Brought to you by:
Paper

Quantum synchronization of the Schrödinger–Lohe model

and

Published 18 August 2014 © 2014 IOP Publishing Ltd
, , Citation Sun-Ho Choi and Seung-Yeal Ha 2014 J. Phys. A: Math. Theor. 47 355104 DOI 10.1088/1751-8113/47/35/355104

1751-8121/47/35/355104

Abstract

We present a quantum synchronization estimate of the Schrödinger–Lohe (S–L) model introduced by Lohe (2010 J. Phys. A: Math. Theor. 43 465301). The S–L model describes the dynamics of quantum oscillators on the nodes of a quantum network. When the coupling strength is positive and the maximal L2 distances between normalized initial wave functions are smaller than $\frac{1}{2}$, we show that the L2 distances between wave functions converge to zero exponentially fast. Our result generalizes earlier work by Chi et al (2014 J. Math. Phys. 55 052703) for the Lohe model.

Export citation and abstract BibTeX RIS

1. Introduction

The spontaneous emergence of synchronous behaviors in ensembles of weakly coupled oscillators is ubiquitous in our biological, physical, and human complex systems (e.g., two pendulum clocks suspended from the same bar, the flashing of fireflies, heart beating by pacemaker cells, the singing of crickets, and hand clapping in a concert hall; see [13, 22, 24]). These collective phenomena were first reported by Christiaan Huygens in the scientific literature during the middle of the 17th century. Since then, synchronization of weakly coupled oscillators has been studied intermittently by researchers until the middle of the 20th century. However, the first rigorous mathematical study of synchronization was begun only 40 years ago by Winfree [27] and Kuramoto [16, 17]. They introduced simple phase models for weakly coupled oscillators and showed how synchronization can emerge from the competing mechanisms between intrinsic randomness and nonlinear couplings. Recently, the study of the synchronization process of weakly coupled oscillators has emerged as an active research area because of its diverse engineering applications in, e.g., the formation and control of automatic robot systems [1, 23, 26].

Consider a quantum complete network consisting of N quantum nodes and quantum channels connecting all possible pairs of quantum nodes. We can view each quantum node as a component of a physical system interacting via quantum channels. For instance, atoms at nodes can have effective spin–spin interactions generated by single-photon pulses traveling along the channels [10] (see [15] for a short survey). If we assume that quantum oscillators are distributed over all quantum nodes and that the quantum states of oscillators are registered by the wave functions, then, in this setting, our motivating question is the following:

'Can we make wave functions at each node synchronize so that all wave functions become identical copies of each other?'

To fix the idea, let ψi be the wave function of the identical quantum oscillator at the ith node and the wave function ψ for the whole system be given by the tensor product $\otimes _{i=1}^{N}{{\psi }_{i}}$. In the absence of potential force, the Schrödinger–Lohe (S–L) model describes the spatial–temporal evolution of ψ and can be written as follows:

Equation (1.1)

Here $\langle \cdot ,\cdot \rangle $ is the standard L2 inner product in an infinite-dimensional Hilbert space, K is the positive coupling strength, and the Planck constant ℏ is assumed to be unity.

The S–L model (1.1) was first introduced by Lohe [18] as an infinite state generalization of the Lohe matrix model [19]. Note that the Lohe coupling term on the right-hand side (rhs) of (1.1) is sublinear so that the well-posedness of (1.1) can be studied by using the standard argument. Thus, in this paper, we will focus on the large-time behavior of (1.1). More precisely, we are interested in the question of whether or not the system (1.1) will exhibit a kind of synchrony asymptotically. Quantum synchronization has received much attention from the quantum optics community with its possible applications in quantum computing and quantum information [11, 12, 15, 20, 21, 25, 28, 29]. In the classical regime, synchronization for the oscillatory system has been studied via phase models, e.g., the Winfree model [26, 27] and Kuramoto model [16, 17]. In contrast, in the quantum mechanical regime, we do not yet have a well-developed theory. Recently, Lohe [18, 19] introduced a first-order matrix-valued ordinary differential equation (ODE) system that can be regarded as one possible quantum generalization of the Kuramoto phase model. In the Lohe model, all quantum oscillators have the same finite state space, whereas our S–L model can have an infinite state space.

The main result of this paper is to provide a framework that guarantees the asymptotic synchronization of the S–L model. More precisely, if the coupling strength K is positive and the maximal L2 distance between initial wave functions is smaller than $\frac{1}{2}$, we will show that

This certainly suggests the formation of quantum synchronization for the model (1.1) (see theorem 3.2).

The rest of this paper is organized as follows. In section 2, we briefly discuss the motivation for the model (1.1) and some basic properties of the S–L model. In section 3, we present our quantum synchronization estimates. Finally, section 4 is devoted to a brief summary of the main results and a discussion of future directions.

Notation: Let f and g be complex-valued $L_{x}^{2}({{\mathbb{R}}^{d}})$ functions. We set

and, for $a,b\in \mathbb{C}$, we set $a\cdot b=\bar{a}b$.

2. Preliminaries

In this section, we provide a brief introduction to the S–L model (1.1) following [18] and a key estimate for the solution splitting via the free Schrödinger equation and the quantum Lohe system that will be used in the quantum synchronization estimate in section 3.

2.1. The S–L model

Let Ui and $U_{i}^{+}$ be a unitary $d\times d$ matrix and its Hermitian conjugate, respectively, and let Hi be the $d\times d$ Hermitian matrix whose eigenvalues correspond to the natural frequencies of the oscillator at node i. Then, the Lohe model [18, 19] for quantum synchronization can be written as follows:

Equation (2.1)

For two special cases, the Lohe model (2.1) can be related to the Kuramoto model and the Schrödinger equation. For example, consider the one-dimensional case, d = 1. In this case, we write Uj and Hj as

Then, it is easy to see that (2.1) reduces to the Kuramoto model

However, for zero coupling, $K=0$, system (2.1) becomes the Schrödinger equation

Thus, the Lohe model (2.1) is a quantum generation of the Kuramoto model and the Schrödinger equation with finite quantum states. A natural question to ask then is 'If we allow each quantum particle to take an infinite quantum state, what happens to the system (2.1)?' In this case, we need to use the standard Schrödinger equation instead of the finite-dimensional one. Note that our purpose is to present a quantum model exhibiting the synchronization property. Thus, in some sense, we need to introduce a synchronization enforcing term into the standard Schrödinger equation. In this paper, we employ the Lohe synchronization for this flocking mechanism

Equation (2.2)

Note that, on the manifold $||{{\psi }_{i}}|{{|}_{2}}=1$, system (2.2) can be rewritten as

which has structural similarity to (2.1). For additional details, we refer to sections 3 and 6 of [18].

2.2. Solution splitting property of the S–L model

In this subsection, we show that the solution to the S–L system (1.1) can be given by the successive composition of the free Schrödinger equation and the Lohe system.

Consider the Cauchy problem for the free Schrödinger equation

Equation (2.3)

Define a one-parameter family of solution operator S(t) to (2.3) with an L2 initial datum ψ0 as follows:

Equation (2.4)

Then, it is easy to see that $S(t){{\psi }_{0}},\;t\gt 0,$ is an analytic function. In the following two lemmas, we study the basic properties of the solution operator S(t).

Lemma 2.1. Let $\psi =S(t){{\psi }_{0}}$ be the solution to (2.3) with L2x initial datum ψ0. Then, we have

Proof. It is the result of the Plancherel theorem. By the Fourier transform, we have

It follows from the above relation that

Therefore, we have

We next study the unitary property of the solution operator S(t).

Lemma 2.2.  (Unitary property of S(t)) Let S(t) be a solution operator given by (2.4). Then for any complex-valued $L_{x}^{2}({{\mathbb{R}}^{d}})$ functions ϕ10 and ϕ20, we have

Proof. Since equation (2.3) is linear, $S(t)({{\psi }_{10}}-{{\psi }_{20}})$ is the solution of (2.3) with initial datum ${{\psi }_{10}}-{{\psi }_{20}}$. Then, by lemma 2.1, we have

We take the square of the above relation and invoke L2 conservation to obtain the desired result.

We next consider L2 conservation of the space-dependent Lohe system

Equation (2.5)

Note that the spatial variable x acts as a parameter in (2.5). For a solution $\Psi =({{\psi }_{1}},...,\;{{\psi }_{N}})$ to (2.5) with initial data ${{\Psi }_{0}}={{\Psi }_{0}}(x)$, we set the solution operator L(t) to the system (2.5) as follows:

Lemma 2.3. The solution operator L(t) conserves an L2 norm:

Proof. Note that

We next state the essential key tool for our later analysis of the solution decomposition relation for the wave function that denotes the scattering property of system (1.1).

Proposition 2.1. Let $\Psi =({{\psi }_{1}},...,{{\psi }_{N}})$ be a solution to system (1.1) with L2 initial data Ψ 0 . Then, the solution ψj can be decomposed as the successive composition of S and L operators to initial datum ψj0 , i.e.,

Proof. Let $\Psi =({{\psi }_{1}},...,{{\psi }_{N}})$ be a solution to (1.1). Then, we first use $S{{(t)}^{-1}}=S(-t)$ to see that

Thus, to prove the desired result, it suffices to check that $S(-t){{\psi }_{j}}(x,t)$ satisfies the Lohe system (2.5). For this, we apply the linear solution operator $S(-t)$ to both sides of (1.1) to obtain

Equation (2.6)

  • (Estimate of the left-hand side): Note that the representation formula (2.4) yields
    We differentiate the above relation with respect to t to obtain
    Equation (2.7)
    where we used the fact that $K(x,-t)$ satisfies the backward free Schrödinger equation
  • (Estimate of the rhs): Since the solution operator $S(-t)$ is linear and unitary, we have
    Equation (2.8)

Finally, we combine the estimates (2.6)–(2.8) to obtain

However, since $S(-t){{\psi }_{i}}{{\mid }_{t=0}}={{\psi }_{i0}}$, by the uniqueness of (2.5), we have

2.3. From the S–L model to the Kuramoto model

We next study the relation between the S–L model (2.2) and the Kuramoto model which was observed in [18].

Consider normalized initial data with the same amplitude and different phases, i.e.,

Equation (2.9)

Then, with initial phases ${{\Theta }_{0}}=({{\theta }_{10}},...,{{\theta }_{N0}})$, we solve the Cauchy problem for the Kuramoto model

We then set

Equation (2.10)

Below, we verify that ${{\psi }_{j}}(x,t)$ constructed by relation (2.10) is a smooth solution to the S–L system. Because of proposition 2.1, it suffices to show that the functions ${{\tilde{\psi }}_{j}}(x,t)=\sqrt{\rho (x)}{{{\rm e}}^{{\rm i}{{\theta }_{j}}(t)}}$ satisfy the Lohe system (2.5).

It follows from direct calculations that ${{\tilde{\psi }}_{j}}$ satisfies

Equation (2.11)

and

Equation (2.12)

where we used $||\rho |{{|}_{1}}=1$ in (2.9).

Finally, we combine (2.11) and (2.12) to obtain

Thus, ${{\psi }_{j}}(t)=S(t){{\tilde{\psi }}_{j}}$ satisfies (1.1).

Note that in the above calculation, the order of ψj and ψk in the coupling term in (2.2) plays a key role. If we take the coupling term as

then the rhs of (2.12) becomes zero. Thus, (1.1) are equivalent with ${{\dot{\theta }}_{i}}=0$, i.e., there is no synchronization.

3. Quantum synchronization estimates

In this section, we present quantum synchronization estimates to the S–L model. Our quantum synchronization estimates yield that all wave functions converge to a common wave function with an exponential decay rate in L2 and Hk norms depending on the regularity of the initial data.

3.1.  L2 synchronization

We first introduce a Lyapunov functional measuring the degree of quantum synchronization. Given an ordered N-tuple of wave functions $\Psi =({{\psi }_{1}},...,{{\psi }_{N}})$, we set

Then, it is easy to see that the functional $D(\Psi )$ is Lipschitz continuous, so it is almost everywhere differentiable.

Definition 3.1. Let $\Psi =({{\psi }_{1}},...,{{\psi }_{N}})$ be a solution to system (1.1). Then the system (1.1) exhibits time-asymptotic quantum synchronization if and only if the following estimate holds

Equation (3.1)

Remark 3.1. Note that our quantum synchronization given by the relation (3.1) corresponds to the phase synchronization in classical case [13].

Before we proceed to estimate quantum synchronization, we make two remarks. First, in the absence of coupling, i.e., K = 0, the L2 distance between solutions to the S–L model is constant, i.e.,

Second, because of lemma 2.1 and proposition 2.1, we have

Thus, quantum synchronization of the S–L model reduces to synchronization of the Lohe system (2.5).

As mentioned, in this section, we consider the space-homogeneous S–L system. We next derive a Gronwall-type inequality for $D(\Psi )$ using conservation of the L2 norm along the flow (1.1).

Lemma 3.1. Suppose that the coupling strength and initial data satisfy

Then, for any solution $\Psi =({{\psi }_{1}},...,{{\psi }_{N}})$ to (2.5), we have

Proof. It follows from lemma 2.3 that

Then ψi satisfies

Therefore, we have

To estimate the rhs of the above relation, we divide it into four parts as follows:

We next provide the estimates for ${{\mathcal{I}}_{1i}},\;i=1,...,4$ separately.

  • (Estimates for ${{\mathcal{I}}_{11}}$ and ${{\mathcal{I}}_{13}})$: By direct calculation, we have
  • (Estimates for ${{\mathcal{I}}_{12}}$): Since $||{{\psi }_{i}}|{{|}_{2}}=1$, we have
  • (Estimates for ${{\mathcal{I}}_{14}}$): Similarly, we have

Finally, we combine all these estimates to obtain

Lemma 3.2. Suppose that the coupling strength and initial data satisfy

Then, for any solution $\Psi =({{\psi }_{1}},...,{{\psi }_{N}})$ to (2.5), $D(\Psi )$ satisfies

Proof. For $t\in {{\mathbb{R}}_{+}}$, we can take an index pair $({{i}_{t}},{{j}_{t}})$ such that

We now use lemma 3.1 to obtain

It follows from the definition of $D(\Psi )$ that

Then, lemma 3.2 yields the quantum synchronization estimate as follows.

Theorem 3.1. Suppose that the coupling strength and initial data satisfy

Then, for any solution $\Psi =({{\psi }_{1}},...,{{\psi }_{N}})$ to (2.5), the diameter $D(\Psi )$ satisfies

Proof. It follows from lemma 3.2 that we have

By the comparison principle of ODEs, we have

Note that the condition $D({{\Psi }_{0}})\lt \frac{1}{2}$ is needed to exclude the finite-time blowup of $D(\Psi (t))$.

Remark 3.2. (1). In section 6 of [18], two finite-dimensional reductions to the S–L model were discussed. For example, for some special class of initial data with the common amplitudes, the S–L model can be reduced to the Kuramoto model which exhibit synchronization for large coupling strength (see section 2.2). However, theorem 3.1 can be applied to initial data with distinct amplitudes.

(2). For the Kuramoto model, exponential synchronization has been extensively studied in [57, 9, 8, 13, 14]. As can be seen in section 2.3, the quantum model (1.1) can be reduced to the Kuramoto model with zero natural frequency for some class of initial configurations given by (2.9). There are two important concepts of synchronization for the Kuramoto model. The first one is 'phase synchronization' which means that all relative phases go to zero asymptotically. Our definition 3.1 corresponds to this phase synchronization in classical case, and the second one is the 'complete synchronization' which means that all relative phase velocities (frequencies) go to zero asymptotically. For the Kuramoto model with zero natural frequencies, phase synchronization will occur only for some restricted class of initial configurations which can be confined to the half circle. For example, consider a bipolar configuration $({{\theta }_{1}},{{\theta }_{2}})=(0,\pi )$. This configuration is clearly an equilibrium for the Kuramoto model with zero natural frequency so that there will be no phase synchronization in this case. In contrast, the complete synchronization will occur for a generic initial configurations, although the complete proof for this is not yet available in literature (e.g., for identical oscillators, this can be found in [8]).

3.2.  ${{H}^{k}}$ synchronization

In this subsection, we study higher order synchronization of the wave functions. For this, we set

Lemma 3.3. Suppose that $\Psi =({{\psi }_{1}},...,{{\psi }_{N}})$ be a global solution to (2.5) satisfying

Then, we have

Proof. We differentiate (2.5) with respect to t and obtain

Equation (3.2)

where we used the fact that $\langle {{\psi }_{k}},{{\psi }_{i}}\rangle $ is only dependent on t. Then we use (3.2) to obtain

For $t\in {{\mathbb{R}}_{+}}$, we can take an index it such that

By the above relation, we have

It follows from performing elementary calculations on the above equation that

This implies

We integrate the above inequality to obtain the desired result.

Theorem 3.2. Suppose that the coupling strength K and initial data satisfy

Then, for any solution $\Psi =({{\psi }_{1}},...,{{\psi }_{N}})$ to equation (2.5), we have

where C2 is the positive constant depending on initial data ψi0 and ψj0.

Proof. Below, for notational simplicity, we suppress the x dependence in ψi, i.e.,

Note that $\partial _{x}^{\alpha }({{\psi }_{i}}(t)-{{\psi }_{j}}(t))$ satisfies

Equation (3.3)

We multiply (3.3) by $\partial _{x}^{\alpha }({{\psi }_{i}}(t)-{{\psi }_{j}}(t))$ and integrate the resulting relation with respect to x to obtain

Equation (3.4)

We next estimate the terms ${{\mathcal{I}}_{2i}},\;i=1,2,$ as follows.

  • (Estimate of ${{\mathcal{I}}_{21}}$): Note the relation
    Equation (3.5)

Here we used the assumption that $D({{\Psi }_{0}})\lt 1/2$ and lemma 3.1 to see the monotonicity of $D(\Phi )$. Relation (3.5) implies

Equation (3.6)

  • (Estimate of ${{\mathcal{I}}_{22}}$): In this case, we use theorem 3.1, lemma 3.3, and
    to derive
    Equation (3.7)

Finally, we combine the estimates (3.4), (3.6), and (3.7) to derive the first-order differential inequality for $||\partial _{x}^{\alpha }({{\psi }_{i}}(t)-{{\psi }_{j}}(t)|{{|}_{2}}$:

Then, Gronwallʼs lemma implies

This yields the desired result.

4. Conclusion

In this paper, we presented a quantum synchronization estimate to the S–L model for identical quantum oscillators. The S–L model was first introduced by Lohe [18] as a partial differential equation model describing the synchronous dynamics of identical quantum oscillators on a quantum network. However, a rigorous analysis of the emergence of quantum synchronization has not been addressed in previous works [18, 19]. We introduced a Lyapunov functional measuring the degree of quantum synchronization, which is the L2-diameter of the one-particle wave functions. For a positive coupling strength and some class of initial wave functions, we verified that L2 distances between wave functions approach zero exponentially fast so that asymptotically the wave function for the system becomes the tensor product of identical one-particle wave functions.

Acknowledgements

The work of S-Y Ha is partially supported by NRF-2009-0083521 (SRC). The authors appreciate D-P Chi for helpful discussion on the Lohe model.

Please wait… references are loading.
10.1088/1751-8113/47/35/355104