Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Generating a 4-photon tetrahedron state: toward simultaneous super-sensitivity to non-commuting rotations

Open Access Open Access

Abstract

It is often thought that the super-sensitivity of a quantum state to an observable comes at the cost of a decreased sensitivity to other non-commuting observables. For example, a squeezed state squeezed in position quadrature is super-sensitive to position displacements, but very insensitive to momentum displacements. This misconception was cleared with the introduction of the compass state [Nature 412, 712 (2001) [CrossRef]  ], a quantum state equally super-sensitive to displacements in position and momentum. When looking at quantum states used to measure spin rotations, N00N states are known to be more advantageous than classical methods as long as they are aligned to the rotation axis. When considering the estimation of a rotation with unknown direction and amplitude, a certain class of states stands out with interesting properties. These states are equally sensitive to rotations around any axis, are second-order unpolarized, and can possess the rotational properties of Platonic solids in particular dimensions. Importantly, these states are optimal for simultaneously estimating the three parameters describing a rotation. In the asymptotic limit, estimating all d parameters describing a transformation simultaneously rather than sequentially can lead to a reduction of the appropriately weighted sum of the measured parameters’ variances by a factor of d. We report the experimental creation and characterization of the lowest-dimensional such state, which we call the “tetrahedron state” due to its tetrahedral symmetry. This tetrahedron state is created in the symmetric subspace of four optical photons’ polarization in a single spatial and temporal mode, which behaves as a spin-2 particle. While imperfections due to the hardware limited the performance of our method, ongoing technological advances will enable this method to generate states which out-perform any other existing strategy in per-photon comparisons.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Quantum metrology [13] employs fundamentally quantum systems as measurement probes to overcome the “standard quantum limit” [48], a statistical bound on the precision of any measurement with uncorrelated particles. Already, ideas from this field of research have been implemented in gravitational wave detectors [4,9] and could become important in a range of precision measurement device such as magnetometers [1012] and atomic clocks [13,14].

One of the classic problems used to demonstrate the advantage of quantum metrology is the measurement of a phase difference $\theta$ between the two arms $a$ and $b$ of a balanced interferometer. For the best measurement scheme using $N$ separable photons, the error in the estimate of $\theta$, $\delta \theta$, scales as $\delta \theta \propto 1/\sqrt {N}$. This is the “standard quantum limit" [4,5] in this problem, which is reflective of the measurement process being independent for each photon. On the other hand, the best general measurement strategy for this estimation is to send all the photons in a superposition of being all in path $a$ and all in path $b$. This highly entangled configuration is called the “$N00N$ state" [1520] and with it, one can in theory achieve the Heisenberg scaling [5,21] of $\delta \theta \propto 1/N$, which is the best scaling allowed by quantum mechanics, although its performance is quickly ruined by loss [22]. Low photon number $N00N$ states have been realized in the laboratory multiple times [2330]. It is often thought that super-sensitivity of this type comes at a price and that a quantum state super-sensitive to a given transformation is necessarily less sensitive to transformations generated by non-commuting operators. For example, a $N00N$ state where $N$ photons are in a superposition of all being in $+z$ eigenstates and all being in $-z$ eigenstates has stellar sensitivity to rotation around the $z$ axis. However, its sensitivity to rotation around the $x$ and $y$ axes is bounded by the “shot-noise limit" of $1/\sqrt {N}$. The 2-photon $N00N$ state is an extreme case which is completely insensitive to rotations around the $y$ axis. The idea of the trade-off in super-sensitivity is a misconception which was first cleared with the introduction of the compass state [31,32], a state of the quantum harmonic oscillator equally super-sensitive to displacement in position $\hat {x}$ and momentum $\hat {p}$.

In this work, we experimentally generate a counterpart of the compass state in the polarization of a light beam, the “tetrahedron state.” Unlike in the $N00N$ state example, we no longer consider the characterization of a single rotation, described by the action of the group U(1) and generated by the angular momentum operator $\hat {J}_z$, but instead we must look at general rotations described by the actions of the group SU(2), which can be generated by a triad of angular momentum operators $\{\hat {J}_x, \hat {J}_y, \hat {J}_z\}$. The tetrahedron state exhibits super-sensitivity for rotations generated by any linear combination of these three operators, corresponding to rotations around any axis, despite the non-commutativity of these generators. Beyond polarization rotations, SU(2) operations include the action of a generalized interferometer (with arbitrary transmission and reflection) on two modes of the electromagnetic field, the action of arbitrary magnetic fields on spins, reference-frame alignment, and many other processes. We will consider scenarios where the axis of the rotation is known but not necessarily aligned with the $z$-axis, as well as scenarios where the axis itself is also unknown. The main difference present in the latter is that the characterization of such processes requires the estimation of a minimum of three parameters instead of a single one [33,34]. The multi-parameter quantum estimation has proven to be more intricate than its single-parameter counterpart [5,3537], with nuances such as parameter compatibility [3840], and is important for many real-world problems [41,42]. This makes for a richer problem and requires us to look in different corners of the Hilbert space for suitable entangled probe states.

The tetrahedron state on which we report has several intriguing properties. It was initially studied as being the opposite of SU(2)-coherent states, earning the epithet “anticoherent” [4345], making it in some sense the most quantum state. Using the Hilbert–Schmidt norm as the distance measure between states, it is the furthest state from the closest SU(2)-coherent state of the same dimension. Therefore, it has been termed as the “queen of quantumness” [46]. In the context of polarization, it manifests “hidden polarization” [47], as its classical (first-order) polarization properties are ignorant of higher-order polarization features [48]. It also minimizes the cumulative multi-poles of its polarization distribution, for which it has earned the moniker “king of quantumness” [49,50], and is the optimal state for rotation sensing, explaining the sobriquet “quantum rotosensor” [51,52]. The tetrahedron state has numerous other extremal properties [53], including maximizing delocalization measures in SU(2) phase space such as the Wehrl entropy [54]. It has also been studied for its entanglement properties [5559], which have been generalized to multi-dimensional systems with multiple qudits [60]. While its polarization-sensing properties require the tetrahedron state to be in light’s polarization degree of freedom [61], it can also exist in light’s spatial degrees of freedom [6264], and can be envisioned in other symmetric multi-qubit systems. We report on creating this unique quantum state in the polarization degree of freedom of light.

2. Theory

2.1 Quantum Fisher Information

The quantum Fisher information (QFI) [65] is a powerful tool that, in the context of multi-parameter estimation, can provide a lower bound on the covariance matrix of the estimated parameters for any measurement scheme. Let ${\tilde {\boldsymbol{\theta}}}$ be the estimator for a set of parameters $\boldsymbol {\theta}$ describing a specific SU(2) rotation $\hat {R}(\boldsymbol {\theta})$; then, the covariance matrix

$$\Sigma_{l,m} = E[(\theta_l -\tilde{\theta}_{l})(\theta_m -\tilde{\theta}_{m})]$$
is bounded from below by the quantum Cramér–Rao bound (QCRB) [36]
$$\boldsymbol{\Sigma} \succeq \boldsymbol{I}_{\rho, \boldsymbol {\theta}}^{{-}1},$$
where $\boldsymbol{I}_{\rho, \boldsymbol {\theta}}$ is the quantum Fisher information matrix with components
$$[\boldsymbol{I}_{\rho, \boldsymbol {\theta}}]_{l,m} = \text{Tr}\left[\hat{\rho}_{\boldsymbol {\theta}}\left\{\hat{L}_{\theta_l},\hat{L}_{\theta_m}\right\}\right],$$
matrix inequalities $A\succeq B$ imply that $A-B$ is positive definite, and the expectation values $E[\cdot ]$ are taken over the probability distribution for the data on which ${\tilde {\boldsymbol{\theta}}}$ is based. The operator $\hat{\rho} _{\boldsymbol {\theta}}$ is the density matrix $\hat {R}\left (\boldsymbol {\theta}\right )\hat{\rho} \hat {R}\left (\boldsymbol {\theta}\right )^\dagger$ after it has been transformed by the action of the SU(2) rotation and the operators ${\hat {\boldsymbol {L}}}$ are the symmetric logarithmic derivative operators defined by
$$\partial_{\theta_l}\hat{\rho}_{\boldsymbol {\theta}} = \frac{1}{2} \left(\hat{\rho}_{\boldsymbol {\theta}}\hat{L}_{\theta_l} + \hat{L}_{\theta_l}\hat{\rho}_{\boldsymbol {\theta}}\right).$$
To find the optimal state to characterize the SU(2) rotation, it is important to have a scalar bound to optimize. Equation (2) can be easily turned into such a bound with the introduction of a cost matrix $\boldsymbol{C}$. The cost function to optimize then becomes the weighted sum of the elements of the covariances
$$\operatorname{Tr}[\boldsymbol{C \Sigma}]\geq \operatorname{Tr}(\boldsymbol{C I}_{\rho, \boldsymbol {\theta}}^{{-}1}) =: \mathcal{C}_{\rho, \boldsymbol {\theta}},$$
which is in turn bounded by $\mathcal {C}$, a scalar version of the QCRB (s-QCRB). Equation (5) cannot always be saturated for multi-parameter estimation but can indeed be saturated using the tetrahedron state [66]. In general, the matrix $\boldsymbol{C}$ can be arbitrary, providing different weightings to the different elements of the covariance matrix. When it comes to characterizing the elements of a group responsible for the transformation, a natural choice of the cost matrix is the Cartan metric of the group [67]. With the cost metric so chosen, Eq. (5) becomes completely independent of the parametrization of the group. It can be shown [68] that, for pure states $|{\psi}\rangle$ with this natural metric,
$$\mathcal{C} = \frac{1}{4}\operatorname{Tr}[\operatorname{\mathbf{Cov}}^{{-}1}_{\psi}({\hat {\boldsymbol {J}}})],$$
where ${\hat {\boldsymbol {J}}}$ are the generators of the Lie algebra associated with the group and $\operatorname {\mathbf{Cov}}_\psi$ contains the symmetrized covariances of those operators, with matrix elements
$$[\operatorname{\mathbf{Cov}}_{\psi}({\hat {\boldsymbol {J}}})]_{l.m} =\frac{1}{2} \left\langle\hat{J}_l\hat{J}_m+\hat{J}_m\hat{J}_l\right\rangle - \left\langle\hat{J}_l\right\rangle\left\langle\hat{J}_m\right\rangle$$
and expectation values $\langle \cdot \rangle$ taken with respect to $|\psi \rangle$.

It is noteworthy that the bound in Eq. (6) is independent of both the parameters to estimate $\boldsymbol {\theta}$ and of the initial rotation of the state (i.e., $|{\psi}\rangle$ and $\hat {R}\left (\boldsymbol{\phi }\right )|{\psi}\rangle$ will share the same bound for all $\boldsymbol{\phi }$). Furthermore, if we choose a Cartesian parametrization for the rotation

$$\hat{R}(\boldsymbol {\theta}) = e^{{-}i\boldsymbol {\theta}\cdot{\hat {\boldsymbol {J}}}},$$
then the s-QCRB with the Cartan metric in Eq. (6) coincides with the s-QCRB with $\boldsymbol{C} = \mathbb{1}$ when the rotation is in the neighborhood of identity. Then, $\mathcal {C}$ is a lower bound for $\sum _i \operatorname {Var}\left [\theta _i\right ]$, the identically weighted sum of the variances of the three rotation angles.

2.2 Polarization as a Spin

A single photon’s transverse polarization can be represented as a spin-$1/2$ particle, with the polarization states $|{H}\rangle\; (m=1/2)$ and $|{V}\rangle \;(m=-1/2)$ forming an orthonormal basis for the Hilbert space. In general, a collection of $N$ photons can be decomposed into the tensor sum of particles with spin $\leq N/2$:

$$\left(\text{spin-}1/2\right)^{\bigotimes N} = \bigoplus_{j = [0, 1/2]}^{N/2}d_j\left(\text{spin-}j\right),$$
where the coefficients $d_j$ represent the multiplicities of the different spin sectors. The highest-dimensional spin sector is the $\text {spin-}N/2$ and is unique ($d_{N/2}=1$). It represents the polarization states that are full-symmetric under the pairwise exchange of any particles. If there are no other degrees of freedom to distinguish photons, the bosonic nature of the photons would constrain their polarization state to this sector. We can therefore identify the basis element of the polarization state of a $N$ photon in a single mode simply by $n_H$ and $n_V$, the number of H and V photons it contains. For example, for a 3-photon state, the following are equivalent:
$$\begin{aligned} & |{j = 3/2, m = 1/2}\rangle = |{2_H, 1_V}\rangle \\ & = \frac{1}{\sqrt{3}} \left(|{HHV}\rangle + |{HVH}\rangle + |{VHH}\rangle\right). \end{aligned}$$
One way to represent the generators of the SU(2) Lie algebra in this space is
$${\hat {\boldsymbol {J}}} = \bigoplus_{j= [0,1/2]}^{N/2} \underbrace{{\hat {\boldsymbol {J}}}^{(j)}\oplus\cdots\oplus {\hat {\boldsymbol {J}}}^{(j)}}_{d_j\,\mathrm{times}}$$
where ${\hat {\boldsymbol {J}}}^{(j)} = \left (\hat{J}_x^{(j)},\hat{J}_y^{(j)},\hat{J}_z^{(j)}\right )$ are the usual spin angular momentum operator in a $\text {spin-}j$ space. It is easy to prove [69] that in this Hilbert space, the state minimizing $\mathcal {C}$ is contained in the full-symmetric $\text {spin-}N/2$ subspace.

2.3 The Majorana Representation

A helpful way to visualize states in the full-symmetric $\text {spin-}N/2$ sector is with the Majorana representation [70,71]. The idea is that one can always write these states as a product of single-particle creation operators over the two polarization modes acting on the vacuum state:

$$\begin{aligned} \sum_{m=0}^N\psi_m|{m_H,\left(N-m\right)_V}\rangle &\propto\prod_{k=1}^N \hat{a}_{\theta_k,\phi_k}^\dagger |{0_H,0_V}\rangle,\\ \hat{a}_{\theta,\phi}^\dagger&{=}\cos\frac{\theta}{2}\hat{a}_H^\dagger{+}e^{i\phi}\sin\frac{\theta}{2}\hat{a}_V^\dagger. \end{aligned}$$
When considered individually, these creation operators $\hat {a}_{\theta,\phi }^\dagger$ each produce a qubit state when acting on the vacuum. Moreover, this overall state is equivalent to the symmetric superposition of the tensor product of all $N$ such qubit states. Hence, one can represent full-symmetric states of the polarization of $N$ photons by $N$ points on the Bloch sphere labeled by their angular coordinates $(\theta _k,\phi _k)$. The action of a SU(2) rotation is simply to rigidly rotate the representation by the corresponding three-dimensional (3D) rotation of the sphere, which makes this visualization an integral tool to gain intuition about the action of rotation on different states.

2.4 Choosing the Optimal State for Multi-Parameter Estimation

2.4.1 The Spin Coherent States

Spin coherent states, depicted in Fig. 1(a), are notable separable states in the symmetric subspace of $N$ photons. They consist of all $N$ photons sharing the same polarization properties. One of those states is

$$|{\psi_{coh}}\rangle = |{N_H, 0_V}\rangle,$$
where all the photons are horizontally polarized, and the rest can be obtained by rotating that state. The spin covariance matrix for this state is given by
$$\operatorname{\mathbf{Cov}}_{\psi_{coh}}({\hat {\boldsymbol {J}}}) = \begin{pmatrix} N/4 & 0 & 0\\ 0 & N/4 & 0\\ 0 & 0 & 0 \end{pmatrix},$$
and is not invertible, which leads to a divergence in the bound s-QCRB and an infinitely bad performance at fully characterizing SU(2) rotations. Intuitively, this is due to the state being completely undisturbed by any rotation around the polarization axis of each of the separable particles. It is therefore incapable of estimating one of the three parameters necessary to characterize the rotation, which leads to a divergence in the weighted sum of the variance of those parameters. In the example above, all the polarizations of all photons were aligned with H, which leads to a variance of $0$ for $\hat {J}_z$, the spin operator aligned with that axis.

 figure: Fig. 1.

Fig. 1. Majorana representations of (a) spin coherent state, (b) 4-photon $N00N$ state, and (c) the tetrahedron state. (d) Conceptual illustration of the tetrahedron’s state creation process. The H polarization of a 3-photon $N00N$ state (three transparent red points on the vertices of an equilateral triangle on the equator) is partially attenuated resulting in the triangle of points with lower latitude (red). Simultaneously, an H polarized heralded single-photon is probabilistically added (green). The result is a 4-photon state represented by four points on the vertices of a tetrahedron.

Download Full Size | PDF

To remedy this situation, we could imagine dividing our available photons into three different spin-coherent states aligned with the $x$, $y$, or $z$ axes and using them sequentially. Because the quantum Fisher information is additive for uncorrelated measurements [36], the bound in Eq. (6) can be found adding up the different spin covariance matrices for each of these states. Assuming $N$ is divisible by three, the resulting matrix and the corresponding s-QCRB are

$$\sum_i\operatorname{\mathbf{Cov}}_{\psi_{coh}^{(i)}}({\hat {\boldsymbol {J}}}) = \frac{N}{6} \times \mathbb{1}, \quad \mathcal{C} = \frac{9}{2N},$$
where $\mathbb {1}$ is the $3\times 3$ identity matrix. As we can see, by splitting our photons into three batches, we retain sensitivity to every parameter and recover the expected shot-noise scaling from a classical state.

2.4.2 The $N00N$ States

The $N00N$ state is an entangled state of $N$ photons. The particular $N00N$ state aligned with the $z$ axis can be written as

$$|{\psi_{N00N}}\rangle = \frac{1}{\sqrt{2}}(|{N_H,0_V}\rangle + |{0_H, N_V}\rangle).$$
As mentioned previously, it is notable for being the most sensitive state for measuring the phase difference between two arms of an interferometer, or equivalently for estimating the angle of a rotation around a known axis. The $N00N$ state aligned as above, for example, would be especially well suited to estimate the angle of a rotation around the $z$-axis. When it comes to estimating all parameters of a rotation, however, we see from the following spin covariance matrix and its corresponding bound,
$$\begin{aligned} \operatorname{\mathbf{Cov}}_{\psi_{N00N}}({\hat {\boldsymbol {J}}}) &= \begin{pmatrix} N/4 & 0 & 0\\ 0 & N/4 & 0\\ 0 & 0 & N^2/4 \end{pmatrix}, \\ \mathcal{C} &= \frac{2}{N} + \frac{1}{N^2}, \end{aligned}$$
that the performance of the $N00N$ state still exhibits a classical scaling. This is because the $N00N$ state fails at performing the estimation of the other two parameters with a quantum advantage and their contributions dominate in the high $N$ limit.

We can attain a quantum scaling, however, by splitting the $N00N$ state into three batches, as with the spin coherent state. If the photons are split equally between $N00N$ states aligned with the $x$, $y$, and $z$ axes, we get

$$\begin{aligned} \sum_i\operatorname{\mathbf{Cov}}_{\psi_{N00N}^{(i)}}({\hat {\boldsymbol {J}}}) &= \frac{1}{36}\left(N\left(N+6\right)\right) \times \mathbb{1}, \\ \mathcal{C} &= \frac{27}{N(N+6)}. \end{aligned}$$
While splitting photons into three batches recovers the advantageous Heisenberg scaling, it causes us to pick up a factor of three in the leading order of $N$. Informally, we can think of the scaling as $O(3 \times 1/ (N/3)^2) = O(3^3/N^2)$.

2.4.3 The Second-Order Unpolarized States

It was proven in Refs. [53,66] that the optimal states for the task at hand are pure states that exhibit the following properties:

$$\langle{{\hat {\boldsymbol {J}}}}\rangle = 0 \quad \langle{\hat{J}_l\hat{J}_m}\rangle =\frac{N}{6}\left(\frac{N}{2} + 1\right) \delta_{l,m}.$$
These states have the property that they are isotropic up to the second moment of the spin angular-momentum operators, hence their name. Indeed, no initial rotation would change the two properties above for any state that satisfies them. These states do not exist in all dimensions. It is, for example, impossible to satisfy this condition with one, two, or three photons. There is one solution for four photons but then again none for five photons. The dimensionality of the spaces for which these states exist still constitutes a topic of research [45,72] that may be intimately connected to spherical $t$-designs [49,7375] and other problems for distributing points on the surface of a sphere [7680]. Multiple names mentioned above have already been used to describe these states and their usefulness has been theoretically demonstrated for a range of applications.

We can quickly see that states satisfying Eq. (19) outperform all the other schemes imagined here by looking at their spin covariance matrix and the s-QCRB:

$$\begin{aligned} \operatorname{\mathbf{Cov}}_{\psi_{plat}}({\hat {\boldsymbol {J}}}) &= \frac{1}{12}\left(N\left(N+2\right)\right) \times \mathbb{1}, \\ \mathcal{C} &= \frac{9}{N(N+2)}. \end{aligned}$$
The first thing to notice is that, in the leading order of $N$, these states recover the factor of three that was lost from splitting the $N00N$ state photons into three batches. Indeed, our cost function is now bounded by something of the order of $O(9/N^2)$ instead of $O(27/N^2)$. This is in line with many recent studies on comparing simultaneous multi-parameter estimation and sequential parameter estimation in the quantum regime. It has been found that optimal simultaneous schemes often outperform sequential schemes by a factor of the number of parameters in the large $N$ limit [35,68]. This is precisely $3$ in the case of the SU(2) rotation. The advantageous rotation-sensing performance of states with these properties has recently been demonstrated using light’s orbital angular momentum degree of freedom, where operations that are mathematically equivalent to rotations are applied using multi-plane light conversion [64].

2.5 The Tetrahedron State

The second-order unpolarized state with the smallest photon number ($N=4$) is the one we call the “tetrahedron state." It is named as such because it has the same symmetries as the tetrahedron under rotation, which is apparent in its Majorana representation [depicted in Fig. 1(c)], where the points correspond to the vertices of a regular tetrahedron inscribed in the Bloch sphere. The state can be written as

$$|{\psi_{tetra}}\rangle= \sqrt{\frac{1}{3}}|{4_H, 0_V}\rangle + \sqrt{\frac{2}{3}}|{1_H, 3_V}\rangle.$$
In this work, we create the tetrahedron state in the polarization state of four photons in the same spatial and temporal mode, which ideally guarantees the polarization to be full-symmetric.

3. Methods and Results

3.1 Experimental Apparatus

Our overall approach to our tetrahedron state creation scheme is intuitive in the Majorana picture. When we rewrite the tetrahedron state as a product of creation operators acting on the vacuum and we re-arrange the terms we get the following expression:

$$|{\psi_{tetra}}\rangle = \frac{1}{3}\left(\left(\frac{\hat{a}_H^\dagger}{\sqrt{2}}\right)^3 + \hat{a}_V^{{\dagger}^3}\right)\hat{a}_H^\dagger |{0}\rangle.$$
As can be seen, the tetrahedron consists of a combination of a 3-photon $N00N$ state in the H/V basis for which the H polarization has been attenuated by a factor of two and a horizontally polarized single photon, respectively.

We create these two components independently with SPDC. The $N00N$ state is created by an interference between a weak single-mode squeezed vacuum generated through SPDC, and a coherent state [26,81,82]. The relevant section experimental apparatus is depicted in Fig. 2(a). We use a pulsed Ti-sapphire laser with a pulse width of 140 fs, a center wavelength of 807 nm, a repetition rate of $80$ MHz, and an average intensity of $3.5$ W. With it, we perform SHG in a $1$ mm BBO crystal to get a beam of over $500$ mW at $404$ nm. This beam, blue in Fig. 2, serves as a pump for both our SPDC processes. In the $N00N$ state source, we perform SPDC with a $2$ mm BBO crystal. This SPDC source is optimized in Type 1 collinear configuration. We measure a pair creation rate of $53$ kHz with a symmetric coupling efficiency of $13.4{\% }$. These numbers include all forms of loss in the experiment, including detector inefficiencies, but are corrected to account for the necessity to probabilistically split the photons in a collinear geometry. These photon pairs interfere with a coherent state in a polarization interferometer with a Mach–Zehnder geometry. The coherent state is light from the attenuated original laser beam, left over from the SHG process.

 figure: Fig. 2.

Fig. 2. Experiment overview. (a) An 808 nm pulsed beam (red) is used to generate a 404 nm beam (blue) through second harmonic generation (SHG) in a 1 mm barium borate (BBO) crystal. Part of this blue beam is sent to generate co-propagating pairs of horizontally polarized 808 nm photons (red) through type-I spontaneous parametric downconversion (SPDC) in a 2 mm BBO crystal while the red beam is attenuated and vertically polarized. The two beams are recombined on a polarizing beamsplitter (PBS) and polarization rotated to form a 3-photon $N00N$ state in the H/V basis. (b) Another blue beam then a pumps a second SPDC (2 mm BBO) to generate another photon pair. Here, after a 3 nm narrowband filter, a detection at a single-photon counting module (SPCM) labeled “T" heralds the presence of a photon in the upper path. (c) Both the $N00N$ state and heralded single photon enter a displaced polarization Sagnac interferometer. The heralded photon strictly takes the clockwise path while the $N00N$ state is split equally amongst the paths. The half-waveplate (HWP) in the counter-clockwise path is set at $45 ^{\circ }$ to transmit all the light in the upwards output while the HWP in the clockwise path is set at $22.5 ^{\circ }$ to attenuate the V polarization of the $N00N$ state by half and transmit the heralded single-photon half the time (see Fig. 1 for context). After being recombined, all the photons pass through a 3 nm narrowband filter to make sure their spectra are the same. (d) The state is transmitted via single-mode (SM) fiber to a tomography apparatus for projection onto an arbitrary polarization axis. Each output path is coupled to a multi-mode fiber splitter to allow for partial photon number counting capabilities (a maximum of three detections per polarization). Five-fold detection events between T and any four SPCMs in the tomography setup are recorded.

Download Full Size | PDF

Importantly, the phase between these two beams should be stable throughout the experiment. We actively stabilize this phase using the locking port in Fig. 2(a). The squeezed vacuum and the coherent states are minutely off from their intended polarizations so that we can collect a bit of light from the locking port. The superposition of two horizontally polarized single photons from the squeezed vacuum and two vertically polarized single photons from the coherent state forms a 2-photon $N00N$ state since those two terms are mostly indistinguishable in any other degrees of freedom. We measure 2-photon coincidence events on a basis where the correct phase between two terms would result in a signal that is in the middle of a fringe. Because of the drifts in the rates, we re-calibrate the set point every 8–12 hr of data collection.

The horizontally polarized single photon, is also created through SPDC, and depicted in Fig. 2(c). This time we use a non-colinear Type I geometry. One of the photons of the pair is detected immediately and acts as a herald, while the other goes to be part of the tetrahedron state. Throughout this experiment, this SPDC source produced photon pairs at a rate of $32$ kHz and with a heralding efficiency of $11.5{\% }$. Using a heralded single photon here as opposed to a low-intensity coherent state was necessary to suppress the contamination from higher-order terms where more than one photon would be created in this mode. In order to ensure that the photons from these different sources are as indistinguishable as possible, all the photons in the tetrahedron state are fed into the same single-mode fiber right after transmission through a $3$ nm narrowband filter. The resulting loss is already included in the efficiency figures above. The herald photon is similarly filtered and collected.

As we will discuss in Section 4, the state we generated is notably far from ideal. The main source of imperfections is the contributions of the higher-order terms from the coherent state and the SPDC sources. The 4-photon tetrahedron state is created in a single spatial mode and a temporal mode coincidentally with a single “herald” photon in another spatial mode. It is only when we detect these five photons simultaneously that we consider that a tetrahedron state has been created and detected. Ideally, in the absence of photon loss of any kind, the number of detected photons is the same as the number of created photons. In a laboratory setting, however, photon loss is unavoidable. As a result, some events with successful post-selection have more than the five intended photons and some photons are lost. The polarization state of four of those remaining photons is not in a tetrahedron state. A detailed model for the loss in this experiment is described in Supplement 1 S1 and S2. To calibrate the model, we performed sets of measurements to find the values of the relevant parameters. These measurements are described in Supplement 1 S3. The parameters include $\eta$ and $\mu$, the squeezing parameters of collinear and non-collinear SPDC sources, respectively. For both sources, the coincidence rate would be proportional to the absolute value squared of these squeezing parameters. To these are added $t$ and $\tau$, the transmission rates for the collinear SPDC and non-collinear SPDC photons, respectively. These include the detection efficiency. The coherent state amplitude, $\alpha$ (the single photon rate is proportional to $|\alpha |^2$), is set to $\sqrt {2\eta t}$ to create the 3-photon $N00N$ state [81]. Loss present in the coherent state path is already captured by this value of $\alpha$. The measured values for $\eta$, $\mu$, $t$, and $\tau$ are 0.078, 0.14, 0.16, and 0.12, respectively. Transmission rates are quite low and this causes the other undesired events to become non-negligible.

We calculated analytical expressions for the probabilities of possible 5-photon events in Supplement 1 S4 and Supplement 1 S5. For example, the probability of the desired 5-photon event that generates a tetrahedron state is proportional to $\eta ^3 t^3 \mu ^2 \tau ^2$. One of the biggest sources of imperfection is measuring an extra photon from the coherent state used to produce the $N00N$ state when one of the photons from the SPDC sources is lost. In these cases, the probability of the case when one of the photons from the collinear SPDC is lost is proportional to $\eta ^4 t^3(1-t) \mu ^2 \tau ^2$, and the probability of the case when the horizontally polarized photon from the non-collinear SPDC is lost is proportional to $\eta ^4 t^4 \mu ^2 \tau (1-\tau )$. All the cases mentioned above result in a 5-photon coincidence and we do not have a method to distinguish them. Substituting the values for our experimental setup, the two undesired events mentioned above have probabilities roughly of the same order of magnitude compared with the probability of generating a tetrahedron state. Several undesired events like the two mentioned above reduce the fidelity and the metrological potential of our state. One possible solution to lower the contributions from the undesired terms was to lower the squeezing parameters and the coherent state amplitude by reducing the single-photon rates. However, the data collection process, explained in Section 3.2, already took eight days even with the current single-photon rates because of the rarity of the 5-photon events and the necessity of regular calibrations every 8–12 hr. Due to these impracticalities, we decided to collect data with the current rates.

3.2 Tomography with Hidden Differences

The Hilbert space for 4-photons’ polarization is $\left (\text {spin-}1/2\right )^{\bigotimes 4} = \left (\text {spin-}2\right ) \bigoplus 3\left (\text {spin-}1\right )\bigoplus 2\left (\text {spin-}0\right )$. Ideally, in this case, the photons are indistinguishable, which constrains their polarization to be full-symmetric and contained in the $\text {spin-}2$ sector. In practice, however, we can expect some small mismatch in the modes of the photons from different sources, which will allow other spin sectors to be populated. Importantly, the parasitic distinguishability between the photons cannot be accessed, i.e., we have no practical way of favorably selecting photons from one source over another, which means that the only measurements we can perform have projectors symmetric under particle exchange. As a result, only the parts of the density matrix depicted in Fig. 3(a) can be reconstructed. They specifically consist of the spin sectors in the tensor sum decomposition of the Hilbert space. The coherences between these sectors are inaccessible. Furthermore, we are only sensitive to the sum of the different spin sectors with the same spin value and choose to report their average in each of the different sectors. Importantly, although our knowledge of the full state of the system is fundamentally limited in this way, the accessible information is enough to predict the result of any symmetric measurement done on the same state.

 figure: Fig. 3.

Fig. 3. (a) Layout of the full density matrix. Here $\rho _\text {sym}$ is the fully symmetric spin-2 sector (blue). Spin-1 (red) and spin-0 (green) sectors are also shown. (b) The full tomographically reconstructed density matrix, $\rho _\text {exp}$. In hidden differences tomography, coherences between spin sectors are neglected and sectors of the same spins are chosen to be identical. (c) Real and imaginary parts of the reconstructed $\rho _\text {sym}$. The density matrix of the ideal tetrahedron state is shown for comparison (dashed). Elements of the density matrix are labeled with spin notation $| {j, m}\rangle$. (d) Real and imaginary parts of a spin-1 sector. The population in each spin-0 sector is less than $10^{-10}$.

Download Full Size | PDF

The technique we use to reconstruct what part we need of the density matrix is called “tomography with hidden differences” [8385]. It consists of a sequence of measurements on the polarization of our state akin to Stern–Gerlach measurements for a spin. The relevant section of the experimental apparatus is depicted in Fig. 2(d). We send photons through a PBS and then through networks of beamsplitters (BS) with every output leading to a SPCM. When four photons are simultaneously detected in coincidence with the herald photon in Fig. 2(b), we consider it a successful event and count the number of photons transmitted versus reflected at the PBS. We perform this polarization measurement in 13 different bases, chosen to be roughly uniformly spread on the Bloch sphere by rotating the quarter-waveplate (QWP) and HWP before the PBS. With only three SPCMs on each output of the PBS, we cannot detect two of the five potential 4-photon outcomes, when the photons are all transmitted, or all reflected. As for the other projectors, they need to be weighted by the likelihood that each photon lands on a different detector. A detailed discussion of this reweighing is included in Supplement 1 S2. With 13 measurement bases, however, we can still perform 39 different projections, which is enough to estimate the 35 linearly independent parameters of the accessible section of the density matrix using a maximum likelihood method.

Overall, 2434 successful events were recorded in 85 hr of counting time. These counting hours were distributed over the course of eight days. Every 8–12 hr, minor realignments were made to readjust the couplings in single-mode fibers and to reset the set-point of the locked interferometer.

Figure 3 shows the tomographic reconstruction of the density matrix. For ease of comparison with the theoretical tetrahedron state, we rotate our reconstructed density matrix by an optimal angle $\phi$ around the H/V axis in post-processing. We set $\phi$ by maximizing the fidelity ($F = \langle{\psi _{tetra}}|e^{-i\hat {J}_z\phi }\hat{\rho} e^{i\hat {J}_z\phi }| {\psi _{tetra}}\rangle$) between the unrotated reconstructed state and the tetrahedron state and find a value of $\phi = 0.135$ for a fidelity with the tetrahedron state of $(0.46 \pm 0.02)$.

Error bars on the all the entries of the density matrix were determined by a Monte Carlo simulation of the tomographic reconstruction process, where the simulated number of detection events for each projection was drawn from a Poisson distribution centered on the actual measurement results.

4. Discussion

Despite the low fidelity between the theoretical tetrahedron state and our reconstructed density matrix, it retains many of the notable features of the former.

As expected, most of the population ($87 \pm 4{\% }$) is in the full-symmetric, spin-2 subspace [$\rho _{\mathrm {sym}}$ in Fig. 3(a)]. The excess population in the other sectors is small but still non-zero, indicating that there were small mode mismatches between the different photons. Indeed, the visibility of a 2-photon $N00N$ state created with a very low rate ($\sim 1kHz$) was measured to be around $88{\% }$, indicating some mode mismatch between the SPDC photons and coherent state photons. This measurement is described in more detail in Supplement 1 S6. As intended, the two basis elements with the highest population in $\rho _{\mathrm {sym}}$ are $|{2,2}\rangle$ and $|{2,-1}\rangle$. The ratio between these populations, however, is $(0.73 \pm 0.12)$, which is significantly larger than the theoretical value of $1/2$. We unfortunately measure significant population ($\approx 10{\% }$) in the states $|{2,0}\rangle$ and $|{2,-2}\rangle$ that are meant to be empty. Finally, the largest coherence measured is between $|{2,2}\rangle$ and $|{2,-1}\rangle$. Theoretically, this is the only non-zero coherence and it takes a value of $\sqrt {2}/3\approx 0.47$. Here, we measure the significantly lower value of $(0.13 \pm 0.02)$, also much lower than $(0.32 \pm 0.02)$, the maximum possible value given the corresponding population. The coherence quoted has no imaginary part specifically because we allowed an optimal rotation around the H/V axis in the reconstructed density matrix. The purity of the measured state is $\text {Tr}[\rho _{\text {exp}}^2] = (0.32 \pm 0.02)$ and the purity of the symmetric part of the density matrix is $(0.43 \pm 0.02)$.

Most of these discrepancies can be explained by considering the contributions of terms created with six or more photons but detected as a successful five-fold coincidence due to photon loss at the detectors or in the optical apparatus. Using the model described in Supplement 1 S1, we compute the theoretical contribution of higher-order photon creation events with the presence of loss in our apparatus, and compare them with our experimental results. Because loss acts symmetrically on every photon, the model does not predict the moderate portion of the state not contained in the full-symmetric sector. It is, however, a great predictor for our results within said sector. When calibrated for our apparatus, we predict a state which has a $91{\% }$ fidelity with the symmetric portion of our experimental state. The state in question can be seen in Supplement 1 S4. This result indicates that the imperfection of our schemes are very well understood and that we can reliably infer the technical requirements required for the creation of a higher quality state. In particular, while keeping the same tetrahedron state detection rate, but while imagining switching the cascade of BS terminated with SPCMs (with assumed detection efficiency of $\sim 65{\% }$) for state-of-the-art photon-number-resolving detectors with near-unity efficiency [86], our model predicts a state with a fidelity of 0.80 with the ideal tetrahedron state, a notable improvement over the fidelity of 0.54 for the state predicted with our current parameters or over the fidelity of 0.53 for the re-normalized fully symmetric sector of our experimental density matrix.

Despite these imperfections, the main eigenstate of the reconstructed density matrix is very similar to the theoretical tetrahedron state with $92{\% }$ fidelity. We can interpret this to mean that our method succeeds in preparing a near-optimal tetrahedron state around half the time, but fails for the remaining half, in good part due to events when more photons were created than intended. A comparison of their Majorana representations is depicted in Fig. 4(a).

 figure: Fig. 4.

Fig. 4. (a) Majorana representations for the ideal tetrahedron state (red) and the eigenstate of the reconstructed density matrix with the largest eigenvalue (blue). (b) Longitude–latitude map of the Wigner representation for the fully symmetric portion of the ideal tetrahedron state. Points on the tetrahedron representation correspond to those on the map. The $y$-axis is the polar angle and the $x$-axis is the azimuthal angle in spherical coordinates. Polar angle is defined to be $\pi /2$ at the top corner and azimuthal angle is defined to be 0 at the leftmost corner. (c)–(f) The reconstructed Wigner representation of the fully symmetric portion of the state (not of the main eigenstate) when a $2\pi /3$ rotation is applied about each of the four points of the tetrahedron.

Download Full Size | PDF

Another feature of note is that our state keeps some of the tetrahedral symmetry that gives the tetrahedron state its name. This can be readily seen in the Wigner function [8789] of our reconstructed state plotted in Fig. 4. Because this Wigner function has the domain of a sphere, we display four different projections, each centered on a different vertex of a tetrahedron. These four different projections of the Wigner function of the theoretical tetrahedron state are the same [shown in Fig. 4(b)] by virtue of its symmetry. The different projections of the symmetric portion of the reconstructed state are plotted in Figs. 4(c)–4(f). Although the features are not as pronounced, we notice that, the troughs and the peaks of the Wigner function align with those of the theoretical state. Indeed, despite the contamination of the noise coming from higher-order events, the reconstructed Wigner function qualitatively keeps, with some imperfections, the theoretical symmetry of the tetrahedron state. We note that a deviation from the optimally chosen phase $\phi$ would simply rotate the Wigner function around the $z$-axis and would not change the symmetry of the results.

We can also directly assess how well our state satisfies the unpolarized properties of Eq. (19). One direct measure of this property is by expanding the state’s quasiprobability distribution on a sphere in terms of multi-pole moments. Unpolarized states have vanishing first and second moments. Our state’s first and second moments are $1.6\times 10^{-2}$ and $3.1\times 10^{-2}$, respectively. This is much smaller than for a comparable “classical” state, with first- and second-order multi-pole moments summing to $2/5$ and $2/7$, respectively, and is even much smaller than for a state whose Majorana constellation is chosen randomly [90], with first- and second-order moments summing to $2.8\times 10^{-1}$ and $2.3\times 10^{-1}$, respectively.

The qualitative features can be further observed in Fig. 5, which compares the performance of the experimental state and the main eigenstate with the ideal tetrahedron state for the measurement of single rotations angles, that is the measurement of $\theta _x$, $\theta _y$, or $\theta _z$ when knowing that the other two angles take a value of zero. We plot the numerically calculated result of a projection on the original tetrahedron state following the rotation. It can be seen that both for the experimental state and the main eigenstate the visibility is not as high as the ideal tetrahedron state since they do not have perfect overlap with the tetrahedron state, especially the experimental state has very low visibility due to low fidelity and low purity. However, all axes still have three maxima per rotation, a signature of tetrahedral symmetry. For example, the 4-photon spin coherent state would have a single maximum, while the 4-photon $N00N$ state would exhibit four maxima for the $z$-axis, and two for the $x$-axis and the $y$-axis. The main experimental eigenstate closely approximates the tetrahedron state with similar levels of visibility for all rotations shown in the figure.

 figure: Fig. 5.

Fig. 5. Numerically calculated projection on the original tetrahedron state after rotations around the $x,y$, and $z$ axes on Bloch sphere for the following states: ideal tetrahedron state (dashed orange); experimental density matrix obtained from the tomography (solid black); and the main eigenstate of the experimental density matrix (solid red). The $x$-axis of the figure is the rotation amplitude in radians and the $y$-axis is the overlap between the rotated state and the ideal tetrahedron state.

Download Full Size | PDF

Finally, returning to multi-parameter estimation in Fig. 6, we compare the s-QCRB calculated for our experimental state with the theoretical estimation strategies outlined in Section 2.4. As mentioned in that section, spin-coherent states have shot-noise scaling. Although $N00N$ states have Heisenberg scaling for one axis, other axes have shot-noise scaling and they dominate the resulting s-QCRB. To obtain Heisenberg scaling with $N00N$ states, one has to use three batches aligned in the $x$, $y$, and $z$ axes sequentially. Since one does not need a sequential strategy for Platonic states, they yield an s-QCRB three times better than sequential $N00N$ states. The figure has two data points, one for the experimental density matrix reconstructed from the tomography, and one for the main eigenstate of the density matrix. The s-QCRB for them are calculated theoretically from the reconstructed density matrix and the error bars are obtained using Monte Carlo simulations. Having an s-QCRB of $0.91 \pm 0.09$, the full density matrix narrowly beats the strategy with coherent states aligned on different axes (which has an s-QCRB of $1.125$), but fails to surpass the strategy of using sequential $N00N$ states (which has an s-QCRB of 0.675), indicating that a quantum advantage can be obtained using our state but that the experimental fidelity needs to be improved before a better-than-$N00N$-state advantage can be harnessed in a real-world metrological task. The dominant eigenstate on the other hand has an s-QCRB of $0.4 \pm 0.05$, very close to the ideal value of 0.375, beating all other strategies under discussion.

 figure: Fig. 6.

Fig. 6. Scalar quantum Cramér–Rao bound $\mathcal {C}$ as a function of total number of photons. Solid lines represent the different strategies enumerated in Section 2.4: spin-coherent states (purple triangles); $N00N$ states (red squares for simultaneous; orange pentagons for sequential); and Platonic states (yellow circles). Sequential strategies use three batches of states aligned in the $x, y$, and $z$ axes, whereas simultaneous strategies use only a single batch of identical states. The s-QCRB decreases with shot-noise scaling for sequential coherent states and simultaneous $N00N$ states, and decreases with Heisenberg scaling for sequential $N00N$ states and Platonic states. The s-QCRB of the experimentally reconstructed density matrix of the tetrahedron state is represented by a cross symbol. The black star represents the s-QCRB of its main eigenstate.

Download Full Size | PDF

Improvements in existing technology can render the creation of Platonic states more feasible. In particular, as we discussed, multi-photon emissions are one of the main reasons for our low fidelity. These could be reduced by reducing the loss in our experiment. The loss upon detection, in particular, has a great impact on our result as we have seen. Another avenue to reduce the contributions of these terms would be to reduce their probability of creation in the first place, prior to the loss. One approach for doing this is to change our scheme to rid ourselves of the coherent state. The coherent state always leaves the possibility of a single extra photon being created, a relatively likely event which an experiment relying only on SPDC would not allow. Recent work on automating the process of finding quantum optics experiments have yielded alternate schemes for the creation of the tetrahedron state which rely only on parametric downconversion [91]. A particular challenge in this experiment is that the photons from different sources need to be spectrally indistinguishable. To achieve this we use narrowband filters in our experiment. However, this comes at the cost of decreasing the coupling efficiency; thus, increasing the multi-photon contamination. These challenges could be simultaneously overcome by the use of modern quantum dot single photon sources [92,93]. Such sources have recently been used for multi-photon interference experiments with up to 20 photons [94]. With these on-demand photon sources one can imagine building the Platonic states using the methods of Refs. [23,85,95]. In this approach, independent photons are non-deterministically combined into the same spatial mode one at a time, each with a polarization state corresponding to a Majorana point. Although the state-creation methods of Ref. [95] are probabilistic, given the high degree of indistinguishability and almost-zero multi-photon emission rates of quantum dot sources, the post-selected multi-photon states created from this method should not suffer the low fidelity we face here. Thus, as quantum technology continues to improve it should be possible to harness the full potential of the Platonic states.

5. Conclusion

In conclusion, we demonstrated a scheme for the creation of the tetrahedron state, a 4-photon polarization state optimal in quantum mechanics for the estimation of polarization rotations. While a correctly aligned $N00N$ state can measure the rotation angle of a rotation with a known axis with Heisenberg scaling, one would need three $N00N$ state copies, each aligned with the different Cartesian axis to measure the three parameters of a spin rotation with the same scaling. This resource increase leads to a degradation of the asymptotic performance by a factor of three in the appropriately weighted sum of the measured parameters’ variances. The tetrahedron state, and other second-order unpolarized states, can be used to simultaneously measure all of the parameters of the rotation with Heisenberg scaling and are therefore not affected by this performance hit. Generalizations of these states tailored to transformations described with a minimum of $d$ parameters would yield $d$-fold improvements in the weighted sum of variances of said $d$ parameters. In this work, we created the tetrahedron state in the polarization of four photons by combining three photons themselves in a $N00N$ state attenuated along the horizontal axis with a horizontally polarized single photon. Implementing this required interfering the output of two SPDC sources and a coherent state. The main source of noise in our experiment comes from events where more photons were created than those needed for the tetrahedron state. These events could be made less significant by reducing the optical losses, which in our experiment could be achieved with better sources and detectors. Furthermore, in principle, the rate of these insidious events can always be made proportionally arbitrarily small by decreasing the overall tetrahedron state creation rate and increasing the counting time accordingly. In part because of the low 4-photon rate of our experiment, we were unable use the quantum state we created to measure parameters of a polarization rotation. Nevertheless, at the proof-of-concept level, we have created and characterized a state that shares many of the interesting qualitative features and symmetries of the optimal tetrahedron state, showing that our method is fundamentally sound. We also have shown that we have a good understanding of the technical limitations of our method and have discussed how to improve upon them, paving the way for future experiments. Continued progress in developing methods to robustly engineer states in high-dimensional Hilbert spaces, both in optics and in other physical media, ensure that it will soon be possible to generate with high fidelity states such as these which will enable significant advantages in many multi-parameter metrology applications, ranging from optical ellipsometry to 3D magnetometry.

Funding

Natural Sciences and Engineering Research Council of Canada (RGPIN-2020-05767, ALLRP 578468 – 22, PDF); Canadian Institute for Advanced Research; Fetzer Franklin Fund (FQXiRFP-1819); Austrian Science Fund (FWF) through BeyondC (F7113).

Acknowledgments

We acknowledge the support of the Natural Sciences and Engineering Research Council of Canada (NSERC), via Discovery grant (RGPIN-2020-05767), via support of the QuEnSi quantum alliance (ALLRP 578468 – 22), and via their support of A.Z.G. through a postdoctoral fellowship. This work was also supported by CIFAR and the Austrian Science Fund (FWF), through BeyondC (F7113). It made use of some equipment purchased through grant FQXiRFP-1819 from the Foundational Questions Institute and Fetzer Franklin Fund, a donor-advised fund of Silicon Valley Community Foundation. H.F., Y.B.Y., N.L.-G., A.O.T.P., and A.M.S. wish to acknowledge that the University of Toronto operates on the traditional land of the Huron-Wendat, the Seneca, and the Mississaugas of the Credit. K.B.-F. and A.Z.G. acknowledge that the NRC headquarters is located on the traditional unceded territory of the Algonquin Anishinaabe and Mohawk people.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. V. Giovannetti, S. Lloyd, and L. Maccone, “Quantum-enhanced measurements: beating the standard quantum limit,” Science 306, 1330–1336 (2004). [CrossRef]  

2. V. Giovannetti, S. Lloyd, and L. Maccone, “Quantum metrology,” Phys. Rev. Lett. 96, 010401 (2006). [CrossRef]  

3. V. Giovannetti, S. Lloyd, and L. Maccone, “Advances in quantum metrology,” Nat. Photonics 5, 222–229 (2011). [CrossRef]  

4. C. M. Caves, “Quantum-mechanical noise in an interferometer,” Phys. Rev. D 23, 1693–1708 (1981). [CrossRef]  

5. M. G. A. Paris, “Quantum estimation for quantum technology,” Int. J. Quantum Inform. 07, 125–137 (2009). [CrossRef]  

6. M. F. Riedel, P. Böhi, Y. Li, et al., “Atom-chip-based generation of entanglement for quantum metrology,” Nature 464, 1170–1173 (2010). [CrossRef]  

7. C. F. Ockeloen, R. Schmied, M. F. Riedel, et al., “Quantum metrology with a scanning probe atom interferometer,” Phys. Rev. Lett. 111, 143001 (2013). [CrossRef]  

8. R. J. Sewell, M. Koschorreck, M. Napolitano, et al., “Magnetic sensitivity beyond the projection noise limit by spin squeezing,” Phys. Rev. Lett. 109, 253605 (2012). [CrossRef]  

9. L. Collaboration, “Enhanced sensitivity of the LIGO gravitational wave detector by using squeezed states of light,” Nat. Photonics 7, 613–619 (2013). [CrossRef]  

10. W. Wasilewski, K. Jensen, H. Krauter, et al., “Quantum noise limited and entanglement-assisted magnetometry,” Phys. Rev. Lett. 104, 133601 (2010). [CrossRef]  

11. M. Koschorreck, M. Napolitano, B. Dubost, et al., “Sub-Projection-Noise Sensitivity in Broadband Atomic Magnetometry,” Phys. Rev. Lett. 104, 093602 (2010). [CrossRef]  

12. J. B. Brask, R. Chaves, and J. Kolodynski, “Improved quantum magnetometry beyond the standard quantum limit,” Phys. Rev. X 5, 031010 (2015). [CrossRef]  

13. A. D. Ludlow, M. M. Boyd, J. Ye, et al., “Optical atomic clocks,” Rev. Mod. Phys. 87, 637–701 (2015). [CrossRef]  

14. M. Schulte, C. Lisdat, P. O. Schmidt, et al., “Prospects and challenges for squeezing-enhanced optical atomic clocks,” Nat. Commun. 11, 5955 (2020). [CrossRef]  

15. B. C. Sanders, “Quantum dynamics of the nonlinear rotator and the effects of continual spin measurement,” Phys. Rev. A 40, 2417–2427 (1989). [CrossRef]  

16. A. N. Boto, P. Kok, D. S. Abrams, et al., “Quantum interferometric optical lithography: exploiting entanglement to beat the diffraction limit,” Phys. Rev. Lett. 85, 2733–2736 (2000). [CrossRef]  

17. H. Lee, P. Kok, and J. P. Dowling, “A quantum Rosetta stone for interferometry,” J. Mod. Opt. 49, 2325–2338 (2002). [CrossRef]  

18. J. P. Dowling, “Quantum optical metrology - the lowdown on high-N00N states,” Contemp. Phys. 49, 125–143 (2008). [CrossRef]  

19. S. Slussarenko, M. M. Weston, H. M. Chrzanowski, et al., “Unconditional violation of the shot-noise limit in photonic quantum metrology,” Nat. Photonics 11, 700–703 (2017). [CrossRef]  

20. C. You, M. Hong, P. Bierhorst, et al., “Scalable multiphoton quantum metrology with neither pre-nor post-selected measurements,” Appl. Phys. Rev. 8, 1 (2021). [CrossRef]  

21. J. P. Dowling, “Correlated input-port, matter-wave interferometer: quantum-noise limits to the atom-laser gyroscope,” Phys. Rev. A 57, 4736–4746 (1998). [CrossRef]  

22. G. Gilbert, M. Hamrick, and Y. S. Weinstein, “Use of maximally entangled N-photon states for practical quantum interferometry,” J. Opt. Soc. Am. B 25, 1336–1340 (2008). [CrossRef]  

23. M. W. Mitchell, J. S. Lundeen, and A. M. Steinberg, “Super-resolving phase measurements with a multiphoton entangled state,” Nature 429, 161–164 (2004). [CrossRef]  

24. P. Walther, J.-W. Pan, M. Aspelmeyer, et al., “De Broglie wavelength of a non-local four-photon state,” Nature 429, 158–161 (2004). [CrossRef]  

25. T. Nagata, R. Okamoto, J. L. O’Brien, et al., “Beating the standard quantum limit with four-entangled photons,” Science 316, 726–729 (2007). [CrossRef]  

26. I. Afek, O. Ambar, and Y. Silberberg, “High-NOON states by mixing quantum and classical light,” Science 328, 879–881 (2010). [CrossRef]  

27. Y. Israel, I. Afek, S. Rosen, et al., “Experimental tomography of NOON states with large photon numbers,” Phys. Rev. A 85, 022115 (2012). [CrossRef]  

28. Y. Israel, S. Rosen, and Y. Silberberg, “Supersensitive polarization microscopy using NOON states of light,” Phys. Rev. Lett. 112, 103604 (2014). [CrossRef]  

29. T. Monz, P. Schindler, J. T. Barreiro, et al., “14-Qubit entanglement: creation and coherence,” Phys. Rev. Lett. 106, 130506 (2011). [CrossRef]  

30. W.-B. Gao, C.-Y. Lu, X.-C. Yao, et al., “Experimental demonstration of a hyper-entangled ten-qubit Schrödinger cat state,” Nat. Phys. 6, 331–335 (2010). [CrossRef]  

31. W. H. Zurek, “Sub-Planck structure in phase space and its relevance for quantum decoherence,” Nature 412, 712–717 (2001). [CrossRef]  

32. F. Toscano, D. A. R. Dalvit, L. Davidovich, et al., “Sub-Planck phase-space structures and Heisenberg-limited measurements,” Phys. Rev. A 73, 023803 (2006). [CrossRef]  

33. M. Szczykulska, T. Baumgratz, and A. Datta, “Multi-parameter quantum metrology,” Adv. Phys.: X 1, 621–639 (2016). [CrossRef]  

34. A. Z. Goldberg and D. F. V. James, “Quantum-limited Euler angle measurements using anticoherent states,” Phys. Rev. A 98, 032113 (2018). [CrossRef]  

35. P. C. Humphreys, M. Barbieri, A. Datta, et al., “Quantum enhanced multiple phase estimation,” Phys. Rev. Lett. 111, 070403 (2013). [CrossRef]  

36. J. S. Sidhu and P. Kok, “Geometric perspective on quantum parameter estimation,” AVS Quantum Sci. 2, 014701 (2020). [CrossRef]  

37. R. Demkowicz-Dobrzanski, W. Górecki, and M. Guta, “Multi-parameter estimation beyond quantum Fisher information,” J. Phys. A: Math. Theor. 53, 363001 (2020). [CrossRef]  

38. H. Zhu, “Information complementarity: A new paradigm for decoding quantum incompatibility,” Sci. Rep. 5, 14317 (2015). [CrossRef]  

39. T. Heinosaari, T. Miyadera, and M. Ziman, “An invitation to quantum incompatibility,” J. Phys. A: Math. Theor. 49, 123001 (2016). [CrossRef]  

40. S. Ragy, M. Jarzyna, and R. Demkowicz-Dobrzanski, “Compatibility in multiparameter quantum metrology,” Phys. Rev. A 94, 052108 (2016). [CrossRef]  

41. F. Albarelli, M. Barbieri, M. G. Genoni, et al., “A perspective on multiparameter quantum metrology: from theoretical tools to applications in quantum imaging,” Phys. Lett. A 384, 126311 (2020). [CrossRef]  

42. E. Polino, M. Valeri, N. Spagnolo, et al., “Photonic quantum metrology,” AVS Quantum Sci. 2, 024703 (2020). [CrossRef]  

43. J. Zimba, “Anticoherent spin states via the Majorana representation,” Electronic Journal of Theoretical Physics 3, 143–156 (2006).

44. D. Baguette, F. Damanet, O. Giraud, et al., “Anticoherence of spin states with point-group symmetries,” Phys. Rev. A 92, 052333 (2015). [CrossRef]  

45. D. Baguette and J. Martin, “Anticoherence measures for pure spin states,” Phys. Rev. A 96, 032304 (2017). [CrossRef]  

46. O. Giraud, P. Braun, and D. Braun, “Quantifying quantumness and the quest for Queens of Quantum,” New J. Phys. 12, 063005 (2010). [CrossRef]  

47. D. N. Klyshko, “Multiphoton interference and polarization effects,” Phys. Lett. A 163, 349–355 (1992). [CrossRef]  

48. P. de la Hoz, G. Björk, A. B. Klimov, et al., “Unpolarized states and hidden polarization,” Phys. Rev. A 90, 043826 (2014). [CrossRef]  

49. G. Björk, M. Grassl, P. D. L. Hoz, et al., “Stars of the quantum Universe: extremal constellations on the Poincaré sphere,” Phys. Scr. 90, 108008 (2015). [CrossRef]  

50. G. Björk, A. B. Klimov, P. de la Hoz, et al., “Extremal quantum states and their Majorana constellations,” Phys. Rev. A 92, 031801 (2015). [CrossRef]  

51. C. Chryssomalakos and H. Hernández-Coronado, “Optimal quantum rotosensors,” Phys. Rev. A 95, 052125 (2017). [CrossRef]  

52. J. Martin, S. Weigert, and O. Giraud, “Optimal detection of rotations about unknown axes by coherent and anticoherent states,” Quantum 4, 285 (2020). [CrossRef]  

53. A. Z. Goldberg, A. B. Klimov, M. Grassl, et al., “Extremal quantum states,” AVS Quantum Sci. 2, 044701 (2020). [CrossRef]  

54. A. Baecklund and I. Bengtsson, “Four remarks on spin coherent states,” Phys. Scr. T163, 014012 (2014). [CrossRef]  

55. M. Aulbach, D. Markham, and M. Murao, “The maximally entangled symmetric state in terms of the geometric measure,” New J. Phys. 12, 073025 (2010). [CrossRef]  

56. J. Martin, O. Giraud, P. A. Braun, et al., “Multiqubit symmetric states with high geometric entanglement,” Phys. Rev. A 81, 062347 (2010). [CrossRef]  

57. W. Ganczarek, M. Kus, and K. Zyczkowski, “Barycentric measure of quantum entanglement,” Phys. Rev. A 85, 032314 (2012). [CrossRef]  

58. A. Burchardt, J. Czartowski, and K. Zyczkowski, “Entanglement in highly symmetric multipartite quantum states,” Phys. Rev. A 104, 022426 (2021). [CrossRef]  

59. A. Z. Goldberg, M. Grassl, G. Leuchs, et al., “Quantumness beyond entanglement: the case of symmetric states,” Phys. Rev. A 105, 022433 (2022). [CrossRef]  

60. C. Chryssomalakos, L. Hanotel, E. Guzmán-González, et al., “Symmetric multiqudit states: stars, entanglement, and rotosensors,” Phys. Rev. A 104, 012407 (2021). [CrossRef]  

61. A. Z. Goldberg, P. D. L. Hoz, G. Björk, et al., “Quantum concepts in optical polarization,” Adv. Opt. Photonics 13, 1–73 (2021). [CrossRef]  

62. F. Bouchard, P. D. L. Hoz, G. Björk, et al., “Quantum metrology at the limit with extremal Majorana constellations,” Optica 4, 1429–1432 (2017). [CrossRef]  

63. R. Gutiérrez-Cuevas and M. A. Alonso, “Platonic Gaussian beams: wave and ray treatment,” Opt. Lett. 45, 6759–6762 (2020). [CrossRef]  

64. M. Eriksson, A. Goldberg, M. Hiekkamäki, et al., “Sensing rotations with multiplane light conversion,” Phys. Rev. Appl. 20, 024052 (2023). [CrossRef]  

65. C. Helstrom and R. Kennedy, “Noncommuting observables in quantum detection and estimation theory,” IEEE Trans. Inform. Theory 20, 16–24 (1974). [CrossRef]  

66. A. Z. Goldberg, A. B. Klimov, G. Leuchs, et al., “Rotation sensing at the ultimate limit,” J. Phys. Photonics 3, 022008 (2021). [CrossRef]  

67. S. Helgason, Differential Geometry, Lie Groups, and Symmetric Spaces (Academic Press, 1979).

68. A. Z. Goldberg, L. L. Sánchez-Soto, and H. Ferretti, “Intrinsic sensitivity limits for multiparameter quantum metrology,” Phys. Rev. Lett. 127, 110501 (2021). [CrossRef]  

69. P. Kolenderski and R. Demkowicz-Dobrzanski, “Optimal state for keeping reference frames aligned and the platonic solids,” Phys. Rev. A 78, 052333 (2008). [CrossRef]  

70. E. Majorana, “Atomi orientati in campo magnetico variabile,” Nuovo Cimento 9, 43–50 (1932). [CrossRef]  

71. I. Bengtsson and K. Zyczkowski, Geometry of Quantum States: An Introduction to Quantum Entanglement, 2nd ed (Cambridge University Press, 2017).

72. “Extremal Polarization States,”.

73. R. H. Hardin and N. J. A. Sloane, “McLaren’s improved snub cube and other new spherical designs in three dimensions,” Discrete Comput. Geom. 15, 429–441 (1996). [CrossRef]  

74. J. Crann, R. Pereira, and D. W. Kribs, “Spherical designs and anticoherent spin states,” J. Phys. A: Math. Theor. 43, 255307 (2010). [CrossRef]  

75. E. Bannai and M. Tagami, “A note on anticoherent spin states,” J. Phys. A: Math. Theor. 44, 342002 (2011). [CrossRef]  

76. J. Thomson, “XXIV. On the structure of the atom: an investigation of the stability and periods of oscillation of a number of corpuscles arranged at equal intervals around the circumference of a circle; with application of the results to the theory of atomic structure,” London, Edinburgh, and Dublin Philos. Mag. and J. Sci. 7, 237–265 (1904). [CrossRef]  

77. P. M. L. Tammes, “On the origin of number and arrangement of the places of exit on the surface of pollen-grains,” Recueil des travaux botaniques néerlandais 27, 1–84 (1930).

78. L. L. Whyte, “Unique arrangements of points on a sphere,” Am. Math. Mon. 59, 606–611 (1952). [CrossRef]  

79. J. H. Conway, R. H. Hardin, and N. J. A. Sloane, “Packing lines, planes, etc.: packings in Grassmannian spaces,” Exp. Math. 5, 139–159 (1996). [CrossRef]  

80. E. B. Saff and A. B. J. Kuijlaars, “Distributing many points on a sphere,” Math. Intell. 19, 5–11 (1997). [CrossRef]  

81. H. F. Hofmann and T. Ono, “High-photon-number path entanglement in the interference of spontaneously down-converted photon pairs with coherent laser light,” Phys. Rev. A 76, 031806 (2007). [CrossRef]  

82. L. A. Rozema, D. H. Mahler, R. Blume-Kohout, et al., “Optimizing the Choice of Spin-Squeezed States for Detecting and Characterizing Quantum Processes,” Phys. Rev. X 4, 041025 (2014). [CrossRef]  

83. R. B. A. Adamson, L. K. Shalm, M. W. Mitchell, et al., “Multiparticle State Tomography: Hidden Differences,” Phys. Rev. Lett. 98, 043601 (2007). [CrossRef]  

84. R. B. A. Adamson, P. S. Turner, M. W. Mitchell, et al., “Detecting hidden differences via permutation symmetries,” Phys. Rev. A 78, 033832 (2008). [CrossRef]  

85. L. K. Shalm, R. B. A. Adamson, and A. M. Steinberg, “Squeezing and over-squeezing of triphotons,” Nature 457, 67–70 (2009). [CrossRef]  

86. A. E. Lita, A. J. Miller, and S. W. Nam, “Counting near-infrared single-photons with 95% efficiency,” Opt. Express 16, 3032–3040 (2008). [CrossRef]  

87. E. Wigner, “On the quantum correction for thermodynamic equilibrium,” Phys. Rev. 40, 749–759 (1932). [CrossRef]  

88. J. P. Dowling, G. S. Agarwal, and W. P. Schleich, “Wigner distribution of a general angular-momentum state: applications to a collection of two-level atoms,” Phys. Rev. A 49, 4101–4109 (1994). [CrossRef]  

89. R. Schmied and P. Treutlein, “Tomographic reconstruction of the Wigner function on the Bloch sphere,” New J. Phys. 13, 065019 (2011). [CrossRef]  

90. A. Z. Goldberg, J. L. Romero, A. S. Sanz, et al., “Random Majorana constellations,” (2021).

91. C. Ruiz-Gonzalez, S. Arlt, J. Petermann, et al., “Digital discovery of 100 diverse quantum experiments with PyTheus,” Quantum 7, 1204 (2023). [CrossRef]  

92. N. Tomm, A. Javadi, N. O. Antoniadis, et al., “A bright and fast source of coherent single photons,” Nat. Nanotechnol. 16, 399–403 (2021). [CrossRef]  

93. H. Wang, Y.-M. He, T.-H. Chung, et al., “Towards optimal single-photon sources from polarized microcavities,” Nat. Photonics 13, 770–775 (2019). [CrossRef]  

94. H. Wang, J. Qin, X. Ding, et al., “Boson sampling with 20 input photons and a 60-mode interferometer in a 1014-dimensional hilbert space,” Phys. Rev. Lett. 123, 250503 (2019). [CrossRef]  

95. J. Fiurášek, “Conditional generation of N-photon entangled states of light,” Phys. Rev. A 65, 053818 (2002). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Document Noise model details; Section S1-S6

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. Majorana representations of (a) spin coherent state, (b) 4-photon $N00N$ state, and (c) the tetrahedron state. (d) Conceptual illustration of the tetrahedron’s state creation process. The H polarization of a 3-photon $N00N$ state (three transparent red points on the vertices of an equilateral triangle on the equator) is partially attenuated resulting in the triangle of points with lower latitude (red). Simultaneously, an H polarized heralded single-photon is probabilistically added (green). The result is a 4-photon state represented by four points on the vertices of a tetrahedron.
Fig. 2.
Fig. 2. Experiment overview. (a) An 808 nm pulsed beam (red) is used to generate a 404 nm beam (blue) through second harmonic generation (SHG) in a 1 mm barium borate (BBO) crystal. Part of this blue beam is sent to generate co-propagating pairs of horizontally polarized 808 nm photons (red) through type-I spontaneous parametric downconversion (SPDC) in a 2 mm BBO crystal while the red beam is attenuated and vertically polarized. The two beams are recombined on a polarizing beamsplitter (PBS) and polarization rotated to form a 3-photon $N00N$ state in the H/V basis. (b) Another blue beam then a pumps a second SPDC (2 mm BBO) to generate another photon pair. Here, after a 3 nm narrowband filter, a detection at a single-photon counting module (SPCM) labeled “T" heralds the presence of a photon in the upper path. (c) Both the $N00N$ state and heralded single photon enter a displaced polarization Sagnac interferometer. The heralded photon strictly takes the clockwise path while the $N00N$ state is split equally amongst the paths. The half-waveplate (HWP) in the counter-clockwise path is set at $45 ^{\circ }$ to transmit all the light in the upwards output while the HWP in the clockwise path is set at $22.5 ^{\circ }$ to attenuate the V polarization of the $N00N$ state by half and transmit the heralded single-photon half the time (see Fig. 1 for context). After being recombined, all the photons pass through a 3 nm narrowband filter to make sure their spectra are the same. (d) The state is transmitted via single-mode (SM) fiber to a tomography apparatus for projection onto an arbitrary polarization axis. Each output path is coupled to a multi-mode fiber splitter to allow for partial photon number counting capabilities (a maximum of three detections per polarization). Five-fold detection events between T and any four SPCMs in the tomography setup are recorded.
Fig. 3.
Fig. 3. (a) Layout of the full density matrix. Here $\rho _\text {sym}$ is the fully symmetric spin-2 sector (blue). Spin-1 (red) and spin-0 (green) sectors are also shown. (b) The full tomographically reconstructed density matrix, $\rho _\text {exp}$. In hidden differences tomography, coherences between spin sectors are neglected and sectors of the same spins are chosen to be identical. (c) Real and imaginary parts of the reconstructed $\rho _\text {sym}$. The density matrix of the ideal tetrahedron state is shown for comparison (dashed). Elements of the density matrix are labeled with spin notation $| {j, m}\rangle$. (d) Real and imaginary parts of a spin-1 sector. The population in each spin-0 sector is less than $10^{-10}$.
Fig. 4.
Fig. 4. (a) Majorana representations for the ideal tetrahedron state (red) and the eigenstate of the reconstructed density matrix with the largest eigenvalue (blue). (b) Longitude–latitude map of the Wigner representation for the fully symmetric portion of the ideal tetrahedron state. Points on the tetrahedron representation correspond to those on the map. The $y$-axis is the polar angle and the $x$-axis is the azimuthal angle in spherical coordinates. Polar angle is defined to be $\pi /2$ at the top corner and azimuthal angle is defined to be 0 at the leftmost corner. (c)–(f) The reconstructed Wigner representation of the fully symmetric portion of the state (not of the main eigenstate) when a $2\pi /3$ rotation is applied about each of the four points of the tetrahedron.
Fig. 5.
Fig. 5. Numerically calculated projection on the original tetrahedron state after rotations around the $x,y$, and $z$ axes on Bloch sphere for the following states: ideal tetrahedron state (dashed orange); experimental density matrix obtained from the tomography (solid black); and the main eigenstate of the experimental density matrix (solid red). The $x$-axis of the figure is the rotation amplitude in radians and the $y$-axis is the overlap between the rotated state and the ideal tetrahedron state.
Fig. 6.
Fig. 6. Scalar quantum Cramér–Rao bound $\mathcal {C}$ as a function of total number of photons. Solid lines represent the different strategies enumerated in Section 2.4: spin-coherent states (purple triangles); $N00N$ states (red squares for simultaneous; orange pentagons for sequential); and Platonic states (yellow circles). Sequential strategies use three batches of states aligned in the $x, y$, and $z$ axes, whereas simultaneous strategies use only a single batch of identical states. The s-QCRB decreases with shot-noise scaling for sequential coherent states and simultaneous $N00N$ states, and decreases with Heisenberg scaling for sequential $N00N$ states and Platonic states. The s-QCRB of the experimentally reconstructed density matrix of the tetrahedron state is represented by a cross symbol. The black star represents the s-QCRB of its main eigenstate.

Equations (22)

Equations on this page are rendered with MathJax. Learn more.

Σ l , m = E [ ( θ l θ ~ l ) ( θ m θ ~ m ) ]
Σ I ρ , θ 1 ,
[ I ρ , θ ] l , m = Tr [ ρ ^ θ { L ^ θ l , L ^ θ m } ] ,
θ l ρ ^ θ = 1 2 ( ρ ^ θ L ^ θ l + L ^ θ l ρ ^ θ ) .
Tr [ C Σ ] Tr ( C I ρ , θ 1 ) =: C ρ , θ ,
C = 1 4 Tr [ C o v ψ 1 ( J ^ ) ] ,
[ C o v ψ ( J ^ ) ] l . m = 1 2 J ^ l J ^ m + J ^ m J ^ l J ^ l J ^ m
R ^ ( θ ) = e i θ J ^ ,
( spin- 1 / 2 ) N = j = [ 0 , 1 / 2 ] N / 2 d j ( spin- j ) ,
| j = 3 / 2 , m = 1 / 2 = | 2 H , 1 V = 1 3 ( | H H V + | H V H + | V H H ) .
J ^ = j = [ 0 , 1 / 2 ] N / 2 J ^ ( j ) J ^ ( j ) d j t i m e s
m = 0 N ψ m | m H , ( N m ) V k = 1 N a ^ θ k , ϕ k | 0 H , 0 V , a ^ θ , ϕ = cos θ 2 a ^ H + e i ϕ sin θ 2 a ^ V .
| ψ c o h = | N H , 0 V ,
C o v ψ c o h ( J ^ ) = ( N / 4 0 0 0 N / 4 0 0 0 0 ) ,
i C o v ψ c o h ( i ) ( J ^ ) = N 6 × 1 , C = 9 2 N ,
| ψ N 00 N = 1 2 ( | N H , 0 V + | 0 H , N V ) .
C o v ψ N 00 N ( J ^ ) = ( N / 4 0 0 0 N / 4 0 0 0 N 2 / 4 ) , C = 2 N + 1 N 2 ,
i C o v ψ N 00 N ( i ) ( J ^ ) = 1 36 ( N ( N + 6 ) ) × 1 , C = 27 N ( N + 6 ) .
J ^ = 0 J ^ l J ^ m = N 6 ( N 2 + 1 ) δ l , m .
C o v ψ p l a t ( J ^ ) = 1 12 ( N ( N + 2 ) ) × 1 , C = 9 N ( N + 2 ) .
| ψ t e t r a = 1 3 | 4 H , 0 V + 2 3 | 1 H , 3 V .
| ψ t e t r a = 1 3 ( ( a ^ H 2 ) 3 + a ^ V 3 ) a ^ H | 0 .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.