Abstract
The Kitaev surface code model is the most studied example of a topologically ordered phase and typically involves four-spin interactions on a two-dimensional surface. A universal signature of this phase is topological entanglement entropy (TEE), but due to low signal to noise, it is extremely difficult to observe in these systems, and one usually resorts to measuring anyonic statistics of excitations or non-local string operators to reveal the order. We describe a continuous-variable analog to the surface code using quantum harmonic oscillators on a two-dimensional lattice, which has the distinctive property of needing only two-body nearest-neighbor interactions for its creation. Though such a model is gapless, it satisfies an area law and the ground state can be simply prepared by measurements on a finitely squeezed and gapped two-dimensional cluster-state without topological order. Asymptotically, the continuous variable surface code TEE grows linearly with the squeezing parameter and a recently discovered non-local quantity, the topological logarithmic negativity, behaves analogously. We also show that the mixed-state generalization of the TEE, the topological mutual information, is robust to some forms of state preparation error and can be detected simply using single-mode quadrature measurements. Finally, we discuss scalable implementation of these methods using optical and circuit-QED technology.
Export citation and abstract BibTeX RIS
Content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.
1. Introduction
Topological order describes a phase of matter whose correlations satisfy an area law while maintaining long-range entanglement and being robust to local perturbations. These properties make such systems attractive candidates for stable quantum memories or processors [1]. However, the lack of a local order parameter makes measuring topological order an experimentally onerous task. Some possibilities include measuring non-local string operators [2] or the statistics of anyonic excitations above the ground state, as has been demonstrated experimentally with small photonic networks [3, 4]. Unfortunately, due to finite correlation lengths of local operators [2, 5], these methods suffer from low visibility if the system is not prepared in a pure phase with vanishing two-point correlations.
An alternative is to study properties of the state itself that are robust to small changes in the correlation length. For a topologically ordered phase, the entanglement entropy of a subsystem in state is the von Neumann entropy
where , is the boundary size of A, and for , [6–8]. The parameter γ is termed the topological entanglement entropy (TEE) [1], which is an intrinsically non-local quantity that characterizes topological phases in a variety of systems, including spin lattices such as the qubit surface code [6, 9], bosonic spin liquids [10], and fermionic Laughlin states [11].
While useful for numerics, actually measuring TEE in a physical system is a daunting task since extracting the von Neumann entropy requires knowledge of the complete spectrum of the reduced state. A different option is to instead measure the Rényi entropy since it was shown [12] that γ is the same when replacing with . The value is an attractive choice since the purity is observable via a simple swap-test measurement on two copies of the state [13]. For example, the qudit (d-level spins) surface code state [14], has , meaning [15]. In contrast, the purity of another area law state with no TEE, such as the qudit cluster-state [16], is . Thus, even using Rényi entropy one still requires a number of measurements exponential in the size to distinguish the two phases.
In this work we study, for the first time, topological order in a continuous-variable (CV) Gaussian state [17] analog of the discrete-variable surface code state. In section 2 we introduce Gaussian states and the graphical calculus used to describe Gaussian transformations including measurements. Next, in section 3 we describe how to prepare CV surface codes efficiently using an intermediate mapping to the ideal, i.e. infinitely squeezed, CV cluster-state. Then in section 4 we show how to extend the mapping for finitely squeezed cluster-states [18–25]. We then show that unlike its qubit (or qudit) counterparts, the CV surface code state has a parent Hamiltonian that is gapless in the thermodynamic limit. Nonetheless we prove that the state exhibits topological order and remarkably it can be revealed simply by computing the TEE from quadrature measurements. Other gapless models such as quantum loop gases with topological order have been investigated in different contexts [26, 27]. In section 5, we present the main results of the paper. It is shown that the CV surface code has a non-zero TEE that asymptotically grows linearly with the squeezing parameter, and that an analogous quantity, the topological logarithmic negativity, behaves in the same way. In section 6 we analyze the stability of the CV topological order against two forms of noise: thermalization in the case of preparation by cooling, and noisy input states in the case of active construction. We conclude in section 7 by proposing experimental realisations for this model that are accessible with todayʼs technology.
2. CV generalities and Gaussian states
We start generalizing the Pauli group of single qubit gates [28] to the Weyl–Heisenberg group of phase-space displacements [29]. For a single qubit, this generalization is most easily accomplished by thinking of the qubit and gates as implementing one-unit cyclic shifts in 'position'
and 'momentum'
The CV analogs of the qubit gates are the translation (position-shift) operator and boost (momentum-shift) operator , with . While for qubits the shift is by an element of the cyclic group , i.e. the gates transform a basis element into the other, in the CV case, one may implement a shift in position or momentum by any real valued amount. In fact, the Weyl–Heisenberg group for CVs is a continuous Lie group, whose generators are the elements of the Lie algebra spanned by the identity operator and the canonical self-adjoint quadrature operators , , which satisfy the canonical commutation relation [28] (with ).
Specifically, the displacement operators are equal to
with group commutator
The action of the displacement operators on the eigenstates (or continuous computational basis) of and is given by [30]
where the subscript means that and . Then, the Pauli group for CV quantum computation on N oscillators is simply given by the set with generating algebra for .
The transformations we will use in the following discussion belong to the Clifford group for CV [29]. This is the group of transformations that preserve under conjugation, i.e. that given any , then for every . We will also describe quantum states in terms of certain special operators called stabilizers. For a generic state we call stabilizer any operator such that .
The stabilizer formalism is a powerful tool to describe topologically ordered systems [1], and typically one resorts to construct the defining Hamiltonian as a linear combination of the elements composing the ground state subspace stabilizer set of the system. Most importantly, the stabilizer formalism allows for a simpler description of the evolution of the state. Given the stabilizer condition, under a unitary transformation of the state, , transforms as in order to preserve its stabilizer status . Note that the transformation under the action of is opposite from the Heisenberg evolution of the observables under the same unitary . In fact, when we evolve stabilizers we are not modeling the evolution of observables, but rather evolving the old stabilizers into new stabilizers for the new state. Hence, the unitary evolution applied to the stabilizer must counteract that applied to the state in order to maintain the stabilizerʼs role as such. For these reasons restricting state transformations to elements of the CV Clifford group enforces the requirement that every operation performed on the state transforms stabilizer operators into stabilizer operators.
Note that in the context of CV systems there exists an equivalent way to express the stabilizer relations by using a class of operators known as nullifiers [29, 31]. In analogy with stabilizers, an operator is called a nullifier for a state when the relation holds. When the generators of the stabilizer set are elements of a Lie group [28], then the elements of the Lie algebra that generates the Lie group compose the nullifier set of the state. Significantly, nullifiers transform under the same transformation rule of the stabilizers.
2.1. Gaussian states and Gaussian transformations
In the following discussion we only consider Gaussian states. A zero-mean, N-mode Gaussian state can be described conveniently and completely by an easy algebraic formalism that follows from the form of its characteristic function [32], which is solely a function of the vector of the first statistical moments and the matrix that carries the information about the second moments. Recall that is the covariance matrix of the Gaussian state defined as
where is the -dimensional column vector of the Hermitian quadrature operators of the N modes. The information contained in the covariance matrix completely determines the entanglement properties of a Gaussian state [33]. Explicit calculations of Gaussian states entanglement entropy are performed making use of the symplectic spectrum of . Let us introduce the symplectic form ,
which is a skew-symmetric matrix that incapsulates the canonical commutation relations of the quadrature operators. For a Gaussian state ρ with covariance matrix , the positive elements of the N pairs of eigenvalues of the matrix product are called symplectic eigenvalues. The entropy for an NA -mode Gaussian subsystem is
calculated using the reduced symplectic spectrum obtained deleting the complementary rows and columns from the covariance matrix.
We define a Gaussian operation any unitary transformation that maps a Gaussian state onto a Gaussian state [33]. These are generated by Hamiltonians at most quadratic in the quadrature operators. Note that all CV Clifford operations are Gaussian operations. Each Gaussian operation has a correspondent matrix representation that belongs to the Symplectic group Sp , while quadratures measurements have a well-defined action on the covariance matrix of the state, which is described below.
A symplectic transformation preserves the canonical commutation relations as follows
while the action of a Gaussian transformation on the quadratures can be expressed by
where the right-hand side corresponds to a matrix multiplication on the quadratures vector. At the level of the covariance matrix, this is reflected in the transformation rule
A state of N independent vacua is described by the covariance matrix
After a Gaussian unitary transformation represented by the matrix is applied to Γ0, the resulting Gaussian state is
Exploiting the decomposition properties of symplectic matrices [31, 34], the product is uniquely specified by , with
where, for an N-mode state, both and are symmetric matrices, and . Hence, the complex linear combination
offers an alternative description for a pure Gaussian state. The graph shows up directly in the position-space wavefunction for an N-mode Gaussian state :
where is a column vector of position-space variables. For this state, from equation (15) can be rewritten as
The matrix has a simple transformation rule under a symplectic transformation . If is decomposed into block form
then the matrix associated to the transformed state is given by
The usefulness of this approach lies in the simple transformation rules of for the most common laboratory procedures corresponding to Gaussian unitary transformations, as listed in [31]. This permits the study of Gaussian state evolution simply in terms of appropriate transformations on .
Measurements also have a straightforward translation in the transformation rules language. A measurement on the kth mode is equivalent to deleting the kth row and column of the matrix, while a measurement is equivalent to applying a phase shift on the kth mode and then measuring the quadrature. At the level of the matrix, any measurement deletes the measured mode while generating new connections among its nearest neighbors.
3. Ideal CV codes
We start the description of the CV models introducing the so-called ideal codes [19, 35]. These codes are ideal, or infinitely squeezed, because they represent unphysical, non-normalizable states of infinite energy. Although these codes are mere mathematical approximation without an equivalent physical state, they play the role of theoretical CV analog of the qubit cluster-state and Kitaevʼs surface codes, being also the limiting case of the physical states that will be introduced in the next section.
3.1. The ideal CV cluster-states
The CV surface code can be easily described and efficiently prepared using an intermediate mapping to the CV cluster-state via a simple pattern of quadrature measurements. Intuitively, the CV cluster-state is the CV analog of its qubit-based cousin [36]. There are many ways to construct physical CV cluster-states [18–24], all of which give slightly different states in the finitely squeezed case [31]. Each has important differences that manifest when using them for measurement-based quantum computation. In the ideal, infinitely squeezed case, however, these differences become largely irrelevant, and since infinitely squeezed states are unphysical anyway, we are free to choose for their analysis the method that is simplest. For this reason we choose the canonical method [18]: despite its inefficiency when used in practice [37], this is the most straightforward generalization of the qubit cluster-state preparation [38].
Given a square lattice defined on a graph , we first substitute the N qubits on the vertices with N qumodes (quantum modes—i.e., harmonic oscillators) initialized in the infinitely-squeezed zero-momentum eigenstate , which is equivalent to the state for qubits. The global state is therefore , and this is stabilized by the single mode operators in the sense that
Alternatively, the same state is nullified, by the set of generators of the stabilizer group since,
As in the qubit case, the ideal CV cluster-state is the result of the pairwise application of controlled-Z gates upon all the nearest-neighbor modes of the initial state as depicted by the graph, explicitly
How does the application of these gates affect the stabilizer set? Under the evolution, the quadrature operators transform as [39]
and thus the initial state stabilizers are changed into the cluster-state stabilizers
which have the same form of the qubit cluster-state stabilizers [40], where indicates the nearest neighbors of the jth mode. We can also express these stabilizers by [31]
equivalent to define the elements of the CV cluster-state nullifier set as
All the (and linear combinations of them) commute and are the elements of the algebra that generates the stabilizer group of . In analogy with the qubit case, we can construct a Hamiltonian whose ground state is the CV cluster-state by imposing an energy penalty for violating any of the nullifier conditions:
Since all nullifiers commute and have a continuous spectrum of eigenvalues , this Hamiltonian also has a continuous spectrum , and is therefore gapless [19].
Even at the ideal level there is some additional freedom in the CV cluster-state construction procedure. In particular, ideal CV cluster-state graphs can have any non-zero real-valued weight associated with each edge. This modifies the strength of the gate represented by that edge: . These weights were first introduced in [20] as a way to enable new methods of construction. They have shown themselves to be very important for the computational properties of these states [41] and when considering efficient construction of cluster-states with very large graphs [22, 23, 31]. For the purposes of all future derivations we set g = 1, but when describing possible experimental implementations of the CV codes at the end of this manuscript, we will show how the results obtained with this assumption apply indistinctly to CV cluster-states constructed with non-unit—but still uniform—weight g.
3.2. The ideal CV surface code
In this section we present the description of the ideal CV surface code. We start introducing the stabilizer description for a general lattice implementation and then look at the special case of a planar square lattice with defined orientations.
The ideal CV cluster-state on a square lattice can be transformed into the corresponding ideal CV surface code by a simple scheme of quadrature measurements [35] inspired by the dynamical mapping of qubit cluster-states to surface codes [42]. In short, start with the CV cluster-state and label vertices by row and column. Then measure in those modes on rows and columns that are both odd and in those that are both even, as in figure 1 for the case of a square lattice. This scheme is equivalent up to translation and/or inversion of the , measurements. Since the CV cluster-state is Gaussian, and quadrature measurements are Gaussian operations [31], after the measurements, we are left with a new Gaussian state. The measurements change the form of the stabilizers [31] and therefore after the measurements we are left with a state described by a new set of nullifiers that corresponds exactly to the CV analog of Kitaevʼs surface codes [9].
To derive the post-measurement nullifier set, it is convenient to borrow notation from the qudit version of surface codes [14] that describes the nature of the coupling involved in the nullifiers in terms of a surface code graph . We assume the graph is oriented and that every face f inherit this orientation. Each quantum mode reside on an edge , with the orientation of any edge determined by for the base of the edge starting at vertex v and the head at vertex .
The derivation of the nullifier set for the CV surface code implies knowledge of the nullifiers transformation rules under quadrature measurements. The explicit form of the surface code nullifiers for a square lattice (while allowing for smooth or open boundaries) are obtained by taking linear combination of neighboring nullifiers of the CV cluster-state. A quadrature measurement on a qumode removes it from the cluster [19]. Given a set of exact nullifiers for a Gaussian state [31], one can obtain new nullifiers by a three-step process:
- (1)Given a quadrature measurement to be made on mode j, where , using linear combinations of the original nullifiers, write a new set of nullifiers (remember, they need to commute) such that the canonically conjugate local quadrature (where appears in only one nullifier in the new set.
- (2)In each new nullifier, replace with the real-valued measurement outcome.
- (3)Eliminate the nullifier that contains .
We always assume that the outcome of the measurement is zero because any other outcome would merely result in the same state up to displacements in phase space. These displacements can always be undone by local unitaries and therefore do not change any entanglement measure we might want to calculate [33].
Then, in this very general case, the stabilizers for the CV surface code are equal to
where the symbols in these expressions denote
and the dot (·) stands for any vertex. By construction, the stabilizers commute:
The CV surface-code subspace is the co-eigenspace of the stabilizers and , similarly to Kitaevʼs surface codes [1].
To simplify the discussion, in the following we specialize to the case of the code graph being a toroidal square lattice. We then fix the edge orientations such that at any vertex v, all incident edges point toward v, or all point away from v and the faces inherit equal counterclockwise orientation; see figure 2. Under these assumptions, the symbol is a constant for any vertex v, thus it only amounts to a sign flip on t in equation (30), which has no effect on a stabilizerʼs role as such, and we can ignore it. In this case, the CV stabilizers function as the CV analogs of the surface code stabilizers for qubits [9] (up to a swap)
and explicitly become
where the stabilizers generators, i.e the nullifiers, are:
Download figure:
Standard image High-resolution imageIf, on the boundary, one of the edges is missing, then that mode is not included in the nullifiers. It should be clear that the choice of the orientation is purely conventional and does not change the state. Hence, if we were to flip the orientation of every edge, simply becomes , which is still a nullifier.
The ideal CV surface code is the non-normalizable ground state of the quadratic Hamiltonian [35]
In analogy with the ideal cluster-state Hamiltonian, the spurious mode operators and are actually Hermitian quadrature operators that all commute and do not obey the canonical commutation relations. Consequently, because of
an extensive number of gapless modes exist. Therefore this Hamiltonian has a fully continuous spectrum and is gapless for any number of systems, even on a square lattice with boundary. Anyonic braiding in this model was studied in [35, 43, 44].
4. Physical CV codes
Ideal states are mathematical representations of states that cannot be produced in the laboratory. With experimental realization in mind we are required to speak about states that have actual physical significance and can be implemented experimentally. In this section, we consider physical realizations of the ideal states, or more specifically, we analyze the case of finite squeezing.
4.1. Physical CV cluster-state
Similarly to the previous section, we start considering the CV cluster-state first [18]. To produce the CV cluster-state in the finite-squeezing case by the canonical method one starts from N vacuum modes . The elements of the initial nullifier set are simply the dimensionless single-mode annihilation operators , since
The vacua are then all squeezed by : squeezing is performed by the unitary operator , with , where is traditionally known as the squeezing parameter. In the Heisenberg picture, , and , such that the variance of (of ) after squeezing is a factor of (of s2) times its original value. Analogously to the ideal case, the squeezed states are then sent the through pairwise controlled-Z gates in accord with an undirected, unweighted graph with one qumode per vertex. Under the action of the squeezing and the couplings, the nullifier set for the CV cluster-state is , with
These operators satisfy the canonical commutation relations for normal-mode operators, and , and therefore the CV cluster-state Hamiltonian is [24]
For finite s, the system has a gap of . The prefactor provides for finite energy even in the limit of infinite squeezing.
4.2. Physical CV surface code
Using the same measurement pattern shown in figure 1, the finitely squeezed CV cluster-state can be mapped to the finitely squeezed CV surface code. The procedure to obtain the nullifier sets in this case is more elaborate: taking linear combinations of neighboring cluster-state nullifiers—specifically sums of neighboring nullifiers around the -measured nodes—and alternating signed cyclic sums around -measured modes, one finds the general form of the surface code nullifiers for a generic lattice with possibly incomplete vertices and faces [39]:
with , V(v) valence of vertex v and boundary size of the lattice face. It is important to realize that now there is a dependence upon the position on the lattice of the vertex or face that we consider. The form of the commutation relations for these nullifiers is not as straightforward as in the previous cases. They are determined by the distance between the vertices and the boundary size of the faces as
Here is the Euclidean distance between the two vertices v and , where the edge lengths of the graph are unit length. Then the most generic CV physical surface code Hamiltonian is given by
and the squeezing dependence of the prefactors for the vertex and face parts ensures the Hamiltonian has finite energy in the infinitely squeezed limit:
Here we used the fact that for infinite squeezing, each vertex nullifier involves a sum of ʼs around that vertex and its four neighboring vertices, and since they all commute, the parent Hamiltonian is simply the squared sum of ʼs around each vertex.
In the simpler case of a square lattice with toroidal boundary conditions, the nullifiers become
where . The state defined by these nullifiers is the CV surface code. Note that now the next-nearest neighbors contribute to the structure of the nullifiers, in contrast to the ideal case, as illustrated in figure 3. Without loss of generality we can consider a square lattice: Then the nullifiers commutation relations from equation (42) can be rewritten as
where and are the Euclidean distance between vertices and faces, respectively on the unit-edge-length lattice and dual lattice. The functions w and x are given by:
and
Using these nullifiers we can construct a Hamiltonian for the physical CV surface code:
The squeezing dependence of the prefactors is done to ensure the Hamiltonian has finite energy for . Unlike the discrete-variable case [6], this Hamiltonian is gapless in the thermodynamic limit. This arises because the nullifiers do not define normal modes. Rather, neighboring nullifiers have non-trivial commutation relations, which allow for low-energy mode excitations.
Download figure:
Standard image High-resolution imageOn the torus, the specifics of the graph are , , and . We first focus on the case where is odd, so that there are independent nullifiers spanning the space of all the physical-mode annihilation operators. To diagonalize the Hamiltonian we introduce the normal-mode operators
where the vertices at the lattice sites have coordinates , and the faces at the dual lattice sites have coordinates . In this base the Hamiltonian is
To find the normal-mode frequencies, we need to solve the equations
where the equality is ensured by , and
with . If we introduce the generic vertex label state and the face label state the two linear equations above can be vectorized and rewritten in these two basis respectively as
where and are the vectorized form of the operators and that result from equation (50). In the following we define operators for a periodic square lattice that make the problem solvable. First, define the shift operator , whose action on the generic vertex label state is:
The action of the shift operator on the face basis is completely analogous. This allows to rewrite the matrices and , whose elements are non-zero according to the form of the commutation relations in equation (46), in the following elegant way:
The linear equations in equation (54) can be solved in the Fourier basis via , where
and the nullifiers in the Fourier basis are decomposed as
In this way, the solutions written in the basis and are
treating the normal mode index as a collective index. Then, the (squeezing-dependent) gap energy is the lowest-frequency mode energy:
For large system sizes, i.e. , and choosing freely that , the gap is equal to
and in the thermodynamic limit the gap goes to zero.
If n and m are even, then not all the face nullifiers are independent. To see this simply bicolor all the lattice faces and assign a plus sign to face operators of one color and a minus sign to faces of the other, then add them to get zero. Thus, the Hamiltonian is underconstrained, and there exists an exact gapless zero mode. For a square lattice with planar boundaries (and not toroidal as we discussed so far), there are boundary effects, but these make only a small modification to the gap, which still scales like the inverse of the system size.
Hence, in distinction to the cluster-state Hamiltonian , the surface code Hamiltonian is gapless in the thermodynamic limit, though for infinite squeezing both models are gapless. Note that neither nor is assumed to be a physical Hamiltonian governing the evolution of the system in question. One may choose to create the initial cluster-state by cooling a physical instance of , but this is not assumed.
5. Detecting topological order in the CV surface code
To study the topological properties of the CV surface-code state, we make use of two alternative (but closely related) definitions of TEE for two dimensional systems. The first was introduced by Kitaev and Preskill (KP) [7],
and the second by Levin and Wen (LW) [8],
with regions shown in figures 4(a) and (b), respectively. Both definitions assume area law behaviour (as per equation (1)) of the entanglement entropy and that the regions are chosen large compared to the correlation length. If the system is not topologically ordered, these combinations of entropies sum to zero exactly. Thus, we say that a model is topologically ordered only when . We emphasize that topological order is a property of the state and not of any background Hamiltonian.
Download figure:
Standard image High-resolution image5.1. Quadrature correlations on the lattice
As a consequence of the CV surface code being Gaussian, to evaluate each subsystem entropy, SX , one only needs the covariance matrix of the post measurement state. To obtain this we find it convenient to use the graphical calculus introduced in section 2.1. Thanks to this formalism it is possible to define the matrix for the CV cluster-state directly, solely making use of the adjacency matrix that describes the square-lattice pattern of connections among the modes. Explicitly [31]:
with squeezing parameter and the unit matrix. This is illustrated in figure 5. To derive the graph of the finitely squeezed CV cluster-state with open boundary conditions, we start with and perform the measurement scheme as shown in figure 1, using the graph transformation rules explained previously.
Download figure:
Standard image High-resolution imageThen the matrix transforms into the CV finitely squeezed surface code state , shown in figure 6. In this case, it is a purely imaginary matrix whose entries are given by
where
and is the unweighted adjacency matrix of the surface code (without self-loops, see figure 6). Note that the component of the matrix, see equation (16), is zero for the surface code. Using the connection between matrix and covariance matrix from equation (18), we derive the covariance matrix of the physical CV surface code state:
Thanks to the formula in equation (9), complete knowledge of the covariance matrix makes calculations of entanglement entropy for different regions of the system straightforward.
Download figure:
Standard image High-resolution imageUsing the TEE to characterize the system is meaningful only if the CV surface code state obeys an area law for the entropy. For this assumption to be true, the quadrature correlations among the modes must decay exponentially with lattice separation [45]. From the form of the CV surface code covariance matrix, we see immediately that – correlations (determined by ) have range at most 1, and – correlations are zero. On the other hand, the -correlations require a more elaborate analysis because they are determined by the matrix , which is more complicated. However, it is possible to provide analytical bounds on the -correlations for arbitrary squeezing. To prove this, first we show that the spectral range of , denoted , satisfies where and . For the minimum eigenvalue, note that for finite squeezing the matrix is positive definite but for , some – correlations become infinite. This indicates that the matrix is singular in that limit and the smallest eigenvalue is therefore zero. Adding finite squeezing shifts the spectrum of by so the minimum eigenvalue of is .
To derive the largest eigenvalue of , observe that for the surface code on a lattice with periodic boundaries, the graph associated with the adjacency matrix is regular with degree 6, i.e. each node connects to six others. The largest eigenvalue of the adjacency matrix of a regular graph is equal to the degree with an associated eigenvector [46], therefore the maximum eigenvalue of is 6. From these considerations it follows that the largest eigenvalue of is . For lattices with open boundary conditions, the eigenvalues of are upper bounded by the maximal degree with corrections that fall off with the system size, so the spectrum of lies in the interval .
On a generic lattice, the matrix is block tridiagonal with n identical matrices on the diagonal and identical on the immediate upper and lower blocks. Now the matrix coordinates (i, j) correspond to Euclidean coordinates on the lattice where for and , etc. It is convenient to define a graph distance between coordinates and . Since away from the edges, the bulk of the graph is the union of a square graph with a graph having two diagonal edges passing through every other face, the graph distance is the number of edges on the shortest path between and and it satisfies
where is the Euclidean distance. The matrix is itself tridiagonal with elements on the main diagonal and on the immediate upper and lower diagonal. The matrix is also tridiagonal with diagonal elements and immediate upper and lower diagonal elements either equal to zero or δ. A theorem of Demko, Moss and Smith [47, 48] shows that banded matrices of a certain class have inverses with matrix elements that decay exponentially with the distance from the diagonal. Specifically, for matrices of size and spectral range with ([47] proposition 5.1):
where the decay sets are
and the support sets are
Here and .
The matrices relevant to our problem are in this class. The matrix power is a banded block symmetric matrix with blocks of size and block band width . Furthermore, each such block is banded with band width . Thus the support set is the set of those matrix coordinates (i, j) such that the graph distance , is no more that . Similarly, the decay set is all matrix coordinates outside the support set. The statement in equation (69) is that for nodes separated in graph distance with associated matrix coordinates (i, j) the inverse matrix element falls off exponentially with graph distance as .
This implies that the position correlations between nodes i and j separated in graph distance by satisfy
where the constant is given by
and the correlation length is
For Gaussian states all higher order correlations are generated by the linear and quadratic ones. Thus the CV surface code state obeys an area law, and the use of the TEE formulas is appropriate. Numerically we find for squeezing the correlation length is and the scale factor of entropy with area is . At 5 dB squeezing .
5.2. Evaluation of non-local topological quantities
In our simulations we use the ground state of a mode CV surface code and the squeezing-dependent values of the TEE are calculated selecting from the covariance matrix the reductions corresponding to regions chosen as prescribed by equations (62) and (63). See figure 7 for numerical results. Note that even for , which corresponds to starting (in the CV cluster-state preparation) with vacuum states instead of momentum-squeezed states, the TEE is very small but non-zero. This is not a numerical artifact but a consequence of applying the controlled-Z gates, which introduce additional squeezing [37, 52] to input vacuum states. The values of the TEE calculated using the KP formula (62) and the LW formula (63) are extremely close for relatively large squeezing, and the difference accounts for numerics approximation, as expected. However, on the inset the difference is greater that the absolute values of and : this is due to the different formulas used, which result in different values of γ for small squeezing. As a sanity check, we also calculated the TEE for a mode CV cluster-state, which proves to be zero for any value of the squeezing. This is in complete analogy with the qubit-based cluster-state, which does not exhibit topological order.
Download figure:
Standard image High-resolution imageFurther, we plot the topological log-negativity (TLN) for the KP regions based on recent results that show this quantity to be a good witness of topological order for stabiliser states of Abelian anyon models [53, 54]. For a state with covariance matrix , as in our case, the log-negativity of the reduced state with support on a subsystem A is [55]
where , with PX being the projector onto the modes in region X and λi being the ith eigenvalue of the matrix argument. Then the TLN is constructed replacing the von Neumann entropy with the logarithmic negativity corresponding to the same region,
Remarkably, we find that the TLN is a rather tight upper bound on the TEE with the same asymptotic slope.
To derive this slope we consider the entanglement entropy for one mode of the smallest meaningful portion of the surface code, specifically a three-mode correlated state, see appendix
6. A noise model
We also check the resilience of the CV topological phase against noise in the state preparation. To model noise in the state preparation, we consider a thermal state with respect to the cluster-state Hamiltonian
as the pre-measurement initial state. Such a state could be generated by several different physical mechanisms. It could be generated by engineering the Hamiltonian , which is gapped for finite squeezing, and then waiting until the system reaches equilibrium with an environment at temperature β-1. Alternatively, using e.g. networks of non-interacting photons, one could start with separable modes each in a thermal input state and then generate the thermal cluster-state as before.
For such mixed states we can detect topological order by making use of the topological mutual information (TMI) [15]. The TMI is constructed replacing in equation (62) the von Neumann entropy SX with (half of) the mutual information between a region X and its complement Xc :
Calculations of the values for the TMI still require knowledge of the covariance matrix of the system (together with its reductions) and its symplectic spectrum, but unfortunately for mixed states we cannot use the graphical calculus for Gaussian states to derive the covariance matrix since it only applies to pure states. Then the covariance matrix must be derived in a different way.
In the pure case, under a given sequence of symplectic transformations and homodyne detections, the ground-state covariance matrix of the CV cluster-state Hamiltonian maps to a CV surface code state . For N equal non-interacting thermal modes the initial covariance matrix is given by
and characterized by the parameter
with energy gap of the finitely squeezed CV cluster-state. Then the thermal covariance matrix of at temperature β-1 is just , since any Gaussian operation performed on is easily computed noticing that . Simulating measurements is more intricate. Consider a covariance matrix : if we perform a measurement on the Nth mode (this can be easily generalized to any mode), then is
where is the reduced covariance matrix of the first modes, is the reduced matrix for the measured mode N and is the matrix that keeps track of the intra-modes correlations. A measurement on the N mode results in a new covariance matrix given by [56]
where
and is the pseudo-inverse (or Moore–Penrose inverse [57]) of the singular matrix . For a measurement one has instead [56]
Then , under the same evolution of the pure states case, maps to the surface code state covariance matrix equivalent to the zero-temperature covariance matrix times κ, i.e. . Note this is not a thermal state for because the spectrum of is non-uniform.
In [58] it is shown that the TMI is limited by lower and upper bounds. The TMI for this class of mixed states is lower bounded by the value computed as [59]. This illustrates the maximum extent to which the TMI can sink below the TEE for any given value of the squeezing parameter for this particular construction of the noise model. Numerically the difference between the TEE and the lowest possible value of the TMI is very small. Intrigued by this behavior, we have studied analytically the case and found that all the κ-dependent contributions in equation (75) cancel out. As shown in appendix
where X runs over all the regions in figure 4(a) and their complements, in accordance with equation (75), and the prime on the sum indicates that we need only include the zero-temperature symplectic eigenvalues σi X for which (see figure 7 for the numerical results).
7. Experimental implementations
In this work we consider canonical CV cluster-states [31], so named because they are the states that would result if one were to use the canonical method of constructing them [18, 19, 60]. This method generates Gaussian states with graphs of the form , where the entries of are either 0 or 1. Although this is straightforward theoretically, the gates are experimentally difficult and inefficient in an optical setting, where the most progress has been made [37, 52].
More efficient and scalable optical construction methods exist [20–23, 37, 61] but produce cluster-states with uniform non-unit edge weight. These optical methods can produce medium-sized or even very large states [62–64], including a recently demonstrated 10 000-mode cluster-state with linear topology [51]. Very large square lattices with toroidal [21], cylindrical [21, 23], and planar topology can also be made, as well as higher-dimensional lattices [65]. Similar states might also be created by cooling a circuit-QED system to the ground state of (equation (40)) [24]. In [66] it was shown that using superconducting co-planar waveguides coupled pairwise via dissipative Cooper pair boxes, one can engineer an effective – interaction between the microwave modes in neighbonring waveguides. By changing the location of the box in the waveguides, one can also generate – couplings. While the cluster-state Hamiltonian in equation (40) also has – couplings, there exist parent Hamiltonians for CV cluster-states (up to phase shifts) that consist only of – and – couplings [25, 31]. All of these methods can be used to efficiently produce square-lattice CV cluster-states with uniform edge weight g, with . In the optical case, these states can be very large (thousands of modes) [22] and all produce a Gaussian state with a graph of the form . By squeezing each mode in by a factor of , . Despite being constructed by a completely different method, the resulting state is a canonical CV cluster-state with effective initial squeezing . Since entanglement measures are local-unitary invariant, all of our results apply to these states if we take . Furthermore, we do not need to actively perform the single-mode squeezing before measuring the TEE/TMI/TLN. We can simply rescale the outcomes of measurements on the original (g-weighted) state [41].
8. Conclusions
We have presented a model of correlated quantum harmonic oscillators in a topologically ordered CV surface code state constructed using only Gaussian operations. This system has the remarkable property that its TEE γ can be observed simply by measuring elements of the covariance matrix via quadrature measurements. In contrast to discrete-variable systems, now γ is a continuous function of a system parameter, the squeezing. This is not surprising since string-like symmetries for the model depend on squeezing and have a group commutator valued in the continuous group [59]. The state preparation protocol we present here can be done in constant time (independent of the system size), and the TEE is robust to preparation errors modeled as thermal inputs. This provides a practical way to observe topological order in bosonic systems using current technology.
Acknowledgments
This material is based on research funded in part by the ARC Center of Excellence in Engineered Quantum Systems (EQuS), Project No CE110001013, and by the Singapore National Research Foundation under NRF Award NRF-NRFF2013-01. GKB and TFD thank the KITP where part of this work was completed with support from the National Science Foundation under grant no. NSF PHY11-25915. NCM was supported by the ARC under grant no. DE120102204. GKB thanks M Hastings, Z Wang, R Ionicioiu and S Rebic for helpful discussions. TFD thanks A Brodutch, M Cirio, S Flammia, M Lewenstein, S Singh and D Terno for discussions, comments and support. NCM is grateful to O Pfister, R Alexander, P van Loock, S Flammia, A Doherty and S Bartlett for helpful discussions. TL thanks L Lehman, D Lombardo and V Siddhu for illuminating conversations.
Appendix A.: Entropic bounds
In the main text we have shown how to calculate the TEE γ and the TMI for the physical CV surface code. In this appendix we first compute an upper bound to the TEE, calculating the subsystem entropy of a simpler network of entangled modes and then, analyzing the structure of the contributions to the TMI in the limit of high temperature noisy input states, we derive a lower bound to the TMI.
A.1. TEE upper bound
Here we compute an upper bound on the TEE for the CV surface-code state by calculating the subsystem entropy of a simpler network of entangled modes. To do this, we invoke a the calculation of subsystem entropy appropriate to stabilizer states, which have vanishing two-point correlation functions. Consider stabilizer states that are quantum doubles of a finite group G (such as the toric code with group ). As shown in [6], the TEE is calculated by dividing the system into two subsystems and identifying the redundant gauge transformations defined on the boundary between the two regions. The entanglement entropy of subsystem A is the logarithm of the number of the (all equivalent) Schmidt coefficients of the state. Exploiting the group properties of G allows one to write the entropy as , implying .
For the CV surface codes, it is complicated to extract an analogous exact expression for the entropy of a subsystem because the Schmidt coefficients are not equal as in the discrete case. Furthermore, the TEE is infinite for infinitely squeezed CV surface code states, and the definition of quantum dimension is not so clear for finitely squeezed CV surface code states since we do not yet have a description of this model in terms of a quantum double of a group. Nevertheless, we can go ahead and compute the subsystem entropy in the same way that would be done for the discrete case and treat this as a bound for the TEE of the CV surface code state. It is simply an upper bound because we are ignoring longer-range correlations that degrade the topological order, but since the correlation length is bounded for any finite amount of squeezing, this should be a reasonably tight bound.
A simple configuration to start with is a quantum double model with a discrete group on a lattice with two faces. This can be realised using a graph with just three edges (physical modes) and two vertices, as shown in figures A1 (a), (b). For the toric code the ground state would be the GHZ state since both vertices implement the stabilizer , and one face enforces , while the other face enforces . Identifying two qubits on one of the faces with subsystem B, the subsystem entropy is , where 2 comes form the size of the boundary of region B and therefore . This simplified surface-code network of three modes can be obtained from a finitely squeezed CV cluster-state with six modes after measuring out three of the modes (figure A1(d)). The resultant CV network has a correlation matrix that can be computed exactly. The symplectic spectrum for one subsystem consisting of one mode has two eigenvalues with
thus, using the formula for the entanglement entropy in equation (9), S(A) is
and the upper bound for the TEE can be expressed as
Download figure:
Standard image High-resolution imageA.2. Surface code topological mutual information from high-temperature cluster-states
Here we derive a lower bound for the TMI of the (noisy) CV surface code state introduced in the main text, analyzing the limit (with , gap of the CV physical cluster-state), which corresponds to the strongest possible decrease of the TMI from the TEE for the noise model considered. Recall that for a reduction of a pure state ρ, the von Neumann entropy determines the entanglement entropy of the subsystem with respect to its complement. For Gaussian states, one can use the formula
where is the collection of symplectic eigenvalues associated to the reduced covariance matrix of the subsystem . About the notation, in the following indicates the total number of symplectic eigenvalues associated to a region X, indicates the number of symplectic eigenvalues , and denotes those .
Noise in our scheme has been modeled as beginning with a thermalized CV cluster-state rather than a pure one. Since the normal-mode energies of the CV cluster-state Hamiltonian (equation (40) in the main text) are all equal, this is equivalent to squeezing identical thermal states instead of vacuum states in the canonical construction procedure [18]. The inverse temperature β of the state defines a useful parameter . To detect the topological order of the resulting mixed CV surface-code state, we use the TMI:
The numerics show that the TMI does not decrease significantly with an increment of the initial value of κ. This is interesting and rather unintuitive, hence an analytical expression is required to confirm our findings. First of all, recall that the covariance matrix of the resulting mixed CV surface-code state is equivalent to the pure one (Γ0) times κ, mathematically
As a consequence, the symplectic eigenvalues of (or any reduced section of it) are simply given by the'pure' symplectic eigenvalues multiplied by the overall κ factor, . This simple transformation of is a special case that only arises due to the fact that all normal modes of the CV cluster-state Hamiltonian are identical, resulting in equal symplectic eigenvalues .
For a large value of κ, we find the following asymptotic expression for each eigenvalue contribution to equation (A.4):
We can use this to show the behavior of the TMI as . To start, consider the first term of the TMI formula, equation (A.5):
where the regions used are shown in (A2 ). If the total state ABCD has N modes, then, for region A, we have nA modes and for its complement BCD we have nBCD modes, such that . Hence, the von Neumann entropy for A is given by
In the limit of very high temperature, corresponding to , we can use the approximation (A.7) and rewrite the von Neumann entropy SA as
Divide now the nA symplectic eigenvalues into the two sets and , such that . Hence, we can rewrite SA l as:
The same argument can be repeated for the region BCD and find that
The N symplectic eigenvalues of ABCD are all equal to , consequently:
and the value for the mutual information IA l in the limit is given by
Notice that the κ-contributions cancel out exactly and, asymptotically, IA l is not a function of κ. Using and the area law behavior for the entropy, i.e., for regions sufficiently big, (although this does not mean that the sets and are the same), we can rewrite the mutual information as
The same argument applies for each other contribution to the TMI, for example, for region C we have:
When substituting these expressions for the mutual information into the TMI formula in equation (A.5), all the elements , which depend on the size of the region boundary, sum to zero for the same reason why all the area-dependent elements cancel out in the KP argument [7] for the TEE.
Download figure:
Standard image High-resolution imageConsequently, the lower limit for the TMI is simply given by
where X runs over all the possible combinations of regions, and the function is defined as
for a partitioning of the system as in figure A2. This formula confirms our numerics and prove that that the value of the TMI, for the particular case of the CV surface code state, is not affected by κ in the limit of thermal input states at high-temperature.