27
A Brief on Quantum Systems Theory and Control Engineering

Thomas Schulte‐Herbrüggen1, Robert Zeier1, Michael Keyl2 and Gunther Dirr3

1 Technische Universität München, Department Chemie, Lichtenbergstrasse 4, D‐85747 Garching, Germany

2 Freie Universität Berlin, Dahlem Center for Complex Quantum Systems, Arnimallee 14, 14195 Berlin, Germany

3 Universität Würzburg, Mathematisches Institut II, Emil‐Fischer‐Strasse 40, 97074 Würzburg, Germany

We illustrate a unified frame of quantum systems theory in view of control engineering. In particular, we shift controllability criteria from the well‐known Lie‐algebra rank condition to symmetry conditions that are easy to check yet rigorously rooted in the branching diagrams for simple subalgebras of images . Reachable sets of closed bilinear control systems are linked to the theory of images ‐numerical ranges. In coherently controlled open Markovian systems, the set of reachable directions (Lindblad generators) forms a Lie wedge that generates a Lie semigroup of quantum maps and helps to approximate reachable sets of open systems. Once reachable sets are known, gradient‐flow algorithms can solve the abstract optimization task on the reachable sets. They thus complement numerical algorithms for concrete optimal control problems on the manifold of admissible control amplitudes presented in a unified programming framework in Chapter 28. Finally, we give an outlook on infinite‐dimensional control of atoms coupled to oscillator modes.

How principles turn into practice has meanwhile emerged in a plethora of examples showing applications in solid‐state devices, circuit‐QED, ion traps, NV‐centers in diamond, quantum dots, spin systems, and images ‐level atoms coupled to a (light)field.

27.1 Introduction

This contribution is meant as an invitation to exchange with the vibrant developments in the field of quantum systems and control (1) in view of future technologies (2). These may be triggered by precise controls for, for example, quantum simulation in order to improve the understanding of quantum phase transitions (3) between conducting and superconducting phases, or between ferromagentic and anti‐ferromagnetic phases to name just a few. Needless to say, a thorough picture of these phenomena will boost material design.

An important issue in quantum simulation (48) is to manipulate all pertinent dynamical degrees of freedom of a system images of interest (which, however, all too often is experimentally not fully accessible) by another quantum system images that is in fact well controllable in practice and the dynamics of which are equivalent to those of images . We will show how to characterize this situation algebraically in terms of quantum systems theory.

Besides the practical applications and implications, quantum systems should also be of appeal to the (classical) control engineer, because nearly all systems of interest boil down to the standard form of bilinear control systems (912)

27.1 equation

Here one may take images as linear operators on a (finite‐dimensional) Hilbert space of quantum states images . For images two‐level spin‐images systems, images . More precisely, images denotes the system or drift Hamiltonian images , whereas the images are the control Hamiltonians images governed by typically piece‐wise constant control amplitudes images . Thus Eq. 27.1 captures all scenarios in Table 27.1, where images denotes a unitary operator on images (e.g., used as quantum gate). For open systems, images is a (linear) quantum map that acts on density operators images and whose time evolution is governed by the (super)operator images including relaxation.

Table 27.1 Bilinear quantum control systems.

Here images represents the Hamiltonian commutator superoperator.

Setting and task Drift Controls
images images images
Closed systems:
Pure‐state transfer images images images
Gate synthesis (with specified global phase)  images images images
State transfer images images images
Gate synthesis (with free global phase) images images images
Open systems:
State transfer images images images
Quantum‐map synthesis images images images

While linear control systems images are fully controllable (13) if the reachability matrix images has full rank, bilinear systems of the form Eq. 27.1 are fully controllable on a compact connected Lie group images (with Lie algebra images so images ) if they satisfy the celebrated Lie‐algebra rank condition (1418)

equation

As in open systems images is usually no longer compact, dissipative systems are obviously more subtle and give rise to Lie semigroups in the Markovian case illustrated below.

For closed quantum systems of images spins‐images , one has images and images with images illustrating how dynamic degrees of freedom in quantum systems scale exponentially in system size (as opposed to classical systems, where they scale linearly). Thus assessing controllability via an explicit computation of the Lie closure for the rank condition, though mathematically straightforward, becomes dramatically tedious in large quantum systems, and beyond seven qubits, it is mostly prohibitive.

27.2 Systems Theory of Closed Quantum Systems

Hence here we sketch particularly simple and powerful symmetry arguments for assessing controllability of quantum systems avoiding an explicit calculation of the Lie closure. For extending the symmetry arguments to a frame embracing closed and open systems in more detail, see the recent Ref. (19), from whence parts are extracted here.

Graphical representation of quantum dynamical control systems.

Figure 27.1 Graph representation of quantum dynamical control systems: vertices represent two‐level systems (qubits), where common color and letter code denote joint local action, whereas the edges stand for pairwise coupling interactions. White vertices are qubits that are just coupled to the dynamic system without allowing to be controlled locally. The first and the last graphs show no symmetries and their underlying control system is fully controllable. In contrast, the second and third graphs do exhibit symmetries: the second one has a mirror symmetry, whereas the third one leaves the Pauli operator images on the upper terminal qubit invariant. These constants of the motion clearly preclude full controllability.

27.2.1 Controllability and its Symmetry Conditions

It pays to envisage bilinear control systems by graphs as illustrated in Figure 27.1: vertices represent local qubits as controlled by typical control Hamiltonians images (represented by Pauli matrices images acting on the qubit represented by the respective vertex), edges stand for pair‐wise coupling interactions typically only occurring in the drift term images (two‐component tensor products of Pauli matrices as, for example, images for Ising interaction or images for Heisenberg–images interaction. Here the Pauli operators act on the two qubits connected by the respective edge).

As a central notion in the subsequent arguments, we characterize a quantum bilinear control system by its system Lie algebra images , which results from the Lie closure of taking nested commutators (until no new linearly independent elements are generated)

equation

as well as by its (potential) symmetries, that is, the centralizer images in images to the system algebra images collecting all terms that commute jointly with all Hamiltonian operators

equation

If there are no symmetries, that is, if the centralizer images is trivial, then the system algebra images is irreducible. This can easily be checked by determining the dimension of the nullspace to the corresponding commutator superoperators (of dimension images ) – so it boils down to solving a system of images homogeneous equations in images dimensions.

Image described by caption and surrounding text.

Figure 27.2 Branching diagrams showing all the irreducible simple subalgebras of images with images for images ‐qubit systems with images as given in (20). Note that for odd images , only the two canonical branches with orthogonal (lower) and symplectic (upper) subalgebras occur. In contrast, for even images , there are always unitary spinor‐type subalgebras images and in some instances images . The orthogonal subalgebras are related to fermionic quantum systems, whereas the symplectic ones are related to compact versions of bosonic ones as described in the text and shown in Tables 27.2 and 27.3.

Therefore, a trivial centralizer plus a connected graph imply that the corresponding system algebra is simple. As the largest possible Lie closure is images , the system algebra images of an irreducible connected qubit system has to be a (proper or improper) irreducible simple subalgebra to images . By making heavy use of computer algebra, in Ref. (20), we have classified all these simple subalgebras of images for images with images qubits as summarized by the branching diagrams in Figure 27.2 thus extending the known results from images (21,22) to images .

This figure also illustrates that every images with images has two canonical branches, a symplectic branch (upper branch) starting with images and an orthogonal branch (lower) commencing with images . Actually, for odd images , these are the only ones (and we conjecture that this holds true even beyond 15 qubits). In contrast, for even images , there are always subalgebras images of unitary (spinor) type shown in black plus potential others (observe the instances of images ). – Clearly, if the (nontrivial) system algebra images of a dynamic system in question can be ruled out to be on any of these three branches, then corresponding control system is indeed fully controllable as will be shown next.

To this end, it is convenient to exclude the symplectic and orthogonal subalgebras in the first place. It is a task that can again be readily accomplished (after having made sure images is irreducible) by determining the dimension of the joint null space (over S) to the following equations for each images with images images or in superoperator form images where by Schur's Lemma images (23). If there is a nontrivial solution for the (+)‐variant, images is of orthogonal type, and if there is for the (images )‐variant, images is of symplectic type. Therefore, if the solution space for both cases (images ) is zero‐dimensional (corresponding to the only solution being trivial), then images is neither of orthogonal nor symplectic type. This can conveniently be decided by solving a system of linear equations as done in Algorithm 3 of Ref. (20).

For odd images , this does in fact already ensure full controllability, as only even images allows for unitary (spinor‐type) simple subalgebras. Yet we conjecture that these findings also hold for all images . Finally, for even images , the spinor‐type subalgebras may be excluded by the subsequent theorem of Ref. (20). To prepare for it, observe that for images selfadjointness of images and images entails

equation

hence the projector on images is in the commutant of the tensor‐square representation, that is, images .

This motivates a closer look at the tensor‐square representation

equation

and its commutant referred to as “quadratic symmetries” of images by images that give a powerful single necessary and sufficient symmetry condition for full controllability:

To sum up, a bilinear images ‐qubit control system as in Eq. 27.1 is fully controllable if and only if all of the following conditions are satisfied

  1. the system has no symmetries, that is, images is trivial;
  2. the system has a connected coupling graph;
  3. the system algebra images is neither of orthogonal nor of symplectic type; and
  4. the system algebra is not of any other type, in particular not of unitary spinor‐type or of exceptional type (images ).

While we gave a rigorous proof in Ref. (20) as already mentioned, above key arguments can easily be made intuitive as follows:

  1. symmetries would entail conserved entities (invariant one‐parameter groups) thus precluding full controllability;
  2. coupling graphs with several connected components preclude that these components can be coherently coupled, which obviously is necessary for full controllability;
  3. orthogonal or symplectic subalgebras are proper subalgebras to images (for images ) and do not explore all dynamic degrees of freedom of images ; and
  4. the same holds for unitary spinor‐type or exceptional subalgebras (images ) of images .

Table 27.2 Heisenberg–images spin chains with a single control on one end (or both) can simulate either fermionic or bosonic systems depending on the chain length as summarized in (20).

Local control over two adjacent qubits is required to make the system fully controllable (last row).

System type Fermionic ‘Bosonic’ System algebra
images ‐spins‐images images Levels — Coupling order —
Illustration of Heisenberg-XX spin chains with a single control on one end simulating either fermionic or bosonic systems depending on the chain length: n. images 2 images
Illustration of Heisenberg-XX spin chains with a single control on one end simulating either fermionic or bosonic systems depending on the chain length: n + 1. images 2 images
Illustration of Heisenberg-XX spin chains with a single control on one end simulating either fermionic or bosonic systems depending on the chain length: n.
for images images Up to images images
for images images Up to images images
Illustration of Heisenberg-XX spin chains with a single control on one end simulating either fermionic or bosonic systems depending on the chain length: n. images Up to images Up to images images

By the branching diagrams in Figure 27.2, it is immediately obvious: establishing full controllability boils down to ensuring the dynamic system is governed by a system algebra that is irreducible (no symmetries), and simple (connected coupling graph) and top of the branch. This shifts the paradigm from the Lie‐algebra rank‐condition to easily verifiable symmetry conditions, which can be checked using only the Hamiltonian generators.

Table 27.3 Ising‐images spin chains with joint controls on all the qubits locally can simulate bosonic systems provided the coupling constants of the right and left branches leaving the central qubit have opposite signs as is also summarized in (20).

Note that even physically unavailable three‐body interactions can be simulated by such systems. The system algebras given on the right specify that for a given chain length all systems are dynamically equivalent, which otherwise would be extremely difficult to analyze.

System type ‘Bosonic’ System algebra
images spins‐images images Levels Coupling order images
Illustration of Ising-ZZ spin chains with joint controls on all the qubits locally simulating bosonic systems provided the coupling constants of the right and left branches leaving the central qubit have opposite signs: n =3. images Up to 3 images
Illustration of Ising-ZZ spin chains with joint controls on all the qubits locally simulating bosonic systems provided the coupling constants of the right and left branches leaving the central qubit have opposite signs: n =3. —”— —”— —”—
Illustration of Ising-ZZ spin chains with joint controls on all the qubits locally simulating bosonic systems provided the coupling constants of the right and left branches leaving the central qubit have opposite signs: n =5. images Up to 5 images
Illustration of Ising-ZZ spin chains with joint controls on all the qubits locally simulating bosonic systems provided the coupling constants of the right and left branches leaving the central qubit have opposite signs: n =5. —”— —”— —”—
Illustration of Ising-ZZ spin chains with joint controls on all the qubits locally simulating bosonic systems provided the coupling constants of the right and left branches leaving the central qubit have opposite signs: n =5. —”— —”— —”—
Illustration of Ising-ZZ spin chains with joint controls on all the qubits locally simulating bosonic systems provided the coupling constants of the right and left branches leaving the central qubit have opposite signs: n =5. —”— —”— —”—
Illustration of Ising-ZZ spin chains with joint controls on all the qubits locally simulating bosonic systems provided the coupling constants of the right and left branches leaving the central qubit have opposite signs: n =5. —”— —”— —”—
Illustration of Ising-ZZ spin chains with joint controls on all the qubits locally simulating bosonic systems provided the coupling constants of the right and left branches leaving the central qubit have opposite signs: n =5. —”— —”— —”—
Illustration of Ising-ZZ spin chains with joint controls on all the qubits locally simulating bosonic systems provided the coupling constants of the right and left branches leaving the central qubit have opposite signs: n =5. —”— —”— —”—

27.2.2 Simulability and its Symmetry Conditions

Recall that fermionic quantum systems (with quadratic Hamiltonians) relate to orthogonal system algebras, whereas compact versions of bosonic ones (henceforth written as “bosonic” for short) relate to symplectic system algebras. Then the link from controlled quantum systems to quantum simulation becomes obvious: the branching diagrams of Figure 27.2 also illustrate that an (irreducible and connected) images ‐qubit quantum system is fully controllable if and only if it can simulate bothbosonic” and fermionic systems.

This is because – clearly – a controlled bilinear dynamic system images can simulate another system images if and only if forthe system algebras one has images . Moreover, given a fixed Hilbert space images , images simulates images efficiently (i.e., with least state‐space overhead in images ) if for any interlacing system images with system algebra images satisfying images one must have either images or images or (trivially) both.

For illustration, consider an images ‐qubit nearest‐neighbor coupled Heisenberg–images spin chain with single local controls. Then Table 27.2 shows that a single controllable qubit at one end suffices to simulate a fermionic system with quadratic interactions on images levels (governed by images ), whereas local controls on both ends are required to simulate quadratic fermionic systems on images levels with system algebra images . Most remarkably, if the controllable qubit is shifted to the second position, one gets dynamic degrees of freedom scaling exponentially in the number of qubits in the chain. This is by virtue of the system algebras images or images , which most noticeably depend on the length of the images ‐qubit chain: if images , the system is fermionic (images ), whereas for images , the system is bosonic (images ) (20). It is not until two adjacent qubits can be coherently controlled (as images ) that the Heisenberg–images spin chains become fully controllable (25).

Table 27.3 illustrates the power of classifying dynamic systems by symmetries and thereby in terms of their system Lie algebras: it turns out that joint controls on all the local qubits simultaneously suffice to even simulate effective three‐body interactions (usually never occurring naturally), provided the Ising‐images coupling in odd‐membered spin chains can be designed to have opposite signs on the two branches reaching out from the central spin.

The same methods can be extended to cover system algebras of fermionic systems (obeying the fermionic super‐selections rules) and their simulability by spin systems (26).

Quadratic symmetries that solve the controllability problem in simple subalgebras of images can be carried over to cover also the case of (compact) semi‐simple subalgebras of images : for images , one has equality iff for their generators images and images the quadratic symmetries fulfill images (27), whereas for equality in the general compact case also, the projections onto the linear center have to be of equal dimension to ensure images (28).

Next we illustrate how system algebras images obtained here by symmetry characterization translate into reachable sets taking the form of group orbits images of initial states images . The orbits in turn can be projected onto observables to give all admissible expectation values.

27.2.3 Reachable Sets and Expectation Values of Closed Quantum Systems: Link to Relative images ‐Numerical Ranges

Once the compact system algebra images of a bilinear control system is determined, for example, by symmetry characterization as in the previous section, then the time evolution is brought about by the corresponding group01 images . Therefore, the reachable set for every initial state images is given by the corresponding subgroup orbit images

equation

In other words, the time evolution of the state images is confined to images in the sense images solves the equation of motion 27.1 under Hamiltonian drift images and controls images in the absence of relaxation, that is images .

In quantum dynamics, the expectation value of a Hermitian observable, or more generally a detection operator images , is defined as projection of images onto images by way of the Hilbert–Schmidt scalar product images where images . Recall that the field of values of images is images , whereas for images the images ‐numerical range of images is images . Therefore, if images is a rank‐1 projector, the expectation value is an element of the field of values images , whereas for general images , it is an element of the images ‐numerical range of images , that is, images . The latter is a star‐shaped subset of the complex plane (29,30) and it specializes to a real line segment in case images and images are both Hermitian.

As illustrated in the previous section, different quantum dynamical scenarios come with specific dynamical subgroups images generated by the specific system algebras images . Typical examples for images include images or images or the subgroup of local unitary operations images .

Consequently, in the instances of images , the admissible expectation values typically fill but a proper subset of images , which hence motivates our definition of a restricted or relative images ‐numerical range (31,32) as subgroup orbit images projected onto images

equation

The particular case of local qubit‐wise actions images leads to what we define as local images ‐numerical range. As images is compact and connected, this obviously extends to images . However, note that although being connected, in general images turns out to be neither star shaped nor simply connected (32) in contrast to the usual images ‐numerical range (29).

The largest absolute value of the relative images ‐numerical range is defined as the relative images ‐numerical radius

equation

it obviously plays a significant role for optimizations aiming at maximal expectation values.

With these stipulations, we will discuss recent applications of the local images ‐numerical range in quantum control.

27.2.4 Constrained Optimization and Relative images ‐Numerical Ranges

In quantum control, one may face the problem to maximize the unitary transfer from matrices images to images subject to suppressing the transfer from images to images , or subject to leaving another state images invariant. For tackling those types of problems, in (33), we asked for a “constrained images ‐numerical range of images

equation

which form it takes and – in view of numerical optimization – whether it is a connected set with a well‐defined boundary. Connectedness is central to any numerical optimization approach, because otherwise one would have to rely on initial conditions in the connected component of the (global) optimum.

Now the constrained images ‐numerical range of images is a compact and connected set in the complex plane, if the constraint can be fulfilled by restricting the full unitary group images to a compact and connected subgroup images . In this case, the constrained images ‐numerical range images is identical to the relative images ‐numerical range images and hence the constrained optimization problem is solved within it, for example, by the corresponding relative images ‐numerical radius images .

The new concept of the relative (or restricted) images ‐numerical range has meanwhile become a popular tool, for example, for analyzing entanglement properties, see (34,35) (and references therein).

Therefore, next we optimize by gradient flow methods over the underlying images ‐orbits.

27.2.5 Optimization by Gradient Flows

A pioneering paper by Brockett (36) extended in books by Helmke and Moore (37) and followed by Bloch (38) triggered to apply gradient flows on the unitary orbit of quantum states (39).

Implementing a gradient method for optimization on a smooth “constrained manifold” – such as an unitary orbit – via the Riemannian exponential map, inherently ensures that the discretized flow remains within that manifold. Therefore, gradient flows on manifolds are intrinsic optimization methods (40), whereas extrinsic optimizations on an embedding space require in general nonlinear projection techniques in order to stay on the constrained manifold. In particular, using the differential geometry of matrix manifolds has become a field of active research. For recent developments, however, without exploiting all the Lie structure, see, for example, Ref. (41,42).

Here we sketch an overview (43) for various optimization tasks in quantum dynamical systems in the common framework of gradient flows on Riemannian manifolds. Let images denote a smooth manifold, for example, the unitary orbit of an initial state images . A flow is a smooth map images such that for all states images and times images

equation

hence the flow acts as a one‐parameter group, and for positive times images as a one‐parameter semigroup of diffeomorphisms on images .

Now, let images be a smooth quality function on images . Recall that the differential of images is a mapping (section) images of the manifold to its cotangent bundle images , while the gradient vector field is a mapping images to its tangent bundle images . Therefore, the scalar product images plays a central role as it allows for identifying images with images ; this is why the pair images has to be a Riemannian manifold with Riemannian metric images . Thus one arrives at the gradient flow images determined by

27.2 equation

Formally, its solutions are obtained by integrating Eq. 27.2 to give

equation

Observe this ensures that images does increase along trajectories of images by virtue of following the gradient direction of images . – Gradient flows typically run into some local extremum as in Figure 27.3. Therefore, sufficiently many independent initial conditions may be needed to provide confidence into numerical results. Sometimes, local extrema can be ruled out; prominent examples of this type are discussed in (43) for Brockett's double‐bracket flow ( 36, 37) addressed below.

Graphical representation of Abstract optimization task.

Figure 27.3 Abstract optimization task: By following the gradient flow images on the manifold images (left), the quality function images (right) is driven into a (local) maximum.

27.2.5.1 Discretized Gradient Flows

In the simplest case, where images , gradient flows may be solved by moving along the gradient images in the sense of a Steepest Ascent Method

equation

with step size images . Here, the manifold images coincides with its tangent space images containing images . Clearly, a generalization is required as soon as images and images are no longer identifiable. This gap is filled by the Riemannian exponential map

equation

such as to arrive at an intrinsic Euler step method. It is performed by the Riemannian exponential map, so straight line segments used in the standard method are replaced by geodesics on images in the Riemannian Gradient Method

27.3 equation

where images is a step size ensuring convergence. For matrix Lie groups images with bi‐invariant metric, Eq. 27.3 simplifies to the Gradient Method on a Lie Group (43)

equation

where images is the usual exponential map.

In either case, the iterative procedure can be pictured as follows: at each point images , one evaluates images in the tangent space images . Then one moves via the Riemannian exponential map in direction images to the next point images on the manifold so that the quality function images improves, images , as shown in Figure 27.3.

27.2.5.2 Gradient Flows on Subgroups

For images let images denote the unitary orbit of images . For minimizing the (squared) Euclidean distance images between images and the unitary orbit of images , we give a gradient flow maximizing the target function images over images with the equivalence images Note images is a compact and connected naturally reductive homogeneous space isomorphic to images where images is the stabilizer group of images .

Moreover, the double‐bracket flow to images just defined is brought about by the gradient images where images is the skew‐Hermitian part of images . Therefore, the corresponding gradient flow

27.4 equation

is an isospectral flow on images . The solutions exist for all images and converge to a critical point images of images characterized by images . A detailed discussion for the real case can be found in (37); for an abstract Lie algebraic version, see also (44).

In order to obtain a numerical algorithm for maximizing images , one can discretize the continuous‐time gradient flow 27.4 as

27.5 equation

with appropriate step sizes images . Eq. 27.5 exploits that the adjoint orbit images is a naturally reductive homogeneous space and thus the knowledge on its geodesics.

For images complex Hermitian (real symmetric) and the full unitary (or orthogonal) group or its respective orbit the gradient flow 27.4 is well understood. However, for non‐Hermitian images and images , the nature of the flow and in particular the critical points have not been analyzed in depth, because the Hessian at critical points is difficult to come by. Even for images Hermitian, a full critical point analysis becomes nontrivial as soon as the flow is restricted to a closed and connected subgroup images . Nevertheless, the above techniques can be taken over to establish a gradient flow and a respective gradient algorithm on the orbit images in a straightforward manner.

Likewise the gradient flow of Eq. 27.4 restricts to the subgroup orbit images by taking the respective orthogonal projection images onto the subalgebra images of images instead of projecting onto the skew‐Hermitian part. Thus images With step sizes images , the corresponding discrete integration scheme reads

27.6 equation

In view of unifying the interpretation of unitary networks, for example, for the task of computing ground states of quantum mechanical Hamiltonians images , the double‐bracket flows for complex Hermitian images on the full unitary orbit images as well as on the subgroup orbits images for partitionings by images with images have shifted into focus. Thus we gave the foundations for the recursive schemes of Eqs. 27.5 and 27.6 listed with many more worked examples in the comprehensive overview Ref. (43).

In particular, in (43), we addressed gradient flows for constrained optimization problems. The intrinsic constraints can be accommodated by restricting the dynamic group to proper subgroups images of the unitary group. Beyond that, gradient flows combining intrinsic constraints by restrictions to proper subgroups with extrinsic constraints can be taken care of by Lagrange‐type penalty parameters. Therefore, Ref. (43) provides a full toolkit of gradient‐flow based optimizations alongside ( 41, 42). It has also been very powerful for best approximations by sums of compact group orbits (45).

27.3 Toward a Systems Theory for Open Quantum Systems

We saw that in closed systems, there is a particularly simple characterization of reachable sets by the compact system algebra images generating the Lie group images and the corresponding group orbit, that is, images In open quantum systems, it is more intricate to estimate the reachable sets. We consider bilinear control systems of open quantum systems for quantum maps following the master equation

equation

In unital systems (those with at least one of the fixed points proportional to images ), one finds by the seminal work of (46) and (47) on majorization that

equation

as used recently in (48). However, equality holds only under the assumption that all coherent controls are infinitely fast in the sense of images which from the viewpoint of physics is most often hopelessly idealizing. Therefore, this simple inclusion becomes totally inaccurate in all physically more realistic scenarios, where the drift Hamiltonian images is necessary for full controllability in the sense of images and – even worse – the inaccuracy increases with system size images . For these experimentally more realistic relevant cases, we thus recently characterized (49,50) the dynamic system in terms of the underlying Lie wedge images generating the dynamic system Lie semigroup images of irreversible (Markovian) time evolution. Here the reachable sets can be much more accurately approximated by

equation

where usually few factors suffice to give a good estimate.

This motivates the sketch of basic features of Lie semigroups.

27.3.1 Markovian Quantum Maps as Lie Semigroups

Let us start with the following distinction: A (completely positive) trace‐preserving quantum map images is (infinitely) divisible, if for all images there is a images with images , whereas it is infinitesimally divisible if for all images there is a sequence images with images .

Moreover, a quantum map images is termed time‐(in)dependent if it is the solution of a time‐(in)dependent master equation images with images being time‐(in)dependent.

Now one finds the important characterization:

To sketch the relation to Lie semigroups, the basic vocabulary can be captured in the following definitions along the lines of Ref. (49):

Moreover one has:

In (49), it turned out that the seminal work of Kossakowski and Lindblad on quantum maps can now be recast into the context of Lie semigroups as follows:

There are indeed elements in the connected component images that cannot be exponentially generated and hence fail to be within the Lie semigroup images . Most noteworthy, they are exactly the non‐Markovian quantum maps in images . Thus in this sense, the Markov properties and the Lie properties of quantum maps are 1:1.

Moreover, one finds yet another important distinction:

In summary, there are two demarcations: (i) the borderline between Markovian and non‐Markovian quantum maps is drawn by the Lie‐semigroup property, whereas (ii) the separation between time‐dependent and time‐independent Markovian quantum maps is marked by the generating Lie wedge and its specialization to the form of a Lie semialgebra (49) in the time‐independent case.

27.3.2 Reachable Sets in Dissipatively Controlled Open Systems

As stated in the introductory part, we have recently characterized coherently controlled bilinear open systems (of images spins‐images ) of the form

27.8 equation

(here images constant with images of the form of Eq. 27.7) by their respective Lie wedges images generating the dynamic system Lie semigroup images of irreversible (Markovian) time evolution in Ref. (50). This promises that the reachable sets can conveniently be approximated by images where images with images and where usually few factors suffice to give a good estimate. — For the sequel, suppose the unitary part of the above system is fully controllable in the sense

27.9 equation

27.3.2.1 The Magic of Switchable Noise and Coherent Control

We have currently gone a step further such as to include into a coupled network of two‐level (spin‐images ) systems a single qubit the relaxion amplitude of which shall be switchable in a bang‐bang manner between the two values images with images . The situation corresponds to Eq. 27.8, where images and the relaxation term acts locally on a single qubit while all the remaining qubits undergo no relaxation. This paves the way to entirely new domains, since the reachable sets enlarge dramatically: if in addition to unitary control there is nonunital switchable (amplitude damping) noise on a single spin (images for images of the form of Eq. 27.7), one finds that the controlled system can act (approximately) transitively on the entire set of density operators, whereas for unital (bit‐flip) switchable noise on a single spin (images ), the reachable set fills all density operators majorized by the initial state.

More precisely, one gets the following:

Needless to say, these physically mild extensions by bang‐bang dissipative control on a single qubit on top of unitary control have a significant impact on numerical optimal control of open quantum systems by implementation into our numerical optimal‐control package DYNAMO (56) (see also Chapter 28) giving explicit control sequences (55) for superconducting qubits coupled to an open transmission line in the explicit experimental setting (GMon) of the Martinis group (57), which (by its short bath correlation) complies well with the Markov conditions.

Next we illustrate the distinction between gradient flows for (i) abstract optimizations on (possibly constrained) reachable sets and (ii) dynamic optimal control via experimentally accessible control amplitudes in a given parameterization.

27.4 Relation to Numerical Optimal Control

While in the previous sections optimizations are treated in an abstract manner, that is, over the dynamic group or over the specific state‐space manifold given by the reachable set (as illustrated in Figure 27.3), quantum engineering takes the optimization problems into the concrete parameterization of the actual experimental setup. More precisely, the parameterization is made in terms of the (discretized) control amplitudes, which then steer the quantum system on the state‐space manifold as an intermediate level. This is illustrated in Figure 27.4 in order to show the distinction from Figure 27.3.

Image described by caption and surrounding text.

Figure 27.4 Optimal control task: the quality function images is driven into a (local) maximum on the reachable set images by following an implicit procedure (intermediate panel). It is brought about by a gradient flow images on the level of experimental control amplitudes images (lower traces) where standard gradient‐assisted methods apply as also described in Chapter 28.

Building upon (58,59), recently we have lined up all the principle numerical algorithms into a unified programming framework DYNAMO (56) matched to solve the underlying bilinear control problems: subject to the equation of motion 27.1 a target function images is maximized over all admissible piece‐wise constant control vectors images . This turns a control vector (pulse sequence) from an initial guess into an optimized shape by following first‐order gradients (or second‐order increments) to all the time slices of the control vector as shown in Figure 27.3, which may be done sequentially (6063), or concurrently ( 58, 59) or in the newly unified version DYNAMO allowing hybrids as well as switches on the fly from one scheme to another one (56).

These numerical schemes have been put to good use for steering quantum systems (in the explicit experimental parameter setting) such as to optimize (i) the transfer between quantum states (pure or nonpure) (58), (ii) the fidelity of a unitary quantum gate to be synthesized in closed systems ( 59,64), (iii) the gate fidelity in the presence of Markovian relaxation (65), and (iv) the gate fidelity in the presence of non‐Markovian relaxation (66).

In recent years, examples for spin systems ( 59, 64) as well as Josephson elements (64) have been illustrated in all detail. For optimizing quantum maps in open systems, time‐optimal controls have been compared to relaxation‐optimized controls (65) in the light of an algebraic interpretation (49).

27.5 Outlook on Infinite‐Dimensional Systems

In infinite dimensions, difficulties arise as – by unbounded operators – the group images of unitaries on an infinite‐dimensional separable Hilbert space images is no Lie group if equipped with the strong topology for studying quantum dynamics. Yet images contains infinite‐dimensional subgroups with proper Lie structure – including in particular a Lie algebra images consisting of unbounded operators and a well‐defined exponential map, for example, unitaries with abelian images ‐symmetry, which in the Jaynes–Cummings model relates to a particle‐number operator.

In (67), this infinite‐dimensional system Lie algebra images is exploited for control theory in infinite dimensions in analogy to the finite‐dimensional case. The symmetry of images and its Lie group images thus excludes full controllability, yet this problem is overcome by complementary methods directly on the group level. The approach is paradigmatic and can be generalized in a natural way to other abelian symmetries (i.e., images and images representations with images ).

For several two‐level atoms interacting with one harmonic oscillator (e.g., a cavity mode or a phonon mode), the methods of (67) allow for extending previous results substantially, mainly in two aspects also summarized in Table 27.4: (i) We answer approximate control and convergence questions for asymptotically vanishing control error. (ii) Our results include not only reachability of states but also its operator lift, that is, simulability of unitary gates. To this end, (67) introduces the notion of strong controllability, and shows that all systems under consideration require only a fairly small set of control Hamiltonians for guaranteeing strong controllability, that is, simulability. – Thus we anticipate the methods of (67) briefly sketched here will find wide application to systematically characterize experimental setups of cavity QED and ion‐traps in terms of pure‐state controllability and simulability.

27.5.1 Controllability and its Symmetry Conditions

More precisely, the control of quantum systems poses considerable mathematical challenges when applied to infinite dimensions. Basically, they arise from the fact that the set of anti‐selfadjoint operators (recall Stone's Theorem (68), VIII.4] to see they are generators of strongly continuous, unitary one‐parameter groups) do neither form a Lie algebra nor even a vector space. On the group level, the group of unitaries equipped with the strong operator topology is a topological group yet not a Lie group. Therefore, whenever strong topology has to be invoked, controllability cannot be assessed via a system Lie algebra. Thus, in these cases, we address the challenges on the group level by employing the controlled time evolution of the quantum system in order to approximate unitary operators, the action of which is measured with respect to arbitrary, but finite sets of vectors. This is formalized in the notion of strong controllability introduced in (67) generalizing the notion of pure‐state controllability in the literature. Central are abelian symmetries: assuming that except one, all Hamiltonians observe such an abelian symmetry, the infinite‐dimensional control system can be analyzed in its block‐diagonalized basis. We obtain strong controllability (beyond pure‐state controllability) if one of the Hamiltonian breaks this abelian symmetry and some further technical conditions are fulfilled.

27.5.1.1 Time Evolution

We treat control problems

27.10 equation

where the images with images are selfadjoint control Hamiltonians on an infinite‐dimensional, separable Hilbert space images and images are piecewise‐constant controls. As images is infinite‐dimensional, if not bounded the images are defined on a dense subspace, that is, the domain images .

27.5.1.2 Pure‐state controllability

A key issue is reachability: given two pure states images , one looks for a time images and controls images such that images . In infinite dimensions, this condition is too strong, as there are states that can only be reached in infinite time, if at all. Yet, one may find a reachable state arbitrarily “close by.” Hence images shall be called reachable from images if for all images there is a finite time images and controls images with images . Therefore, system 27.10 is called pure‐state controllable, if every pair of pure states images can be images ‐interconverted04.

27.5.1.3 Strong Controllability

Analogously, for unitaries images , time images , and controls images , one can approximate a target images in the strong sense by

equation

that is, comparing images and images only on a finite set of states with worst‐case error bounded by images . System 27.10 is called strongly controllable if every unitary images can be thus approximated05. Clearly, strong controllability implies pure‐state controllability.

27.5.1.4 The Dynamical Group images

Strong controllability is conceptually related to the strong operator topology [ (68), VI.1] on the group images of unitary operators on images : The sets

equation

form a neighborhood base in the strong topology, hence called (strong) images ‐neighborhood. Therefore, strong controllability says: any images ‐neighborhood of images contains a time‐evolution operator images for appropriate time images and control functions images , or in other words: images is an accumulation point of the set images of all unitaries images . The set of all accumulation points of images (which contains images itself) is a strongly closed subgroup06 of images , which we will call the dynamical group images generated by control Hamiltonians images with images . For piecewise constant controls with only one images different from zero at each time, images is just the smallest strongly closed subgroup of images that contains all images for all images and all images . It contains in particular all unitaries that can be written as a strong limit s‐images . In finite dimensions, images can be calculated via its system algebra images generated by the images , as each images can be written as images for an images .

In infinite dimensions, one can avoid problems of joint domains by going back to finite‐dimensional Lie algebras with a dense set of analytic vectors (70,71) or to study systems with bounded generators in images (72). Yet first way comes at the expense of loosing full controllability, while the second is unphysical. Thus here we take an approach by splitting generators into two classes. The first images generators images admit an abelian symmetry and can be treated – with Lie‐algebra methods – along the lines outlined next. Secondly, the last generator images breaks this symmetry and achieves full controllability by a simple argument.

27.5.1.5 Abelian Symmetries

To study control systems with symmetries, consider the case of a images ‐symmetry07, that is, a (strongly continuous) unitary representation images of the abelian group images on images . It can be written in terms of a selfadjoint operator images with pure point spectrum consisting of (a subset of) images as images . By the eigenprojection of images to the eigenvalue images denoted as images (allow images if images is no eigenvalue of images ), we get a block‐diagonal decomposition of images in the symmetry‐adapted basis

27.11 equation

and we can rewrite images again as images .

Here two assumptions (with substantial loss of generality) facilitate subsequent discussion:

  1. All eigenvalues of images shall be of finite multiplicity, that is, the images be finite‐dimensional.
  2. All eigenvalues of images be non‐negative. (This assumption can be relaxed at certain points.)

By finite‐dimensionality of images , assumption (1) makes the space of finite particle vectors images a “good” domain for basically all unbounded operators in this section and one gets the following theorem:

The idea is to cut off the decomposition 27.11 at sufficiently high images without sacrificing the strong approximation, that is, images increases with decreasing error. This strategy allows for tracing many calculations back to finite‐dimensional Lie algebras.

Next, consider a subgroup of images and its Lie algebra relating unitaries with determinant one to traceless generators. As images need neither be bounded nor positive, general definitions of tracelessness and determinant may run into problems circumvented by the block diagonal decomposition of 27.11, where all images and images are infinite sums of operators08 with images , images , and images denoting the projection onto the images ‐eigenspace images . As all images and images are operators on finite‐dimensional vector spaces, one can define

equation

images is a (strongly closed) subgroup of images and images is a Lie subalgebra of images . The image of images under the exponential map coincides with images , which is effectively an infinite direct product of groups images , not just the “special” subgroup of images .

27.5.1.6 Breaking the Symmetry

For a fully controllable system, one has to leave the group images “represented” block diagonal in Figure 27.5 a by adding control Hamiltonians that break the symmetry by a complementary direct sum decomposition of images , where images , with images are projections onto the subspaces images and should satisfy images . For images , we thus introduce the nonzero projections images , whereas for images , the relation images shall hold.

We write images for the overlap of images and images , which can (in contrast to images ) be equal to zero for all images . The images are projections onto the subspaces images satisfying images .

Image described by caption and surrounding text.

Figure 27.5 (a) Block structure of operators in images (dark gray) and of operators complementary to images (light gray) in the case where the projection images vanishes. (b) Energy diagram for the Jaynes–Cummings model (here two atoms in a cavity under individual controls images and images ) with combined atom–cavity transitions matching the block structure of (a) given in dark gray (see Eqs. (24.11 and 24.13)) as commuting with images or images below, and complementary transitions solely within the atoms given in light gray (see Eqs. (24.12 and 24.14)).

This definition entails an important controllability result:

27.5.2 Application to Jaynes–Cummings Systems

Exploiting controlled dynamics of quantum systems is of increasing importance not only for solving computational tasks but for both quantum communication and simulation ( 4, 8 7376) including many‐body correlations to create “quantum matter.” Ultra‐cold atoms in optical lattices model large‐scale correlations ( 76,77), where tunability and control over system parameters allow for switching between low‐energy states of different quantum phases ( 3,78) or for following real‐time dynamics such from the super‐fluid to the Mott insulator regime (79). Manipulating several atoms in a cavity is a key step to this end (80) posing challenging infinite‐dimensional control problems. While in finite dimensions controllability can readily be assessed by the Lie‐algebra rank condition ( 14 18), infinite‐dimensional systems are more intricate (81), as exact controllability seemed daunting ( 70, 71,82,83), before approximate controllability paved the way to realistic assessments (8486), see also (87) and references therein.

Here we illustrate control systems of two‐level atoms coupled to a cavity mode, that is, the Jaynes–Cummings model (8891). We build on symmetry arguments ( 20, 26) and apply appropriate operator topologies for assessing (i) to which extent pure states can be interconverted and (ii) unitary gates can be approximated with arbitrary precision thus going beyond previous work (9295).

Table 27.4 Controllability results for several two‐level atoms in a cavity as derived in (67).

System Control Hamiltonians Controllability
System Algebra images , Dynamic Group images
One atom images ,images , Eq. 27.12 images ,images Theorem 3.110
images ,images , Eq. 27.12 Strongly controllable09
images , Eq. 27.13 with images Theorem 3.210
images atoms images ,images images and
with individual controls of Eq. 27.14 images Theorem 3.310
images ,images Strongly controllable09
with individual controls Eqs. 27.14 and 27.15 with images Theorem 3.410
images atoms images ,images images and
under collective control of Eq. 27.16 images Theorem 3.510
images ,images images and
under collective control of Eq. 27.18 images Theorem 3.610
images ,images Strongly controllable09
under collective control of Eq. 27.18 with images Theorem 3.710

Here in the strong topology, no system algebra or exponential map exists.

The theorems are given with reference to (67).

27.5.2.1 One Atom

In one atom, the Hilbert space of the system is given by images and the dynamics is described by the well‐known Jaynes–Cummings Hamiltonian (88):

27.12 equation

where images with images are the Pauli matrices (images ), images denote the annihilation and creation operator, and images is the number operator. The charge‐type operator images (determining the block structure) then takes the form images . To get a fully controllable system, one has to add a Hamiltonian that breaks the symmetry, for example, by

27.13 equation

so that transitions in the two‐level system are driven by images in the sense of images ‐pulses.

27.5.2.2 Many Atoms with Individual Controls

In this case, the Hilbert space of the system is readily generalized to images where images denotes the number of atoms. The control Hamiltonians become

27.14 equation

where images and images . As depicted by the dark gray parts in Figure 27.5, all the images are invariant under the symmetry defined by the charge operator images where images denotes again the number operator. To get strong controllability, one has to add again one Hamiltonian, where again a images ‐flip of one atom is sufficient (see the light gray parts in Figure 27.5), since

27.15 equation

is complementary to images .

27.5.2.3 Many Atoms Under Collective Control

Now one may modify the setup from the last section by considering again images atoms interacting with one mode, but assuming that one can control the atoms only collectively rather than individually. Instead of the Hamiltonians images and images with images of Eq. 27.14, one only has their sums

27.16 equation

where images and images , combined with the free evolution

27.17 equation

of the cavity. The best result so far is to replace the operators from Eqs. 27.16 and 27.17 by

27.18 equation

The operators images with images commute with images and generate the Lie algebra images . In addition, we have images , which is complementary to images .

27.6 Conclusion

We have cast a number of recent results into context to sketch an overarching frame of an emerging quantum systems theory. In particular, the unifying Lie picture comes for bilinear control systems of closed and open systems. This is of eminent importance also for control engineering and steering quantum dynamical systems with high precision. In doing so, we have shown how the emerging quantum systems theory links to many applications in quantum simulation and control without sacrificing mathematical rigor. Beyond addressing optimization tasks on reachable sets and state‐space manifolds, we have pointed out how gradient flows form the missing link to numerical optimal control algorithms for explicit steerings (control amplitudes) for manipulating closed and open (Markovian and non‐Markovian) systems in finite dimensions as, e.g., in Chapter 28.

Finally, we gave an outlook on a Lie picture of a systems and control theory in infinite dimensions and its application to Jaynes–Cummings systems, for example, like atoms in a cavity.

Acknowledgments

This work has been supported in part by the EU program SIQS, the exchange with COQUIT, moreover by the Bavarian excellence network ENB via the International Doctorate Programme of Excellence Exploring Quantum Matter (EXQM) as well as by the Deutsche Forschungsgemeinschaft (DFG) in the collaborative research center SFB 631 as well as the international research group FOR 1482 supported via the grant SCHU 1374/2‐1. Moreover, R.Z. was funded by DFG under grant Gl 203/7‐2.

Exercises

  1. 27.1 [] [] [] Lie Algebras in Quantum Dynamics

    A Lie algebra is a vector space images over some field images endowed with a mapping images , where

    1. [images ] is linear, that is, images , for all images ,
    2. [images ] is antisymmetric11, that is, images ,
    3. [images ] obeys Jacobi's identity: images .

    Show that

    1. a set of finite‐dimensional Hamiltonians images of a bilinear control system 27.1 generates the dynamic system Lie algebra images (in the sense of Section 27.2.1) by the linear span over the Lie closure (taking mutual commutators until no new linearly independent matrices are generated) written images ;
    2. the centralizer to images in a larger matrix Lie algebra images , when given via all matrices in images simultaneously commuting with all images , forms a Lie algebra itself, which is written images and is a subalgebra to images ;
    3. the centralizer thus expresses the joint symmetries to all images and to the entire system algebra images in the sense of Section 27.2.1 ;
    4. the system algebra images of the Hamiltoninans images above generates the dynamic system Lie group images , when images is surjective (we assume that images is compact and connected); and thus the reachable set of all density operators under the bilinear control system given by the images takes the form of the subgroup orbit images as in Section 27.2.1 ;
    5. for a bilinear quantum control system, the set of all reachable expectation values to an observable images (or more generally to a non‐Hermitian detection operator images ) geometrically boils down to the projection of the reachable set in (d) onto the operator images since for all images , images where the latter coincides with the relative images ‐numerical range images of Section 27.2.3 (set images ).
  2. 27.2 Lie Algebras in Spin and Pseudospin Systems

    Based on the definition in Exercise 1, show that

    1. the Pauli matrices (images ) generate a Lie algebra that is images ;
    2. (images ) is a Lie algebra isomorphic to images ; interpret the relation between the Bloch equation images and the Liouville equation images ;
    3. the unit quaternions images (with images , images and images ) give rise to a group images that is isomorphic to images ;
    4. the Heisenberg algebra images and the oscillator algebra images , where images are Lie algebras; (images : quantum dynamics expressable by finite‐dimensional Lie algebras can often be solved algebraically, see Wilcox (96) or Sattinger and Weaver (97);
    5. in images spins‐1/2, the entire Lie algebra images can be generated by commutation of the Pauli matrices on each spin and Ising terms images if the nonvanishing images can be represented as vertices of an arbitrary connected graph;
    6. (e) is equivalent to saying that a system of images spins‐1/2 is fully controllable and there is a universal set of quantum gates that can be realized (compare Section 27.2.1 ).
  3. 27.3# 27.3# Spin: Recommended Advanced Reading on Foundations

    In proper terms, the spin is defined as the quantum angular momentum images that has to be added to the orbital angular momentum images so that the total angular momentum images is invariant under Lorentzian—and already Galilean!—transformation.

    Convince yourself by reading:

    1. a first simplified introduction in W. Greiner, Theoretical Physics Vol. 4 (Chap. 13) to see that spin arises naturally from linearizing the equation of motion;
    2. the famous originals of Dirac (98) for the Lorentz invariance;
    3. Lévy‐Leblond (99) and Varadarajan (100), (Chap. IX.8) for Galilei invariance;
    4. the amusing story on Bohr's train trip to Leiden in December 1925 where he discussed spin–orbit coupling with Einstein and Ehrenfest (in: Pais (101) p 303 f).

    Do you now see why spin already follows from Galilei invariance, whereas spin–orbit coupling invokes Lorentz invariance?

References

  1. 1 Dowling, J.P. and Milburn, G. (2003) Quantum technology: the second quantum revolution. Philos. Trans. R. Soc. London, Ser. A, 361, 1655.
  2. 2 Glaser, S.J., Boscain, U., Calarco, T., Koch, C.P., Köckenberger, W., Kosloff, R., Kuprov, I., Luy, B., Schirmer, S., Schulte‐Herbüggen, T., Sugny, D., and Wilhelm, F.K. (2015) Training Schröodinger's cat: quantum optimal control. Eur. Phys. J. D, 69, 279.
  3. 3 Sachdev, S. (1999) Quantum Phase Transitions, Cambridge University Press, Cambridge.
  4. 4 Feynman, R.P. (1982) Simulating physics with computers. Int. J. Theor. Phys., 21, 467.
  5. 5 Abrams, D. and Lloyd, S. (1997) Simulation of many‐body Fermi systems on a quantum computer. Phys. Rev. Lett., 79, 2586.
  6. 6 Bennett, C.H., Cirac, I., Leifer, M.S., Leung, D.W., Linden, N., Popescu, S., and Vidal, G. (2002) Optimal simulation of two‐qubit Hamiltonians using general local operations. Phys. Rev. A, 66, 012305.
  7. 7 Dodd, J.L., Nielsen, M.A., Bremner, M.J., and Thew, R.T. (2002) Universal quantum computation and simulation using any entangling Hamiltonian and local unitaries. Phys. Rev. A, 65, 040301.
  8. 8 Jané, E., Vidal, G., Dür, W., Zoller, P., and Cirac, J. (2003) Simulation of quantum dynamics with quantum optical systems. Quantum Inf. Comput., 3, 15.
  9. 9 Levine, W.S. (ed.) (1996) The Control Handbook, CRC Press, Boca Raton, FL in cooperation with IEEE Press.
  10. 10 Sontag, E. (1998) Mathematical Control Theory, Springer, New York.
  11. 11 Elliott, D. (2009) Bilinear Control Systems: Matrices in Action, Springer, London.
  12. 12 Dirr, G. and Helmke, U. (2008) Lie theory for quantum control. GAMM‐Mitteilungen, 31, 59.
  13. 13 Kalman, R., Falb, P.L., and Arbib, M.A. (1969) Topics in Mathematical System Theory, McGraw‐Hill, New York.
  14. 14 Sussmann, H. and Jurdjevic, V. (1972) Controllability of nonlinear systems. J. Differ. Equ., 12, 95.
  15. 15 Jurdjevic, V. and Sussmann, H. (1972) Control systems on Lie groups. J. Differ. Equ., 12, 313.
  16. 16 Brockett, R.W. (1972) System theory on group manifolds and coset spaces. SIAM J. Control, 10, 265.
  17. 17 Brockett, R.W. (1973) Lie theory and control systems defined on spheres. SIAM J. Appl. Math., 25, 213.
  18. 18 Jurdjevic, V. (1997) Geometric Control Theory, Cambridge University Press, Cambridge.
  19. 19 Schulte‐Herbrüggen, T., Dirr, G. and Zeier, R. (2017) Quantum systems theory viewed from Kossakowski‐Lindblad Lie Semigroups – and Vice Versa, Open Sys. Information Dyn., 24, p. 1740019.
  20. 20 Zeier, R. and Schulte‐Herbrüggen, T. (2011) Symmetry principles in quantum system theory. J. Math. Phys., 52, 113510; note addendum: (2014) J. Math. Phys., 55, 129901.
  21. 21 MacKay, W.G. and Patera, J. (1981) Tables of Dimensions, Indices, and Branching Rules for Representations of Simple Lie Algebras, Marcel Dekker, New York.
  22. 22 Polack, T., Suchowski, H., and Tannor, D.J. (2009) Uncontrollable quantum systems. Phys. Rev. A, 79, 053403.
  23. 23 Obata, M. (1958) On subgroups of the orthogonal group. Trans. Am. Math. Soc., 87, 347.
  24. 24 (a) Dynkin, E.B. (1957) Maximal subgroups of the classical groups. Am. Math. Soc. Transl., Ser. 2, 6, 245; (b) Reprinted in Dynkin, E.B. (2000) Selected Papers of E. B. Dynkin with Commentary, American Mathematical Society and International Press, pp. 37–170.
  25. 25 Burgarth, D., Maruyama, K., Montangero, S., Calarco, T., Noi, F., and Plenio, M. (2009) Scalable quantum computation via local control of only two qubits. Phys. Rev. A, 81, 040303.
  26. 26 Zimborás, Z., Zeier, R., Keyl, M., and Schulte‐Herbrüggen, T. (2014) A dynamic systems approach to fermions and their relation to spins. EPJ Quantum Technol., 1, 11.
  27. 27 Zeier, R. and Zimborás, Z. (2015) On squares of representations of compact Lie algebras. J. Math. Phys., 56, 081702.
  28. 28 Zeier, R., Zimborás, Z., Schulte‐Herbrüggen, T., and Burgarth, D. (2015) Symmetry criteria for quantum simulability of effective interactions. Phys. Rev. A, 92, 042309.
  29. 29 Li, C.‐K. (1994) ‐numerical ranges and ‐numerical radii. Linear Multilin. Algebra, 37, 51.
  30. 30 Cheung, W.‐S. and Tsing, N.‐K. (1996) The C‐numerical range of matrices is star‐shaped. Linear Multilin. Algebra, 41, 245.
  31. 31 Schulte‐Herbrüggen, T., Dirr, G., Helmke, U., Kleinsteuber, M., and Glaser, S. (2008) The significance of the ‐numerical range and the local ‐numerical range in quantum control and quantum information. Linear Multilin. Algebra, 56, 3.
  32. 32 Dirr, G., Helmke, U., Kleinsteuber, M., and Schulte‐Herbrüggen, T. (2008) Relative ‐numerical ranges for applications in quantum control and quantum information. Linear Multilin. Algebra, 56, 27.
  33. 33 Schulte‐Herbrüggen, T. (1998) Aspects and prospects of high‐resolution NMR. PhD thesis. Diss‐ETH 12752, Zürich.
  34. 34 Gawron, P., Puchała, Z., Miszczak, J.A., Skowronek, L., and Życzkowski, K. (2010) Restricted numerical range: a versatile tool in the theory of quantum information. J. Math. Phys., 51, 102204.
  35. 35 Puchała, Z., Miszczak, J.A., Gawron, P., Dunk, C.F., Holbrook, J.A., and Życzkowski, K. (2012) Restricted numerical shadow and geometry of quantum entanglement. J. Phys. A, 45, 415309.
  36. 36 Brockett, R.W. (1988) Dynamical systems that sort lists, diagonalise matrices, and solve linear programming problems. Proceedings of IEEE Decision Control, Austin, Texas, USA, 1988, p. 779; see also: (1991) Linear Algebra Appl., 146, 79.
  37. 37 Helmke, U. and Moore, J.B. (1994) Optimisation and Dynamical Systems, Springer‐Verlag, Berlin.
  38. 38 Bloch, A. (ed.) (1994) Hamiltonian and Gradient Flows, Algorithms and Control, Fields Institute Communications, American Mathematical Society, Providence, NJ.
  39. 39 Glaser, S.J., Schulte‐Herbrüggen, T., Sieveking, M., Schedletzky, O., Nielsen, N.C., Sørensen, O.W., and Griesinger, C. (1998) Unitary control in quantum ensembles: maximising signal intensity in coherent spectroscopy. Science, 280, 421.
  40. 40 Chou, M.T. and Driessel, K.R. (1990) The projected gradient method for least‐squares matrix approximations with spectral constraints. SIAM J. Numer. Anal., 27, 1050.
  41. 41 Absil, P.A., Mahony, R., and Sepulchre, R. (2008) Optimization Algorithms on Matrix Manifolds, Princeton University Press, Princeton, NJ.
  42. 42 Chu, M.T. (2008) Linear algebra algorithms as dynamical systems. Acta Numer., 17, 1.
  43. 43 Schulte‐Herbrüggen, T., Glaser, S.J., Dirr, G., and Helmke, U. (2010) Gradient flows for optimization in quantum information and quantum dynamics: foundations and applications. Rev. Math. Phys., 22, 597.
  44. 44 Brockett, R.W. (1993) Differential geometry and the design of gradient algorithms. Proc. Symp. Pure Math., 54, 69.
  45. 45 Li, C.K., Poon, Y.T., and Schulte‐Herbrüggen, T. (2011) Least‐squares approximation by elements from matrix orbits achieved by gradient flows on compact Lie groups. Math. Comput., 275, 1601.
  46. 46 Uhlmann, A. (1971) Sätze über Dichtematrizen. Wiss. Z. Karl‐Marx‐Univ. Leipzig, Math. Nat. R., 20, 633.
  47. 47 Ando, T. (1989) Majorization, doubly stochastic matrices, and comparison of eigenvalues. Linear Algebra Appl., 118, 163–248.
  48. 48 Yuan, H. (2010) Characterization of majorization monotone quantum dynamics. IEEE. Trans. Autom. Control, 55, 955.
  49. 49 Dirr, G., Helmke, U., Kurniawan, I., and Schulte‐Herbrüggen, T. (2009) Lie semigroup structures for reachability and control of open quantum systems. Rep. Math. Phys., 64, 93.
  50. 50 O'Meara, C., Dirr, G., and Schulte‐Herbrüggen, T. (2012) Illustrating the geometry of coherently controlled unital quantum channels. IEEE Trans. Autom. Control (IEEE‐TAC), 57, 2050; see also largely extended e‐print: http://arXiv.org/pdf/1103.2703 (2011).
  51. 51 Wolf, M.M. and Cirac, J.I. (2008) Dividing quantum channels. Commun. Math. Phys., 279, 147.
  52. 52 Kossakowski, A. (1972) On quantum statistical mechanics of non‐Hamiltonian systems. Rep. Math. Phys., 3, 247.
  53. 53 Gorini, V., Kossakowski, A., and Sudarshan, E. (1976) Completely positive dynamical semigroups of ‐level systems. J. Math. Phys., 17, 821.
  54. 54 Lindblad, G. (1976) On quantum statistical mechanics of non‐Hamiltonian systems. Commun. Math. Phys., 48, 119.
  55. 55 Bergholm, V., Wilhelm, F.K., and Schulte‐Herbrüggen, T. (2016) Arbitrary ‐Qubit State Transfer Implemented by Coherent Control and Simplest Switchable Local Noise. e‐print: http://arXiv.org/pdf/1605.06473 (accessed 10 November 2017).
  56. 56 Machnes, S., Sander, U., Glaser, S.J., de Fouquiéres, P., Gruslys, A., Schirmer, S., and Schulte‐Herbrüggen, T. (2011) Comparing, optimising and benchmarking quantum control algorithms in a unifying programming framework. Phys. Rev. A, 84, 022305.
  57. 57 Chen, Y., Neill, C., Roushan, P., Leun, N., Fang, M., Barends, R., Kelly, J., Campbell, B., Chen, Z., Chiaro, B., Dunsworth, A., Jeffrey, E., Megrant, A., Mutus, J.Y., O'Malley, P.J.J., Quintana, C.M., Sank, D., Vainsencher, A., Wenner, J., White, T.C., Geller, M.R., Cleland, A.N., and Martinis, J.M. (2014) Qubit architecture with high coherence and fast tunable coupling. Phys. Rev. Lett., 113, 220502.
  58. 58 Khaneja, N., Reiss, T., Kehlet, C., Schulte‐Herbrüggen, T., and Glaser, S.J. (2005) Optimal control of coupled spin dynamics: design of NMR pulse sequences by gradient ascent algorithms. J. Magn. Reson., 172, 296.
  59. 59 Schulte‐Herbrüggen, T., Spörl, A.K., Khaneja, N., and Glaser, S.J. (2005) Optimal control‐based efficient synthesis of building blocks of quantum algorithms: a perspective from network complexity towards time complexity. Phys. Rev. A, 72, 042331.
  60. 60 Krotov, V.F. and Feldman, I.N. (1983) Iteration method of solving the problems of optimal control. Eng. Cybern., 21, 123; Russian original: (1983) Izv. Akad. Nauk. SSSR Tekh. Kibern., 52, 162.
  61. 61 Krotov, V.F. (1996) Global Methods in Optimal Control, Marcel Dekker, New York.
  62. 62 Sklarz, S.E. and Tannor, D.J. (2006) Quantum computation via local control theory: direct sum vs. direct product Hilbert spaces. Chem. Phys., 322, 87.
  63. 63 Singer, K., Poschinger, U., Murphy, M., Ivanov, P., Ziesel, F., Calarco, T., and Schmidt‐Kaler, F. (2010) Trapped ions as quantum bits: essential numerical tools. Rev. Mod. Phys., 82, 2609.
  64. 64 Spörl, A.K., Schulte‐Herbrüggen, T., Glaser, S.J., Bergholm, V., Storcz, M.J., Ferber, J., and Wilhelm, F.K. (2007) Optimal control of coupled Josephson qubits. Phys. Rev. A, 75, 012302.
  65. 65 Schulte‐Herbrüggen, T., Spörl, A., Khaneja, N., and Glaser, S.J. (2011) Optimal control for generating quantum gates in open dissipative systems. J. Phys. B, 44, 154013.
  66. 66 Rebentrost, P., Serban, I., Schulte‐Herbrüggen, T., and Wilhelm, F.K. (2009) Optimal control of a qubit coupled to a non‐Markovian environment. Phys. Rev. Lett., 102, 090401.
  67. 67 Keyl, M., Zeier, R., and Schulte‐Herbrüggen, T. (2014) Controlling several atoms in a cavity. New J. Phys., 16, 065010.
  68. 68 Reed, M. and Simon, B. (1990) Methods of Modern Mathematical Physics: Functional Analysis, vol. I, Academic Press, San Diego, CA.
  69. 69 Halmos, P. (1982) A Hilbert Space Problem Book, Springer, New York.
  70. 70 Huang, G.M., Tarn, T.J., and Clark, J.W. (1983) On the controllability of quantum‐mechanical systems. J. Math. Phys., 24, 2608.
  71. 71 Lan, C., Tarn, T.J., Chi, Q.S., and Clark, J.W. (2005) Analytic controllability of time‐dependent quantum control systems. J. Math. Phys., 46, 052102.
  72. 72 Lang, S. (1996) Differential and Riemannian Manifolds, 2nd edn, Springer, New York.
  73. 73 Vidal, G. and Cirac, I. (2002) Optimal simulation of nonlocal Hamiltonians using local operations and classical communication. Phys. Rev. A, 66, 022315.
  74. 74 Wocjan, P., Rötteler, M., Janzing, D., and Beth, T. (2002) Universal simulation of Hamiltonians using a finite set of control operations. Quantum Inf. Comput., 2, 133.
  75. 75 Zeier, R., Grassl, M., and Beth, T. (2004) Gate simulation and lower bounds on the simulation time. Phys. Rev. A, 70, 032319.
  76. 76 Lewenstein, M., Sanpera, A., and Ahufinger, V. (2012) Ultracold Atoms in Optical Lattices: Simulating Quantum Many‐Body Systems, Oxford University Press, Oxford.
  77. 77 Bloch, I., Dalibard, J., and Zwerger, W. (2008) Many‐body physics with ultracold gases. Rev. Mod. Phys., 80, 885.
  78. 78 Carr, L. (ed.) (2010) Understanding Quantum Phase Transitions, CRC Press, Boca Raton, FL.
  79. 79 Greiner, M., Mandel, O., Esslinger, T., Hänsch, T.W., and Bloch, I. (2002) Quantum phase transition from a superfluid to a Mott insulator in a gas of ultracold atoms. Nature, 415, 39.
  80. 80 Haroche, S. and Raimond, J.M. (2006) Exploring the Quantum: Atoms, Cavities, and Photons, Oxford University Press, Oxford.
  81. 81 Li, X. and Yong, J. (1995) Optimal Control Theory for Infinite Dimensional Systems, Birkhäuser, Boston, MA.
  82. 82 Turinici, G. and Rabitz, H. (2003) Wavefunction controllability for finite‐dimensional bilinear quantum systems. J. Phys. A, 36, 2565.
  83. 83 Wu, R.B., Tarn, T.J., and Li, C.W. (2006) Smooth controllability of infinite‐dimensional quantum‐mechanical systems. Phys. Rev. A, 73, 012719.
  84. 84 Adami, R. and Boscain, U. (2005) Controllability of the Schrödinger equation via intersection of eigenvalues. Proceedings of the 44th IEEE Conference on Decision Control CDC CD ROM.
  85. 85 Chambrion, T., Mason, P., Sigalotti, M., and Boscain, U. (2009) Controllability of the discrete‐spectrum Schrödinger equation driven by an external field. Ann. Inst. Henri Poincaré (C), 26, 329.
  86. 86 Boscain, U., Gauthier, J.P., Rossi, F., and Sigalotti, M. (2015) Approximate controllability, exact controllability, and conical eigenvalue intersections for quantum mechanical systems. Commun. Math. Phys., 333, 1225.
  87. 87 Borzi, A. (2011) Quantum optimal control using the adjoint method. Nanoscale Syst. Math. Model. Theory Appl., 1, 93.
  88. 88 Jaynes, E.T. and Cummings, F.W. (1963) Comparison of quantum and semiclassical radiation theories with application to the beam maser. Proc. IEEE, 51, 89.
  89. 89 Tavis, M. and Cummings, F.W. (1968) Exact solution for an N‐molecule‐radiation‐field Hamiltonian. Phys. Rev., 170, 379.
  90. 90 Tavis, M. and Cummings, F.W. (1969) Approximate solution for an N‐molecule‐radiation‐field Hamiltonian. Phys. Rev., 188, 692.
  91. 91 Brecha, R.J., Rice, P.R., and Xiao, M. (1998) two‐level atoms in a driven optical cavity: quantum dynamics of forward photon scattering for weak incident fields. Phys. Rev. A, 59, 2392.
  92. 92 Rangan, C., Bloch, A.M., Monroe, C., and Bucksbaum, P.H. (2004) Control of trapped‐ion quantum states with optical pulses. Phys. Rev. Lett., 92, 113004.
  93. 93 Brockett, R.W., Rangan, C., and Bloch, A.M. (2003) The controllability of infinite quantum systems. Proceedings of the 42nd IEEE Conference on Decision Control CDC, p. 428.
  94. 94 Bloch, A., Brockett, R.W., and Rangan, C. (2010) Finite controllability of infinite‐dimenisonal quantum systems. IEEE Trans. Autom. Control, 49, 1797.
  95. 95 Yuan, H. and Lloyd, S. (2007) Controllability of the coupled spin‐ harmonic oscillator system. Phys. Rev. A, 75, 052331.
  96. 96 Wilcox, R.M. (1967) Exponential operators and parameter differentiation in quantum physics. J. Math. Phys., 8, 962.
  97. 97 Sattinger, D.H. and Weaver, O.L. (1986) Lie Groups and Algebras with Applications to Physics, Geometry and Mechanics, Springer, New York.
  98. 98 Dirac, P.A.M. (1928) The quantum theory of the electron. Proc. R. Soc. London, Ser. A, 117, 610; (1927) 118, 351.
  99. 99 Lévy‐Leblond, J.M. (1967) Nonrelativistic particles and wave equations. Commun. Math. Phys., 6, 286.
  100. 100 Varadarajan, V.S. (1985) Geometry of Quantum Theory, Springer, New York.
  101. 101 Pais, A. (2000) The Genius of Science, Cambridge University Press, Cambridge, p. 303.

Notes

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

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