of spacetime and particles which are unchanged under some transformation, in the context of quantum mechanics, relativistic quantum mechanics and quantum field theory, and with applications in the mathematical formulation of the standard model and condensed matter physics. In general, symmetry in physics, invariance, and conservation laws, are fundamentally important constraints for formulating physical theories and models. In practice, they are powerful methods for solving problems and predicting what can happen. While conservation laws do not always give the answer to the problem directly, they form the correct constraints and the first steps to solving a multitude of problems.

Quantum stochastic calculus is a generalization of stochastic calculus to noncommuting variables. The tools provided by quantum stochastic calculus are of great use for modeling the random evolution of systems undergoing measurement, as in quantum trajectories. Just as the Lindblad master equation provides a quantum generalization to the Fokker–Planck equation, quantum stochastic calculus allows for the derivation of quantum stochastic differential equations (QSDE) that are analogous to classical Langevin equations.

The quantum Fisher information is a central quantity in quantum metrology and is the quantum analogue of the classical Fisher information. The quantum Fisher information of a state with respect to the observable is defined as

References

  1. Quantum State Tomography. "UIUC".
  2. Smithey, D. T.; Beck, M.; Raymer, M. G.; Faridani, A. (1993-03-01). "Measurement of the Wigner distribution and the density matrix of a light mode using optical homodyne tomography: Application to squeezed states and the vacuum". Physical Review Letters. 70 (9): 1244–1247. Bibcode:1993PhRvL..70.1244S. doi:10.1103/physrevlett.70.1244. ISSN   0031-9007. PMID   10054327.
  3. D'Ariano, G Mauro; Laurentis, Martina De; Paris, Matteo G A; Porzio, Alberto; Solimeno, Salvatore (2002-06-01). "Quantum tomography as a tool for the characterization of optical devices". Journal of Optics B: Quantum and Semiclassical Optics. 4 (3): S127–S132. arXiv: quant-ph/0110110 . Bibcode:2002JOptB...4S.127M. doi:10.1088/1464-4266/4/3/366. ISSN   1464-4266. S2CID   17185255.
  4. 1 2 3 4 5 Blume-Kohout, Robin (2010). "Optimal, reliable estimation of quantum states". New Journal of Physics. 12 (4): 043034. arXiv: quant-ph/0611080 . Bibcode:2010NJPh...12d3034B. doi:10.1088/1367-2630/12/4/043034. S2CID   28735241.
  5. 1 2 3 Lvovsky, A.I.; Raymer, M.G. (2009). "Continuous-variable optical quantum state tomography". Reviews of Modern Physics. 81 (1): 299–332. arXiv: quant-ph/0511044 . Bibcode:2009RvMP...81..299L. doi:10.1103/RevModPhys.81.299. S2CID   118928592.
  6. D’Auria, V.; Fornaro, S.; Porzio, A.; Solimeno, S.; Olivares, S.; Paris, M. G. A. (2009-01-13). "Full Characterization of Gaussian Bipartite Entangled States by a Single Homodyne Detector". Physical Review Letters. 102 (2): 020502. arXiv: 0805.1993 . Bibcode:2009PhRvL.102b0502D. doi:10.1103/PhysRevLett.102.020502. ISSN   0031-9007. PMID   19257255. S2CID   21354226.
  7. Vogel, K.; Risken, H. (1989-09-01). "Determination of quasiprobability distributions in terms of probability distributions for the rotated quadrature phase". Physical Review A. 40 (5): 2847–2849. Bibcode:1989PhRvA..40.2847V. doi:10.1103/PhysRevA.40.2847. ISSN   0556-2791. PMID   9902492.
  8. Bradben. "Pauli measurements - Azure Quantum". docs.microsoft.com. Retrieved 2022-04-16.
  9. "Quantum State Tomography — Qiskit Experiments 0.2.0 documentation". qiskit.org. Retrieved 2022-04-11.
  10. "Quantum Tomography — Qiskit 0.36.0 documentation". qiskit.org. Retrieved 2022-04-11.
  11. Online Encyclopedia of Laser Physics and Technology. "Frequency Doubling". Archived from the original on 2016-06-03. Retrieved 2015-08-16.
  12. Lvovsky, A I (2004-06-01). "Iterative maximum-likelihood reconstruction in quantum homodyne tomography". Journal of Optics B: Quantum and Semiclassical Optics. 6 (6): S556–S559. arXiv: quant-ph/0311097 . Bibcode:2004JOptB...6S.556L. doi:10.1088/1464-4266/6/6/014. ISSN   1464-4266. S2CID   15890005.
  13. Řeháček, J.; Hradil, Z.; Ježek, M. (2001). "Iterative algorithm for reconstruction of entangled states". Phys. Rev. A. 63 (4): 040303. arXiv: quant-ph/0009093 . Bibcode:2001PhRvA..63d0303R. doi:10.1103/physreva.63.040303. S2CID   119482766.
  14. Gross, D.; Liu, Y. K.; Flammia, S.; Becker, S.; Eisert, J. (2010). "Quantum State Tomography via Compressed Sensing". Physical Review Letters. 105 (15): 150401. arXiv: 0909.3304 . Bibcode:2010PhRvL.105o0401G. doi:10.1103/PhysRevLett.105.150401. PMID   21230876. S2CID   19029700.
  15. Permutationally Invariant Quantum Tomography. "Pitomography". Archived from the original on 2016-03-04. Retrieved 2013-07-03.
  16. Tóth, G.; Wieczorek, W.; Gross, D.; Krischek, R.; Schwemmer, C.; Weinfurter, H. (2010). "Permutationally Invariant Quantum Tomography". Physical Review Letters. 105 (25): 250403. arXiv: 1005.3313 . Bibcode:2010PhRvL.105y0403T. doi:10.1103/PhysRevLett.105.250403. PMID   21231565. S2CID   21786571.
  17. Moroder, T.; Hyllus, P.; Tóth, G. Z.; Schwemmer, C.; Niggebaum, A.; Gaile, S.; Gühne, O.; Weinfurter, H. (2012). "Permutationally invariant state reconstruction". New Journal of Physics. 14 (10): 105001. arXiv: 1205.4941 . Bibcode:2012NJPh...14j5001M. doi:10.1088/1367-2630/14/10/105001. S2CID   73720137.
  18. Schwemmer, C.; Tóth, G. Z.; Niggebaum, A.; Moroder, T.; Gross, D.; Gühne, O.; Weinfurter, H. (2014). "Experimental Comparison of Efficient Tomography Schemes for a Six-Qubit State". Physical Review Letters. 113 (4): 040503. arXiv: 1401.7526 . Bibcode:2014PhRvL.113d0503S. doi:10.1103/PhysRevLett.113.040503. PMID   25105604. S2CID   26493608.
  19. D'Ariano, G. Mauro; Paris, Matteo G. A.; Sacchi, Massimiliano F. (2003-02-04). "Quantum Tomography". arXiv: quant-ph/0302028 .
  20. Chuang, Isaac L.; Nielsen, M. A. (1997). "Prescription for experimental determination of the dynamics of a quantum black box". Journal of Modern Optics. 44 (11–12): 2455–2467. arXiv: quant-ph/9610001 . Bibcode:1997JMOp...44.2455C. doi:10.1080/09500349708231894. ISSN   0950-0340. S2CID   119497365.
  21. 1 2 Altepeter, J. B.; Branning, D.; Jeffrey, E.; Wei, T. C.; Kwiat, P. G.; Thew, R. T.; O’Brien, J. L.; Nielsen, M. A.; White, A. G. (2003-05-15). "Ancilla-Assisted Quantum Process Tomography". Physical Review Letters. 90 (19): 193601. arXiv: quant-ph/0303038 . Bibcode:2003PhRvL..90s3601A. doi:10.1103/PhysRevLett.90.193601. ISSN   0031-9007. PMID   12785945. S2CID   15307742.
  22. 1 2 Mohseni, M.; Rezakhani, A. T.; Lidar, D. A. (2008-03-13). "Quantum-process tomography: Resource analysis of different strategies". Physical Review A. 77 (3): 032322. arXiv: quant-ph/0702131 . Bibcode:2008PhRvA..77c2322M. doi:10.1103/PhysRevA.77.032322. ISSN   1050-2947. S2CID   40376523.
  23. Shabani, A.; Kosut, R.; Mohseni, M.; Rabitz, H.; Broome, M.; Almeida, M.; Fedrizzi, A.; White, A. (2011). "Efficient Measurement of Quantum Dynamics via Compressive Sensing". Physical Review Letters. 106 (10): 100401. arXiv: 0910.5498 . Bibcode:2011PhRvL.106j0401S. doi:10.1103/PhysRevLett.106.100401. PMID   21469772. S2CID   5543694.