Entanglement in Bipartite Quantum Systems with Fast Local Unitary Control
Abstract.
The well-known Schmidt decomposition, or equivalently, the complex singular value decomposition, states that a pure quantum state of a bipartite system can always be brought into a “diagonal” form using local unitary transformations.
In this work we consider a finite-dimensional closed bipartite system with fast local unitary control.
In this setting one can define a reduced control system on the singular values of the state which is equivalent to the original control system.
We explicitly describe this reduced control system and prove equivalence to the original system.
Moreover, using the reduced control system, we prove that the original system is controllable and stabilizable and we deduce quantum speed limits.
We also treat the fermionic and bosonic cases in parallel, which are related to the Autonne–Takagi and Hua factorization respectively.
Keywords. Bipartite entanglement, reduced control system, local unitary control, controllability, stabilizability, quantum speed limits, singular value decomposition, Autonne–Takagi factorization, Hua factorization
MSC Codes. 81Q93, 15A18, 81P42, 93B05
1. Introduction
1.1. Motivation
Entanglement is one of the distinguishing features of quantum mechanics, and it lies at the core of emerging quantum technologies. Generating sufficient entanglement is a prerequisite for pure state quantum computation [28], it is at the root of quantum cryptography [5, 11] and it can act as a resource for increasing the sensitivity of quantum sensors [7]. At the same time many basic questions about entanglement remain unanswered, and our theoretical understanding of multipartite and mixed state entanglement is limited [4]. One setting in which entanglement is well-understood is the case of a bipartite quantum system in a pure state, where the singular value decomposition (a.k.a. the Schmidt decomposition) can be employed. Similar tools can also be applied to indistinguishable subsystems [15] (bosons and fermions), where the singular value decomposition is to be replaced with the Autonne–Takagi and Hua factorization respectively.
In this paper we apply quantum control theory [8, 6] to closed bipartite quantum systems subject to fast local unitary control. Factoring out the fast controllable degrees of freedom leads to a reduced control system defined on the singular values of the state. Such reduced control systems have been addressed in [1, Ch. 22] under simplifying assumptions. In particular, there, the reduced state space is assumed to be a smooth manifold without singularities. In our case singularities occur where two singular values coincide or one singular value vanishes, and in fact they constitute the main technical difficulty. In [20] this assumption is removed such that we can apply the results therein to the present setting.
Similar ideas have been applied to open Markovian quantum systems with fast unitary control in [25, 31, 23] and by the author et al. in [22]. There, the reduced control system describes the evolution of the eigenvalues of the density matrix (a positive semidefinite matrix of unit trace) representing a mixed quantum state. In quantum thermodynamics, a natural simplification of this system has been explored in [9, 29, 24, 30].
Finally, in an upcoming paper [21] we will apply the methods derived here combined with optimal control theory to derive explicit control solutions for low dimensional systems.
1.2. Outline
We start by defining the bilinear control system describing a closed bipartite quantum system with fast local unitary control in Section 2.1. We consider the case of distinguishable subsystems, as well as that of bosonic and fermionic subsystems. In Section 2.2 we show that these systems are related to certain matrix diagonalizations (which themselves are related to symmetric Lie algebras) and we derive corresponding reduced control systems in Section 2.3. In Section 2.4 we show how some controllability assumptions can be weakened if one neglects the global phase of the quantum state. Then we turn to some applications. In Section 3.1 we use the reduced control system to prove that the full bilinear control system is always controllable and stabilizable. In Section 3.2 we derive some quantum speed limits for the evolution of the singular values. The connection between the present quantum setting and symmetric Lie algebras is drawn in Appendix A, and this allows us to apply the results from [20].
2. Control Systems
First we define what we mean by a closed bipartite quantum system with fast local unitary control before we derive the equivalent reduced control system obtained by factoring out the local unitary action. The reduced control system describes the dynamics of the singular values of the state and due to the normalization of the quantum state the reduced state space turns out to be a hypersphere. The main result of this section is the equivalence, in a precise sense which will be made clear, of the full bilinear control system and the reduced one. In particular there is no loss of information incurred by passing to the reduced control system. A brief discussion of global phases concludes the section.
2.1. Bilinear Control System
Let two finite dimensional Hilbert spaces and of dimensions representing the subsystems be given. The total Hilbert space of the bipartite system is then .111Abstractly the symbol denotes the tensor product, but since we always work with concrete vectors and matrices we interpret as the Kronecker product. This identifies the vector spaces and and similarly for matrices. We denote by the unitary Lie algebra consisting of skew-Hermitian matrices in dimensions. Our goal is to study the full bilinear control system [16, 10] defined by the following controlled Schrödinger equation222Throughout the paper we set and thus write the (uncontrolled) Schrödinger equation as . on :
(B) |
where is the drift Hamiltonian (or coupling Hamiltonian), and are the control Hamiltonians, and and are the corresponding control functions. We make the following key assumptions:
-
(I)
The control functions and are locally integrable, in particular they may be unbounded.
-
(II)
The control Hamiltonians generate the full local unitary Lie algebra:
where is the Lie algebra of the Lie group of local unitary transformations. Put simply, we have fast control over .
Remark 2.1.
One may also define the group of local special unitary operations. Consider the local unitary . Then and hence, if , the value of the applied phase is restricted to a discrete set. Thus, if we do not neglect the global phase of the state , fast control over is not sufficient to generate all local unitary state transfers. To simplify the exposition we assume fast control over , but we will revisit this issue in Section 2.4 to show how this assumption can be weakened.
This covers the case of two distinguishable subsystems. However, we also wish to treat systems composed of two indistinguishable subsystems. In this case both subsystems have the same dimension . In the bosonic case, the state is unchanged by swapping the two subsystems, i.e., , where . These “symmetric” states lie in the space . In the fermionic case, swapping yields a phase factor of , i.e., . Such “skew-symmetric” states are contained in the space . In both cases the set of local unitaries applicable to the system is restricted to symmetric local unitaries . The corresponding Lie algebra is and is isomorphic to .333Note however that the map given by is a double cover with kernel . The set of all coupling Hamiltonians applicable to such systems is the set . Hence, in the case of two indistinguishable subsystems the bilinear control system takes the form:
(B’) |
where and . Assumption I remains unchanged, but Assumption II is slightly modified to state:
-
(III)
The local control Hamiltonians for generate the full Lie algebra .
Note that the only difference between the bosonic and fermionic case is that the initial state of the control system (B’) lies in and respectively.
Remark 2.2.
Similarly to Remark 2.1, one might consider control Hamiltonians of the form with the additional restriction of , defining the Lie algebra . In this case we again lose control over the global phase of the state. Thus, for simplicity, we do not make this assumption here and refer to Section 2.4 for more details.
2.2. Related Matrix Decompositions
In order to derive and understand the reduced control system (which we will introduce in the next section) obtained by factoring out the local unitary action, we must first understand the mathematical structure of this action. In the case of distinguishable subsystems, the local unitary action corresponds the complex singular value decomposition, an thus the only invariants of a state under this action are its singular values, which therefore are the natural choice for the reduced state. The bosonic and fermionic cases correspond to less well-known matrix decompositions called the Autonne–Takagi and Hua factorization respectively. Importantly all of these matrix decompositions also correspond to certain symmetric Lie algebras, and this is the key to applying the results on reduced control systems from [20] to the full bilinear control systems (B) and (B’).
Let and denote the standard orthonormal bases444In the following we will usually omit the index denoting the subsystem, since it will be clear form the order. of and respectively, and let be a state vector. The components of are uniquely given by
Hence every bipartite state can be represented by a matrix.555Our convention is consistent with [4, Sec. 9.2] and [15]. In other contexts one often defines the vectorization operation which turns a matrix into a vector by stacking its columns and satisfies . Our convention is slightly different in that, identifying via the Kronecker product, we may write . More precisely, we have used the canonical isomorphism
(1) |
where denotes the dual space. We will use to denote the matrix corresponding to under this isomorphism and vice versa. For distinguishable subsystems, the matrix representing is an arbitrary complex matrix in (the constraint induced by the normalization of the state will be discussed in Remark 2.3 below). For indistinguishable subsystems, it holds that , and the matrix is symmetric in the bosonic case and is skew-symmetric in the fermionic case.
Let be a local unitary, and set . The coordinates of are then defined by , and similarly . Then . In matrix form this can be rewritten as . Another way to state this is . Note that in the case of indistinguishable subsystems we have . This suggests a connection to certain matrix diagonalizations, namely:
-
•
The complex singular value decomposition in the distinguishable subsystems case (often referred to as the Schmidt decomposition in the context of quantum mechanics). It states that for any complex matrix there exist unitary matrices and such that is real and diagonal. The correspondence is established by
-
•
The Autonne–Takagi factorization in the bosonic case. It states that for any complex symmetric matrix , there is a unitary such that is real and diagonal. The correspondence is then given by
-
•
The Hua factorization in the fermionic case. It states that for any complex skew-symmetric matrix , there is a unitary such that is real and quasi-diagonal in the following sense: if is even, then is block diagonal with blocks of size , if is odd, there is an additional block of size in the lower right corner. Note that the quasi-diagonal matrix is still skew-symmetric, and so the diagonal is zero. The correspondence to local unitary state transformations is as in the bosonic case.
Note that the Autonne–Takagi factorization and the Hua factorization are special cases of singular value decompositions, and hence the resulting (quasi/̄)diagonal matrix will have the singular values on its (quasi/̄)diagonal. In the first case we denote by the subspace of “real diagonal states” corresponding to the set of real diagonal matrices under the isomorphism (1). Clearly has dimension . In the second case we write for the -dimensional subspace corresponding to the real diagonal matrices . Similarly, in the third case we write for the -dimensional subspace of states corresponding to the real (skew-symmetric) quasi-diagonal matrices . We will use the following maps to send the singular values to their corresponding (quasi/̄)diagonal state:
A convenient shorthand notation is resp. . We will always use the standard Euclidean inner product on , and on we will use the real part of the standard inner product. Then, due to the inclusion of the factor it holds that the maps above are -linear isometric isomorphisms. Furthermore we denote666The symbol is a combination of and . by the cone of states where the diagonal elements are non-negative and arranged in non-increasing order, and analogously we write for the quasi-diagonal states where the are non-negative and arranged in non-increasing order. The set resp. is called the Weyl chamber.
Conversely to the diagonal embeddings we also define the following orthogonal projections:
More precisely these are the orthogonal projections on and followed by and respectively.
Remark 2.3.
The normalization of the quantum state entails a normalization of the corresponding singular values. More precisely, the norm of the quantum state equals the Frobenius norm of the matrix , and hence has unit norm if and only if the singular values of satisfy . Hence the singular values define a point on the unit sphere of dimension in the indistinguishable and in the bosonic case. In the fermionic case there are again singular values and they lie on the unit sphere of dimension . However there is an additional restriction as the singular values come in pairs of opposite values. Taking only one singular value of each pair and multiplying it by (and ignoring the singular value in the odd dimensional case) we find that the resulting vector lies on the unit sphere of dimension . In all cases will call this the Schmidt sphere, denoted in the distinguishable case, in the bosonic case, and in the fermionic case. The maps and then yield isometric embeddings of the Schmidt sphere into and respectively. The Schmidt sphere will be the state space of our reduced control system, which we will define in the following section. The Weyl chambers and then yield corresponding Weyl chambers in the Schmidt sphere , , and .
Remark 2.4.
Often one considers the Schmidt values, which are the squares of the singular values, within the standard simplex, which is then called the Schmidt simplex [4, Sec. 16.4]. This is easier to visualize, but in our case would lead to unnatural dynamics, which is why we will remain on the sphere. Note also that if one of the systems is a qubit, then the Schmidt sphere is a circle parametrized by the Schmidt angle [4, p. 440].
2.3. Reduced Control System
Due to Assumptions I and II (resp. III), we can move arbitrarily quickly within the local unitary orbits of the system if we ignore the drift term [10, Prop. 2.7]. With the drift this is still approximately true. In the previous section we have shown that using local unitary transformations, we can always obtain a state of (quasi/̄)diagonal form which is completely determined by the singular values of the state. In particular, within the bilinear control systems (B) and (B’) two states are effectively equivalent if and only if they have the same singular values (up to order and sign). This strongly suggests that there should exist a “reduced” control system, defined on the singular values — or rather the Schmidt sphere (cf. Remark 2.3). This is indeed the case. The reduced control system is defined in greater generality in [20, Sec. 2.1] using symmetric Lie algebras, which unify many well-known matrix diagonalizations, such as the ones encountered in the previous section. No knowledge of symmetric Lie algebras is presupposed here, but the connections are expounded in Appendix A.
Let us briefly motivate the definition of the reduced control system. Let be a solution to the full control system (B). Assume that the corresponding matrix can be diagonalized in a differentiable way as , and that it is regular777We say that is regular if its singular values are distinct and non-zero.. Here is the diagonal matrix with diagonal elements . Then by differentiating (cf. [19, Lem. 2.3]) we obtain that (and analogously or in the bosonic and fermionic cases) where
We call , and the induced vector fields. The collection of induced vector fields is denoted
Note that these are linear vector fields on , and respectively, and hence they can be represented as matrices in the respective standard basis. We will later see that these are indeed skew-symmetric matrices and that the corresponding dynamics preserve the Schmidt sphere. The following proposition gives the explicit expressions.
Proposition 2.5.
Let denote an arbitrary coupling Hamiltonian and let and for be given such that . Then the induced vector field on takes the form888Here denotes the Hadamard (elementwise) product of two matrices. If the (square) matrices are of different size the resulting matrix will have the size of the smaller one. Similarly denotes the elementwise imaginary part.
Now assume additionally that and that is symmetric. For bosonic systems, on , we obtain
For fermionic systems, on , we obtain
Proof.
To simplify notation we first consider a product Hamiltonian . Let be given. Let and compute
The case of general for distinguishable subsystems as well as the bosonic case follow by linearity.
Now let us consider the fermionic case. Again for simplicity we consider a product Hamiltonian since the general result follows by linearity. Then and can be seen as block matrices with blocks of size denoted and (and, in the odd-dimensional case, an additional row and column). Moreover let denote the standard symplectic form. Then we can compute in a similar fashion
This concludes the proof. ∎
Remark 2.6.
The fermionic case can be interpreted as follows. We consider the even dimensional case for simplicity. First define the matrices obtained from and by choosing all the odd indexed rows from and all the even indexed rows from . More precisely, if is odd and if is even. Then we divide the into blocks of size and we define a new matrix of half the size by replacing each block by the imaginary part of its determinant.
We see immediately that if the drift Hamiltonian is local, then the induced vector fields vanish:
Corollary 2.7.
If the drift Hamiltonian is local, meaning that , then for every . The same is true for and whenever and .
Finally we can define the reduced control systems:
Definition 2.8 (Reduced control systems).
Let be an interval of the form with or . We define three reduced control systems:
For distinguishable subsystems we set:
(R) |
A solution is an absolutely continuous path satisfying (R) almost everywhere for some measurable control function .
As mentioned previously, and shown in Lemma 2.11 below, the induced vector fields preserve the Schmidt sphere, and hence we may define the reduced control systems directly on the Schmidt sphere.
Remark 2.9.
There are several slightly different ways of defining the reduced control system which are given in [20, Sec. 2.1]. The most intuitive definition, given above, is to consider the control system where the control function is measurable and the solution is absolutely continuous. A more geometric definition uses the differential inclusion , where denotes the set of achievable derivatives at defined by . The differential inclusion is exactly equivalent to our definition by Filippov’s Theorem, cf. [26, Thm. 2.3]. Often it is convenient to consider a “relaxed” version of the differential inclusion where also convex combinations of achievable derivatives are allowed: . This slightly enlarges the set of solutions, but every solution to the relaxed system can still be approximated uniformly on compact time intervals by solutions to our system, cf. [2, Ch. 2.4, Thm. 2]. Analogous remarks also hold for the symmetric cases () and ().
The main result of [20] is the equivalence of the full bilinear control system (B) resp. (B’) and the reduced control system (R), resp. () or (), proven in Theorems 3.8 and 3.14 therein. In our case this specializes to the following result.
First we need to define the quotient maps and . Given a (possibly not normalized) vector (or ), the map yields the singular values of the corresponding matrix , chosen non-negative and arranged in non-increasing order. Similarly, for , the map yields the singular values of the skew-symmetric matrix , except that we keep only one singular value of each pair and multiply it by to keep the normalization. Note that when restricting the domain of and to (normalized) quantum states, the image will lie in the respective Schmidt sphere, and even in the Weyl chamber resp. and .
Recall that here and throughout the paper we use Assumptions I and II (resp. III), unless stated otherwise.
Theorem 2.10 (Equivalence Theorem).
Let be a solution on to the bilinear control system (B), and let be defined by . Then is a solution to the reduced control system (R).
Conversely, let be a solution to the reduced control system (R) with control function and let denote the corresponding state. Then can be approximated by solutions to the full control system (B) arbitrarily well. More precisely, for every there exists a solution to (B) such that , where denotes the supremum norm.
The analogous results, mutatis mutandis999Most results in this paper hold in all three cases with only minimal differences in notation, which are summarized in Table 1., also hold in the bosonic and fermionic cases, where the full control system is (B’) and the reduced control systems are () and () respectively.
Proof.
The proof is mostly a technicality, as we simply have to show that the full control systems (B) and (B’) and their respective reduced versions (R), () and () can be interpreted as control systems on certain symmetric Lie algebras. We focus on the case of distinguishable subsystems. The corresponding symmetric Lie algebra is that of type AIII. The isomorphisms and defined in Appendix A.1 translate the quantum setting into the Lie algebra setting. The results of the appendix then show that all of the conditions of Proposition A.1 are satisfied and that the reduced control system (R) indeed corresponds to the reduced control system in the Lie algebraic setting. Taken together this proves the equivalence in the distinguishable case. The other cases are entirely analogous. ∎
The Equivalence Theorem 2.10 shows that the full bilinear control system (B) resp. (B’) and the reduced control system (R), resp. () or (), contain essentially the same information. Hence for every control theoretic notion, such as controllability and stabilizability, there is a specialized equivalence result, see [20, Sec. 4] for an overview. As a first consequence we obtain:
Lemma 2.11.
The induced vector fields are skew-symmetric matrices:
In particular the Schmidt sphere in invariant.
Proof.
Let us also recall the equivalence of reachable sets here, which is arguably the most useful consequence. First we give the definitions of reachable sets in the reduced control system (R). The definitions for other control systems are entirely analogous. The reachable set of at time is defined as
for any . By we denote the all time reachable set of , and by we denote the reachable set of up to time .
The following result is an immediate consequence of the Equivalence Theorem 2.10 and [20, Prop. 4.3].
Proposition 2.12.
Let be given and assume that and satisfy101010Here denotes the element of whose elements are the absolute values of the elements of arranged in non-increasing order. . Then is holds that
In particular, the closures coincide:
The analogous result, mutatis mutandis, holds also for the bosonic and the fermionic cases.
The Equivalence Theorem 2.10 guarantees the existence of an approximate lift, but its proof also provides a way to find corresponding control functions. Under some additional assumptions we can give an explicit formula for the controls of an exact lift, see [20, Prop. 3.10]. In particular this requires the solution to be smooth and regular, and the controls of the bilinear system (B) resp. (B’) to linearly span the corresponding Lie algebra. Before stating the result we define some notation.
The group of local unitary operations and its symmetric counterpart act on the state spaces , and respectively. The corresponding infinitesimal action of the Lie algebras and can be determined as in Section 2.2 using the formula . In the special case where the state is regular and diagonal this function (more precisely its negative) gets a special name:
where and denote the (quasi-)diagonal matrices corresponding to and . Although these maps are not bijective, by restricting the domain to the orthocomplement of the kernel and the codomain to the image, inverse maps can be defined. Indeed, this is nothing but the Moore–Penrose pseudoinverse. Explicit expressions are given in Lemmas A.7, A.10 and A.13.
To use these inverse maps we have to understand the images of the maps , , and . It turns out that, for regular resp. , these images are exactly given by the orthocomplement of the diagonal subspaces . We denote the orthogonal projection on with kernel by , and use the same notation on . In the matrix picture, this map simple removes the real part of the diagonal elements of . Similarly, on , the orthogonal projection with kernel is denoted and it removes the real part of the quasi-diagonal of .
With these definitions [20, Prop. 3.10] can be specialized as follows:
Proposition 2.13.
Let be a solution to the reduced control system (R) with control function . Assume that is regular and that is continuously differentiable. Let and let be given by
Then satisfies . The analogous result also holds mutatis mutandis for bosonic and fermionic systems.
We call the second term in the definition of in Proposition 2.13 the compensating Hamiltonian since it compensated for the local Hamiltonian action induced by the drift term .
2.4. Global Phases
Mathematically the quantum state has a global phase which is physically undetectable and hence may be considered irrelevant. We keep the phase throughout this paper for convenience, noting that the global phase is removed automatically in the reduced control system. This section briefly discusses how Assumptions II and III can be slightly weakened by neglecting the global phase. We denote by and the Lie algebras obtained from and by requiring the trace to vanish. Consider the following two weakened controllability assumptions:
-
(II’)
The control Hamiltonians generate the local special unitary Lie algebra:
-
(III’)
The control Hamiltonians generate the symmetric local special unitary Lie algebra:
In both cases adding the (symmetric) local control Hamiltonian is sufficient to obtain the stronger Assumptions II and III respectively. Since this Hamiltonian commutes with everything, the only effect of adding or removing the corresponding term from the control system is a change in the global phase. More concretely, if our control system is (B) (resp. (B’)) but only satisfies Assumptions I and II (resp. III), then we can add the control Hamiltonian so that it satisfies Assumption II (resp. III). Now we can compute any solution in this extended system and we obtain a corresponding solution in the actual system by setting the control function of to zero. The resulting solution will, at all times, be equal to the solution of the extended system up to a global phase.
3. Some Applications
In the previous section we defined the reduced control systems (R), (), and () and proved that they are equivalent to the corresponding full control system (B) in the distinguishable case and (B’) in the bosonic and fermionic cases, cf. Theorem 2.10. In this section we will use the reduced control system to deduce some consequences on controllability, stabilizability and speed limits. Using the Equivalence Theorem 2.10 these results can then be translated to the full control system.
3.1. Controllability and Stabilizability
In this section we show that the reduced control system is always controllable and stabilizable. As a consequence, the full control system is also controllable and all states can be stabilized in a certain sense. For notational simplicity we focus on the case of distinguishable subsystems, noting that the bosonic and fermionic cases are entirely analogous.
First we lift the reduced control system (R) to the Lie group . Unless otherwise noted it is assumed that the initial state is . The operator lift can be defined by
(L) |
and analogously with and in the bosonic and fermionic cases. A solution is absolutely continuous and satisfies (L) almost everywhere for some measurable .
Remark 3.1.
The reduced control system is controllable on if for every two states it holds that and it is controllable on in time if for every two states we have that . Approximate controllability is defined in the same way except that one considers the closures of the respective reachable sets. The analogous definitions also holds for all other control systems.
To understand the properties of the operator lift, we study the set of generators . A key property of is that it is invariant under conjugation by the Weyl group , see [20, Lem. A.2] and Appendix A. This fact allows us to prove the following result.
Proposition 3.2.
The Weyl group acts irreducibly on . In particular, if the coupling Hamiltonian is not local, then the operator lift (L) is controllable. The analogous result holds, mutatis mutandis, in the bosonic and fermionic cases.
Proof.
To show that the Weyl group acts irreducibly, we start with an arbitrary non-zero element and show that the subrepresentation generated by is all of . If this is trivially true since is one-dimensional. So assume that . Consider the basis and let be the coefficients of in this basis. Since is non-zero, at least one of the coefficients is non-zero. Using a permutation in we may assume that . Let be the diagonal matrix whose diagonal equals everywhere except in the -th position, where it equals . Consider the matrix . Then and . Iterating this procedure with we obtain a multiple of , showing that lies in the subrepresentation generated by . From this, using the permutations in , all other basis elements can be obtained. This shows that the representation of is irreducible. Controllability of the operator lift then follows from [20, Prop. 4.19]. ∎
This result can now be lifted to the full bilinear system using the equivalence of the systems.
Theorem 3.3.
If the full control system (B) is controllable in time , then the reduced control system (R) is controllable in time on the Weyl chamber . Conversely, if the reduced control system (R) is controllable in time on the Weyl chamber , then the full control system (B) is controllable in time for all . Moreover there exists a finite time such that both systems are controllable in time . The analogous result holds, mutatis mutandis, in the bosonic and fermionic cases.
Proof.
Clearly if (B) is controllable in time , so is (R) on the Weyl chamber by Proposition 2.12. The same result shows that if (R) is controllable in time on the Weyl chamber, then (B) is approximately controllable in time . Proposition 3.2 shows in particular that (R) is directly accessible at every point (see [20, Sec. 4.6] for the definitions related to accessibility) and hence by [20, Prop. 4.15] the bilinear system (B) is accessible (on the set of normalized states) at every regular state. Then [16, Ch. 3 Thm. 2] implies that (B) is controllable in time for regular initial states. Since regular states are dense and since we can leave the set of non-regular states in an arbitrarily short amount of time , the full control system (B) is controllable in time . That (R) is controllable in finite time follows from Proposition 3.2. ∎
Now we turn to stabilizability. In accordance with the definitions given in [20, Sec. 4.3] we say that a state is stabilizable for (R) if . A direct consequence of Proposition 3.2 is that every state is stabilizable.
Corollary 3.4.
Every state is stabilizable for the reduced control systems.
Proof.
If is local the statement trivially holds. Otherwise choose some non-zero . Consider the uniform combination which is clearly -invariant. If , it is clear that . In higher dimensions -invariance and irreducibility of the action of (Proposition 3.2) again show that . The proof for the bosonic and fermionic cases is the same. ∎
If a state is stabilizable for (R), then in the bilinear control system (B), one can stay close to the local unitary orbit for arbitrary amount of time, cf. [20, Prop. 4.7].
Recall that a point is strongly stabilizable if there is such that . Specializing [20, Prop. 4.5] we obtain the following result.
Proposition 3.5.
Let and and set . If there is some satisfying , then and is strongly stabilizable.
Conversely let be strongly stabilizable and let be such that . Moreover assume that is regular. Then there is some such that where we again set . In fact one can choose
where is given explicitly in Lemma A.7. The analogous result holds, mutatis mutandis, in the bosonic and fermionic cases.
Note that the assumption on regularity is necessary in general, cf. [20, Ex. 3.12].
The local Hamiltonian in the previous result is called a compensating Hamiltonian, and indeed this is a special case of Proposition 2.13. Note that the expressions in Lemmas A.7, A.10 and A.13, and hence the compensating Hamiltonian, blow up as approaches a non-regular state.
The following result yields a simple special case in which strong stabilizability is easy to determine.
Lemma 3.6.
Let and assume that all commute or that all commute. Then there exists such that . In particular, in this case every state is strongly stabilizable. The analogous result holds, mutatis mutandis, in the bosonic and fermionic cases.
3.2. Speed Limit and Control Time
By speed limit we simply mean an upper bound on the velocity that any solution to the given control system can achieve. Note that the full control system (B) (resp. (B’)) does not have any such speed limit, since the controls may be unbounded, but, by construction, the reduced control system (R) (resp. () and ()) always admits a (finite) speed limit, cf. [20, Prop. 4.1].
For any matrix , we write for the largest singular value of . This is exactly the operator norm with respect to the usual Euclidean norm, and hence it is clear that for , the norm corresponds to the largest velocity that achieves on the unit sphere. This immediately yields the following result:
Lemma 3.7.
Let be any solution to (R). Then it holds that almost everywhere.111111The maximum exists and is achieved since the map is continuous on a compact domain. The analogous result holds, mutatis mutandis, in the bosonic and fermionic cases.
Hence we need to find a good upper bound for over all .
Lemma 3.8.
Let . Then
where denotes the Frobenius norm. The same bound holds a fortiori for and .
Proof.
The Frobenius norm and the spectral norm are related by , see [13, Prob. 5.6.P23]. Using the Cauchy-Schwarz inequality we compute for any that
This concludes the proof in the case of distinguishable subsystems. The bound continues to hold in the bosonic and fermionic cases since restricting the drift or the controls cannot lead to faster evolution of the singular values. ∎
To obtain an lower limit on the time needed to reach any target state from any initial state, we also need to know the largest distance between any pair of points. This is the diameter of the space121212Note that distances in the reduced state space are computed as the length of the shortest geodesic joining two points on the sphere. Hence, somewhat unintuitively, the diameter of a unit hypersphere is , which is the distance between two antipodal points., and due to the Weyl group symmetry every state has an equivalent state in the Weyl chamber. Hence we are particularly interested in the diameter of the Weyl chamber, which is given in the following result.
Lemma 3.9.
Consider the unit sphere embedded in and let be the Weyl group acting by coordinate reflections and permutations. Then the corresponding Weyl chamber has diameter .
Proof.
First recall that the shortest distance on the sphere between two points is given by . The maximal distance in the Weyl chamber is achieved by two of its corners. Then it is clear that these points are and for the standard Weyl chamber . The result follows immediately. ∎
The control time of a control system is the shortest (infimum) time sufficient to reach any state from any other state. Using the upper bound on the speed of a solution, the lower bound on the diameter of the Weyl chamber and Theorem 3.3, we can give a lower bound on the control time:
Acknowledgments
I would like to thank Frederik vom Ende, Thomas Schulte-Herbrüggen and Gunther Dirr for valuable and constructive comments during the preparation of this manuscript. The project was funded i.a. by the Excellence Network of Bavaria ENB under the International PhD Programme of Excellence Exploring Quantum Matter (ExQM) and by the Munich Quantum Valley of the Bavarian State Government with funds from Hightech Agenda Bayern Plus.
Appendix A Relation to Symmetric Lie Algebras
In the main text we have shown that the local unitary actions on bipartite quantum states correspond to certain matrix diagonalizations, and we have stated that they themselves are related to certain symmetric Lie algebras. In this appendix we make these relations explicit and give all the relevant formulas. For a compact overview of the relation of symmetric Lie algebras to matrix diagonalizations see [19, Tab. 2].
Since we want to define a reduced control system on the singular values, a key question is how the singular values change in time. More precisely, given a differentiable path of matrices , what can we say about the derivative of the singular values? This question is made more complicated by the fact that the order and signs of the singular values are not unique (and if they are chosen in a unique way, they are not guaranteed to be differentiable). These issues can be resolved, and in fact one can do so in the more general setting of semisimple orthogonal symmetric Lie algebras, see [19] and in particular [20, Ex. 1.1 & 1.2]. We will recall and adapt the pertinent results as necessary.
The reduction of control systems was proven in detail in [20] in the setting of semisimple orthogonal symmetric Lie algebras. In order to rigorously prove the Equivalence Theorem 2.10, we need to show how the control systems considered here can be interpreted as control systems in such symmetric Lie algebras. First we need the following generalization of the equivalence results proven in [20].
Proposition A.1.
Let be an -dimensional real inner product space, let be a compact Lie group with Lie algebra acting on , and let be a linear vector field on . Consider the control system
(2) |
with fast and full control on . Moreover assume that we have a semisimple orthogonal symmetric Lie algebra with associated pair and maximal Abelian subspace . Let be a linear isometric isomorphism and a surjective Lie group homomorphism such that
(3) |
Then the control system is equivalent in the sense of Theorem 2.10 to the following reduced control system on :
(4) |
where , with the orthogonal projection onto , the inclusion, and the pullback.
Proof.
Let is the surjective Lie algebra homomorphism corresponding to . By surjectivity there are such that for all . By differentiating (3) we get . Let be the drift vector field on . If , then
The remainder of the proof is split into two parts, the projection and the lift.
We begin with the projection. Let be a solution to (2). We want to show that , defined by , is a solution to (4). We get that
almost everywhere. Hence is a solution of the corresponding control system on , and we may apply [20, Thm. 3.8] to obtain that is a solution of the reduced control system on , more explicitly, for almost every there is some such that . Next we show that solves (4). Indeed for the same we obtain by linearity of that for some . This concludes the projection part of the proof.
Conversely, assume that we have a solution to the reduced system (4). Again we find for almost all some such that , and so solves the corresponding control system on . Hence there exists a corresponding control function which is measurable. Using [20, Thm. 3.14] we find approximate lifted solutions with . As above we can compute with
and see that is a solution to (2). Since is an isometry, for any measurable lift of along (which exists due to [19, Lem. 2.29]) we get
it is also an -approximation. This concludes the proof. ∎
See Table 1 for an overview of the notation related to the different control systems.
General | Distinguishable | Bosonic | Fermionic | Lindbladian | |
Full System | (2) | (B) | (B’) | (B’) | — |
Ambient space | |||||
State space | |||||
State | |||||
Fast Controls | |||||
Drift | |||||
action | |||||
Reduced System | (4) | (R) | () | () | — |
Ambient space | |||||
State space | |||||
State | |||||
Weyl group | |||||
Inclusion | |||||
Projection | |||||
Quotient map | |||||
Induced vector fields | |||||
Natural wedge | — | ||||
Decomposition | — | Complex SVD | Autonne–Takagi fact. | Hua fact. | Hermitian EVD |
Lie Algebra | — | AIII | CI | DIII | A |
A.1. Complex Singular Value Decomposition (Type AIII)
The complex singular value decomposition is encoded by the symmetric Lie algebra of type AIII, see for instance [17, App. A.7] and [12, Ch. X §2.3]. The standard matrix representation of this Lie algebra is the indefinite special unitary Lie algebra with Cartan decomposition where131313Often one denotes .
A choice of corresponding compact Lie group is .
Remark A.2.
In general we consider (semi)simple orthogonal symmetric Lie algebras and so we also have to provide a “compatible” inner product on . The inner product is (up to some irrelevant scaling) uniquely defined using the Killing form on , cf. [12, Ch. V, Thm. 1.1]. Due to simplicity of the Killing form is (again up to scaling) given by . In the following we will set the inner product on to and on to , and and are orthogonal to each other. Furthermore, we always use the real inner product on states in .
The spaces and are identified using the map
Lemma A.3.
The map is an -linear141414Note that in the Lie algebraic context we always work with real vector spaces, even if their standard representation involves complex numbers. isometric isomorphism. The subspace is maximal Abelian and . The Weyl group is isomorphic to the generalized permutations and is a Weyl chamber.
Proof.
It is clear that is an -linear isomorphism. With the inner product on and defined as in Remark A.2 a simple computation shows that is even an isometry:
That is maximal Abelian is well-known, cf. [19, Tab. 2]. The fact that is an isometry also proves that . That the Weyl group acts by generalized permutations follows from the fact that the singular values are unique up to order and sign and the fact that any generalized permutation can be implemented by choosing and appropriately. ∎
Moreover we define the following maps:
Note that is the derivative of at the identity.
Lemma A.4.
It holds that is a Lie group isomorphism, and so is a Lie algebra isomorphism151515Contrary to , the map is not an isometry with respect to the inner products of Remark A.2.. For , and , the isomorphism and satisfy the compatibility conditions
(5) |
Similarly we have the correspondence of the infinitesimal action and of the induced vector fields where if then . Moreover the map corresponds to the quotient map with image in the Weyl chamber , see [19, Sec. 2.1].
Proof.
Even though does not always lie in , we can choose such that , and this phase disappears in the tensor product and in the adjoint representation. Indeed any satisfying will do. Hence is well defined, and one easily verifies that it is an isomorphism. The compatibility condition (5) follows from a simple computation using the isomorphism of Section 2.2. Similarly the corresponding Lie algebra isomorphism can be written as to show explicitly that it is well defined. By definition . Using the compatibility condition (5) this immediately yields as desired. Recall that we defined and in [20, Sec. 2.1] we defined . (We are slightly abusing notation here by writing .) Using the compatibility condition (5), and , the claim follows from a simple computation. The claim about the quotient map is just a restatement of the uniqueness of the singular values. ∎
Remark A.5.
Explicitly (5) states that . In a semisimple orthogonal symmetric Lie algebra every element in can be mapped into using the group action of , cf. [19, Lem. A.26], which one might call “diagonalization”. In our case this means that can be mapped to some element in . This exactly corresponds to the complex singular value decomposition. Note however that in the Lie algebra setting we have and hence there is an additional restriction on the determinants of and .
Remark A.6.
For regular we have defined the map in the main text and Lemma A.4 shows that it is related to the adjoint representation (hence the name). Denoting by and the orthogonal complement of the commutant of in and respectively, it turns out that the restriction becomes bijective and hence invertible, see [19, Prop. 2.14]. In fact this inverse is simply the Moore–Penrose pseudoinverse. Moreover it holds that is just the orthocomplement of . Hence, the pseudoinverse is defined on with image in . Note however that is not an isometry and so the orthocomplement has to be calculated in . The precise definition is given in the following lemma.
Lemma A.7.
Let be regular. Then it holds that , and the map can be described explicitly as
where and whenever and , and for and we get
(6) |
If resp. we additionally have
Proof.
We only consider the case for simplicity. Consider the map where is the corresponding diagonal matrix. Then . Hence, when we invert the map above, we will assume that . Moreover we need to find the kernel of the map, i.e. solve for (for all or any regular ). This happens if , and for and for . The orthocomplement of the kernel is given by , and for and for . Using, for , that
we find
Finally for we find and for we get . ∎
Note that this lemma uniquely defines , although there is some freedom in the choice of and since we can shift some real multiple of the identity between them.
A.2. Autonne–Takagi Factorization (Type CI)
First discovered by Autonne [3] and Takagi [27], the Autonne–Takagi factorization [13, Sec. 4.4] states that for any complex symmetric matrix there exists a unitary matrix such that is real and diagonal. The diagonal elements are uniquely defined up to order and signs, and they are in fact the singular values of .
The corresponding symmetric Lie algebra is that of type CI, usually represented by the real symplectic Lie algebra , see [17, Sec. 4.3] and again [12, Ch. X §2.3]. The Cartan decomposition is given explicitly by
The corresponding state space isomorphism is given by
Lemma A.8.
The map is an -linear isometric isomorphism. The subspace is maximal Abelian and . The Weyl group is isomorphic to the generalized permutations and is a Weyl chamber.
Proof.
The proof is analogous to that of Lemma A.3 so we just compute the inner product on :
Hence is an isometry and this concludes the proof. ∎
Now consider the following maps
Lemma A.9.
The maps and are Lie isomorphisms satisfying the compatibility conditions
(7) |
As in Lemma A.4 we get the correspondence of infinitesimal action, induced vector fields and quotient map.
Proof.
As in Remark A.5, the relation to the Autonne–Takagi factorization can be seen from (7), which explicitly states that , and from the fact that .
As described in Remark A.6 we can explicitly compute the appropriate inverse of the map . Note that in this case we have .
Lemma A.10.
Let be regular. Then it holds that , and the map can be explicitly described as
where
Proof.
Let . Then we have and . Hence . To invert the second equation we compute
and hence taking sum and difference we get
This concludes the proof. ∎
A.3. Hua Factorization (Type DIII)
A skew-symmetric version of the Autonne–Takagi factorization also exists [13, Coro. 4.4.19]. It is called the Hua factorization, and was originally proven in [14, Thm. 7]. It states that for every skew-symmetric complex matrix there exists a unitary such that is real and block diagonal with skew-symmetric blocks of size . If is odd then there is an additional block containing a zero. We call such matrices quasi-diagonal. Each block is determined by a single real number (and its negative) which taken together yield the singular values of .
This matrix factorization is related to the symmetric Lie algebra of type DIII, usually represented by the , see [17, App. A.6] and again [12, Ch. X §2.3].161616Note that [12] uses a different but isomorphic matrix representation. In this case we have the following Cartan like decomposition
Clearly the state space isomorphism is
Lemma A.11.
The map is an -linear isometric isomorphism. The subspace is maximal Abelian and . The Weyl group is isomorphic to the generalized permutations and is a Weyl chamber.
Proof.
The proof is entirely analogous to that of Lemma A.3. ∎
The isomorphisms on the Lie group and algebra level are:
Lemma A.12.
The maps and are Lie isomorphisms satisfying the compatibility conditions
(8) |
As in Lemma A.4 we get the correspondence of infinitesimal action, induced vector fields and quotient map.
Proof.
Since the map is well-defined, and it is clearly a Lie group isomorphism. The remainder of the proof is analogous to that of Lemma A.4. ∎
The relation to the Hua factorization can be seen form (8), which becomes , and the fact that .
Lemma A.13.
Let be regular. It holds that . The map takes the following form:
where for we get
where and indexes the blocks of the respective matrices, and . By we denote the value satisfying . In the case where is odd we additionally have
Proof.
First consider the even-dimensional case. Let . For we compute the bocks as well as
It follows that , and hence we find that
Here we used that the for to lie in the diagonal blocks must be real multiples of the identity. If is odd there is an additional column at the bottom and row on the right of to be determined. We find that
The claimed results follow immediately. ∎
References
- [1] Agrachev, A., and Sachkov, Y. Control Theory from the Geometric Viewpoint. Encyclopaedia of Mathematical Sciences. Springer, Heidelberg, 2004.
- [2] Aubin, J.-P., and Cellina, A. Differential Inclusions: Set-Valued Maps and Viability Theory. Springer, Berlin Heidelberg, 1984.
- [3] Autonne, L. Sur les Matrices Hypohermitiennes et sur les Matrices Unitaires. Ann. Univ. Lyon Nouvelle Ser. 38, 1 (1915).
- [4] Bengtsson, I., and Życzkowski, K. Geometry of Quantum States: An Introduction to Quantum Entanglement, 2nd ed. Cambridge University Press, Cambridge, 2017.
- [5] Bennett, C. H., and Brassard, G. Quantum Cryptography: Public Key Distribution and Coin Tossing. Theoretical Computer Science 560 (2014), 7–11. Theoretical Aspects of Quantum Cryptography – celebrating 30 years of BB84.
- [6] D’Alessandro, D. Introduction to Quantum Control and Dynamics, 2nd ed. Chapman and Hall/CRC, New York, 2021.
- [7] Degen, C. L., Reinhard, F., and Cappellaro, P. Quantum Sensing. Rev. Mod. Phys. 89 (2017), 035002.
- [8] Dirr, G., and Helmke, U. Lie Theory for Quantum Control. GAMM-Mitteilungen 31 (2008), 59–93.
- [9] Dirr, G., vom Ende, F., and Schulte-Herbrüggen, T. Reachable Sets from Toy Models to Controlled Markovian Quantum Systems. Proc. IEEE Conf. Decision Control (IEEE-CDC) 58 (2019), 2322.
- [10] Elliott, D. Bilinear Control Systems: Matrices in Action. Springer, London, 2009.
- [11] Gisin, N., Ribordy, G., Tittel, W., and Zbinden, H. Quantum Cryptography. Rev. Mod. Phys. 74 (2002), 145–195.
- [12] Helgason, S. Differential Geometry, Lie Groups, and Symmetric Spaces. Academic Press, New York, 1978.
- [13] Horn, R., and Johnson, C. Matrix Analysis, 2nd ed. Cambridge University Press, Cambridge, 2012.
- [14] Hua, L. K. On the Theory of Automorphic Functions of a Matrix Variable. I: Geometrical Basis. Amer. J. Math. 66 (1944), 470–488.
- [15] Huckleberry, A., Kuś, M., and Sawicki, A. Bipartite Entanglement, Spherical Actions, and Geometry of Local Unitary Orbits. J. Math. Phys. 54, 2 (2013), 022202.
- [16] Jurdjevic, V. Geometric Control Theory. Cambridge University Press, Cambridge, 1997.
- [17] Kleinsteuber, M. Jacobi-Type Methods on Semisimple Lie Algebras: A Lie Algebraic Approach to the Symmetric Eigenvalue Problem. PhD thesis, TU Munich, 2006.
- [18] Liu, W. An Approximation Algorithm for Nonholonomic Systems. SIAM J. Control Optim. 35 (1997), 1328–1365.
- [19] Malvetti, E., Dirr, G., vom Ende, F., and Schulte-Herbrüggen, T. Analytic, Differentiable and Measurable Diagonalizations in Symmetric Lie Algebras, 2022. arXiv:2212.00713.
- [20] Malvetti, E., Dirr, G., vom Ende, F., and Schulte-Herbrüggen, T. Reduced Control Systems on Symmetric Lie Algebras, 2023. arXiv:2307.13664.
- [21] Malvetti, E., and van Damme, L. Optimal Control of Bipartite Quantum Systems. In preparation.
- [22] Malvetti, E., vom Ende, F., Dirr, G., and Schulte-Herbrüggen, T. Reachability, Coolability, and Stabilizability of Open Markovian Quantum Systems with Fast Unitary Control, 2023. arXiv:2308.00561, submitted to SIAM J. Control Optim.
- [23] Rooney, P., Bloch, A., and Rangan, C. Steering the Eigenvalues of the Density Operator in Hamiltonian-Controlled Quantum Lindblad Systems. IEEE Trans. Automat. Contr. 63 (2018), 672–681.
- [24] Schulte-Herbrüggen, T., vom Ende, F., and Dirr, G. Exploring the Limits of Open Quantum Dynamics I: Motivation, First Results from Toy Models to Applications. In Proc. MTNS (2022), p. 1069. arXiv:2003.06018.
- [25] Sklarz, S., Tannor, D., and Khaneja, N. Optimal Control of Quantum Dissipative Dynamics: Analytic Solution for Cooling the Three-Level System. Phys. Rev. A 69 (2004), 053408.
- [26] Smirnov, G. Introduction to the Theory of Differential Inclusions. Amer. Math. Soc., Providence, Rhode Island, 2002.
- [27] Takagi, T. On an Algebraic Problem Related to an Analytic Theorem of Carathéodory and Fejér and on an Allied Theorem of Landau. Jap. J. Math. 83, 1 (1925).
- [28] Vidal, G. Efficient Classical Simulation of Slightly Entangled Quantum Computations. Phys. Rev. Lett. 91, 14 (2003), 147902.
- [29] vom Ende, F. Reachability in Controlled Markovian Quantum Systems: An Operator-Theoretic Approach. PhD thesis, TU Munich, 2020.
- [30] vom Ende, F., Malvetti, E., Dirr, G., and Schulte-Herbrüggen, T. Exploring the Limits of Controlled Markovian Quantum Dynamics with Thermal Resources. Open Syst. Inf. Dyn. 30, 1 (2023), 2350005.
- [31] Yuan, H. Characterization of Majorization Monotone Quantum Dynamics. IEEE Trans. Automat. Contr. 55 (2010), 955–959.