This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A dimensionality and purity measure for high-dimensional entangled states

Isaac Nape School of Physics, University of the Witwatersrand, Private Bag 3, Wits 2050, South Africa    Valeria Rodríguez-Fajardo School of Physics, University of the Witwatersrand, Private Bag 3, Wits 2050, South Africa    Feng Zhu School of Engineering and Physical Sciences, Heriot-Watt University, Edinburgh, EH14 4AS, UK    Hsiao-Chih Huang Department of Physics, National Taiwan University, Taipei 106, Taiwan    Jonathan Leach School of Engineering and Physical Sciences, Heriot-Watt University, Edinburgh, EH14 4AS, UK    Andrew Forbes School of Physics, University of the Witwatersrand, Private Bag 3, Wits 2050, South Africa
Abstract

High-dimensional entangled states are promising candidates for increasing the security and encoding capacity of quantum systems. While it is possible to witness and set bounds for the entanglement, precisely quantifying the dimensionality and purity in a fast and accurate manner remains an open challenge. Here, we report an approach that simultaneously returns the dimensionality and purity of high-dimensional entangled states by simple projective measurements. We show that the outcome of a conditional measurement returns a visibility that scales monotonically with entanglement dimensionality and purity, allowing for quantitative measurements for general photonic quantum systems. We illustrate our method using transverse spatial modes of photons that carry orbital angular momentum and verify high-dimensional entanglement over a wide range of state purities. Our approach advances the high-dimensional tool box for characterising quantum states by providing a simple and direct dimensionality and purity measure, even for mixed entangled states.

High-dimensional entangled states are widely used throughout quantum science to increase secure information bandwidth and security bounds for quantum communication cozzolino2019high . Through the precise control of high dimensional photonic states erhard2017quantum ; babazadeh2017high , i.e., time-energy, transverse momentum, spatial degrees of freedom or all of them simultaneously deng2017quantum , the unparalleled benefits of high dimensional state encoding are taking center stage. Recent developments in this direction have displayed the feasibility of quantum information processing that is robustness against optimal quantum cloning machines gisin1997optimal ; bouchard2017high , environmental noise ecker2019overcoming and improved information rates dixon2008gigahertz ; barreiro2008beating , demonstrating a significant advantage in comparison to traditional qubit encoding.

Despite the advantages of high-dimensional quantum states, certifying and quantifying the dimensionality of such systems still remains challenging, particularly in the presence of noise. The intuitive approach of simply measuring the width of the modal spectrum is a necessary but not sufficient condition to determine dimensionality as it fails to account for non-local correlations. Consequently, many techniques have been developed to witness, bound and attempt to quantify high-dimensional quantum states. These include approximating the density matrix via quantum state tomography (QST) with multiple qubit state projections Agnew2011 , using mutually unbiased bases giovannini2013characterization ; bavaresco2018measurements to probe the states, and testing non-local bi-photon correlations by generalised Bell tests in higher dimensions vaziri2002experimental ; groblacher2006experimental ; dada2011experimental ; oemrawsingh2005experimental ; oemrawsingh2006high ; gotte2007quantum ; huang2018various . However, the spectrum measurements do not confirm entanglement, the QST approach scales unfavourably with dimension, only bounds or witnesses are possible with the mutually unbiased bases method and the dimension to be probed must be known a priori (e.g., valid for prime or prime power dimensions), and finally, the high-dimensional Bell tests can fail the fair sampling condition dada2011bell ; romero2013tailored . A further limitation in the present state-of-the-art is that certain dimensionality measurements consider only pure states pors2008shannon ; giovannini2013characterization , yet noise mechanisms always introduce some degree of mixture to the system. This has a detrimental effect on the accuracy of measured dimensions due to the reduced purity zhu2019high . Consequently, no approach allows both the purity and dimensionality of arbitrary high dimensional mixed states to be quantitatively deduced in a simple and accurate manner. Yet, knowing the purity and dimension of the state is crucial for fundamental tests of quantum mechanics as well as for quantum information processing protocols, setting the required violation of inequalities in the former, and the information capacity of the state, the allowed error bounds in secure communication systems, and the requirement for entanglement distillation in the latter.

In this work we will provide a solution to this pressing problem and illustrate it using the topical example of transverse spatial modes of photons carrying orbital angular momentum (OAM), which have been instrumental in realising quantum entanglement beyond qubits fickler2012quantum ; krenn2015twisted ; erhard2017quantum ; erhard2018twisted ; fickler2014interface ; forbes2019quantum ; molina2007twisted . They have emerged as an ideal resource in quantum information processing and communication, including superdense coding wang2005quantum ; barreiro2008beating , multi-photon entanglement Malik2016 , quantum teleportation goyal2014qudit and entanglement swapping zhang2017simultaneous ; bornman2019ghost , ghost imaging jack2009holographic ; chen2014quantum ; bornman2019ghostaa and secure communication in free-space mafu2013higher ; steinlechner2017distribution and optical fibre liu2020multidimensional ; cao2020distribution ; cozzolino2019air ; cozzolino2019orbital , fuelled by their inherent advantages such as robustness against noise ecker2019overcoming and higher information capacity per photon leach2012secure . Further, the toolkit to engineer high-dimensional states is readily available, e.g., by linear zhang2016engineering and non-linear processes torres2003quantum ; miatto2011full ; Romero2012 ; Terriza ; zhang2014simulating , with up to 100×\times100 dimensions already demonstrated krenn2014generation .

With OAM as our example, we present a scheme to simultaneously quantify the dimensionality and purity of a two-particle entangled state. By measuring coincidence fringes from projective measurements using analysers acting on the entangled photons, we are able to accurately measure the dimensionality and purity of our entangled state from the visibility, which is only reproducible by entangled photons. We first outline the concept and theory and then demonstrate it experimentally on states with arbitrary purity. Our quantitative technique is simple, fast and robust, making it ideal for practical implementations (even with undesired noise) of quantum protocols with general high-dimensional photonic quantum entangled states.

I Concept

Refer to caption
Figure 1: (a) Conceptual visualisation of different analysers sampling various portions of a high dimensional discrete Hilbert space. Mode analysers construction for (b) n=3n=3 and (c) n=7n=7 superpositions of fractional OAM states. (d) Schematic of the experimental setup used to measure the dimensionality and purity of a quantum system. (NLC: Nonlinear crystal, f1,2,3,4: lens, BS: 50:50 Beam-splitter, SLM: Spatial light modulator, D: detector, APD: avalanche photo diode coupled to a single mode fiber (SMF), C.C.: coincidence counter.)

The task here is to measure the effective dimensions and purity of an entangled system. In general, the purity of the quantum system, and therefore the entanglement between photon pairs, is reduced due to noise introduced by the environment or detection system, usually background white noise produced by ambient light or high dark counts in single photon detectors or unwanted multiphoton events zhu2019high ; ecker2019overcoming . Such noisy quantum systems can be modelled by an isotropic state following

ρ=p|ΨΨ|+(1pd2)𝕀d2,\rho=p\ket{\Psi}\bra{\Psi}+\left(\tfrac{1-p}{d^{2}}\right)\mathbbm{I}_{d^{2}}, (1)

which considers contributions of both pure, |Ψ\ket{\Psi}, and mixed, 𝕀d2\mathbbm{I}_{d}^{2} (d2-dimensional identity operator), parts. Here pp is a parameter that determines the purity of the isotropic state, and varies from a maximally mixed (separable) state for p=0p=0 to a completely pure (entangled) state for p=1p=1. The dimension of the pure part of the state ρ\rho can be quantified by the Schmidt number, K=1/j|λj|2K=1/\sum_{j}|\lambda_{j}|^{2}, where λj\lambda_{j} are the Schmidt coefficients obtained from the decomposition of |Ψ\ket{\Psi} onto the basis states |ζj2\ket{\zeta_{j}}\in\mathcal{H}^{2}_{\infty}, therefore spanning the two photon infinite dimensional Hilbert space. We aim to find the dimensionality of the pure part of the system while also determining the purity assuming that the noise encroaching on the system is due to white noise from the environment.

The working principle of our technique is visualised in Fig. 1(a), where a set of analysers probe distinct parts of a discrete Hilbert space. We can think of each analyser as a probe that scans a sparse set of modes, reminiscent of a conditional measurement that indicates whether there is entanglement within the subspace or not. By combining the information gathered from a number of such analysers, we infer how many dimensions the state occupies. We will demonstrate this procedure both theoretically and experimentally.

To illustrate this concept, we consider a pair of entangled photons generated from spontaneous parametric down conversion (SPDC). Such photons can be entangled in their polarisation, energy-time, momentum or spatial mode forbes2019quantum such as orbital angular momentum (OAM). Due to the great potential of the latter, particularly for quantum communications, we illustrate and demonstrate our method for OAM entangled states. In this case, the basis states, |ζ||\ket{\zeta_{\ell}}\rightarrow\ket{\ell}\ket{-\ell}, are associated with an azimuthal phase profile exp(iϕ)\text{exp}(i\ell\phi), with \ell\in\mathbb{Z} being the topological charge and \ell\hbar OAM per photon. An OAM entangled pure state can be expressed as

|Ψ==LLλ|A|B,\ket{\Psi}=\sum_{\ell=-L}^{L}\lambda_{\ell}\ket{\ell}_{A}\ket{-\ell}_{B}, (2)

where LL is a positive integer denoting the largest mode index with a nonzero probability (simple to deduce by observing the counts), and |λ|2|\lambda_{\ell}|^{2} is the probability of generating photons in the states |±\ket{\pm\ell} for photons AA and BB, respectively. While in general the state |Ψ\ket{\Psi} can be represented using an unbounded number of eigenmodes, i.e., LL\rightarrow\infty, we truncate |Ψ\ket{\Psi} to the dimensions specified by KK. That is, we choose L=(K1)/2L=(K-1)/2. This validly approximates the state in terms of its effective dimensions, enabling for its representation on a finite Hilbert space.

To gain access to various parts of the Hilbert space, we make use of high dimensional mode projectors that map onto the states

|M,αn=𝒩j=02Lcwj,Mn(α)|j,\ket{M,\alpha}_{n}=\mathcal{N}\sum^{2L}_{j=0}c_{w_{j},M}^{n}(\alpha)\ket{j}, (3)

where 𝒩\mathcal{N} is a normalisation factor, |j\ket{j} are the basis states on the d=2L+1d=2L+1 dimensional space. The coefficients, cwj,Mn(α)c^{n}_{w_{j},M}(\alpha), control the amplitudes and phases of the modes in the superposition (see Methods section). For OAM basis states, the coefficients can be represented accordingly by replacing the index wjw_{j} with the topological charge =jL\ell=j-L. Examples of the phase profiles for two such analysers are shown in Fig. 1 (d) and (c) for n=3n=3 and n=7n=7, respectively, with full details on their construction in the Methods section and Supplementary Information. While nn and MM can be chosen arbitrarily, we find it optimal to set nn as an odd positive integer and M=n/2M=n/2.

Refer to caption
Figure 2: Experimental (points) and theoretical (solid lines) coincidence count rates resulting from projections of photons AA and BB onto the states |M,θn\ket{M,\theta}_{n} and |M,0n\ket{-M,0}_{n}, respectively, as a function of the relative orientation angle θ\theta for (a) n=5n=5 and (b) n=9n=9. Theoretical visibility as a function of the dimensionality (KK) and purity (pp) for (c) n=5n=5 and (d) n=9n=9, exemplifying it increases monotonically with both parameters. The (red) planes intersecting the curves are the experimental visibilities, with the possible solution space for each shown as a red trajectory. The resulting trajectories for n=1,2,,11n=1,2,...,11 are shown in (e), where the thickness of each is due to the uncertainty in the visibility outcome. The dimension and purity of the system are found where they coincide, shown as a dashed red circle. The later corresponds to the optimal (p,K)(p,K) that minimise the function χ2(p,K)\chi^{2}(p,K), or, equivalently, maximizes (e) 1/χ2\sqrt{1/\chi^{2}}, where the minimum of χ2\chi^{2} is now shown as a peak corresponding to (p,K)=(0.45±0.03,22.84±0.62)(p,K)=(0.45\pm 0.03,22.84\pm 0.62).

Next, we project each photon in the isotropic state in Eq. (1), onto identical but conjugated analysers orientated by an angle θ\theta relative to one another. A typical experimental setup for implementing this is sketched in Fig. 1 (d). Entangled photon pairs are generated from SPDC and subsequently projected onto the states |M,θ\ket{M,\theta} and |M,0\ket{-M,0} by means of holograms having the transmission functions Un(ϕ;θ)U_{n}(\phi;\theta) and Un(ϕ;0)U_{n}^{*}(\phi;0), respectively. In the OAM degree of freedom, the holograms correspond to fractional OAM modes gotte2007quantum , known to have a non-integer azimuthal phase gradient. The modulated photons are then coupled into single mode fibers and measured in coincidences. The outcome probability of such a measurement, i.e., 0,M|nθ,M|nρ|M,θn|M,0n\bra{0,-M}_{n}\bra{\theta,M}_{n}\rho\ket{M,\theta}_{n}\ket{-M,0}_{n}, is

Pn(θ;p,K)=pPn(θ,K)+1pK2In(0,K),P_{n}(\theta;p,K)=pP_{n}(\theta,K)+\frac{1-p}{K^{2}}I_{n}(0,K), (4)

where In(0,K)/K2I_{n}(0,K)/K^{2} is the probability resulting from the overlap of the analysers with the maximally mixed state and Pn(θ,K)=|=LLλc,Mn(0)c,Mn(θ)|2P_{n}(\theta,K)=\left|\sum_{\ell=-L}^{L}\lambda_{\ell}c^{n}_{\ell,M}(0)c^{n}_{-\ell,M}(\theta)\right|^{2} is the overlap probability with the pure state, with M=n/2M=n/2 the fractional charge and λ\lambda_{\ell} the initial bi-photon OAM spectrum. For a pure state, the probability curves have a parabolic shape following Pn(θ)=(π(2t1)nθ)2/π2P_{n}(\theta)=(\pi(2t-1)-n\theta)^{2}/\pi^{2}, where t=1,2,nt=1,2,...\ n. In Fig. 2 (a) and (b), we show as solid lines the theoretical probabilities (calculated using Eq. (4)) of such a measurement as function of θ\theta. We choose odd values of nn and M=n/2M=n/2 to ensure a high visibility, which increases monotonically with KK and pp for each analyser (see Supplementary Information). In general both the shape and visibility of the fringes yield information about the state. To make the approach accurate and precise, we measure NN visibilities, VnV_{n} for n=1,2,, 2N1n=1,2,...,\ 2N-1, and infer the state properties by the intersection of their solution spaces.

II Results

The set-up used to demonstrate our scheme is shown conceptually in Fig. 1 (d) with the corresponding detailed description in the Methods section. We measure the coincidences between the signal and idler photons for analyser projections on both arms as a function of the relative rotation angle of the holograms. To achieve this, we encoded the fractional OAM mode analyser on the SLM in the signal arm fixed at an angle 0, while the conjugate mode was encoded in the idler arm and rotated at angles θ[0,2π]\theta\in[0,2\pi].

Refer to caption
Figure 3: Dimensionality and purity measurements assuming the SPDC mode spectrum for low (top) and high noise (bottom) levels. The points are the measured visibilities while the solid lines correspond to the fitted values. Measured spiral spectrum for the (b) low and (c) high noise levels.

To illustrate the operation of our technique we measured the coincidence-rates for six (N=6N=6) analysers with n=1,3,5,7,9n=1,3,5,7,9 and 11, and M=n/2M=n/2, with example outcomes for n=5n=5 and 9 shown as filled circles in Fig. 2(a) and (b), respectively. Here, no accidental count subtraction was performed on the measurements. Importantly, the periodicity in the detected probabilities confirms the azimuthal nn-fold symmetry predicted by our theory (solid curve). Because the visibility is a monotonically increasing function of dimension and purity, a measured visibility returns a range of possible (p,K)(p,K) values, a “trajectory” or curve in the (p,K)(p,K) space. This is illustrated in Fig. 2(c) and (d), where the measured visibility (red horizontal plane) intercepts the visibility function along a curve (red curve) that restricts the possible solutions, KK and pp, to those consistent with the measurement outcome. The set of such curves from measuring many visibilities (each with its own analyser/projection) then restricts the final solution to a narrow region in (p,K)(p,K), whose uncertainty (width) is determined primarily from the uncertainly in the visibility measurement. An example is shown in Fig. 2(e), where each solution trajectory is projected onto the (p,K)(p,K) plane. Final values and uncertainty of (p,K)(p,K) can be determined by an appropriate routine to find the interception of all such trajectories by a minimisation procedure, as shown in Fig. 2(f). Using this approach we infer the purity and dimensionality of the system to be (p,K)=(0.45±0.03,22.84±0.62)(p,K)=(0.45\pm 0.03,22.84\pm 0.62).

In Fig. 3 (a) we show the six measured visibilities as square data points together with the calculated visibility (solid red line) based on the inferred (p,K)(p,K), which clearly match very well. This is confirmation of the minimisation procedure for finding the intercept. In order to assess the procedure under high noise levels, we introduced background noise using a white light source and repeated the measurements, shown as the circle data points and associated the blue dashed line in Fig. 3. The average quantum contrast (see Supplementary Information), measured from the spiral spectrum in Fig. 3 (c) and (d), dropped from Q=19.19Q=19.19 to Q=3.76Q=3.76, resulting in a reduced purity and dimensionality of (p,K)=(0.13±0.01,17.73±0.71)(p,K)=(0.13\pm 0.01,17.73\pm 0.71).

As a form of validation of these results, we estimate values from other techniques, with the comparison given in Table 2. If the dimension and noise of a system is known, then it is possible to calculate the purity following p^=(Q1)/(Q1+d)\hat{p}=(Q-1)/(Q-1+d) where QQ is the quantum contrast and dd the dimension zhu2019high . Likewise, if the state is assumed to be pure and not mixed and background subtraction is done to remove noise, then the spiral spectrum can be used to get an upper bound on the dimension. For the two noise cases in Table 2, low and high, we find purity estimates of p^0.44\hat{p}\approx 0.44 (assuming d=23d=23) and p^0.13\hat{p}\approx 0.13 (assuming d=19d=19), respectively, while estimates of the dimensionality return K^22\hat{K}\approx 22 and K^18\hat{K}\approx 18, respectively. These values are in excellent agreement with our results, which did not require any such assumptions, nor any noise adjustments. The (perhaps) surprisingly low purity in our results is due to the fact that no background subtraction due to accidental counts was performed.

Noise level pp KK QQ p^\hat{p} K^\hat{K}
low 0.45 ±\pm 0.03 22.84 ±\pm 0.62 19.19 0.44 22
high 0.13 ±\pm 0.01 17.73 ±\pm 0.71 3.76 0.13 18
Table 1: Measured purity (pp) and dimensionality (KK), under low and high noise levels, compared to estimates from other methods. Here QQ is the average quantum contrast.

III Discussion and Conclusion

A measure of dimensionality and purity, particularly in the presence of (inevitable) deleterious noise that degrades the purity, is crucial for many quantum protocols and studies. For example, there is a minimum purity needed to witness entanglement in a given dimension peres1996separability ; horodecki1997separability ; horodecki2009quantum , setting the transition from separable to entangled states. Likewise, knowing the purity is important in entanglement distillation processes since it informs whether the noise can be removed for a given dimension vollbrecht2003efficient ; horodecki1999reduction , while in entanglement based quantum communication there is a minimum purity Collins2002 associated with security ekert1991quantum . In turn, the dimensionality sets the information capacity of the state for quantum information processing and the error tolerance in quantum communication protocols, while high-dimensional states are important for fundamental tests of quantum mechanics where qubits will not suffice klyachko2008simple ; lapkiewicz2011experimental . While a full quantum state tomography would also provide the necessary information, such a measurement would take several days for the states studied here, as compared to less than an hour with our technique.

Unlike a conventional Schmidt decomposition, we do not assume the state is pure, and the dimension extracted from our technique is conditioned on the presence of entanglement: a maximally mixed and maximally entangled system cannot yield the same result. While our approach would benefit from knowledge of the modal spectrum, which can be measured very quickly for OAM Pires2010 ; kulkarni2017single , the outcome on purity and dimensionality are only modestly affected by typical spectrum shapes (see Supplementary Information), e.g., in our examples the uncertainty in dimensionality is 5%\approx 5\% with knowledge of the spectrum, increasing to 10%\approx 10\% without. In this work we used OAM as our demonstration example, but reiterate that the projections are applicable to general states in the spatial mode basis.

In summary, we have developed a simple yet powerful technique to measure the dimensionality and purity of high dimensional entangled photonic quantum systems. Our approach is robust, fast, and provides quantitative values rather than bounds or witnesses, and works on both pure and mixed states. Our scheme exploits visibility in fringes after joint projections, making it fast and easy to implement, returning the key parameters of the system in a fraction of the time that a full quantum state tomography would take. We believe that this tool will be useful to the active research in high-dimensional spatial mode entanglement and foster its wide-spread deployment in quantum based protocols.

Acknowledgements

The authors express their gratitude to Bienvenue Ndagano for his inputs. I.N. would like to acknowledge the Department of Science and Technology (South Africa) for funding.

Author contributions

The experiment was performed by I.N. and V.R.F., the theory was developed by I.N., F.Z,. H.H., and J.L., the data analysis was performed by I.N., V.R.F. and A.F. and the experiment was conceived by I.N., V.R.F., and A.F. All authors contributed to the writing of the manuscript. A.F. supervised the project.

Competing Interests statement

The authors declare not competing interests.

Methods

High dimensionsal state projections. Our analysers project onto the high dimensional Hilbert space, d\mathcal{H}_{d}, mapping onto the states in Eq. (11), i.e |M,αn\ket{M,\alpha}_{n}, repeated here as

|M,αn=𝒩j=0d1cwj,Mn(α)|j,\ket{M,\alpha}_{n}=\mathcal{N}\sum^{d-1}_{j=0}c_{w_{j},M}^{n}(\alpha)\ket{j}, (5)

composed from coherent superpositions of basis states |j{|j,j=0,1..d1}\ket{j}\in\{\ket{j},j=0,1..d-1\} with tune-able phases and amplitudes

cwj,Mn(α)=𝒩eiπwj(n1)/nAwjncwj,M(α),c_{w_{j},M}^{n}(\alpha)=\mathcal{N}e^{-i\pi w_{j}(n-1)/n}A_{w_{j}}^{n}c_{w_{j},M}(\alpha), (6)

and where wj=j(d1)/2w_{j}=j-(d-1)/2 and the factors

cwj,M(α)=ieiwjαπ(Mwj).c_{w_{j},M}(\alpha)=-\frac{ie^{-iw_{j}\alpha}}{\pi(M-w_{j})}. (7)

and

Awjn=\displaystyle A_{w_{j}}^{n}= {1,mod{wj,n}=00,otherwise.\displaystyle\left\{\begin{array}[]{cc}1,&\text{mod}\left\{w_{j},n\right\}=0\\[2.84526pt] 0,&\text{otherwise}\\ \end{array}\right.. (10)

Here, cwk,M(α)c_{w_{k},M}(\alpha) controls the relative phases and amplitudes of the eigenmodes and AwjnA_{w_{j}}^{n} modulates the coefficients’ amplitudes while α[0,2π/n]\alpha\in[0,2\pi/n]. The spectrum given by Eq. 6 can be tuned by carefully selecting nn, therefore enabling precise control of the subspaces that will be probed.

In the OAM basis, i.e |d\ket{\ell}\in\mathcal{H}_{d}, the index wjw_{j} can be replaced with the index 𝒵\ell\in\mathcal{Z}. The mode projectors can be constructed from spiral phase profiles having the transmission function

Un(ϕ,α)=k=0n1exp(iΦM(ϕ;βkα)),U_{n}(\phi,\alpha)=\mathcal{M}\sum^{n-1}_{k=0}\exp\left(i\Phi_{M}\left(\phi;\beta{{}_{k}}\oplus\alpha\right)\right), (11)

that is constructed from superpositions of fractional OAM modes (See Supplementary Information) gotte2007quantum ; oemrawsingh2005experimental ,

exp(iΦM(ϕ;α))={eiM(2π+ϕα)0ϕ<αeiM(ϕα)αϕ<2π,\exp\left(i\Phi_{M}(\phi;\alpha)\right)=\begin{cases}e^{iM\left(2\pi+\phi-\alpha\right)}&\quad 0\leq\phi<\alpha\\ e^{iM\left(\phi-\alpha\right)}&\quad\alpha\leq\phi<2\pi\end{cases}, (12)

having the same charge, MM, but rotated by an angle βkα=mod{βk+α,2π}\beta_{k}\oplus\alpha=\text{mod}\left\{\beta_{k}+\alpha,2\pi\right\} for βk=2πnk\beta_{k}=\frac{2\pi}{n}k, as illustrated in Figs. 1 (b) and (c) for n=3n=3 and n=7n=7, respectively. Here, ϕ\phi is the azimuthal coordinate and \mathcal{M} a normalization constant.

Experimental setup. The experimental setup for the generation and measurement of entangled photons is illustrated schematically in Fig. 1 (d). A potassium-titanium-phosphate (PPKTP) type I nonlinear crystal (NLC) was pumped with a 405 nm wavelength diode laser. The crystal temperature was set to obtain co-linear signal and idler entangled SPDC photons centred at a wavelength 810 nm. The photon pairs were then separated in path using a 50:50 beam splitter (BS). Each entangled photon was imaged onto a spatial light modulator (SLM) using a 4f4f telescope (f1f_{1} and f2f_{2} having focal lengths of 100 mm and 500 mm, respectively), then subsequently coupled into single mode fibers with a second 4f4f telescope (lenses f3f_{3} and f4f_{4} having focal lengths of 750 mm and 2 mm, respectively) and finally detected with avalanche photo-detectors (APDs). Signals from each arm were measured in coincidences within a 25 ns coincidence window. The entangled photons were filtered with 10 nm bandpass filters centered at a wavelength of 810 nm. For our experimental demonstration, we restrict our measurements to a specific optical setup and we varied the purity of the entangled state by introducing background noise in the form of white light. To obtain a high purity state (p=0.45 in K=22 dimensions), we had to reduce the laser power using an ND filter such as to reduce multi-photon emission events, known to have an impact on the purity of the SPDC photons zhu2019high .

Holographic fractional OAM mode projections. The conventional method for generating fractional OAM modes involves imprinting an azimuthally dependent phase retardation onto an incoming field with spiral phase plates oemrawsingh2004half or digital holograms leach2004observation . Here, we employed dynamic phase control of liquid crystal spatial light modulators (SLMs) rosales2017shape to generate rotated fractional OAM phase masks to modulate the transverse plane of photons. We achieved this by encoding grey-scale holograms with the phase profile of a transmission function corresponding to the desired projection mode. Accordingly, we prepared analysers with superpositions of multiple rotated fractional spiral phases having a transmission function given by eq. 11, with the azimuthal coordinate ϕ=tan1(yx)\phi=\tan^{-1}(\frac{y}{x}) and (x,y)(x,y) the coordinates of each pixel. We generated holograms with the desired phase and subsequently added them to a blazed grating for a final hologram mapped as

ΦSLM(x,y)=mod{arg(Un(α,ϕ))\displaystyle\Phi_{\text{SLM}}(x,y)=\text{mod}\left\{\text{arg}\left(U_{n}(\alpha,\phi)\right)\right.
+Gxx+Gyy,2π},\displaystyle\left.+G_{x}x+G_{y}y,2\pi\right\}, (13)

where Gx,yG_{x,y} are the grating wavenumbers in the xx and yy directions, respectively. An example of the phase hologram generated of a superposition of three fractional OAM spiral phases (n=3n=3 with M=1.5M=1.5), and its corresponding rotated complex-conjugate are shown as inserts in Fig. 1 (d).

Optimal purity and dimensionality calculation. Using the fact that the visibility obtained for each analyser is affected by the dimensionality and purity of the input state, we describe the procedure for determining their values for a given entangled quantum system, assuming it can be modelled by the isotropic state in Eq. (1). We measure the probability curves for N analysers each with n=1,3,,2N1n=1,3,...,2N-1, and compute their corresponding visibilities Vn:=Vn(p,K)V_{n}:=V_{n}(p,K). This results in a set of NN nonlinear equations that depend on the parameters pp and KK. We then determine the optimal (p,K)(p,K) pair that best fit the function Vn(p,K)V_{n}(p,K) to all NN measured visibilies by employing the method of least squares (LSF), which aims to minimise the objective function

χ2(p,K)=i=1N|V2i1The.(p,K)V2i1Exp.|2.\chi^{2}(p,K)=\sum_{i=1}^{N}|V_{2i-1}^{\text{The.}}(p,K)-V^{\text{Exp.}}_{2i-1}|^{2}. (14)

where are the terms in the summation are the residuals (absolute errors) for each n=2i1n=2i-1 visibility measurement with respect to the theory.

References

  • (1) D. Cozzolino, B. Da Lio, D. Bacco, and L. K. Oxenløwe, “High-dimensional quantum communication: Benefits, progress, and future challenges,” Advanced Quantum Technologies, vol. 2, no. 12, p. 1900038, 2019.
  • (2) M. Erhard, M. Malik, and A. Zeilinger, “A quantum router for high-dimensional entanglement,” Quantum Science and Technology, vol. 2, no. 1, p. 014001, 2017.
  • (3) A. Babazadeh, M. Erhard, F. Wang, M. Malik, R. Nouroozi, M. Krenn, and A. Zeilinger, “High-dimensional single-photon quantum gates: concepts and experiments,” Physical review letters, vol. 119, no. 18, p. 180510, 2017.
  • (4) F.-G. Deng, B.-C. Ren, and X.-H. Li, “Quantum hyperentanglement and its applications in quantum information processing,” Science bulletin, vol. 62, no. 1, pp. 46–68, 2017.
  • (5) N. Gisin and S. Massar, “Optimal quantum cloning machines,” Physical review letters, vol. 79, no. 11, p. 2153, 1997.
  • (6) F. Bouchard, R. Fickler, R. W. Boyd, and E. Karimi, “High-dimensional quantum cloning and applications to quantum hacking,” Science advances, vol. 3, no. 2, p. e1601915, 2017.
  • (7) S. Ecker, F. Bouchard, L. Bulla, F. Brandt, O. Kohout, F. Steinlechner, R. Fickler, M. Malik, Y. Guryanova, R. Ursin, et al., “Overcoming noise in entanglement distribution,” Physical Review X, vol. 9, no. 4, p. 041042, 2019.
  • (8) A. Dixon, Z. Yuan, J. Dynes, A. Sharpe, and A. Shields, “Gigahertz decoy quantum key distribution with 1 mbit/s secure key rate,” Optics express, vol. 16, no. 23, pp. 18790–18797, 2008.
  • (9) J. T. Barreiro, T.-C. Wei, and P. G. Kwiat, “Beating the channel capacity limit for linear photonic superdense coding,” Nature physics, vol. 4, no. 4, p. 282, 2008.
  • (10) M. Agnew, J. Leach, M. McLaren, F. S. Roux, and R. W. Boyd, “Tomography of the quantum state of photons entangled in high dimensions,” Physical Review A, vol. 84, p. 062101, 2011.
  • (11) D. Giovannini, J. Romero, J. Leach, A. Dudley, A. Forbes, and M. J. Padgett, “Characterization of high-dimensional entangled systems via mutually unbiased measurements,” Physical Review Letters, vol. 110, no. 14, p. 143601, 2013.
  • (12) J. Bavaresco, N. H. Valencia, C. Klöckl, M. Pivoluska, P. Erker, N. Friis, M. Malik, and M. Huber, “Measurements in two bases are sufficient for certifying high-dimensional entanglement,” Nature Physics, pp. 1745–2481, 2018.
  • (13) A. Vaziri, G. Weihs, and A. Zeilinger, “Experimental two-photon, three-dimensional entanglement for quantum communication,” Physical Review Letters, vol. 89, no. 24, p. 240401, 2002.
  • (14) S. Gröblacher, T. Jennewein, A. Vaziri, G. Weihs, and A. Zeilinger, “Experimental quantum cryptography with qutrits,” New Journal of Physics, vol. 8, no. 5, p. 75, 2006.
  • (15) A. C. Dada, J. Leach, G. S. Buller, M. J. Padgett, and E. Andersson, “Experimental high-dimensional two-photon entanglement and violations of generalized bell inequalities,” Nature Physics, vol. 7, no. 9, p. 677, 2011.
  • (16) S. Oemrawsingh, X. Ma, D. Voigt, A. Aiello, E. t. Eliel, J. Woerdman, et al., “Experimental demonstration of fractional orbital angular momentum entanglement of two photons,” Physical review letters, vol. 95, no. 24, p. 240501, 2005.
  • (17) S. Oemrawsingh, J. de Jong, X. Ma, A. Aiello, E. Eliel, J. Woerdman, et al., “High-dimensional mode analyzers for spatial quantum entanglement,” Physical Review A, vol. 73, no. 3, p. 032339, 2006.
  • (18) J. B. Götte, S. Franke-Arnold, R. Zambrini, and S. M. Barnett, “Quantum formulation of fractional orbital angular momentum,” Journal of Modern Optics, vol. 54, no. 12, pp. 1723–1738, 2007.
  • (19) H.-C. Huang, “Various angle periods of parabolic coincidence fringes in violation of the bell inequality with high-dimensional two-photon entanglement,” Phys. Rev. A, vol. 98, p. 053856, Nov 2018.
  • (20) A. C. Dada and E. Andersson, “On bell inequality violations with high-dimensional systems,” International Journal of Quantum Information, vol. 9, no. 07n08, pp. 1807–1823, 2011.
  • (21) J. Romero, D. Giovannini, D. Tasca, S. Barnett, and M. Padgett, “Tailored two-photon correlation and fair-sampling: a cautionary tale,” New Journal of Physics, vol. 15, no. 8, p. 083047, 2013.
  • (22) J. Pors, S. Oemrawsingh, A. Aiello, M. Van Exter, E. Eliel, J. Woerdman, et al., “Shannon dimensionality of quantum channels and its application to photon entanglement,” Physical review letters, vol. 101, no. 12, p. 120502, 2008.
  • (23) F. Zhu, M. Tyler, N. H. Valencia, M. Malik, and J. Leach, “Are high-dimensional entangled states robust to noise?,” arXiv preprint arXiv:1908.08943, 2019.
  • (24) R. Fickler, R. Lapkiewicz, W. N. Plick, M. Krenn, C. Schaeff, S. Ramelow, and A. Zeilinger, “Quantum entanglement of high angular momenta,” Science, vol. 338, no. 6107, pp. 640–643, 2012.
  • (25) M. Krenn, J. Handsteiner, M. Fink, R. Fickler, and A. Zeilinger, “Twisted photon entanglement through turbulent air across vienna,” Proceedings of the National Academy of Sciences, vol. 112, no. 46, pp. 14197–14201, 2015.
  • (26) M. Erhard, R. Fickler, M. Krenn, and A. Zeilinger, “Twisted photons: new quantum perspectives in high dimensions,” Light: Science & Applications, vol. 7, no. 3, pp. 17146–17146, 2018.
  • (27) R. Fickler, R. Lapkiewicz, M. Huber, M. P. Lavery, M. J. Padgett, and A. Zeilinger, “Interface between path and orbital angular momentum entanglement for high-dimensional photonic quantum information,” Nature communications, vol. 5, no. 1, pp. 1–6, 2014.
  • (28) A. Forbes and I. Nape, “Quantum mechanics with patterns of light: Progress in high dimensional and multidimensional entanglement with structured light,” AVS Quantum Science, vol. 1, no. 1, p. 011701, 2019.
  • (29) G. Molina-Terriza, J. P. Torres, and L. Torner, “Twisted photons,” Nature Physics, vol. 3, no. 5, pp. 305–310, 2007.
  • (30) C. Wang, F.-G. Deng, Y.-S. Li, X.-S. Liu, and G. L. Long, “Quantum secure direct communication with high-dimension quantum superdense coding,” Physical Review A, vol. 71, no. 4, p. 044305, 2005.
  • (31) M. Malik, M. Erhard, M. Huber, M. Krenn, R. Fickler, and A. Zeilinger, “Multi-photon entanglement in high dimensions,” Nature Photonics, vol. 10, pp. 248–252, 2016.
  • (32) S. K. Goyal, P. E. Boukama-Dzoussi, S. Ghosh, F. S. Roux, and T. Konrad, “Qudit-teleportation for photons with linear optics,” Scientific reports, vol. 4, p. 4543, 2014.
  • (33) Y. Zhang, M. Agnew, T. Roger, F. S. Roux, T. Konrad, D. Faccio, J. Leach, and A. Forbes, “Simultaneous entanglement swapping of multiple orbital angular momentum states of light,” Nature Communications, vol. 8, no. 1, p. 632, 2017.
  • (34) N. Bornman, M. Agnew, F. Zhu, A. Vallés, A. Forbes, and J. Leach, “Ghost imaging using entanglement-swapped photons,” npj Quantum Information, vol. 5, no. 1, pp. 1–6, 2019.
  • (35) B. Jack, J. Leach, J. Romero, S. Franke-Arnold, M. Ritsch-Marte, S. M. Barnett, and M. J. Padgett, “Holographic ghost imaging and the violation of a bell inequality,” Phys. Rev. Lett., vol. 103, p. 083602, 2009.
  • (36) L. Chen, J. Lei, and J. Romero, “Quantum digital spiral imaging,” Light: Science & Applications, vol. 3, no. 3, pp. e153–e153, 2014.
  • (37) N. Bornman, S. Prabhakar, A. Vallés, J. Leach, and A. Forbes, “Ghost imaging with engineered quantum states by hong–ou–mandel interference,” New Journal of Physics, vol. 21, no. 7, p. 073044, 2019.
  • (38) M. Mafu, A. Dudley, S. Goyal, D. Giovannini, M. McLaren, M. J. Padgett, T. Konrad, F. Petruccione, N. Lütkenhaus, and A. Forbes, “Higher-dimensional orbital-angular-momentum-based quantum key distribution with mutually unbiased bases,” Physical Review A, vol. 88, no. 3, p. 032305, 2013.
  • (39) F. Steinlechner, S. Ecker, M. Fink, B. Liu, J. Bavaresco, M. Huber, T. Scheidl, and R. Ursin, “Distribution of high-dimensional entanglement via an intra-city free-space link,” Nature communications, vol. 8, p. 15971, 2017.
  • (40) J. Liu, I. Nape, Q. Wang, A. Vallés, J. Wang, and A. Forbes, “Multidimensional entanglement transport through single-mode fiber,” Science advances, vol. 6, no. 4, p. eaay0837, 2020.
  • (41) H. Cao, S.-C. Gao, C. Zhang, J. Wang, D.-Y. He, B.-H. Liu, Z.-W. Zhou, Y.-J. Chen, Z.-H. Li, S.-Y. Yu, et al., “Distribution of high-dimensional orbital angular momentum entanglement over a 1 km few-mode fiber,” Optica, vol. 7, no. 3, pp. 232–237, 2020.
  • (42) D. Cozzolino, E. Polino, M. Valeri, G. Carvacho, D. Bacco, N. Spagnolo, L. K. Oxenløwe, and F. Sciarrino, “Air-core fiber distribution of hybrid vector vortex-polarization entangled states,” Advanced Photonics, vol. 1, no. 4, p. 046005, 2019.
  • (43) D. Cozzolino, D. Bacco, B. Da Lio, K. Ingerslev, Y. Ding, K. Dalgaard, P. Kristensen, M. Galili, K. Rottwitt, S. Ramachandran, et al., “Orbital angular momentum states enabling fiber-based high-dimensional quantum communication,” Physical Review Applied, vol. 11, no. 6, p. 064058, 2019.
  • (44) J. Leach, E. Bolduc, D. J. Gauthier, and R. W. Boyd, “Secure information capacity of photons entangled in many dimensions,” Physical Review A, vol. 85, no. 6, p. 060304, 2012.
  • (45) Y. Zhang, F. S. Roux, T. Konrad, M. Agnew, J. Leach, and A. Forbes, “Engineering two-photon high-dimensional states through quantum interference,” Science Advances, vol. 2, no. 2, p. e1501165, 2016.
  • (46) J. Torres, A. Alexandrescu, and L. Torner, “Quantum spiral bandwidth of entangled two-photon states,” Physical Review A, vol. 68, no. 5, p. 050301, 2003.
  • (47) F. M. Miatto, A. M. Yao, and S. M. Barnett, “Full characterization of the quantum spiral bandwidth of entangled biphotons,” Physical Review A, vol. 83, no. 3, p. 033816, 2011.
  • (48) J. Romero, D. Giovannini, S. Franke-Arnold, S. M. Barnett, and M. J. Padgett, “Increasing the dimension in high-dimensional two-photon orbital angular momentum entanglement,” Phys. Rev. A, vol. 86, p. 012334, Jul 2012.
  • (49) G. Molina-Terriza, J. P. Torres, and L. Torner, “Management of the angular momentum of light: Preparation of photons in multidimensional vector states of angular momentum,” Phys. Rev. Lett., vol. 88, p. 013601, Dec 2001.
  • (50) Y. Zhang, M. Mclaren, F. S. Roux, and A. Forbes, “Simulating quantum state engineering in spontaneous parametric down-conversion using classical light,” Optics Express, vol. 22, no. 14, pp. 17039–17049, 2014.
  • (51) M. Krenn, M. Huber, R. Fickler, R. Lapkiewicz, S. Ramelow, and A. Zeilinger, “Generation and confirmation of a (100×\times 100)-dimensional entangled quantum system,” Proceedings of the National Academy of Sciences, vol. 111, no. 17, pp. 6243–6247, 2014.
  • (52) A. Peres, “Separability criterion for density matrices,” Physical Review Letters, vol. 77, no. 8, p. 1413, 1996.
  • (53) P. Horodecki, “Separability criterion and inseparable mixed states with positive partial transposition,” Physics Letters A, vol. 232, no. 5, pp. 333–339, 1997.
  • (54) R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki, “Quantum entanglement,” Reviews of modern physics, vol. 81, no. 2, p. 865, 2009.
  • (55) K. G. H. Vollbrecht and M. M. Wolf, “Efficient distillation beyond qubits,” Physical Review A, vol. 67, no. 1, p. 012303, 2003.
  • (56) M. Horodecki and P. Horodecki, “Reduction criterion of separability and limits for a class of distillation protocols,” Physical Review A, vol. 59, no. 6, p. 4206, 1999.
  • (57) D. Collins, N. Gisin, N. Linden, S. Massar, and S. Popescu, “Bell inequalities for arbitrarily high-dimensional systems,” Phys. Rev. Lett., vol. 88, p. 040404, Jan 2002.
  • (58) A. K. Ekert, “Quantum cryptography based on bell’s theorem,” Physical review letters, vol. 67, no. 6, p. 661, 1991.
  • (59) A. A. Klyachko, M. A. Can, S. Binicioğlu, and A. S. Shumovsky, “Simple test for hidden variables in spin-1 systems,” Physical review letters, vol. 101, no. 2, p. 020403, 2008.
  • (60) R. Lapkiewicz, P. Li, C. Schaeff, N. K. Langford, S. Ramelow, M. Wieśniak, and A. Zeilinger, “Experimental non-classicality of an indivisible quantum system,” Nature, vol. 474, no. 7352, pp. 490–493, 2011.
  • (61) H. D. L. Pires, H. C. B. Florijn, and M. P. van Exter, “Measurement of the spiral spectrum of entangled two-photon states,” Phys. Rev. Lett., vol. 104, p. 020505, 2010.
  • (62) G. Kulkarni, R. Sahu, O. S. Magaña-Loaiza, R. W. Boyd, and A. K. Jha, “Single-shot measurement of the orbital-angular-momentum spectrum of light,” Nature communications, vol. 8, no. 1, pp. 1–8, 2017.
  • (63) S. Oemrawsingh, E. Eliel, J. Woerdman, E. Verstegen, J. Kloosterboer, et al., “Half-integral spiral phase plates for optical wavelengths,” Journal of Optics A: Pure and Applied Optics, vol. 6, no. 5, p. S288, 2004.
  • (64) J. Leach, E. Yao, and M. J. Padgett, “Observation of the vortex structure of a non-integer vortex beam,” New Journal of Physics, vol. 6, no. 1, p. 71, 2004.
  • (65) C. Rosales-Guzmán and A. Forbes, How to shape light with spatial light modulators. SPIE Press, 2017.

IV Schmidt number

The Schmidt number (KK) quantifies the minimum number of basis states taken from an overcomplete set, required to fully describe a quantum system. For example, in the OAM basis, the product state, ||\ket{\ell}\ket{-\ell}, for some integer \ell, can be used to represent a system of two photons,

|Ψ==LLλ||,\ket{\Psi}=\sum_{\ell=-L}^{L}\lambda_{\ell}\ket{\ell}\ket{-\ell}, (15)

where |λ|2|\lambda_{\ell}|^{2} is the probability of detecting the biphoton state ||\ket{\ell}\ket{-\ell}. The Schmidt number of such a state can be obtained from

K=(|λ|2)2|λ|4.K=\frac{\left(\sum_{\ell}|\lambda_{\ell}|^{2}\right)^{2}}{\sum_{\ell}|\lambda_{\ell}|^{4}}. (16)

Another convenient measure is the width, Δ\Delta\ell, of the distribution as two times the square root of its variance (second central moment), given by

Δ=2||2|λ|2|λ|2.\Delta\ell=2\sqrt{\frac{\sum_{\ell}|\ell|^{2}\ |\lambda_{\ell}|^{2}}{\sum_{\ell}|\lambda_{\ell}|^{2}}}. (17)
Refer to caption
Figure 4: Examples of various OAM (\ell) distributions for a quantum source possessing OAM entanglement.

Both KK and Δ\Delta\ell depend on the shape of the distribution |λ|2|\lambda_{\ell}|^{2}. Examples of various types of distributions for |λ|2|\lambda_{\ell}|^{2} are shown in Fig. 4 for K=21K=21. The distributions are: a square distribution, corresponding to a uniform (maximally entangled state) within a given \ell-range,

λ=1/2L+1,||L;\lambda_{\ell}=1/\sqrt{2L+1},\ |\ell|\leq L; (18)

a Gaussian (normal) distribution

|λg|2exp[||2γG2],|\lambda_{\ell_{g}}|^{2}\propto\exp\left[-\frac{|\ell|^{2}}{\gamma_{G}^{2}}\right], (19)

where γG\gamma_{G} scales with the width of the distribution; a SPDC source miatto2011full ; torres2003quantum

|λS|2(2γS21+γS2)||,|\lambda_{\ell_{S}}|^{2}\propto\left(\frac{2\gamma_{S}^{2}}{1+\gamma_{S}^{2}}\right)^{|\ell|}, (20)

where γS\gamma_{S} is determined by the experimental conditions; and a Lorentz distribution

|λL|21πγL(1+2γL2),|\lambda_{\ell_{L}}|^{2}\propto\frac{1}{\pi\gamma_{L}\left(1+\frac{\ell^{2}}{\gamma_{L}^{2}}\right)}, (21)

where γL\gamma_{L} is a scaling parameter.

Note that all distributions in Fig. 4 have the same KK value, but differ in Δ\Delta\ell, with Δ=12,11.8,14.6 and 41.45\Delta\ell=12,11.8,14.6\text{ and }41.45 for the square, Gaussian, SPDC and Lorentz distributions, respectively.

In this work, we make use the SPDC and normal distributions. For convenience, we relate the the Schmidt number to the scaling parameters as, γS(K1)/4\gamma_{S}\approx\sqrt{(K-1)/4} and γG2.5066K\gamma_{G}\approx 2.5066K for the SPDC and normal distributions, respectively.

V Single fractional OAM analyser

A single fractional OAM mode analyser can be represented on the high dimensional Hilbert space using the OAM basis modes |\ket{\ell}\in\mathcal{H}_{\infty} as

|M,α==c,M(α)|,\ket{M,\alpha}=\sum_{\ell=-\infty}^{\infty}c_{\ell,M}(\alpha)\ket{\ell}, (22)

where the complex coefficients, c,M(α)c_{\ell,M}(\alpha), are computed from the overlap integral, eiϕeiΦM(ϕ;α)𝑑ϕ\int e^{-i\ell\phi}e^{i\Phi_{M}(\phi;\alpha)}\ d\phi. Here eiΦM(ϕ;α)e^{i\Phi_{M}(\phi;\alpha)} is the azimuthally dependent mode characterizing the analyser orientated at an angle α\alpha. Note that a complete decomposition would require an expansion onto a complete basis that includes the radial component. For brevity, we restrict ourselves to the azimuthal degree of freedom, consistent with gotte2007quantum .

By computing the overlap integral, one arrives at complex amplitudes

c,M(α)=ieiαsin(μπ)π(M),c_{\ell,M}(\alpha)=-\frac{ie^{-i\ell\alpha}\sin{\left(\mu\pi\right)}}{\pi(M-\ell)}, (23)

with μ\mu representing the fractional part of the total charge MM. The detection probability for each OAM mode with charge \ell is therefore

P=|c,M(α)|2=sin2(μπ)π2(M)2,P_{\ell}=|c_{\ell,M}(\alpha)|^{2}=\frac{\sin^{2}(\mu\pi)}{\pi^{2}(M-\ell)^{2}}, (24)

consistent with probability amplitudes computed in gotte2007quantum for fractional OAM states.

VI Superpositions of fractional OAM analysers

We have shown that fractional OAM modes project onto the high dimensional state space of OAM modes with complex amplitudes given by Eq. (23). Next, we tailor new amplitudes and phases by superimposing rotated fractional OAM modes

|M,αn=𝒩k=0n1|M,βkα,\ket{M,\alpha}_{n}=\mathcal{N}\sum^{n-1}_{k=0}\ket{M,\beta{{}_{k}}\oplus\alpha}, (25)

where 𝒩\mathcal{N} is a normalization constant. Each fractional mode in this superposition has the same charge, MM, but is rotated by an angle βkα=mod{βk+α,2π}\beta_{k}\oplus\alpha=\text{mod}\left\{\beta_{k}+\alpha,2\pi\right\}, with βk=2πnk\beta_{k}=\frac{2\pi}{n}k. In the OAM basis, Eq. (25) becomes

|M,αn=\displaystyle\ket{M,\alpha}_{n}= 𝒩k=0n1{c,M(βkα)|},\displaystyle\mathcal{N}\sum^{n-1}_{k=0}\left\{\sum_{\ell}c_{\ell,M}(\beta{{}_{k}}\oplus\alpha)\ket{\ell}\right\},
=\displaystyle= 𝒩c,Mn(α)|,\displaystyle\mathcal{N}\sum_{\ell}c^{n}_{\ell,M}(\alpha)\ket{\ell}\,, (26)

where the coefficients c,Mn(α)c^{n}_{\ell,M}(\alpha) are computed from

c,Mn(α)=k=0n1c,M(βkα).\displaystyle c^{n}_{\ell,M}(\alpha)=\sum^{n-1}_{k=0}c_{\ell,M}(\beta{{}_{k}}\oplus\alpha). (27)

Using Eq.  (23) and the condition mod{βkα,2π}=0\mod\{\beta_{k}\oplus\alpha,2\pi\}=0, we obtain

c,Mn(α)=c,M(α)k=0n1eiβk.\displaystyle c^{n}_{\ell,M}(\alpha)=c_{\ell,M}(\alpha)\sum^{n-1}_{k=0}e^{i\beta_{k}\ell}. (28)

Since the summation can be evaluated as a geometric series, after some simplification it results in

k=0n1eiβk=eiπ(n1)/ncsc(πn)sin(π).\sum_{k=0}^{n-1}e^{i\beta_{k}\ell}=e^{-i\pi\ell(n-1)/n}\csc\left(\frac{\pi\ell}{n}\right)\sin(\pi\ell).

Therefore the coefficients can be written as

c,Mn(α)=\displaystyle c^{n}_{\ell,M}(\alpha)= eiπ(n1)/nAnc,M(α),\displaystyle e^{-i\pi\ell(n-1)/n}A^{n}_{\ell}\ c_{\ell,M}(\alpha)\,, (29)

where

An=\displaystyle A^{n}_{\ell}= csc(πn)sin(π),\displaystyle\csc\left(\frac{\pi\ell}{n}\right)\sin\left(\pi\ell\right),
=\displaystyle= {0mod{,n}01mod{,n}=0.\displaystyle\left\{\begin{array}[]{cc}0&\text{mod}\left\{\ell,n\right\}\neq 0\\[5.69054pt] 1&\text{mod}\left\{\ell,n\right\}=0\\ \end{array}\right.. (32)

Consequently, the overlap probabilities are P,n=|𝒩Anc,M(α)|2P_{\ell,n}=|\mathcal{N}\ A^{n}_{\ell}c_{\ell,M}(\alpha)|^{2}. Importantly, the probabilities are independent of α\alpha. Accordingly, the new spectrum has the amplitudes |c,M||c_{\ell,M}|, but following the selection rule AnA^{n}_{\ell}. Indeed, this new spectrum can be tuned by carefully selecting nn, therefore enabling control of the OAM subspaces.

VII Spiral imaging of fractional OAM analysers

Our fractional OAM analysers can be decomposed into the OAM basis using entangled photons through digital spiral imaging. In this scheme, one photon from an entangled pair interacts with the analyser while its twin is decomposed in the OAM basis. The entangled photon pair has a biphoton state

|Ψ==LLλ||,\ket{\Psi}=\sum_{\ell=-L}^{L}\lambda_{\ell}\ket{\ell}\ket{-\ell}, (33)

as defined in Eq.  (15). The probability amplitude for detecting the mmth OAM mode, given a M charged fractional mode of nn superpositions, is

c~mn(α)=\displaystyle\tilde{c}^{n}_{m}(\alpha)= m|M,α||nΨ,\displaystyle\mathcal{M}\bra{m}\braket{M,\alpha}{{}_{n}}{\Psi},
=\displaystyle= =LLλm|M,α||n.\displaystyle\mathcal{M}\ \sum_{\ell=-L}^{L}\lambda_{\ell}\braket{m}{\ell}\braket{M,\alpha}{{}_{n}}{-\ell}. (34)

where \mathcal{M} is a normalisation constant such that m|c~mn(α)|2=1\sum_{m}|\tilde{c}^{n}_{m}(\alpha)|^{2}=1. Due to the orthonormality of the OAM basis, the overlap m|\braket{m}{\ell} is simply the Kronecker delta function δm,\delta_{m,\ell}, which evaluates as 0 if m\ell\neq m or 1 if =m\ell=m. Since from Eq.  (29) we know the expansion coefficients for the analyser in terms of the OAM basis, M,α||n\braket{M,\alpha}{{}_{n}}{\ell} evaluates as

c~mn(α)=\displaystyle\tilde{c}^{n}_{m}(\alpha)= =LLδm,λ[𝒩cn(α)]\displaystyle\mathcal{M}\ \sum_{\ell=-L}^{L}\delta_{m,\ell}\lambda_{\ell}\ \left[\mathcal{N}c^{n}_{-\ell}(\alpha)\right]^{*}
=\displaystyle= 𝒩λm[cmn(α)].\displaystyle\mathcal{M}\mathcal{N}\ \lambda_{m}\ \left[c^{n}_{-m}(\alpha)\right]^{*}. (35)

These new weightings are simply the original coefficients of the analysers modulated by the spectrum of the entangled system. For a maximally entangled state, we obtain the expression |c~mn(α)|2=|cmn(α)|2|\tilde{c}^{n}_{m}(\alpha)|^{2}=|c^{n}_{-m}(\alpha)|^{2}, being the original weightings of the analyser, as desired.

In Fig. 5 we show the measured weightings for our SPDC system which has a normally distribution of OAM modes with Δ=11\Delta\ell=11 centered at =0\ell=0. We show results for |M,αn=|0.5,01,|1.5,03,|2.5,05\ket{M,\alpha}_{n}=\ket{0.5,0}_{1},\ket{1.5,0}_{3},\ket{2.5,0}_{5} for analysers n=1,3n=1,3 and 55 in Fig. 5(a),(b) and (c), respectively. It can be seen that the theory (points) and experiment (bars) are in good agreement. To obtain these results, two photons where generated from an SPDC source and modulated with SLMs (see experimental setup in the Methods section in the main text). One SLM was encoded with a fractional OAM mode projecting onto the state, |M,0n\ket{M,0}_{n}, while the second SLM was encoded with OAM basis modes, |\ket{\ell}.

Refer to caption
Figure 5: Measured (bars) and theoretical spectrum (points) for fractional OAM analysers (a) |M,αn=|0.5,01\ket{M,\alpha}_{n}=\ket{0.5,0}_{1}, (b) |M,αn=|1.5,03\ket{M,\alpha}_{n}=\ket{1.5,0}_{3}, and (c) |M,αn=|2.5,05\ket{M,\alpha}_{n}=\ket{2.5,0}_{5} resulting from digital spiral imaging with entangled photons. Here the weightings are modulated by the OAM spectrum of the entanglement source according to Eq.  (35).

VIII Detection probability from relative rotations of the analysers

Refer to caption
Figure 6: Simulated normalised probability curves for (a) nn = 1, (b) nn = 3, (c) nn = 5 and (d) nn = 7, with fractional OAM M=n2M=\frac{n}{2}, as a function of the relative orientation θ\theta between the two analysers and the dimensions, KK, of the entanglement state. The second row of panels (e)-(h) are probability curves for specific KK values for each analyser. The normalisation to unity was performed to illustrate the impact of the dimensions on the visibility. Here, the OAM spectrum shape was assumed to follow a normal (Gaussian) distribution.

Given a bipartite system of the form of Eq. (33), we want to know what the detection probability is, due to the relative rotations of our fractional OAM analysers acting the entangled photons. Suppose the first analyser projects onto the state |M,θ1n\ket{M,\theta_{1}}_{n}, and the second analyser projects on the state |M,θ2n\ket{-M,\theta_{2}}_{n}. A joint measurement on a two photon system using the two analysers is characterized by the product state |M,θ1n|M,θ2n\ket{M,\theta_{1}}_{n}\ket{-M,\theta_{2}}_{n}. The probability amplitude resulting from such a measurement is

Cn(θ1,θ2)=\displaystyle C_{n}(\theta_{1},\theta_{2})= θ2,M|nθ1,M||nΨ\displaystyle\bra{\theta_{2},-M}_{n}\braket{\theta_{1},M}{{}_{n}}{\Psi}
=\displaystyle= =λθ1,M||nθ2,M||n.\displaystyle\sum_{\ell=-\infty}^{\infty}\lambda_{\ell}\ \braket{\theta_{1},M}{{}_{n}}{\ell}\ \braket{\theta_{2},-M}{{}_{n}}{-\ell}. (36)

Therefore we only need to know how to decompose each of the analysers in the OAM basis to obtain the detection probability for the joint measurements. Using Eq. (29), it follows that

Cn(θ1,θ2)\displaystyle C_{n}(\theta_{1},\theta_{2})\propto =λSPDC[c,Mn(θ1)analyserc,Mn(θ2)analyser].\displaystyle\sum^{\infty}_{\ell=-\infty}\underbrace{\lambda_{\ell}}_{\text{SPDC}}\ [\ \underbrace{c_{\ell,M}^{n}(\theta_{1})}_{\text{analyser}}\ \underbrace{c_{-\ell,-M}^{n}(\theta_{2})}_{\text{analyser}}\ ]^{*}. (37)

We use this approach to numerically calculate the detection probabilities |Cn(θ1,θ2)|2|C_{n}(\theta_{1},\theta_{2})|^{2} by simply calculating the probability amplitudes for each analyser in the OAM basis with a desired rotation θ1,2\theta_{1,2} and multiplying them with the coefficients λ\lambda_{\ell} that determine the quantum system being probed.

An alternative approach, can be to compute the overlap integral by considering the modal overlaps in the azimuthal degree of freedom, ϕ\phi, following

θ,M||n=12πexp(iΦM(ϕ;θ))×exp(iϕ)𝑑ϕ,\braket{\theta,M}{{}_{n}}{\ell}=\frac{1}{2\pi}\int\exp(-i\Phi_{M}(\phi;\theta))\times\exp(i\ell\phi)\ d\phi,

with ΦM(ϕ;θ)/2π\Phi_{M}(\phi;\theta)/\sqrt{2\pi} being the transmission function of the fractional OAM analyser projects onto the state |M,θn\ket{M,\theta}_{n}. We can rewrite probability amplitude Cn(θ1,θ2)C_{n}(\theta_{1},\theta_{2}) as an overlap integral given

Cn(θ1,θ2)=14π=(λeiΦM(ϕ1;θ1)eiϕ1×eiΦM(ϕ2;θ2)eiϕ2dϕ1dϕ2),C_{n}(\theta_{1},\theta_{2})=\frac{1}{4\pi}\sum_{\ell=-\infty}^{\infty}\left(\lambda_{\ell}\iint e^{-i\Phi_{M}(\phi_{1};\theta_{1})}e^{i\ell\phi_{1}}\right.\\ \left.\vphantom{\iint e^{-i\Phi_{M}(\phi_{1};\theta_{1})}}\times e^{-i\Phi_{-M}(\phi_{2};\theta_{2})}e^{-i\ell\phi_{2}}\ d\phi_{1}d\phi_{2}\right), (38)

where Φ±M(ϕ1,2,θ1,2)\Phi_{\pm M}(\phi_{1,2},\theta_{1,2}) are the phases of the fractional OAM analysers. Since eiΦM(ϕ1;θ1)e^{-i\Phi_{M}(\phi_{1};\theta_{1})} has no \ell dependence, we can introduce the summation into the second integral resulting in

Cn(θ1,θ2)=12πeiΦM(ϕ1;θ1)(eiΦM(ϕ2;θ2)×12π=λei(ϕ1ϕ2)dϕ2)dϕ1.C_{n}(\theta_{1},\theta_{2})=\frac{1}{2\pi}\int e^{-i\Phi_{M}(\phi_{1};\theta_{1})}\ \left(\vphantom{\sum_{\ell=-\infty}^{\infty}}\int e^{-i\Phi_{M}(\phi_{2};\theta_{2})}\right.\\ \left.\times\frac{1}{2\pi}\sum_{\ell=-\infty}^{\infty}\lambda_{\ell}e^{i\ell\left(\phi_{1}-\phi_{2}\right)}d\phi_{2}\right)d\phi_{1}. (39)

It is convenient to define the periodic function

Λ(ϕ1ϕ2)=12π=λei(ϕ1ϕ2),\Lambda(\phi_{1}-\phi_{2})=\frac{1}{2\pi}\sum_{\ell=-\infty}^{\infty}\lambda_{\ell}e^{i\ell\left(\phi_{1}-\phi_{2}\right)},

with angular harmonics ei(ϕ1ϕ2)e^{i\ell\left(\phi_{1}-\phi_{2}\right)} determined by the coefficients λ\lambda_{\ell}, and use it to rewrite Cn(θ1,θ2)C_{n}(\theta_{1},\theta_{2}) as

Cn(θ1,θ2)=12πeiΦM(ϕ1;θ1)(eiΦM(ϕ2;θ2)Λ(ϕ1ϕ2)𝑑ϕ2)dϕ1.C_{n}(\theta_{1},\theta_{2})=\frac{1}{2\pi}\int e^{-i\Phi_{M}(\phi_{1};\theta_{1})}\\ \left(\int e^{-i\Phi_{M}(\phi_{2};\theta_{2})}\Lambda(\phi_{1}-\phi_{2})d\phi_{2}\right)d\phi_{1}. (40)

Notice that the second integral is a convolution between Λ(ϕ1ϕ2)\Lambda(\phi_{1}-\phi_{2}) and the second analyser. As a example, we consider a maximally entangled state (λ:=constant\lambda_{\ell}:=\text{constant}). In this case, Λ(ϕ1ϕ2)=δ(ϕ1ϕ2)\Lambda(\phi_{1}-\phi_{2})=\delta(\phi_{1}-\phi_{2}) and therefore

Cn(θ1,θ2)=12πeiΦM(ϕ;θ1)eiΦM(ϕ;θ2)𝑑ϕ.C_{n}(\theta_{1},\theta_{2})=\frac{1}{2\pi}\int e^{-i\Phi_{M}(\phi;\theta_{1})}\ e^{-i\Phi_{-M}(\phi;\theta_{2})}d\phi.

The integral now only depends in the transmission functions of the analysers with an analytical solution found in huang2018various .

We now calculate the probability Pn(θ1,θ2)=|Cn(θ1,θ2)|2P_{n}(\theta_{1},\theta_{2})=\left|C_{n}(\theta_{1},\theta_{2})\right|^{2} as a function of relative orientation θ=(θ1θ2)\theta=(\theta_{1}-\theta_{2}) between the two analysers and the dimensions, KK, of an entangled system with some given OAM spectrum |λ|2|\lambda_{\ell}|^{2}. The latter is embedded in the function Λ(ϕ1ϕ2)\Lambda(\phi_{1}-\phi_{2}). Figures 6(a)-(d) show examples of the probability surfaces assuming a normal (Gaussian) spectrum |λ|2|\lambda_{\ell}|^{2} for superposition states n=1,3,5n=1,3,5 and 77. In the second row of, Fig. 6 (e)-(h), we show examples of the probability curves normalised to unity for several values of dimensionality KK. Here, it can be seen that the frequency of the probabilities as a function of θ\theta increases with nn, owing to the n-fold symmetry in the phase profiles of the analysers.

Refer to caption
Figure 7: Contour plots of visibility vs dimensionality (KK) vs nn (number of fractional OAM superpositions) for the (a) Normal, (b) SPDC theory and (c) maximally uniform distribution ( or maximally entangled pure state). Here we demonstrate the sensitivity of the analysers to the dimensions of a OAM entanglement. The visibilities from the maximally entangled state demonstrates the minimum number of modes required to have a visibility V=1V=1.
Refer to caption
Figure 8: Visibility as a function of purity (p) and dimensions (K) for the (a) Normal, (b) SPDC theory and (c) the uniform (maximally entangled state) obtained for fractional OAM projections corresponding to n=1,5,7,51n=1,5,7,51.

Crucially, the exact shape and visibility of the curves depends on both the dimensions (KK) of the state being probed and the number of superpositions (nn). For all nn’s, the visibility for a specific KK shows a decreasing trend as the number of superpositions nn are increased. Therefore the analysers are sensitive to the dimensions of the system.

We also found that the shape of the spectrum affects the measured probabilities, as illustrated in Fig. 7(a)-(d) for square (maximally entangled), normal (Gaussian) and SPDC, respectively.

Now that we have shown how the detected probabilities depended on the dimensions and superposition states measured, in the following section we study the relation between the visibility and dimensions quantitatively.

IX Visibility for different spectra

The visibilities are calculated from detection probabilities resulting from the projections of an entangled state with an initial OAM distribution |λ|2|\lambda_{\ell}|^{2} onto the states |M,0n|M,θn\ket{M,0}_{n}\ket{-M,\theta}_{n}, where θ[0,2π]\theta\in[0,2\pi] is their relative rotation.

For example, for a square (uniform) (KK\rightarrow\infty) distribution and nn superpositions of fractional modes the probability is given by huang2018various

P(θ1,θ2)=\displaystyle P(\theta_{1},\theta_{2})= |C(θ1,θ2)|2\displaystyle|C(\theta_{1},\theta_{2})|^{2}
=\displaystyle= asin2(Mπn)+cos2(Mπn),\displaystyle a\sin^{2}\left(\frac{M\pi}{n}\right)+\cos^{2}\left(\frac{M\pi}{n}\right), (41)

with a=(π(2t1)nθ)2/π2a=(\pi(2t-1)-n\theta)^{2}/\pi^{2} for 2πn(t1)θ2πn(t),t=1,,n,\frac{2\pi}{n}(t-1)\leq\theta\leq\frac{2\pi}{n}(t),\quad t=1,...,n, where tt indexes each 2π/n2\pi/n period over the range of 0θ<2π0\leq\theta<2\pi and θ=θ1θ2\theta=\theta_{1}-\theta_{2}. This oscillating function results in fringes with a visibility function given by

Vn(M)=1cos2(Mπn)1+cos2(Mπn).V_{n}(M)=\frac{1-\cos^{2}\left(\frac{M\pi}{n}\right)}{1+\cos^{2}\left(\frac{M\pi}{n}\right)}.

For n=1n=1, parabolic fringes with perfect visibility occur when M=+0.5M=\ell+0.5 for all OAM integer charges \ell. In contrast, when n>1n>1 high visibility fringes occur for only specific choices of nn and MM. That is, parabolic fringes with high visibility (V=1V=1) are expected when nn is odd and mod{Mn2,n}=0\text{mod}\left\{M-\frac{n}{2},n\right\}=0.

Contour plots of the visibilities with changing dimensions (KK) and fractional OAM superpositions (nn) for various OAM spectral shapes (Normal, SPDC, Uniform) are shown in Fig. 7 (a)-(c) for pure states. As shown, the assumed spectrum can affect the visibility that is measured for various superpositions (nn). The visibilities for each analyser (nn) and spectrum shape are monotonic with increasing KK. In particular, for the uniform spectrum (maximal entanglement in K dimensions) the visibility is 1 above some K=dnK=d_{n} and zero below this. We further exploit this property to determine the dimensionality of an entanglement system.

X Visibility of mixed states

The visibilities that can be measured with our analysers are not only dependent on the effective dimensions of the system but also the purity. In particular, we consider the isotropic state,

ρp=p|ΨdΨd|+1pd2𝕀d2,\rho_{p}=p\ket{\Psi_{d}}\bra{\Psi_{d}}+\frac{1-p}{d^{2}}\mathbb{I}_{d^{2}}, (42)

which can decomposed into the high-dimensional entangled state, |Ψd\ket{\Psi_{d}}, and the separable and mixed state, 1/d2𝕀d2=1/d2,=L,=L||||1/d^{2}\mathbb{I}_{d^{2}}=1/d^{2}\sum_{\ell,\ell^{\prime}=-L}^{\ell,\ell^{\prime}=L}\ket{\ell}\ket{\ell^{\prime}}\bra{\ell^{\prime}}\bra{\ell}, where 𝕀d2\mathbb{I}_{d^{2}} is the identity operator. Such states model quantum systems that have noise contributions from the environment. Here pp can be associated with the purity of the state ranging from a maximally mixed (p=0p=0) to a pure state (p=1p=1). Interestingly, the isotropic state is separable for p1/(d+1)p\leq 1/(d+1) and entangled otherwise. Importantly the generalised Bell inequality can also be violated when p>2/Sdp>2/S_{d} where SdS_{d} is the Bell parameter Collins2002 . We show that both pp and dd can be measured using our analysers. For convenience, we assume dKd\approx K, where K is the effective dimensionality of the entanglement. We will demonstrate that we can measure both pp and KK using our analysers.

Firstly, we calculate the detection probabilities from the overlap, Pn(θ;K,p)=Trace(M^ρp)P_{n}(\theta;K,p)=\text{Trace}(\hat{M}\rho_{p}) where M^\hat{M} projects onto the states |M,0n|M,θn\ket{M,0}_{n}\ket{-M,\theta}_{n}. As a result, the detection probability can be written as

Pn(θ;p,K)=pPn(θ;K)+1pK2In(0;K),P_{n}(\theta;p,K)=pP_{n}(\theta;K)+\frac{1-p}{K^{2}}I_{n}(0;K), (43)

where Pn(θ,K)=|=LLλcn(0)cn(θ)|2P_{n}(\theta,K)=\left|\sum_{\ell=-L}^{L}\lambda_{\ell}c^{n}_{\ell}(0)c^{n}_{-\ell}(\theta)\right|^{2} and In(0;K)=|=LL|cn(0)|2|2I_{n}(0;K)=\left|\sum_{\ell=-L}^{L}\left|c^{n}_{\ell}(0)\right|^{2}\ \right|^{2} is the overlap of the analysers with the maximally mixed state. Since the functions are periodic and obtain maximum and minimum values for θ=0\theta=0 and π/n\pi/n, respectively, we obtain the expression

ΔPn(p,K)\displaystyle\Delta P_{n}(p,K) =Pn(0;p,K)Pn(π/n;p,K)\displaystyle=P_{n}(0;p,K)-P_{n}(\pi/n;p,K)
=pΔPn(K),\displaystyle=p\Delta P_{n}(K), (44)

where ΔPn(K)=Pn(0,p=1,K)Pn(π/n,p=1,K)\Delta P_{n}(K)=P_{n}(0,p=1,K)-P_{n}(\pi/n,p=1,K). The visibilities can be calculated from

Vn(p,K)\displaystyle V_{n}(p,K) =ΔPn(p,K)Pn(π/n;p,K)+Pn(0;p,K),\displaystyle=\frac{\Delta P_{n}(p,K)}{P_{n}(\pi/n;p,K)+P_{n}(0;p,K)}, (45)

We show the dependence of the visibilities on the dimensions (KK) and purity (pp) in Fig. 8 (a-c) for the Normal, SPDC and uniform distribution, respectively. Each panel shows the visibilties from various analysers depending on the number of superposition (nn). As shown the visibilities increase monotonically with increasing dimensions (KK) as well as purity pp for each analyser where p=1p=1 obtains a maximal visibility. However, as nn increases the visibilities decrease for all pp and KK. Since the visibilities are monotonic in both pp and KK as well as nn, we can exploit this property to map the dimensions of a quantum system. We favour this approach since the visibilities can be easily measured and require few measurements (peak and trough).

XI Comparison between a known and guessed spectrum

Noise level pSPDCp^{SPDC} KSPDCK^{SPDC} pnormalp^{\text{normal}} KnormalK^{\text{normal}} QQ K^\hat{K} p^\hat{p}
low 0.45 ±\pm 0.03 22.84 ±\pm 0.62 0.42 ±\pm 0.02 20.00 ±\pm 0.32 19.19 22 0.44
high 0.13 ±\pm 0.01 17.73 ±\pm 0.71 0.13 ±\pm 0.01 17.18 ±\pm 0.34 3.76 18 0.13
Table 2: Measured purity (pp) and dimensionality (KK), under low and high noise levels, compared to estimates from other methods. Here QQ is the average quantum contrast.

Using our procedure we measured the dimensions and purity of SPDC photons with varying noise levels (low and high). The results are summarised in Table 2. In the second and third column, we know what the input spectrum shape (SPDC) is and can therefore accurately optimise for the dimensions (KK) and purity (pp) of the state (see Results section). Further, if we guess the spectrum based on its shape (symmetry) we also obtain values that are similar to the expected results, with a relative error of up to 13%\approx 13\%. This was done using the normal distribution as the function modelling the mode spectrum. Next, we verify our result using the values extracted from the spiral bandwidth.

To calculate the the expected dimensions, K^\hat{K}, we used the coincidences from the spiral spectrum in the OAM basis, i.e CA,mBC_{\ell_{A},m_{B}}, where A\ell_{A} denotes the mode index of photon A and mBm_{B} for photon B. Since we want the Schmidt number of the pure part of the state, we subtracted the accidentals and then used Eq. (16), yielding results with a low relative error of 3%3\%, validating our results. Subsequently, we estimated the purity p^\hat{p} (see Results section). Note that no accidental subtraction was performed in this case. Accordingly, to estimate the purity we measured the quantum contrast using

Q=C¯/C¯,Q=\bar{C}/\bar{C}^{\prime}, (46)

taken from the ratio between the average coincidences in the anti-diagonal entries, C¯=C,\bar{C}=\sum_{\ell}C_{\ell,-\ell} and the average noise contribution from coincidences excluding the anti-diagonal entries, i.e C¯=1d2d(CTdC¯)\bar{C}^{\prime}=\frac{1}{d^{2}-d}\left(C_{T}-d\ \bar{C}\right). Here CTC_{T} corresponds to the total coincidences CT=,mC,mC_{T}=\sum_{\ell,m}C_{\ell,m}. Indeed, using the quantum contrast we obtained a purity that is comparable to that obtained from our method showing a relative error of only up to 2%2\%.