28
Quantum Computing Implemented via Optimal Control: Application to Spin and Pseudospin Systems

Thomas Schulte‐Herbrüggen1, Andreas Spörl2, Raimund Marx1, Navin Khaneja3 John Myers4, Amr Fahmy5, Samuel Lomonaco6, Louis Kauffman7 and Steffen Glaser1

1 Technische Universität München, Department Chemie, Lichtenbergstr. 4, 85747 Garching, Germany

2 Deutsches Zentrum für Luft‐ und Raumfahrt, German Aerospace Center, Münchener Str. 20, 82234 Wessling, Germany

3 Systems and Control Engineering, IIT Bombay, Powai 400076, India

4 Harvard University, Gordon McKay Laboratory, Division of Engineering and Applied Sciences, Cambridge, MA 02138, USA

5 Biological Chemistry and Molecular Pharmacology, Harvard Medical School, 240 Longwood Ave, Boston, MA 02115, USA

6 University of Maryland, Department of Computer Science and Electrical Engineering, Baltimore County, 1000 Hilltop Circle, Baltimore, MD 21250, USA

7 Department of Mathematics, Statistics and Computer Science, 851 South Morgan Street, University of Illinois at Chicago, Chicago, IL 60607‐7045, USA

28.1 Introduction

In this chapter, we discuss algorithmic and experimental aspects of quantum control of spin and pseudospin systems in view of realizing quantum algorithms or quantum simulations (1,2) at minimal cost, in particular in a minimum amount of time. For example, we will see that the time required for implementing a quantum module experimentally is a most natural measure of cost, whereas the number of standard elementary gates, that is, the network complexity, often does not allow for a simple one‐to‐one translation into the actual time complexity. Further typical cost functions may include relaxative losses or sensitivity to experimental imperfection. In view of future technologies, considerable recent progress of steering experimental quantum systems (3) is due to combining the tools of two mature research disciplines: (i) magnetic resonance (4) with its ample arsenal of methodology (5) for manipulating quantum systems and (ii) optimal control theory (6,7), nowadays an indispensable tool in system theory (8) and engineering (9). Optimal control can readily be extended to quantum systems (10) and has become a field of growing interest (1113).

Although the main source of examples presented is liquid‐state nuclear magnetic resonance (NMR), the techniques shown here are in no way confined to ensemble quantum computing but hold for single‐spin solid‐state quantum computing (14), electron spin resonance (ESR), and techniques beyond spin dynamics such as charge or flux qubits in a Josephson element (15). Methods of geometric control on Lie groups (16) apply to all quantum systems whose dynamics are governed by finite‐dimensional Lie algebras within the framework of spin or pseudospin systems (at least to sufficient approximation).

Quantum computational qubit systems may be implemented with particular convenience by nuclear spins‐1/2, since spin degrees of freedom are largely isolated from their environment. Moreover, the isotropic overall tumbling of the molecules in a liquid sample decouples the, say, images spins within each molecule from all their surrounding ones, and the spins can readily be represented by a density matrix. It carries the ensemble average over all molecules, each with images spins (4). Thus, the spin degrees of freedom span a Hilbert space of dimension images as desired in systems of images qubits.

It is the ease of experimental control and theoretical setting that gave NMR a head start in the experimental realization of fundamental concepts in quantum computing (1719).

As a liquid NMR sample contains an ensemble of many spin systems of the same kind, one can neither manipulate nor detect individual ones, thus precluding the preparation of pure states. Nevertheless, in order to use the usual quantum algorithms designed for initial conditions in terms of pure states, one may transcribe the density operator of a spin ensemble to a so‐called pseudopure state (18). Furthermore, ensemble‐averaged expectation values are detected rather than observables of individual spin systems. Hence, NMR quantum processors are examples of expectation‐value quantum computers (EVQC), where the outcome of a given quantum algorithm can be extracted from the resulting NMR spectra. Under mild conditions, spin‐1/2 systems are fully controllable in the sense that every unitary transform can be realized experimentally (vide infra), and thus universal sets of elementary quantum gates (20) can be put into practice. Then the basic steps of quantum algorithms can be implemented by spin‐selective radio‐frequency pulses (see Figure 28.1) and, for example, CNOT (controlled NOT) gates. In this manner, many algorithms were carried out with NMR experiments, such as the Deutsch–Jozsa algorithm for two (21,22), three (23), and five (24) qubits; variations of Grover's algorithm for two (25,26) and three qubits (27); the period‐finding algorithm for five qubits (28); and a pioneering version of Shor's algorithm for seven qubits (29). So NMR emulation of quantum computing has also been extensively reviewed, for example, in ( 3 3042); in turn, for the specific importance of optimal control in NMR, see also the reviews ( 3,43).

Scheme for cavity QED setup using circular Rydberg atoms.

Figure 28.1 Schematic representation of a liquid‐state NMR experiment. The sample comprises an ensemble of about images molecules in an external magnetic field of some 10–25 T. After preparing the density operator representing the nuclear spin ensemble of the molecules in an initial state, unitary transformations of a quantum algorithm are applied as a sequence of separate radio‐frequency pulses and delays or as shaped pulses. Finally, the outcome of the experiment is deduced from acquired NMR spectra.

28.2 From Controllable Spin Systems to Suitable Molecules

28.2.1 Reachability and Controllability

Neglecting decoherence, a quantum system is said to be fully controllable (4449) or operator controllable (50), if for any arbitrary initial state represented by its density operator images the entire unitary orbit images can be reached or, in equivalent terms of systems theory, if images is the reachability set to the initial state images . In systems of images qubits (e.g., spins‐1/2), this is the case under the following mild conditions ( 49,51,52): (i) the qubits have to be inequivalent, that is, distinguishable and selectively addressable and (ii) they have to be pairwise coupled (e.g., by Ising interactions), where the coupling topology may take the form of any connected graph. Of course, fully controllable qubit systems are equivalent to those in which at least one universal and all local quantum gates may be realized by admissible controls (53).

Controllability and reachability are important basic concepts in quantum systems theory. They are discussed on a much more general level in Chapter 27 of this book.

28.2.2 Molecular Hardware for Quantum Computation

From a chemical perspective, compounds with suitable spin systems require molecules with images coupled spins‐1/2. To this end, not all of the spins have to be mutually coupled, but they have to form a connected coupling topology, so there should be no working qubits without any coupling to the other ones. In particular for large qubit systems, linear spin chains with coupling topologies of nearest‐neighbor interactions (images ) are far more realistic than complete coupling topologies (images ) as shown in Figure 28.2. If controlled individually, arbitrary coupling terms can be used, such as combinations of isotropic and dipolar couplings.

Graphical illustration of Experimental quantum Rabi oscillation.

Figure 28.2 (a) Schematic representation of a system consisting of images mutually coupled spins‐1/2 (qubits). (b) Spin chains are sufficient for an images ‐qubit quantum computer.

In order to perform a large number of basic computational steps in a quantum algorithm, the time for each quantum gate must be considerably smaller than the relaxation time of the qubits. Moreover, it is highly desirable to strive for time‐optimal implementations of quantum algorithms or their modules in order to avoid unnecessary decoherence. The particular strength of optimal control for achieving this goal will be shown in Section 28.5.

Currently used sample preparations for liquid‐state NMR quantum computers result in nuclear spin relaxation times of up to several seconds. Characteristic spin–spin coupling constants are of the order of 10–images Hz, resulting in a typical duration of two‐qubit quantum gates between directly coupled spins of images s. Hence, sequences of up to images images two‐qubit quantum gates are feasible based on current liquid‐state NMR technology, and even more quantum gates may be possible by increasing the spin–spin coupling constants, for example, by using dipolar couplings in liquid crystalline media (54) and by further increasing the relaxation times. Compared to two‐qubit operations, single‐spin quantum gates such as NOT or Hadamard gates are very short. For example, in heteronuclear spin systems, typical single‐spin gate durations are of the order of images s. The minimum time required for a given single‐spin quantum gate not only depends on the maximum amplitude of radio‐frequency pulses but also on the smallest frequency difference of the nuclear spins in a given molecule (24).

Graphical illustration of single resonant atom prepares a coherent superposition of two large fields with different phases.

Figure 28.3 Schematic representation of BOC‐images images N‐images ‐glycine‐fluoride with the coupled five‐spin system (represented schematically by white arrows) that forms the molecular basis of a five‐qubit NMR quantum computer (24). The atoms that form the spin system of interest are shown as spheres. The rest of the molecule is shown in a stick representation.

For the first NMR quantum computers with up to three qubits, readily available compounds were used, such as 2,3‐dibromothiophene ( 17,55), images C‐chloroform (22), 2,3‐dibromopropanoic acid ( 22,56), and images ‐alanine (57). For the realization of the first five‐qubit NMR quantum computer, the compound BOC‐(images images N‐images ‐Gly)‐F was synthesized ( 24,58) (see Figure 28.3). If the deuterium spins are decoupled, the nuclear spins of images , images N, images , images (i.e., images ), and images F form a coupled spin system consisting of five spins‐1/2. The images coupling constants range between 13.5 Hz (images ) and 366 Hz (images ), and for a magnetic field of 9.4 T, the smallest frequency differences are 12 kHz (images ). Further synthetic five‐qubit systems are (i) a perfluorobutadienyl iron complex as entirely homonuclear spin system consisting of five coupled images F spins (28) and (ii) images images N‐diethyl‐(dimethylcarbamoyl)fluoromethylphosphonate as fully heteronuclear spin system allowing for fast selective pulses (59). A carbon‐labeled analog to (i) has been used as a seven‐qubit molecule for implementing a variant of Shor's algorithm (29). Another seven‐qubit molecule suggested for NMR quantum computing applications is images ‐crotonic acid (60). The design and synthesis of molecules with suitable spin systems for 10–20 qubits is not a trivial chemical challenge. An alternative way of realizing a molecular architecture with more than 10 coupled spins is the synthesis of polymers with a repetitive unit consisting of three or more spins (53). This approach is appealing because only a small number of resonances have to be addressed selectively. However, in such an architecture the implementation of quantum algorithms will require an additional overhead, which poses new challenges for the efficient implementation of quantum gates.

28.3 Scalability

28.3.1 Scaling Problem with Pseudopure States

The density operator of a spin system at thermal equilibrium is proportional to images , where images is the spin Hamiltonian of the images ‐spin molecule used as a quantum register, images is Boltzmann's constant, and images is the temperature. As the usual magnetic fields in NMR are of the order of 10 T, the so‐called high‐temperature approximation is valid above temperatures of some 10 mK, and the thermal density operator can be given by the first two terms in the Taylor expansion

28.1 equation

where images is the angular frequency of the images th nucleus and images is defined by a tensor product over all images spins in which all the factors are unit operators except for images as the images th factor of the tensor product.

Although highly mixed, this state can be transformed into a so‐called pseudopure state, ( 17, 18,61) resulting in an initial density operator of the form

28.2 equation

for some (usually small) coefficient images . With the identity operator images being invariant under any similarity transform and all spin observables being traceless, pseudopure states form handy starting points for NMR implementations of quantum algorithms. However, this convenience comes at a high cost: the coefficient images decreases exponentially with the number of qubits images (62). Hence, the spectroscopic signal decreases as well, and severe signal‐to‐noise problems are expected for experiments with more than about 10 qubits.

The exponential signal loss is often thought to impose a fundamental limit on the scalability of ensemble NMR (63), although hyperpolarization techniques (6466) come very close to pure states. Fortunately, the purity problem can also be circumvented by several approaches avoiding pseudopure states altogether.

28.3.2 Scalable Quantum Computing on Thermal Ensembles

One attractive approach avoiding pure or pseudopure states altogether is to design ensemble quantum computing algorithms based on the thermal density operator instead of a pure state, an early example being a scalable version of the Deutsch–Jozsa algorithm (67,68). At the expense of an extra qubit and a modified oracle, balanced functions can be distinguished from constant ones using an initial state obtained merely by a hard images images ‐pulse applied to the thermal state. This requires neither pseudopure states of Eq. 28.2 nor temporal averaging. Let images denote the number of levels in an images ‐qubit system. Then, for an Oracle of a function images , one implements a substitute images for images instead of images . To this end, relate images to images by

equation

Given the Oracle images , a scalable NMR quantum computer can readily discriminate balanced functions from constant ones. Note that resolving the output spectra (67) does not build upon any demands growing exponentially with the number of qubits. For example, for the constant function images and the balanced function images , the scalable version of the Deutsch–Jozsa algorithm requires an additional qubit (images ) and the implementation of images for images images and of images images . For the five‐qubit system BOC‐(images images N‐images ‐Gly)‐F ( 24, 58), the resulting spectra (68) of images are shown in Figure 28.4a,b for images and images , respectively. Constant and balanced functions can be easily distinguished by the presence or absence of the signals.

Scheme for Experimental setup of an optical cavity QED experiment.

Figure 28.4 Experimental spectra (68) representing the result of the new version of the Deutsch–Jozsa algorithm (67) based on the thermal density operator for a constant (a) and a balanced (b) test function.

(Panel (a): Fahmy et al. 2008 (68). Copyright 2008, American Physical Society. Panel (b): Myers et al. 2001 (67). Copyright 2001, American Physical Society.)

It is important to note that for this version of the algorithm, the number of molecules in the ensemble does not have to increase exponentially with the number of qubits images within the molecule. These favorable scaling properties are at variance to a previous alternative ensemble implementation of the Deutsch–Jozsa algorithm (69,70).

A more recent approach is to perform quantum algorithms designed to work on ensembles as addressed in the worked example of Section 28.6: It is DQC1‐type algorithm to classify knots by topological invariants inferred from measuring spin ensembles.

However, with an increasing number of qubits, not only the synthetic requirements grow, but also the control demands with respect to NMR instruments and pulse‐sequence design, in order to cope or circumvent experimental imperfections (such as rf inhomogeneity) or relaxation. This asks for optimal control methods (3) at large.

28.4 Algorithmic Platform for Quantum Control Systems

In practice, quantum control problems amount to steering a dynamic system such as to maximize a given figure of merit subject to the constraint of following a given equation of motion. In (finite‐dimensional) quantum dynamics, the pertinent equations of motion are typically linear both in the state as well as in the control terms, and dynamic systems of this form are known as bilinear control systems ( 9,71,72)

equation

with “state” images , drift images , controls images , and control amplitudes images thus defining the images as effective generators. This setting captures all of the following important scenarios:

  1. controlled Schrödinger equation01
    equation
  2. quantum gate for closed system
    equation
  3. quantum state in open quantum system
    equation
  4. quantum map of open quantum system
    equation

Moreover, the quality function may be expressed via the scalar product as the overlap between the final state (or operator) of the controlled system at time images and the target state so that the common options amount to

equation

Define the boundary conditions as images , images and fix the total time images . For simplicity, we henceforth assume equal discretized time spacing images for all timeslices images . So images . Then, the total generator (i.e., Hamiltonian images or Lindbladian images ) governing the evolution in the time interval images shall be labeled by its final time images as

equation

which governs the controlled time evolution in the timeslice images . Then, gradient‐based and second‐order optimal control algorithms such as GRAPE (7375) or KROTOV type ( 11 7679) proceed in the following basic steps entering the unified modular platform DYNAMO (74) described in more detail below.

  1. initialize with a random (or guessed) control vector (pulse sequence) consisting of the piecewise‐constant control amplitudes images ;
  2. exponentiate images for all images with images ;
  3. calculate forward‐propagation 1 images ;
  4. calculate back‐propagation images
  5. evaluate fidelity say images , where images ;
  6. evaluate gradients for all images : with images of Eq. 28.3 or 28.4 below and images images ;
  7. update amplitudes for all images , for example, by images for a quasi‐Newton second‐order increment, where images is a suitable step size and images is (e.g., an LBFGS‐approximation to) the inverse Hessian;
  8. reiterate up to terminal condition (e.g., images for all images ).
Graphical illustration of Photon statistics of a deterministic single-photon source.

Figure 28.5 Overview on the update schemes of gradient‐based optimal control algorithms unified in the DYNAMO platform. They all turn initial guesses for pulse shapes (i.e., piece‐wise constant control amplitudes) into optimized shapes. In GRAPE (a) all the timeslices are updated concurrently. In contrast, sequential update schemes of KROTOV type (b) update a single timeslice. Hybrid versions (c) can be implemented such as to update a subset of different timeslices before moving to the next (disjoint) set of timeslices. Optimizations may take total time, power, robustness, smoothness, or excitation bandwidth into account and may be executed for closed systems or open systems with known relaxation parameters.

Here, the exact derivative in closed systems (or unital open systems characterized by their normal Lindblad generators) can be read element‐wise from the eigendecomposition (with eigenvectors images to the eigenvalues images )

28.3 equation

while in nonunital open systems other methods apply like

28.4 equation

as long as the digitization by images is sufficient to satisfy images , or one will have to resort to finite‐differences, and so on (see ( 73 75)). This scheme covers all the optimal control problems specified earlier.

Recently, we have provided a unified MATLAB‐based programming frame DYNAMO (74) designed in a modular way such that to the above set of bilinear control problems it embraces the different algorithmic approaches known in the literature and shown in Figure 28.5. While the GRAPE algorithm (gradient‐assisted pulse engineering) (73) updates all timeslices in the pulse sequence (control vector) concurrently, another type of well‐established algorithms of KROTOV type ( 11 76 79) do so sequentially. It has turned out that for optimizing unitary gate synthesis for quantum information, concurrent updates of GRAPE type overtake sequential algorithms of KROTOV type well before reaching qualities in the order of the error‐correction threshold, while for spectroscopy purposes lower fidelities that KROTOV may reach faster often suffice. This is due to the fact that the recursive scheme (BFGS) to approximate the inverse Hessian pays when a constant set of time slices is updated as in GRAPE, while sequential updates preclude full profit from such recursions for second‐order methods, and their first‐order variants naturally loose power in the vicinity of critical points. In DYNAMO, one may easily change between different schemes on the fly during an optimization run, whenever needed to save computation time. Moreover, DYNAMO can readily be kept state of the art with respect to future developments such as, for example, improved preconditioning, further Newton‐type algorithms, or including incoherent degrees of freedom as control parameters.

28.5 Applied Quantum Control

Although one can decompose any quantum computing algorithm into a series of single‐spin operations and two‐spin gates between directly coupled spins, some fundamental questions remain: they are of both theoretical and practical interest. What is the minimum time required to realize a given unitary transformation in a given coupling topology of a spin system with a required fidelity? Which controls (pulse sequences) achieve the task in minimal time?

In addition to numerical approaches ( 73, 74), where by repeating controlled state transfer with decreasing final times images up to a minimal time images still allowing to get full coherence transfer (see ( 1012, 73, 80)) optimal control theory has also provided analytical approaches and solutions for time‐optimal quantum transfers and gates (8191).

Moreover, one may characterize time‐optimal pulse sequences algebraically by geometric optimal control (16) showing that the problem reduces to finding geodesics (i.e., shortest paths) between cosets (81), as will be demonstrated in Section 28.5.1.3.

But before that we focus on quantum gate control ( 10, 73 9294).

28.5.1 Regime of Fast Local Controls: the NMR Limit

Firstly, we choose the limit of fast local controls (by strong pulses), the timescale of which can safely be neglected as compared to the time‐limiting coupling interactions of the Ising type. Not only is this regime typical of NMR with weak scalar couplings, it also lends itself for a theoretical understanding in Lie‐algebraic terms. Here, the images ‐couplings are assumed to be uniform in the following examples, thus allowing to express the time required in units of images . However, the numerical algorithms are of course general and can cope with coupling types and strengths directly matching the experimental settings, and even finite durations for local controls can be dealt with as shown in Section 28.5.2.

28.5.1.1 The Quantum Fourier Transform

The quantum Fourier transform (QFT) is in the core of all quantum algorithms of Abelian hidden subgroup type (95,96) such as, for example, the algorithms of Deutsch–Jozsa's, Simon's, and Shor's. In order to speed up quantum modules and minimize decoherence, the QFT should be implemented in the fastest way. Clearly, the time required for realizing the QFT in images ‐qubit systems depends on the coupling topology and the interaction type and strength of the pertinent experimental setting. Figure 28.6 demonstrates how in linear spin chains (images ) with the nearest‐neighbor Ising interactions, numerical time‐optimal control provides a decomposition of the QFT that is much faster than the corresponding decomposition into standard gates would impose: in six qubits, for instance, the speedup is more than eightfold and in seven qubits approximately ninefold.

Graphical illustration of Time-resolved quantum-beat experiment with two single-photon pulses impinging simultaneously on a beam splitter.

Figure 28.6 The QFT in linear coupling topologies images : (a) gate complexity by standard‐gate decomposition (images ) (97) and optimized scalable gate decomposition (images ) (98) and (b) time complexity of the QFT. Upper traces give analytical times associated with the decompositions of part (a): standard‐gate decompositions (images ) (97) and optimized scalable gate decompositions (images ) (98); (images ) gives a special (i.e., nonscalable) five‐qubit decomposition into standard gates obtained by simulated annealing (98). Lowest trace: fastest realizations (94) currently obtained by numerical time‐optimal control (rounded to images ) giving trace fidelities images (images ) and images for the last point.

28.5.1.2 Multiply Controlled NOTs

Analogously the images NOT‐gate can be decomposed in a time‐optimized way. Interestingly, in a complete coupling topology of images qubits, the algorithmic complexity was described by Barenco et al. (99) as growing exponentially up to six qubits, whereas the increase from seven qubits onward was said to be quadratic. Again, time‐optimal control provides a dramatic speedup in this case, see Figure 28.7.

Plots for Cn-1NOT gate on complete coupling topologies Kn.

Figure 28.7 The images NOT gate on complete coupling topologies images : (a) network complexity (99) and (b) time complexity. Upper trace: analytical times for decomposition into standard gates (images ) (99). Lowest trace: fastest realizations (94) currently obtained by numerical time‐optimal control (rounded to images ) giving trace fidelities images (images ) and images for the last point.

28.5.1.3 Geometry of Time‐Optimal Gates

In NMR, the markedly different timescales for fast local controls (pulses) versus slow coupling evolutions lend themselves for making use of Cartan decomposition of real semisimple Lie algebras images (where images ). The goal of time‐optimal realizations then reduces to finding constrained shortest paths in the cosets images . For images spins‐1/2, images , images , and the coset images takes the form of a Riemannian symmetric space. Thus, time‐optimal trajectories between points in images correspond to Riemannian geodesics. For images , the cosets images are no longer Riemannian symmetric spaces, so the time‐optimal trajectories in images denote sub‐Riemannian geodesics.

Yet, in the sub‐Riemannian geometry of three spins, there are examples that can be fully understood: for example, the time‐optimal simulation of three‐spin‐interaction Hamiltonians of the form images where images , images , images can be images , images or images and images is an effective trilinear coupling constant. We considered a linear coupling topology consisting of a chain of three heteronuclear spins‐1/2 with the coupling constants images , images and the coupling term images Here, the time‐optimal realization of the trilinear coupling term images , see Figure 28.8, can be derived by means of geometric control (100).

Image described by caption and surrounding text.

Figure 28.8 Time‐optimal pulse sequences for synthesizing the propagator images with images . (a) In the scheme for time‐optimal (geodesic) pulse sequence derived on algebraic grounds, the radio‐frequency amplitude images of the hatched pulse is images (100). (b) Pulse sequence found by numerical optimal control for images . Only the images channel is used, amplitudes on the other spins are of negligible amplitude, thus matching the theoretically predicted controls in (a).

Compared to conventional approaches (101), the time‐optimal synthesis of effective three‐spin propagators of the form images has a duration of only images (100) compared to the duration images of conventional implementations (101) and hence provides significant time savings as shown in Figure 28.9. Further results on geometric control can be found in ( 81 91).

Image described by caption and surrounding text.

Figure 28.9 (a) Times required for simulating the trilinear coupling Hamiltonian images in the conventional and the time‐optimal way (100). In the lower trace, the solid line (‐) is calculated on algebraic grounds, while numerical results (73) are inserted point‐wise: full circles (images ) represent times achieving full trace fidelity, while empty ones (images ) denote times giving trace fidelities between images and images . The time resolution for the numerical calculations was images . (b) The speedup factors are most prominent for small flip angles images .

28.5.2 Regime of Finite Local Controls: Beyond NMR

28.5.2.1 CNOT and TOFFOLI Gates for Charge Qubits

Clearly the optimal control methods presented thus far can be generalized such as to hold for systems with finite times for local controls as long as one has finite degrees of freedom allowing for a pseudospin formulation in terms of closed Lie algebras. Suffice it to mention that the standard CNOT gate can be realized in two coupled charge qubits of a solid‐state Josephson device some five times faster than in the pioneering setting of Nakamura (102). One easily obtains (103) a trace fidelity beyond 0.99999. With the same fidelities one finds realizations of the TOFFOLI‐gate in three linearly coupled charge qubits that are some nine times faster than by standard‐gate decomposition and approximately 13 times faster than one would infer from the CNOT in Ref. (102).

28.6 Worked Example: Unitary Controls for Classifying Knots by NMR

Many of the well‐established quantum algorithms operate by solving the hidden subgroup problem in an efficient way (104,105). Moreover they do so by resorting to the circuit model with its experimentally challenging accuracy demands (error‐correction threshold). In search for different and more robust classes of quantum algorithms, topological quantum computing with anyonic quasi‐particles brought up relations to braid groups (106108). This is because anyonic world lines in a three‐dimensional model of spacetime (comprising two spatial and one temporal dimension) form braids that can be exploited as quantum gates. These gates have the power of the circuit model with the advantage of being more robust. When establishing the relation between topological and ordinary quantum computation, it turned out that unitary representations of braid groups useful for anyonic topological quantum computing can also be used to compute invariants of knots and links such as the Jones polynomial.

Thus, there is a fruitful interplay between topological and circuit‐based algorithms mediated via braid groups of knots, that is, by unitary representations of the braid operations. In order to implement these unitaries experimentally, optimal control is pertinent again.

Resuming (109,110), in this section we illustrate how thermal ensembles can be used for approximating the trace of a unitary matrix (68) in order to classify knots by their topologies. This paves the way to a recent class of quantum algorithms related to knot theory, because it allows for efficiently evaluating Jones polynomials over a range of parameters. Since knots with different Jones polynomials are clearly inequivalent (while the converse does not hold), efficient quantum algorithms determining the trace of unitaries can be of great help (in the cases distinguishable by the Jones polynomials) to solve the classically NP‐hard decision problem whether two knots are equivalent in the sense they can be transformed into one another by using only Reidemeister moves and trivial moves, that is, those which do not change the number of crossings.

More precisely, while a knot is defined as an embedding of the circle in three‐space up to ambient isotopy, a link is an analogous embedding of several disjoint circles again up to isotopy. Now a knot invariant is any function that remains invariant under Reidemeister (and trivial) moves mentioned already. The Jones polynomial is a special form of Laurent polynomial (i.e., a polynomial with terms of both positive and negative degrees forming a ring) which itself has a degree that grows at most linearly with the number of crossings in the corresponding link. Note there is an important relation to braid groups established by Alexander's theorem. It says that any link can be constructed as a plat closure of some braid, namely by moving “return” strands back into the braid, see, for example, Ref. (111) for details.

Now the algorithm of Aharonov et al. (107,112,113) takes the trace of some unitary representation of the corresponding braid group to give the Jones polynomial. Here the braid group with images strands, images , is generated by its images generators representing right‐handed twists images . For evaluating the trace, it is most convenient to exploit the connection to the Temperley‐Lieb algebra (114,115) and its unitary representation images by

equation

where images is of modulus 1 and images is real symmetric, while images is the generator of the braid group associated to the knot of interest.02

Next, we focus on the three‐stranded braid group images generated by the elements images . It comprises the well‐known standard knots Trefoil (up to addition of a circle disjoint from the knot), Figure‐Eight, and the Borromean Rings shown in Figure 28.10.

Scheme for Standard knots that relate to the braid group with three strands B3.

Figure 28.10 Standard knots and links that relate to the braid group with three strands images . (a) The Trefoil knot can be represented by the braid group element images , (b) the Figure‐Eight knot by images , and (c) the Borromean rings by images .

In the unitary (path model) representation of images one ends up with the following unitaries that contain images (related to the variable images of the bracket and Jones polynomial).

equation

Now, in order to get hold of the trace of images by a quantum measurement, we follow Ref. (68) and enlarge the quantum register by one ancilla qubit. Then, the unitary images is translated into a controlled unitary with respect to the ancilla in the sense

equation

Based on the thermal ensemble state images with images , it is routine (here on the molecule chloroform by images H saturation followed by gradient filters) to prepare the suitable initial state images with the images ‐magnetization on the natural abundance images C used as qubit. With these stipulations it is easy to proceed in three final steps

  1. prepare images .
  2. let images evolve under the signature sequence images (vide infra) of images 's specific to the knot in question. This gives the total images .
  3. measure the expectation value of the phase‐sensitive ensemble detection operator03 images as to give images .

In simple cases it is well known how to translate unitary operators into NMR pulse sequences. In the more intricate case here, similar recipes apply: by combining algebraic approaches with numerical ones (e.g., by GRAPE), one arrives at the pulse sequences shown in Figure 28.11, which are specifically designed to continuously depend on the variable images via

equation

so that they can be implemented over a range of values of images .

Image described by caption and surrounding text.

Figure 28.11 NMR pulse sequences implementing the set of controlled unitaries images corresponding to the generators of the three‐strand braid group images encapsulating the Trefoil knot, the Figure‐Eight knot, and the Borromean rings.

Now, for the Trefoil knot the NMR pulse sequence for images has to be applied thrice images , while for the Figure‐Eight knot it is images and for the Borromean Rings images to be read from right to left to give the respective images and images . As shown in Figure 28.12, the Jones polynomial was experimentally evaluated for each knot or link at 31 values of images distributed over a continuous part of the domain accessible by the quantum algorithm. This approach (110) readily discriminates the three‐stranded knots or links by two qubits, while in Ref. (116) only single values of images were used. Note that the experimental data nicely follow the theoretical prediction, and the functional dependence is so different that the predictive power of distinguishing knots or links is higher than by mere evaluation of single points.

Image described by caption and surrounding text.

Figure 28.12 Experimental results (110) with real and imaginary parts of images , from whence the Jones polynomial of the various knots are calculated as functions of images . For the Trefoil, data are given in (a), for the Figure Eight in (b), and for the Borromean rings in (c). The respective traces compare experimental results, theoretical predictions, and simulated experiments, where realistic imperfections like relaxation, images ‐field inhomogeneity, and finite length of the pulses are included.

Yet both experimental demonstrations include an evaluation of the Jones polynomial at a root of unity and thus implement a DQC1‐complete quantum algorithm (see (117)). In Ref. (116), only links that contain disjoint circles were evaluated. As already mentioned, a much simpler quantum calculation using fewer qubits (here two qubits for a two‐strand braid representation) can calculate the Jones polynomials of the given links equally well. In contrast, the evaluations for the Figure‐Eight knot and the Borromean rings cannot do with fewer than three strands and two qubits as shown in Ref. (109).

Even moderately intricate molecular hardware with several qubits and realistic coupling topologies goes beyond pulse sequences as easy as in Fig. 28.11 for the two‐qubit molecule chloroform. Already the four‐carbon architecture used in (116) required the GRAPE algorithm to be implemented experimentally. Hence, control algorithms will play a role for future algorithms inspired by topological quantum computation.

28.7 Conclusions

Apart from low temperatures, hyperpolarization techniques, or in situ reactions with para‐hydrogen (where ensemble states of high purity can be obtained), even thermal ensemble states may be used for NMR implementations of quantum algorithms, which are in principle scalable (118). Beyond the early example of an ensemble variant to the Deutsch–Jozsa algorithm (63), we discussed a DQC1 quantum algorithm ( 109, 116, 117) to classify knots by their topological invariants (Jones polynomials) easily read from spin ensembles.

28.7.1 Optimal Control as a Quantum CISC Compiler

On a general scale, exploiting quantum control is not only for constructing the standard restricted instruction set of local unitaries and two‐qubit universal unitary gates (in the sense of a quantum RISC compiler), but may also readily go beyond to the complex instruction set of many‐qubit unitaries, from which entire algorithms may recursively be assembled. In this sense, quantum control lends itself as a quantum CISC compiler (119121).

Concomitantly, the algorithmic network complexity counting standard RISC building blocks can be complemented by the more realistic time complexity (i.e., the duration of the time‐optimal CISC gates) as the experimentally relevant cost: it allows for exploiting the continuous differential geometry of the unitary Lie‐groups as well as the power of quantum control for getting constructive upper bounds to the time complexity by (numerical approximations to) time‐optimal controls perfectly matching the experimental setting (94).

28.7.2 Generalization and Further Applications

Geometric and optimal quantum control are most powerful tools for optimizing experimental implementations of quantum computing, whenever the quantum degrees of freedom can be described in closed Lie‐algebraic form. This means that the quantum system in question can be treated as a spin or pseudospin system. Important applications include cavity QED (122), trapped ions (123), superconducting and Josephson devices (103), NV centers in diamond (124126), or the Jaynes–Cummings model of atoms in a cavity (127).

Recent extensions of numerical control algorithms include CPU‐node‐ (128) and time‐parallelized versions (129), algorithms for discrete‐valued controls (130), and tracking algorithms, for example, for decoupling sequences (131,132). Further applications of quantum control extend to pattern recognition by quantum neural networks (133).

Therefore, we anticipate that the tools sketched here await broad application.

Acknowledgments

This work was supported in part by the integrated EU‐programme SIQS as well as by DFG (Deutsche Forschungsgemeinschaft) in the collaborative research centre SFB 631 on solid‐state‐based quantum computation.

Exercises

  1. 28.1 Spin Polarization
    1. Set images . Verify for a single spin‐1/2 that at images K and a magnetic field images T no more than roughly 1 in images spins of an ensemble contributes to the spin polarization. (Hint: use images , where images and images MHz images , images J images , images J s.)
    2. Show that the “high‐temperature” approximation 28.1 holds for images mK. Give the population difference images at images and 300 K.
  2. 28.2 von Neumann Entropy of Spin Ensembles

    Using results from Exercise 1, show that at almost all practical temperatures

    1. in von Neumann's entropy images one finds images ;
    2. for images near images , the relative entropy becomes images ;
    3. images for both images near images .

    (Hint: use images to first order and set images for the expansion to hold even for eigenvalues images ).

  3. 28.3 Unitary Equivalence versus Entropy Conservation

    Show that for two density operators to be unitarily similar, conservation of von Neumann's entropy is a necessary but not a sufficient condition.

    (Hint: use the moments images and the series of Exercise 2.) Give a sufficient condition.

References

  1. 1 Feynman, R.P. (1982) Int. J. Theor. Phys., 21, 467.
  2. 2 Feynman, R.P. (1996) Feynman Lectures on Computation, Perseus Books, Reading, MA.
  3. 3 Glaser, S.J., Boscain, U., Calarco, T., Koch, C., Köckenberger, W., Kosloff, R., Kuprov, I., Luy, B., Schirmer, S., Schulte‐Herbrüggen, T., Sugny, D., and Wilhelm, F.K. (2015) Eur. Phys. J. D, 69, 279.
  4. 4 Ernst, R.R., Bodenhausen, G., and Wokaun, A. (1987) Principles of Nuclear Magnetic Resonance in One and Two Dimensions, Clarendon Press, Oxford.
  5. 5 Grant, D.M. and Harris, R.K. (eds) (1996) Encyclopedia of Nuclear Magnetic Resonance, Vols I–IX, John Wiley & Sons, Ltd, Chichester.
  6. 6 Pontryagin, L.S., Bol'tanskii, V.G., Gamkrelidze, R.S., and Mischenko, E.F. (1964) The Mathematical Theory of Optimal Processes, Pergamon Press, New York.
  7. 7 Lee, E.B. and Markus, L. (1967) Foundations of Optimal Control Theory, John Wiley & Sons, Inc., New York.
  8. 8 Sastry, S. (1999) Non‐Linear Systems: Analysis, Stability and Control, Springer, New York.
  9. 9 Levine, W.S. (ed.) (1996) The Control Handbook, CRC Press, Boca Raton, FL.
  10. 10 Butkovskiy, A.G. and Samoilenko, Yu.I. (1990) Control of Quantum‐Mechanical Processes and Systems, Kluwer, Dordrecht.
  11. 11 Tannor, D.J. and Rice, S.A. (1985) J. Chem. Phys., 83, 5013.
  12. 12 Shi, S. and Rabitz, H. (1989) J. Chem. Phys., 92, 364.
  13. 13 [special volume on optimal control in chemical physics] (2001) Chem. Phys., 267.
  14. 14 Kane, B.E. (1998) Nature (London), 393, 133.
  15. 15 Makhlin, Yu., Schön, G., and Shnirman, A. (2001) Rev. Mod. Phys., 73, 357.
  16. 16 Jurdjevic, V. (1997) Geometric Control Theory, Cambridge University Press, Cambridge.
  17. 17 Cory, D.G., Fahmy, A.F., and Havel, T.F. (1996) Proceedings of the 4th Workshop on Physics and Computation, New England Complex Systems Institute, Boston, MA, pp. 87–91.
  18. 18 Cory, D.G., Fahmy, A.F., and Havel, T.F. (1997) Proc. Natl. Acad. Sci. U.S.A., 94, 1634.
  19. 19 Gershenfeld, N.A. and Chuang, I.L. (1997) Science, 275, 350.
  20. 20 Deutsch, D. (1985) Proc. R. Soc. London, Ser. A, 400, 97.
  21. 21 Jones, J.A. and Mosca, M. (1998) J. Chem. Phys., 109, 1648.
  22. 22 Chuang, I.L., Vandersypen, L.M.K., Zhou, X.L., Leung, D.W., and Lloyd, S. (1998) Nature (London), 393, 143.
  23. 23 Linden, N., Barjat, H., and Freeman, R. (1998) Chem. Phys. Lett., 296, 61.
  24. 24 Marx, R., Fahmy, A.F., Myers, J.M., Bermel, W., and Glaser, S.J. (2000) Phys. Rev. A, 62, 012310.
  25. 25 Chuang, I.L., Gershenfeld, N., and Kubinec, M.G. (1998) Phys. Rev. Lett., 80, 3408.
  26. 26 Jones, J.A., Mosca, M., and Hansen, R.H. (1998) Nature (London), 393, 344.
  27. 27 Vandersypen, L.M.K., Steffen, M., Sherwood, M.H., Yannoni, C.S., Breyta, G., and Chuang, I.L. (2000) Appl. Phys. Lett., 76, 468‐887.
  28. 28 Vandersypen, L.M.K., Steffen, M., Breyta, G., Yannoni, C.S., Cleve, R., and Chuang, I.L. (2000) Phys. Rev. Lett., 85, 5452.
  29. 29 Vandersypen, L.M.K., Steffen, M., Breyta, G., Yannoni, C.S., Sherwood, M.H., and Chuang, I.L. (2001) Nature (London), 414, 883‐887.
  30. 30 Mehring, M. (1999) Appl. Magn. Reson., 17, 141.
  31. 31 Jones, J.A. (2000) Fortschr. Phys., 48, 909‐924.
  32. 32 Cory, D.G., Laflamme, R., Knill, E., Viola, L., Havel, T.F., Boulant, N., Boutis, G., Fortunato, E., Lloyd, S., Martinez, R., Negrevergne, C., Pravia, M., Sharf, Y., Teklemariam, G., Weinstein, Y.S., and Zurek, W.H. (2000) Fortschr. Phys., 48, 875‐907.
  33. 33 Havel, T.F., Somaroo, S.S., Tseng, C.H., and Cory, D.G. (2000) Appl. Algebra Eng. Commun., 10, 339‐374.
  34. 34 Jones, J.A. (2001) Prog. NMR Spectrosc., 38, 325‐360.
  35. 35 Jones, J.A. (2001) Phys. Chem. Commun., 11, 11.
  36. 36 Glaser, S.J. (2001) Angew. Chem., 113, 151‐153; (2001) Angew. Chem. Int. Ed., 40, 147‐149.
  37. 37 Vandersypen, L.M.K., Chuang, I.L., and Suter, D. (2010) Liquid‐state NMR quantum computing, in Encyclopedia of Nuclear Magnetic Resonance (eds D.M. Grant and R.K. Harris), John Wiley & Sons, Ltd, Chichester.
  38. 38 Havel, T.F., Cory, D.G., Lloyd, S., Boulant, N., Fortunato, E.M., Pravia, M.A., Teklemariam, G., Weinstein, Y.S., Bhattacharyya, A., and Hou, J. (2002) Am. J. Phys., 70, 345‐362.
  39. 39 Ramanathan, C., Boulant, N., Chen, Z., Cory, D.L., Chuang, I.L., and Steffen, M. (2004) Quantum Inf. Process., 3, 15.
  40. 40 Vandersypen, L.M.K. and Chuang, I.L. (2004) Rev. Mod. Phys., 76, 1037.
  41. 41 Jones, J.A. (2011) Prog. NMR Spectrosc., 59, 91.
  42. 42 Takui, T., Berliner, L., and Hanson, G. (eds) (2016) Electron Spin Resonance (ESR) Based Quantum Computing, Springer, New York.
  43. 43 Nielsen, N.C., Kehlet, C., Glaser, S.J., and Khaneja, N. (2010) Optimal control methods in NMR spectroscopy, in Encyclopedia of Nuclear Magnetic Resonance (eds D.M. Grant and R.K. Harris), John Wiley & Sons, Ltd, Chichester.
  44. 44 Sussmann, H. and Jurdjevic, V. (1972) J. Differ. Equ., 12, 95.
  45. 45 Jurdjevic, V. and Sussmann, H. (1972) J. Differ. Equ., 12, 313.
  46. 46 Brockett, R.W. (1972) SIAM J. Control, 10, 265.
  47. 47 Brockett, R.W. (1973) SIAM J. Appl. Math., 25, 213.
  48. 48 Boothby, W.M. and Wilson, E.N. (1979) SIAM J. Control Optim., 17, 212.
  49. 49 Schulte‐Herbrüggen, T. (1998) Aspects and prospects of high‐resolution NMR. PhD thesis. Diss‐ETH 12752, Zürich,
  50. 50 Albertini, F. and D'Alessandro, D. (2003) IEEE Trans. Autom. Control, 48, 1399.
  51. 51 Glaser, S.J., Schulte‐Herbrüggen, T., Sieveking, M., Schedletzky, O., Nielsen, N.C., Sørensen, O.W., and Griesinger, C. (1998) Science, 280, 421.
  52. 52 Zeier, R. and Schulte‐Herbrüggen, T. (2011) J. Math. Phys., 52, 113510.
  53. 53 Lloyd, S. (1993) Science, 261, 1569.
  54. 54 Yannoni, C.S., Sherwood, M.H., Miller, D.C., Chuang, I.L., Vandersypen, L.M.K., and Kubinec, M.G. (1999) Appl. Phys. Lett., 75, 3563.
  55. 55 Cory, D.G., Price, M.D., and Havel, T.F. (1998) Physica D, 120, 82‐101.
  56. 56 Mádi, Z.L., Brüschweiler, R., and Ernst, R.R. (1998) J. Chem. Phys., 109, 10603‐10611.
  57. 57 Cory, D.G., Mass, W., Price, M., Knill, E., Laflamme, R., Zurek, W.H., Havel, T.F., and Somaroo, S.S. (1998) Phys. Rev. Lett., 81, 2152‐2155.
  58. 58 Marx, R., Fahmy, A.F., Myers, J.M., Bermel, W., and Glaser, S.J. (2000) in Quantum Computing, Proceedings of SPIE, vol. 4047 (eds E. Donkor and A.R. Pirich), SPIE, pp. 131‐138.
  59. 59 Marx, R., Pomplun, N., Bermel, W., Zeiger, H., Engelke, F., Fahmy, A.F., and Glaser, S.J. (2015) Magn. Reson. Chem., 53, 442.
  60. 60 Knill, E., Laflamme, R., Martinez, R., and Tseng, C.H. (2000) Nature (London), 404, 368.
  61. 61 Wei, D., Chang, Y., Glaser, S.J., and Yang, X. (2014) Appl. Phys. Lett., 104, 242409.
  62. 62 Warren, W.S. (1997) Science, 277, 1688.
  63. 63 Braunstein, S.L., Caves, C.M., Jozsa, R., Linden, N., Popescu, S., and Schack, R. (1999) Phys. Rev. Lett., 83, 1054‐1057.
  64. 64 Hübler, P., Bargon, J., and Glaser, S.J. (2000) J. Chem. Phys., 113, 2056.
  65. 65 Verhulst, A.S., Liivak, O., Sherwood, M., Vieth, H.M., and Chuang, I.L. (2001) Appl. Phys. Lett., 79, 2480.
  66. 66 Griesinger, C., Bennati, M., Vieth, H.M., Luchinat, C., Parigi, G., Höfer, P., Engelke, F., Glaser, S.J., Denysenkov, V., and Prisner, T.F. (2012) Prog. NMR Spectrosc., 64, 4.
  67. 67 Myers, J.M., Fahmy, A.F., Glaser, S.J., and Marx, R. (2001) Phys. Rev. A, 63, 032302.
  68. 68 Fahmy, A.F., Marx, R., Bermel, W., and Glaser, S.J. (2008) Phys. Rev. A, 78, 022317.
  69. 69 Woodward, F.M. and Brüschweiler, R. (2000) Solution of the Deutsch‐Josza Problem by NMR Ensemble Computing without Sensitivity Scaling. quant‐ph/0006024.
  70. 70 Arvind and Collins, D. (2003) Phys. Rev. A, 68, 052301.
  71. 71 Sontag, E. (1998) Mathematical Control Theory, Springer, New York.
  72. 72 Elliott, D. (2009) Bilinear Control Systems: Matrices in Action, Springer, London.
  73. 73 Khaneja, N., Reiss, T.O., Kehlet, C., Schulte‐Herbrüggen, T., and Glaser, S.J. (2005) J. Magn. Reson., 172, 296.
  74. 74 Machnes, S., Sander, U., Glaser, S.J., de Fouquiéres, P., Gruslys, A., Schirmer, S., and Schulte‐Herbrüggen, T. (2011) Phys. Rev. A, 84, 022305.
  75. 75 de Fouquiéres, P., Schirmer, S., Glaser, S.J., and Kuprov, I. (2011) J. Magn. Reson., 212, 412.
  76. 76 Krotov, V.F. (1996) Global Methods in Optimal Control, Marcel Dekker, New York.
  77. 77 Krotov, V.F. and Feldman, I.N. (1983) Eng. Cybern., 21, 123; Russian original: (1983) Izv. Akad. Nauk. SSSR Tekh. Kibern., 52, 162.
  78. 78 Konnov, A.I. and Krotov, V.F. (1999) Autom. Remote Control, 60, 1427; Russian original: (1999) Avtom. Telemekh., 1999, 77.
  79. 79 Palao, J.P. and Kosloff, R. (2003) Phys. Rev. A, 68, 062308.
  80. 80 Ohtsuki, Y., Turinici, G., and Rabitz, H. (2004) J. Chem. Phys., 120, 5509.
  81. 81 Khaneja, N., Brockett, R., and Glaser, S.J. (2001) Phys. Rev. A, 63, 032308.
  82. 82 Khaneja, N. and Glaser, S.J. (2001) Chem. Phys., 267, 11.
  83. 83 Reiss, T.O., Khaneja, N., and Glaser, S.J. (2002) J. Magn. Reson., 154, 192‐195.
  84. 84 Khaneja, N., Heitmann, B., Spörl, A., Yuan, H., Schulte‐Herbrüggen, T., and Glaser, S.J. (2007) Phys. Rev. A, 75, 012322.
  85. 85 Yuan, H., Khaneja, N., and Glaser, S.J. (2007) Phys. Rev. A, 76, 012316.
  86. 86 Nimbalkar, M., Zeier, R., Neves, J.L., Elavarasi, S.B., Yuan, H., Khaneja, N., Dorai, K., and Glaser, S.J. (2012) Phys. Rev. A, 85, 012325.
  87. 87 Assémat, E., Lapert, M., Sugny, D., and Glaser, S.J. (2013) Math. Control Relat. Fields, 3, 375.
  88. 88 Garon, A., Glaser, S.J., and Sugny, D. (2013) Phys. Rev. A, 88 (4). doi: 10.1103/physreva.88.043422.
  89. 89 Yuan, H., Wei, D., Zhang, Y., Glaser, S.J., and Khaneja, N. (2014) Phys. Rev. A, 89, 042315.
  90. 90 van Damme, L., Zeier, R., Glaser, S.J., and Sugny, D. (2014) Phys. Rev. A, 90, 013409.
  91. 91 Köcher, S., Heydenreich, T., Zhang, Y., Reddy, G.N.M., Caldarelli, S., Yuan, H., and Glaser, S.J. (2016) J. Chem. Phys., 144, 164103.
  92. 92 Paolo, J.P. and Kosloff, R. (2002) Phys. Rev. Lett., 89, 188301.
  93. 93 Sklarz, S.E. and Tannor, D.J. (2006) Chem. Phys., 322, 87.
  94. 94 Schulte‐Herbrüggen, T., Spörl, A.K., Khaneja, N., and Glaser, S.J. (2005) Phys. Rev. A, 72, 042331.
  95. 95 Nielsen, M.A. and Chuang, I.L. (2000) Quantum Computation and Quantum Information, Cambridge University Press, Cambridge.
  96. 96 Alber, G., Beth, T., Horodecki, M., Horodecki, P., Horodecki, R., Rötteler, M., Weinfurter, H., Werner, R., and Zeilinger, A. (2000) Quantum Information: An Introduction to Basic Concepts and Experiments, Springer Tracts in Modern Physics, vol. 173, Springer‐Verlag, Heidelberg.
  97. 97 Saito, A., Kioi, K., Akagi, Y., Hashizume, N., and Ohta, K. (2000) Actual Computational Time‐Cost of the QFT in a Quantum Computer Using Nuclear Spins. quant‐ph/0001113.
  98. 98 Blais, A. (2001) Phys. Rev. A, 64, 022312.
  99. 99 Barenco, A., Bennett, C.H., Cleve, R., DiVincenzo, D.P., Margolus, N., Shor, P., Sleator, T., Smolin, J.A., and Weinfurter, H. (1995) Phys. Rev. A, 52, 3457.
  100. 100 Khaneja, N., Glaser, S.J., and Brockett, R. (2002) Phys. Rev. A, 65, 032301.
  101. 101 Tseng, C.H., Somaroo, S., Sharf, Y., Knill, E., Laflamme, R., Havel, T.F., and Cory, D.G. (2000) Phys. Rev. A, 61, 012302.
  102. 102 Yamamoto, T., Pashkin, Yu.A., Astaviev, O., Nakamura, Y., and Tsai, J.S. (2003) Nature (London), 425, 941.
  103. 103 Spörl, A., Schulte‐Herbrüggen, T., Glaser, S.J., Bergholm, V., Storcz, M.J., Ferber, J., and Wilhelm, F.K. (2007) Phys. Rev. A, 75 (1), 12302.
  104. 104 Ettinger, M., Høyer, P., and Knill, E. (2004) Inf. Process. Lett., 91, 43.
  105. 105 Kauffman, L.H. and Lomonaco, S.J. (2008) Quantum hidden subgroup problems: an algorithmic toolkit, in Mathematics of Quantum Computation and Quantum Technology, Chapman & Hall/CRC, Boca Raton, FL, pp. 3‐46.
  106. 106 Kauffman, L.H. and Lomonaco, S.J. (2006) Proc. SPIE, 6244, 62440Z.
  107. 107 Aharonov, D., Jones, V., and Landau, Z. (2006) Proceedings of the 38th ACM Symposium on Theory of Computing (STOC 2006), vol. 2006, p. 427, https://dl.acm.org/citation.cfm?id=1132516; https://www.cs.washington.edu/stoc06/.
  108. 108 Kauffman, L.H. and Lomonaco, S.J. (2008) Quantum computing and quantum topology, in Mathematics of Quantum Computation and Quantum Technology, Chapman & Hall/CRC, Boca Raton, FL, pp. 409–514.
  109. 109 Marx, R., Fahmy, A., Kauffman, L.H., Lomonaco, S.J., Spörl, A., Pomplun, N., Schulte‐Herbrüggen, T., Myers, J., and Glaser, S.J. (2010) Phys. Rev. A, 81, 032319.
  110. 110 Schulte‐Herbrüggen, T., Marx, R., Fahmy, A., Kauffman, L., Lomonaco, S., Khaneja, N., and Glaser, S.J. (2012) Philos. Trans. R. Soc. London, Ser. A, 380, 4651.
  111. 111 Shor, P. and Jordan, S. (2008) Quantum Inf. Comput., 8, 0681.
  112. 112 Aharonov, D., Jones, V., and Landau, Z. (2009) Algorithmica, 55, 395.
  113. 113 Aharonov, D. and Arad, I. (2011) New J. Phys., 13, 035019.
  114. 114 Abramsky, S. (2008) Temperley‐Lieb algebra: from knot theory to logic and computation via quantum mechanics, in Mathematics of Quantum Computation and Quantum Technology, Chapman & Hall/CRC, Boca Raton, FL, pp. 515‐558.
  115. 115 Aharonov, D., Arad, I., Eban, E., and Landau, Z. (2007) Polynomial Quantum Algorithms for Additive approximations of the Potts model and other Points of the Tutte Plane. arXiv.org/quant‐ph/0702008.
  116. 116 Passante, G., Moussa, O., Ryan, C., and Laflamme, R. (2009) Phys. Rev. Lett., 103, 250501.
  117. 117 Jordan, S. and Wocjan, P. (2009) Quantum Inf. Comput., 9, 0264.
  118. 118 Laflamme, R., Cory, D.G., Negrevergne, C., and Viola, L. (2002) Quantum Inf. Comput., 2, 166.
  119. 119 Gradl, T., Spörl, A.K., Huckle, T., Glaser, S.J., and Schulte‐Herbrüggen, T. (2006) Parallelising matrix operations on clusters for an optimal control‐based quantum compiler, in Euro‐Par 2006 Parallel Processing. Euro‐Par 2006, Lecture Notes in Computer Science, vol. 4128 (eds W.E. Nagel, W.V. Walter, and W. Lehner), Springer‐Verlag, Berlin, Heidelberg, p. 751.
  120. 120 Schulte‐Herbrüggen, T., Spörl, A., Waldherr, K., Gradl, T., Glaser, S.J., and Huckle, T. (2008) Using the HLRB cluster as quantum CISC compiler: matrix methods and applications for advanced quantum control by gradient‐flow algorithms on parallel clusters, in High‐Performance Computing in Science and Engineering, Garching 2007 (eds S. Wagner, M. Steinmetz, A. Bode, and M. Brehm), Springer‐Verlag, Berlin, pp. 517–533.
  121. 121 Schulte‐Herbrüggen, T., Spörl, A., and Glaser, S.J. (2007) Quantum CISC compilation by optimal control and scalable assembly of complex instruction sets beyond two‐qubit gates, arXiv.org/quant‐ph/0712.3227.
  122. 122 Fisher, R., Helmer, F., Glaser, S.J., Marquardt, F., and Schulte‐Herbrüggen, T. (2010) Phys. Rev. B, 81, 085328.
  123. 123 Tomoney, N., Elman, V., Glaser, S.J., Weiss, C., Johanning, M., Neuhauser, W., and Wunderlich, C. (2008) Phys. Rev. A, 77, 052334.
  124. 124 Dolde, F., Bergholm, V., Wang, Y., Jakobi, I., Naydenov, B., Pezzagna, S., Meijer, J., Jelezko, F., Neumann, P., Schulte‐Herbrüggen, T., Biamonte, J., and Wrachtrup, J. (2014) Nat. Commun., 5, 3371.
  125. 125 Waldherr, G., Wang, Y., Zaiser, S., Jamali, M., Schulte‐Herbrüggen, T., Abe, H., Ohshima, T., Isoya, J., Du, J.F., Neumann, P., and Wrachtrup, J. (2014) Nature, 506, 204.
  126. 126 Zaiser, S., Rendler, T., Jakobi, I., Wolf, T., Lee, S.‐Y., Wagner, S., Bergholm, V., Schulte‐Herbüggen, T., Neumann, P., and Wrachtrup, J. (2016) Nat. Commun., 8, 12279.
  127. 127 Keyl, M., Zeier, R., and Schulte‐Herbrüggen, T. (2014) New J. Phys., 16, 065010.
  128. 128 Waldherr, K., Huckle, T., Auckenthaler, T., Sander, U., and Schulte‐Herbrüggen, T. (2010) Fast 3D block parallelisation for the matrix multiplication prefix problem: application in quantum control, in High‐Performance Computing in Science and Engineering, Garching 2009 (eds S. Wagner, M. Steinmetz, A. Bode, and M. Muller), Springer‐Verlag, Berlin, pp. 39‐50.
  129. 129 Riahi, M.K., Salomon, J., Glaser, S.J., and Sugny, D. (2016) Phys. Rev. A, 93, 043410.
  130. 130 Dridi, G., Lapert, M., Salomon, J., Glaser, S.J., and Sugny, D. (2015) Phys. Rev. A, 92, 043417.
  131. 131 Schilling, F., Warner, L.R., Gershenzon, N.I., Skinner, T.E., Sattler, M., and Glaser, S.J. (2014) Angew. Chem. Int. Ed., 53, 4475.
  132. 132 Neves, J.L., Heitmann, B., Khaneja, N., and Glaser, S.J. (2009) J. Magn. Reson., 201, 7.
  133. 133 Neigovzen, R., Neves, J.L., Sollacher, R., and Glaser, S.J. (2009) Phys. Rev. A, 79, 042321.

Notes

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset