Structured light has become topical of late, where controlling light in all its degrees of freedom has offered novel states of light long predicted, enhanced functionality in applications, and a modern toolbox for probing fundamental science. Structuring light as single photons and entangled states allows the spatial modes of light to be used to encode a large alphabet, accessing high dimensional Hilbert spaces for fundamental tests of quantum mechanics and improved quantum information processing tasks. In this tutorial, we outline the basic concepts of high dimensional quantum states expressed in a basis of spatial modes (structured light) and explain how to create, control, and detect such quantum states in the laboratory with a focus on transverse spatial modes such as the orbital angular momentum and pixel (position) modes. Finally, we highlight some example applications of such quantum structured light, from communications to imaging.
I. INTRODUCTION
It has become topical of late to tailor light in all its degrees of freedom, looking beyond the intensity profile alone for so-called structured light.1 Traditionally this has been performed with classical light, for example, amplitude, phase, and polarization control in space2 and wavelength/frequency manipulation for temporal control.3 Blending space and time together has given rise to exotic forms of spatiotemporal light, long been predicted but only recently observed,4–6 fueling a movement toward high-dimensional forms of structured light.7 Creation and detection tools are numerous,8 including directly from lasers, for compact sources of structured light.9 These tailored optical fields with complex structures have found a myriad of applications, which have been extensively reviewed to date.10,11 Perhaps the most topical example is that of optical orbital angular momentum (OAM). The fact that light could carry OAM has been known since at least the days of atomic physics (accounting for the rare quadrupole transitions in atomic states), but it was only 30 years ago (in 1992) that OAM was directly connected to the spatial structure of light through an azimuthal phase of the form exp(iℓϕ), winding the phase ℓ times around the azimuth (ϕ) for ℓℏ of OAM per photon.12 The explosion of activity since has been captured in recent commemorative perspectives and reviews.11,13
The control of structured quantum states of light is far less developed, having appeared in the spatial domain with OAM just 20 years ago.14 In this seminal work by Nobel Laureate Anton Zeilinger, the conservation of OAM was confirmed down to the single photon level and demonstrated as qubit entangled states in several OAM subspaces, the first demonstration of spatial mode entanglement. As there are an infinite number of spatial modes on any given basis, this form of quantum structured light has the potential to realize d dimensional states15–17 for a large encoding alphabet that scales as log2 d bits/photon and enhanced security with cloning fidelity that scales as , making this approach attractive for quantum information processing. Although d can be increased by degrees of freedom such as path and time, here we will consider only spatially structured quantum light. One good reason for the interest in this avenue is that the classical properties of spatial modes often translate to benefits in the quantum realm, for instance, self-healing of Bessel structured quantum states18 and enhanced control by mixing polarization and spatial modes for hybrid entanglement,19–21 analogous to vectorial structured light.
For a long time, spatial mode entanglement remained at the qubit (d = 2) level, mimicking polarization states, although with multiple options for the two-dimensional subspaces through many combinations of orthogonal spatial modes. Indeed, this analogy allowed the easy transition of the quantum toolkit from polarization qubits to structured light qubits, for instance, quantum state tomography (QST)22,23 and bell violation tests,24 enabled by computer generated holograms. It is only in the last decade that the true high-dimensional nature of structured quantum light has come to the fore, both as entangled states and as heralded single photons. Seminal contributions have demonstrated high-dimensional Bell tests,25 high-dimensional QSTs,26 fast and accurate entanglement tests27 and witnesses,28 high-dimensional quantum interference,29 and sophisticated high-dimensional quantum state creation tools.30–34 The application of structured quantum light in high-dimensional quantum information processing has seen quantum key distribution up to d = 7 in free space35,36 and d = 6 in optical fibers,37,38 quantum secret sharing up to d = 11 with Bessel photons,39 entanglement swapping up to d = 2 with OAM,40 and teleportation up to d = 3 with linear optics41,42 and d = 15 with non-linear optics.43
In this tutorial, we start by briefly covering the formalism used in working with high-dimensional structured quantum light, concentrating on practical tools using OAM and pixel (position) modes as examples. We then outline some of the basics needed in order to get started, extending from the creation step in bulk crystals to aligning, manipulating, detecting, and characterizing the output state, along with some caveats needed for these considerations. Finally, we look at some example applications ranging from imaging to secure communications, where the advantages and challenges that come with harnessing high-dimensional states are outlined.
II. THEORETICAL CONCEPT
A. Structured light and its degrees of freedom
In this section, we introduce the reader to the concepts of internal degrees of freedom of photons and thereafter discuss ways to obtain high dimensional encoding in quantum structured light.
High dimensional structured photons. (a) A photon field can be ascribed to a state that characterizes its polarization and spatial and temporal degrees of freedom. (b) By selecting one of the degrees of freedom, say polarization, a qubit space can be formed consisting of a two dimensional (d = 2) basis formed by the right |R⟩ and left |L⟩ circular polarization states. The two-dimensional space can be visualized geometrically using the Bloch sphere (equivalent to the Poincaré sphere for polarization). (c) The temporal and transverse spatial degree of freedom are higher dimensional since more than two states are required to describe each of them. This can be seen by partitioning each degree of freedom into discrete, non-overlapping cells, where each cell contains information about the state of the photon. These states are commonly referred to as the time-bin and pixel basis states for the temporal and spatial components, respectively. (d) Besides the decomposition of an arbitrary spatial profile using the discrete pixel basis modes (top panel), one can also use spatial structured modes, where the arbitrary field can now be expressed using a complete set of spatially structured patterns (bottom panel). These modes also constitute the modal basis for quantum structured light.
High dimensional structured photons. (a) A photon field can be ascribed to a state that characterizes its polarization and spatial and temporal degrees of freedom. (b) By selecting one of the degrees of freedom, say polarization, a qubit space can be formed consisting of a two dimensional (d = 2) basis formed by the right |R⟩ and left |L⟩ circular polarization states. The two-dimensional space can be visualized geometrically using the Bloch sphere (equivalent to the Poincaré sphere for polarization). (c) The temporal and transverse spatial degree of freedom are higher dimensional since more than two states are required to describe each of them. This can be seen by partitioning each degree of freedom into discrete, non-overlapping cells, where each cell contains information about the state of the photon. These states are commonly referred to as the time-bin and pixel basis states for the temporal and spatial components, respectively. (d) Besides the decomposition of an arbitrary spatial profile using the discrete pixel basis modes (top panel), one can also use spatial structured modes, where the arbitrary field can now be expressed using a complete set of spatially structured patterns (bottom panel). These modes also constitute the modal basis for quantum structured light.
The polarization degree of freedom was initially the best candidate for photon information processing due to its ease of control with conventional linear optical elements and was used to demonstrate numerous fundamental tests of quantum mechanics (Bell inequality violations44 and quantum erasers45) and the initial demonstrations of quantum communication and cryptography (quantum key distribution,46 teleportation,47 and superdense coding48). However, the polarization states of a single photon are restricted to a two level system that can be composed of the canonical right (|0⟩ ≡ |R⟩) and left (|1⟩ ≡ |L⟩) circular polarization eigenstates as basis modes. Here, the states |0(1)⟩ represent the logical (standard) basis.
The two level system can be visualized using the Bloch sphere, as shown in Fig. 1(b), where any two modes on the opposite ends of the sphere can be used to form a logical basis. For example, if we look at the equator of the sphere, the superposition states |R⟩ ± |L⟩, corresponding to the horizontal |0⟩ ≡ |H⟩ and vertical |1⟩ ≡ |V⟩ linear polarization states, form another logical basis for expressing polarization states, respectively. The same is true for the rectilinear basis states, |R⟩ ± i|L⟩, corresponding to the diagonal |0⟩ ≡ |D⟩ and anti-diagonal |1⟩ ≡ |A⟩ linear polarization states, respectively. Therefore, for polarization states, the encoding basis, , can only contain two orthogonal states, |0⟩ and |1⟩, at a time. For higher dimensional encoding, we require that the encoding basis has d > 2 states, i.e., having the elements |0⟩, |1⟩, …, |d − 1⟩.
In the section that follows, we explore how the remaining degrees of freedom (time and space) can be used to obtain higher dimensional states. In particular, we will focus on the spatial-momentum basis, although some of the techniques can be transferred to the temporal basis.
B. High dimensional (d > 2) structured photons
Our remaining degrees of freedom, the temporal, spectral, and spatial components, are continuous. For example, the spatial components span , which includes the transverse coordinates (x, y) and the longitudinal component z. Similarly, the conjugate basis for the spatial components, which corresponds to the momentum basis, can also be separated into its continuous transverse and longitudinal parts. Because we would like to obtain a discrete basis, it is pertinent to ask, how can we obtain discrete higher dimensional photon states given the continuous nature of the spatial basis?
While this approach works well, the pixel dimensionality can be limited by the performance of the optical system and must thus be chosen appropriately.52 The choice of pixel sizes can, for example, affect the quality of quantum imaging experiments if not chosen in conjunction with the performance of the optical setup (angular resolution, point spread function), as well as taking into account the joint probability amplitude (correlation length) of the correlated photons.52,53 Put simply, more pixel states do not imply higher resolution; the performance is highly dependent on the quality of the photon sources and optical elements. Furthermore, for encoding purposes, pixel modes are not stable on propagation due to diffraction and can present challenges in applications that require long propagation distances, especially if the pixels are encoded with arbitrary amplitudes and phases and the resulting fields are not modes of free space.
Laguerre–Gaussian basis modes. Intensity profiles for the spatial distributions described by LGℓ,p modes with indices ℓ = [−2, 2] and p = [0, 3] with the phase profiles given as insets in the top-right corner of each mode. An example of the two dimensional state space that can be created using the LG mode basis for (b) ℓ = ±1 and p = 0 (c) ℓ = 2, 3 and p = 3.
Laguerre–Gaussian basis modes. Intensity profiles for the spatial distributions described by LGℓ,p modes with indices ℓ = [−2, 2] and p = [0, 3] with the phase profiles given as insets in the top-right corner of each mode. An example of the two dimensional state space that can be created using the LG mode basis for (b) ℓ = ±1 and p = 0 (c) ℓ = 2, 3 and p = 3.
Furthermore, OAM modes (LGℓ,p=0) have been shown to have an analogous representation on the Bloch sphere, similar to polarization states in Fig. 1(b).59 We illustrate our OAM Bloch sphere in Fig. 2(b), where the poles of the sphere contain the |±ℓ⟩ = |LG±ℓ,p=0⟩ modes, and the equally weighted superpositions are contained at the equator. While this is only shown for ℓ = ±1, multiple spheres can be constructed using arbitrary ℓ values. Furthermore, our illustration in Fig. 2(c) shows that the radial modes, in combination with the OAM modes, can be used to construct arbitrary two dimensional spheres. Infinitely many such spheres can be constructed from the LG modes. As with the polarization qubits, given any two OAM modes, numerous experiments have emerged where the OAM qubits are used as a computational basis for processing quantum information with further advances made to harness them for high dimensional encoding schemes for single photons,36 entangled two photon states,60 and three photon states33 showing their significance in quantum science.
Next, we explore two photon states and introduce the concept of entanglement.
C. Entangled structured light
So far, our modes of structured light have been discussed in the context of single photons. A quantum system may entail more than a single photon, and in some instances, the state of one single photon may depend on that of another (entanglement). In order to elucidate this, we describe how to write the state of more than one photon and what it means in cases where there is entanglement present.
The implications of this are that one can no longer describe either photon individually (in each basis alone) but rather requires both at the same time. Accordingly, if |ℓ1⟩ was measured on photon A, photon B would then be in the state |ℓ2⟩, and if |ℓ2⟩ was measured on photon A, photon B would then be in the state |ℓ1⟩. This is always true despite any arbitrary distance of separation. This prompted Einstein to question the reality,63 calling this “spooky action at a distance,” which has since been termed quantum entanglement.64 Consequently, this property is a direct consequence of the inability to factorize the state into its subsystems. As such, composite states for which one is not able to write as a product state, i.e., |ψ⟩AB ≠ |ψ⟩A ⊗ |ψ⟩B, can be called entangled.
Density matrices of pure and mixed high dimensional entangled states. Exemplary density matrices of photons entangled in high dimensions for (a) d = 2, (b) d = 3, and (c) d = 5. The orbital angular momentum basis modes for each state are shown above each density matrix. Isotropic mixed states with p = 0.4 for the same entangled states in (d) d = 2, (e) d = 3, and (f) d = 5.
Density matrices of pure and mixed high dimensional entangled states. Exemplary density matrices of photons entangled in high dimensions for (a) d = 2, (b) d = 3, and (c) d = 5. The orbital angular momentum basis modes for each state are shown above each density matrix. Isotropic mixed states with p = 0.4 for the same entangled states in (d) d = 2, (e) d = 3, and (f) d = 5.
So far, we see that the potential benefits of using high dimensional quantum structured light are coming to the fore. Next, we introduce measures that quantify the dimensions, information capacity, and purity of our quantum states.
III. KEY EQUATIONS FOR HIGH DIMENSIONAL STATES
Here, we outline some of the key equations that are used to qualitatively and quantitatively characterize high dimensional quantum states, ranging from the amount of information that can be packed into quantum states to the quality of the quantum states that are produced.
A. Information capacity
Suppose Alice and Bob wish to share information using a simple prepare and measure protocol by encoding photons with high dimensional structured modes. Each state, on its chosen basis, is associated with a message (symbol). Here, Alice prepares states on this basis, and Bob performs measurements on the same basis. We aim to quantify the amount of information that can be transmitted between the two parties.
Information capacity. (a) The scattering probability matrix for a d = 4 dimensional basis of spatial modes (shown as insets) above the diagonal. The matrix resembles an identity matrix indicating that all prepared states can be identified. The average fidelity is F = 1. (b) The mutual information (capacity) as a function of dimensions (d) in units of bits per photon. (c) Scattering probabilities for a given system that has high crosstalk (noisy). The fidelity for this case is F = 0.47. (d) Mutual information vs average fidelity for selected dimensions. Because errors can reduce channel fidelity, the information capacity of the quantum channel will also be affected.
Information capacity. (a) The scattering probability matrix for a d = 4 dimensional basis of spatial modes (shown as insets) above the diagonal. The matrix resembles an identity matrix indicating that all prepared states can be identified. The average fidelity is F = 1. (b) The mutual information (capacity) as a function of dimensions (d) in units of bits per photon. (c) Scattering probabilities for a given system that has high crosstalk (noisy). The fidelity for this case is F = 0.47. (d) Mutual information vs average fidelity for selected dimensions. Because errors can reduce channel fidelity, the information capacity of the quantum channel will also be affected.
Mutual information is a measure that is commonly employed to quantify the amount of information that can be transmitted between Alice and Bob. For a perfect channel, this is given by IAB = log2(d). In Fig. 4(b), we show the channel capacity as a function of the d dimensions. Provided that the channel is perfect, we can pack more information into the photons by using more spatial modes. For qubits (d = 2), the mutual information is restricted to 1 bit per photon and can be doubled by using d = 4 dimensional encoding. As the encoding dimensions are increased, the mutual information also increases.
B. Dimensionality
Subsection III A highlights an important point: to overcome noise, it makes more sense to use higher dimensional basis states. Moreover, it has been proven that even higher dimensional quantum entangled structured photons can be robust against noise.62 For this reason, it is imperative to have a consistent way of characterizing the dimensions of a quantum system.
For single photon channels, it should be easy to see that the dimensionality is set by the number of modes that can be transmitted through the channel and detected successfully while maintaining high mutual information. Further restrictions on d can be set by the apertures of the system as well as the angular spectrum and resolution of the generation and detection techniques;67 this means that modes requiring high spatial resolution will incur noise, so fewer modes can be used.
For the two photon case, we return to our decomposition of the OAM entangled photons in Eq. (10), where the coefficient for each Schmidt basis state is given by cℓ. The dimensions, d, of this state can be characterized using the Schmidt number witness,68 which can be estimated from projections with two mutually unbiased bases.28 For an entangled state with cℓ = , the Schmidt number is exactly d. This has a significant implication in high-dimensional entangled systems, as this quantifies the “amount” of entanglement in these systems.
C. Linear entropy
IV. CREATION AND DETECTION
In this part of the tutorial, we introduce the reader to techniques that are commonly used to produce, detect, and characterize high dimensional quantum structured light in the laboratory. We focus on the spatial degree of freedom of photons that are generated from SPDC. We subsequently discuss basic procedures for aligning a simple quantum experiment with SPDC photons and executing spatial projective measurements. We end the section by discussing methods for optimizing the generation and measurement steps.
A. Photon pair generation with spontaneous parametric down conversion
1. Spontaneous parametric down conversion from nonlinear crystals
One of the most commonly used approaches for generating quantum structured light uses crystals with a second-order susceptibility via three-wave mixing through a process called spontaneous parametric down-conversion (SPDC), and so we will use this as our source throughout this tutorial (for other possibilities, see recent reviews73,74). Examples of crystals that enable this process are beta-barium borate (BBO), lithium niobate (LN), potassium dihydrogen phosphate (KDP), and potassium titanyl phosphate (KTP).
Figure 5(a) shows a cartoon description of the SPDC process where a pump (p) beam of angular frequency, ωp, interacts with a non-linear crystal, resulting in the generation of two daughter photons of ωs and ωi, respectively. Due to the conservation of energy, the angular frequencies of said signal (s) and idler (i) are governed by the relation ωp = ωs + ωi. This relation is shown in the energy level diagram in Fig. 5(b). Accordingly, the entangled biphotons need not have the same energy (ωs = ωi), termed degenerate SPDC, but can be non-degenerate (ωs ≠ ωi).
Spontaneous parametric down-conversion. (a) Generation of SPDC photons from nonlinear crystals (NC) that are characterized by a second order nonlinear susceptibility tensor, χ(2). In this process, a single photon with a frequency, ωp, is absorbed by the crystal, and twin photons with frequencies, ωs and ωi, are produced. (b) Energy diagram for the three photons involved in SPDC, showing energy conservation. (c) Momentum conservation in the SPDC process. The momentum vectors of the emitted s and i photons must add up to those of the pump photon. The phase mismatch occurs in the longitudinal direction. The emission cones can be altered by changing the phase matching conditions. This can be achieved via (d) angle tuning in bulk crystals and (e) temperature tuning in periodically poled (PP-) crystals. (f) Far-field intensity profiles of SPDC photons showing non-collinear and collinear geometries where the entangled photons are found in the regions highlighted with dashed lines.
Spontaneous parametric down-conversion. (a) Generation of SPDC photons from nonlinear crystals (NC) that are characterized by a second order nonlinear susceptibility tensor, χ(2). In this process, a single photon with a frequency, ωp, is absorbed by the crystal, and twin photons with frequencies, ωs and ωi, are produced. (b) Energy diagram for the three photons involved in SPDC, showing energy conservation. (c) Momentum conservation in the SPDC process. The momentum vectors of the emitted s and i photons must add up to those of the pump photon. The phase mismatch occurs in the longitudinal direction. The emission cones can be altered by changing the phase matching conditions. This can be achieved via (d) angle tuning in bulk crystals and (e) temperature tuning in periodically poled (PP-) crystals. (f) Far-field intensity profiles of SPDC photons showing non-collinear and collinear geometries where the entangled photons are found in the regions highlighted with dashed lines.
Linear momentum is also conserved in the SPDC process; the momentum of the produced photons should add to that of the pump, resulting in kp = ks + ki, shown pictorially in Fig. 5(c), expressed using the wave vectors, k, of the respective pump, signal, and idler photons. If energy and momentum conservation are met, one has perfect phase matching. On the contrary, phase mismatch occurs and is proportional to Δkz, in the longitudinal momentum components, as illustrated in Fig. 5(c). In order to alter the phase-matching conditions, most bulk crystals require angle tuning, as shown in Fig. 5(d), where the angle of incidence of the pump mode with respect to the crystal is varied. Alternatively, one can use periodically poled crystals [shown in Fig. 5(e)] with engineered domain switching of a period of Λ.75,76 For these crystals, the phase matching conditions can be altered by changing the temperature of the crystal.
There are two types of phase-matching geometries that can be identified based on the matching of linear momentum. In Fig. 5(f), on the top panel, the output field propagates off-axis with respect to the pump, but with equal and opposite trajectories on either side; this is known as non-collinear SPDC. The emitted photons, in this case, are anti-correlated in transverse momentum and, therefore, the entangled photons are located on any two opposite ends of the SPDC ring. For collinear SPDC, shown in the bottom panel of Fig. 5(f), the output field wave-vectors are also anti-correlated, except when ks = ki = 0.
Due to the polarization sensitivity and phase-matching conditions, there are different phase-matching regimes that can be engineered. This results in different polarization pairings with active research considering exotic fabrications and crystal pairings, allowing one to engineer unique geometries and output states. A good review covering this can be found in Ref. 77. For this tutorial, we will consider the generic cases commonly available and summarized in Table I.
Phase matching types. Summary of various phase-matching regimes for second order non-linear interactions that differ for input and output field polarization with respect to parallel orientation to the crystal's ordinary (o) or extraordinary (e) axes.
Type . | Pump pol. . | Signal pol. . | Idler pol. . |
---|---|---|---|
0 | o | o | o |
I | e | o | o |
II | e | o | e |
Type . | Pump pol. . | Signal pol. . | Idler pol. . |
---|---|---|---|
0 | o | o | o |
I | e | o | o |
II | e | o | e |
The distinction between the different types relies on the relative polarizations with respect to the crystal's ordinary (o) and extraordinary (e) axes. The examples shown in Table I are special cases but can, in general, be summarized as follows: in type 0, the signal and idler photons carry the same polarization as the pump, whereas in type I, their polarization is orthogonal to that of the pump; for type II, however, the signal and idler photons are orthogonally polarized with respect to each other, leading to polarization entanglement. We note that these characteristics are all determined by the nonlinear electric tensor of the crystals.
2. Spatial mode entanglement from SPDC
In the SPDC process, photon pairs can be entangled in their energy-time,78,79 path,80 and transverse spatial degrees of freedom24,49,52,81–83 (see Ref. 84). In this tutorial, we focus on discrete spatial modes and thus consider the entanglement generated therein.
B. Alignment and back-projection
Typically, quantum optical setups can be aligned both forwards and backward (backprojection) in order to ensure both entangled photons sent through the system travel the desired path and coincide well with the detection system. Accordingly, we will first look at typical forward alignment strategies for different types of SPDC, which are illustrated in Fig. 6, before seeing how back-alignment works and, ultimately, how one may view the system for simulation.
Generic SPDC alignment strategies. Typical strategies for alignment of quantum systems with collinear and non-collinear SPDC sources and how they may differ for the degenerate and non-degenerate cases. Irises (black triangles) are typically placed in the beam paths for each arm in order to mark the path of the photons.
Generic SPDC alignment strategies. Typical strategies for alignment of quantum systems with collinear and non-collinear SPDC sources and how they may differ for the degenerate and non-degenerate cases. Irises (black triangles) are typically placed in the beam paths for each arm in order to mark the path of the photons.
The alignment of the SPDC depends both on the type of trajectory of the entangled photon pair one wishes to measure (collinear or non-collinear) as well as the distinguishing properties between them. The simplest configuration lies with collinear SPDC. As the linear momentum of the biphotons lies in the same direction as the pump, one may use the pump laser as a guide throughout the optical setup. Separation of the biphotons into the desired arms of the system can then be achieved by passing them through a 50:50 beamsplitter when they hold the same polarization and wavelength, as shown in the top illustration of Fig. 6. Alternatively, if the SPDC is non-degenerate, as in the second illustration in Fig. 6, one may employ a dichroic mirror (DM) to separate the wavelengths. In this case, the side coated for the reflected wavelength should be the incident face. When considering the pump beam as a guide, two parallel beams are traditionally reflected as a result of incidence on the front and back sides of the optic, as illustrated in the inset. If orientated correctly, the first reflection of the pump beam should indicate the path of the reflected photons. Depending on the cutoff wavelength between the reflected and transmitted light, the pump beam will be brighter in either transmission or reflection. For cases where the pump is not bright enough to see both the transmission and reflected intensities through the DM, another wavelength in the complementary wavelength band may be aligned with the pump beam beforehand, or one may consider relying on back alignment alone.
For non-collinear SPDC, the alignment varies appreciably as the biphotons no longer travel in the direction of the pump beam but rather in opposite directions to either side. This corresponds to selecting the photons on opposite ends of the cone shown in Fig. 5(f), which we see illustrated in the horizontal plane of the last illustration of Fig. 6. The pump may no longer act as a guide in this case. This results in the need to see the SPDC cone produced with a sensitive CCD and align through two apertures in each arm to opposite transverse sides of the cone for planes with an appreciable distance between them. These apertures then allow one to use back alignment in order to align the rest of the optical elements from these points.
For back alignment, one must consider the optical system in its entirety. For instance, considering the typical system in Fig. 7(a), after traversing the desired optical elements [such as the lenses and spatial light modulators (SLMs)], the biphotons are each focused into single mode fibers (SMF), which relay them to sensitive detectors like avalanche photodiodes (APD). Back alignment then considers this system in reverse. One can envision that if the photons can travel to the detector in the forward direction, light traveling in the opposite direction from the same point in the detector must travel the same path. As such, an additional laser is used where the light is sent back through the detection end of the fiber (replacing the detectors in the figure) such that it traverses the elements backward toward the crystal face. By using this backpropagating light, one may then align the couplers and optical elements as desired. Here, one may be aided by using the same wavelength as the SPDC to ensure the optical operations observed in back alignment correspond well and are not shifted due to the difference in wavelength.
Klyshko model for a quantum system. Demonstration of the Klyshko model for quantum systems where the non-linear crystal (NC) in the optical system (a) can be seen as a mirror (M) in the model (b). Experimentally, the setups are almost identical. In the quantum setup (a), a pump beam is incident on an NC, which produces our SPDC state consisting of entangled photon pairs, which are then spatially separated into the two arms of the experiment. The SPDC state is imaged onto the SLMs with lenses (L) and measured using single photon detectors (Det). In the Klyshko back-projection model, one of the detectors is replaced by a source, and a single beam is sent through the setup by replacing the NC with a mirror.
Klyshko model for a quantum system. Demonstration of the Klyshko model for quantum systems where the non-linear crystal (NC) in the optical system (a) can be seen as a mirror (M) in the model (b). Experimentally, the setups are almost identical. In the quantum setup (a), a pump beam is incident on an NC, which produces our SPDC state consisting of entangled photon pairs, which are then spatially separated into the two arms of the experiment. The SPDC state is imaged onto the SLMs with lenses (L) and measured using single photon detectors (Det). In the Klyshko back-projection model, one of the detectors is replaced by a source, and a single beam is sent through the setup by replacing the NC with a mirror.
When the forward alignment overlaps well with the back alignment, one should see the signal at the detector, allowing one to further improve the setup and take measurements. For the non-collinear geometry discussed earlier, forward alignment ends at the initial apertures. The alignment of the backpropagating laser through the initial aperture then sets the path such that alignment of the optics is performed with this laser as a guide instead of the pump. Fine tuning of the phase matching may then also be achieved by aligning the detection SMF back through the apertures used for the pump and checking that the maximal counts lies at the correct crystal parameters, be it temperature or angle.
An additional method exists that allows one to experimentally simulate and probe a quantum optical system using classical light. Here, when looking at the setup in a retrodictive manner, the quantum system at any point in time may simply be described by evolving the measured states backward in time.90 Based on a rigorous formulation, it was shown by Kyshko in 1988 that one may employ back-propagation to generate a classical analog that predicts the measured quantum correlations produced from the optical system being traversed by the biphotons. Known as the Klyshko advanced wave model, the nonlinear crystal is replaced with a mirror, and in one of the detection arms, the detector is replaced with a laser. This formulation is illustrated in Fig. 7(b). As can be seen, the light then passes through the system, bounces off a mirror positioned where the crystal was, and traverses the other arm in the forward direction before being detected in place of the correlated SPDC photon. The corresponding intensity measurements then reflect the coincidences that would be seen when detecting the biphotons in the equivalent quantum setup. For example, the measurements yield the conditional probability of detecting a photon in one arm, given that another photon is detected in the other arm.
Furthermore, the effect of the pump beam on the system can also be modeled by simply adjusting the properties of the mirror substituting for it. For instance, when the pump beam is not a plane wave in the crystal, the phase curvature is equivalent to the mirror being curved, or when the pump is angled, it equates to the mirror being tilted at the related angle.91–93
It may be noted that, in addition to the rigorous theoretical formulation, it has been experimentally verified, showing excellent agreement in comparison to the quantum outcomes.94,95 Consequently, this not only provides a good way to numerically model the system but also serves as a good probe in order to determine and correct the physical parameters that affect the desired system.
C. Coincidences
Detection of the entangled photons in a quantum experiment is traditionally achieved by distinguishing time correlations in the signals received by the detectors. As each of the entangled photons is “birthed” at the same time in the crystal, they should arrive at their respective detectors at exactly the same time (within the uncertainty principle) or after a set delay related to traveling different distances before reaching the detectors. This yields a way to distinguish them from other non-correlated detection events (photons that are not entangled) or noise, with a pair being detected in the appropriate interval being called a coincidence.
Practically, these can be detected using event timers that log the time of arrival for each photon signal and looking at the difference in the time delay between each photon detected in arm 1 and every photon detected in arm 2, or vice versa. Subsequently, these entangled photons are identified by taking a histogram of the detected photons where all the respective time differences (Δt) between the signals form values in discretized timebins (e.g., 40 ps) from each of the arms, as illustrated in Fig. 8. Due to the aforementioned time correlation, the entangled photons will be found to have the same time difference, and thus the counts or signal in that time bin of the histogram will increase, such as in the second time bin of the histogram. Conversely, stray or uncorrelated light that does not have a set correlation will be spread randomly across all the other bins and thus not have any specific signal buildup. This then gives a degree of noise that depends on the ratio of entangled photons being detected to stray or uncorrelated light. It may be noted that there is a degree of uncertainty in the processing and propagation times of devices like the event timer and APD. As a result, this time delay correlation will also “spill over” or spread into adjacent time delay bins of the histogram if the timebins are made on the order of the jitter in the detectors and instruments.
Coincidence detection. A typical biphoton quantum setup utilizes two detectors (APDs in this case) to detect individual photon events, and a timer observes such events over a certain time period. Correlations are then drawn between events in both channels according to delay time, Δtn. (b) For a series of events detected by channels 1 and 2 shown in the 1000 ps time window, events are (c) time-binned according to relative time delays between events in each channel. For example, if an event is detected at time t1 by channel 1, the time tagger compares this event to all other detected events in channel 2 in the same time window and orders events in timebins corresponding to the relative time delays between t1 and all other detected events in channel 2. This is performed for all detected events in channel 1. Once this process is completed, one observes a peak around the expected delay time between correlated photon pairs from the SPDC process.
Coincidence detection. A typical biphoton quantum setup utilizes two detectors (APDs in this case) to detect individual photon events, and a timer observes such events over a certain time period. Correlations are then drawn between events in both channels according to delay time, Δtn. (b) For a series of events detected by channels 1 and 2 shown in the 1000 ps time window, events are (c) time-binned according to relative time delays between events in each channel. For example, if an event is detected at time t1 by channel 1, the time tagger compares this event to all other detected events in channel 2 in the same time window and orders events in timebins corresponding to the relative time delays between t1 and all other detected events in channel 2. This is performed for all detected events in channel 1. Once this process is completed, one observes a peak around the expected delay time between correlated photon pairs from the SPDC process.
D. Noise
E. Detection and measurement of quantum structured light
1. Detection using projective measurements
High-dimensional states require a reliable method of detection where projections into some orthogonal basis are needed. If we consider the SPDC state given by Eq. (23), then a detection scheme is required that can distinguish between different OAM states. In fact, what we need is a coupled detection system that only accepts a single OAM state, |ℓDet⟩. For the example of OAM used in this tutorial, we use what is known as a spatial light modulator (SLM) coupled with an SMF. It is worth noting that this detection scheme works just as well for any chosen spatial basis, such as the Hermite–Gaussian, Ince–Gaussian, and pixel mode bases, to name but a few and, therefore, extends past OAM, making it an ideal scheme to be used to generate and detect spatial modes of single photons and measure correlations between entangled photons of high-dimensional quantum states.
The detection of spatial modes with SLMs borrows from methods that are commonly used for creating arbitrary light fields in beam shaping.103 In beam shaping an incoming light field, usually, the fundamental Gaussian mode (|0⟩ ≡ |LG0,0⟩) is modulated in phase and amplitude using an SLM. As for the detection, the reciprocal nature of light is employed, where one is able to “reverse” the modulation of a particular field back to the fundamental Gaussian state, which is subsequently coupled to a single mode fiber and a photon counting module.104,105 This approach flattens the field. The concept is illustrated in Fig. 9(a), where the fundamental mode is converted into a desired spatial pattern, Φ(r), and can be detected in reverse. Consequently, the fiber and the SLM form an optical correlation measurement that is analogous to performing an inner product measurement of the form |⟨Φ|Ψ⟩|2 for some input field Ψ(r) and some target (detection) mode Φ(r).105 Furthermore, to maximize coupling efficiency, the creation and detection beam sizes need to be comparable. As outlined, two elements are at play: (i) the detection mode that is encoded as a hologram , and (ii) the probability of the demodulated mode coupling into the single mode fiber. In this and the proceeding sections, we place emphasis on this coupled detection scheme, focusing on the different optical inner-product measurements that overlap the state |Ψ⟩ with the state |Φ⟩ using SLMs and fibers. Later, we extend this to entangled photons.
Detection scheme for high dimensional states. (a) The detection scheme for an input field Ψ(r) impinging on an SLM encoded with a hologram Φ*(r). The SMF then accepts only the LG0,0 component of the output field. Reciprocally, we see the usual SLM modulation scheme where the SMF now acts as our source producing the mode LG0,0, which is incident on the SLM encoded with Φ(r), thereby producing the desired mode (or, depending on the encoding scheme, a superposition of modes is created with the desired mode being among them). (b) Examples of encoded phase holograms (no grating) with encoded hologram functions taking the form H(Φ(r)) = exp(iℓϕ) for ℓ = 1, 3, −1. (c) Examples of complex-amplitude holograms (grating included) with encoded hologram functions taking the form H(Φ(r)) = |LGℓ,0|sin(arg(LGℓ,0) + Λgrating) for ℓ = 1, 3, −1. (d) Schematic of the detection scheme for a biphoton state |Ψ⟩AB.
Detection scheme for high dimensional states. (a) The detection scheme for an input field Ψ(r) impinging on an SLM encoded with a hologram Φ*(r). The SMF then accepts only the LG0,0 component of the output field. Reciprocally, we see the usual SLM modulation scheme where the SMF now acts as our source producing the mode LG0,0, which is incident on the SLM encoded with Φ(r), thereby producing the desired mode (or, depending on the encoding scheme, a superposition of modes is created with the desired mode being among them). (b) Examples of encoded phase holograms (no grating) with encoded hologram functions taking the form H(Φ(r)) = exp(iℓϕ) for ℓ = 1, 3, −1. (c) Examples of complex-amplitude holograms (grating included) with encoded hologram functions taking the form H(Φ(r)) = |LGℓ,0|sin(arg(LGℓ,0) + Λgrating) for ℓ = 1, 3, −1. (d) Schematic of the detection scheme for a biphoton state |Ψ⟩AB.
Spatial light modulators typically come in two forms. One is a phase-only modulator, and the other is an amplitude-only modulator. With a few clever tricks, both of these devices can be used to manipulate or modulate the amplitude as well as the phase profiles of your input photons. In this tutorial, we will focus on phase-only modulation devices. For more information on using amplitude-only devices, see Ref. 106.
We can consider the action of the SLM on the input state, |Ψ⟩, to be given by a phase modulation function , commonly called the hologram. As will be seen, by incorporating a grating function, one is able to encode the amplitude in addition to the phase of the optical field with phase only modulation.107 Accordingly, this means that the hologram has a phase transmittance function, .107
The other modes in Eq. (28) can also be attributed to artifacts of the modulation technique used on the SLM. Ultimately, regardless of the technique used, modes besides the fundamental mode are filtered out when passing through an SMF. The probability of obtaining the fundamental mode is given by η ∝|⟨Φ|Ψ⟩|2. How can we encode the holograms as a function that achieves this mapping?
Considering that we have control over the individual LCs of the SLM and that we can tailor the phase of arbitrary fields, the next step is to consider how we may use it to map the amplitude information, f(Φ(r)), if we wish to measure complete amplitude and phase. The simplest approach is to modulate the blaze grating with the field amplitude [see example holograms in Fig. 9(c)]. This approach might be efficient but may come at a loss of accuracy since the field amplitude does not map linearly with conversion efficiency. Alternative methods can be found in Refs. 107, 109, and 110. Furthermore, we direct the reader to Ref. 111 for a comparison of commonly used techniques. One relation for f(·) that can produce the best outcome was reported in Ref. 107, where the amplitude function is encoded as, . Here, is the Bessel function of the first kind. The phase function is altered as .
With these phase modulation techniques, it is possible to prepare complex amplitude holograms that accurately encode (or reciprocally detect) the target mode Φ(r). Next, we unpack the final detection step, where the fiber is used as a tool to achieve the mapping in Eq. (28) in order to filter out the desired fundamental mode once the hologram has modulated the photon field.
2. Influence of optical fibers on the measurements
In Fig. 10(a), we show the detection of OAM modes from SPDC, where each photon is projected into OAM modes ΦA,B(r) ∼ exp(iℓsiϕ) with corresponding charges of ℓs,i for photons A (signal) and B (idler), respectively. Figure 10(b) shows the corresponding spectrum for ℓs = −ℓi taken from Fig. 10(a). This configuration produces a fixed number of modes . The number of generated and detected modes can be affected by the source parameters (input mode and crystal parameters) and the quality of the detection system. Next, we show how the source and detection systems can be optimized to achieve such higher dimensional encoding.
SPDC generational OAM bandwidth. Example of measured (a) SPDC OAM coincidence rate (counts per second) for different OAM projections on the signal (ℓs) and idler (ℓi) photons, showing anti-correlations in OAM. (b) Coincidences taken for ℓs = −ℓi. From this plot, the spiral bandwidth (dimensionality), K = 21, was measured. (c) Spiral bandwidth optimization as a function of the pump beam waist size wp and the non-linear crystal length L computed from Keff in Eq. (34).
SPDC generational OAM bandwidth. Example of measured (a) SPDC OAM coincidence rate (counts per second) for different OAM projections on the signal (ℓs) and idler (ℓi) photons, showing anti-correlations in OAM. (b) Coincidences taken for ℓs = −ℓi. From this plot, the spiral bandwidth (dimensionality), K = 21, was measured. (c) Spiral bandwidth optimization as a function of the pump beam waist size wp and the non-linear crystal length L computed from Keff in Eq. (34).
3. Optimizing the measurement of quantum structured light
While the phase flattening measurement techniques we established earlier work as a highly sensitive single photon detection system, the nature of the SMF results in an additional modal modulation with the Gaussian distribution describing the coupling of the flattened light into the fiber.71,105 As a result, the measured bandwidth needs to account for the SPDC state and the overlap thereof with both of the chosen detection modes for either photon.
SPDC measurable OAM bandwidth. Example of measurable bandwidth (K ≡ Keff) as a function of the pump and detection beam sizes for a (a) L = 5 mm crystal and a (b) L = 1 mm crystal. The dotted line indicates the optimal ratio for maximum bandwidth between the detection and signal beam waists.
SPDC measurable OAM bandwidth. Example of measurable bandwidth (K ≡ Keff) as a function of the pump and detection beam sizes for a (a) L = 5 mm crystal and a (b) L = 1 mm crystal. The dotted line indicates the optimal ratio for maximum bandwidth between the detection and signal beam waists.
F. Other special elements for generation and detection
Besides the modulation and projective measurement approaches that were discussed earlier, spatial modes can be created and measured using a variety of other techniques and tools. Here, we explore a few that are commonly used for quantum structured light. We focus on components that couple the polarization and spatial components of photons since they have been instrumental in demonstrating quantum walks,115 quantum cryptography,21 and creating hybrid entangled states.116 Furthermore, we also explore the use of refractive elements for mode sorting since they show the capability of measuring high dimensional quantum structured light.117–120
1. Dynamic and geometric phase control of photons
In an effort to gain control over and manipulate the spatial structures of photons, significant efforts have been made toward developing additional elements in addition to the vanilla spatial light modulators. This, in an effort to allow specialized control in the manipulation and subsequent use of quantum light, further extends the toolbox for high dimensional applications. Here we consider three significant devices based on OAM, which, since their demonstration, have facilitated interesting applications for using structured quantum light and hold further potential.
An alternative principle also lends itself to the generation of phases, but instead by using the “memory” of the transformations applied. Here, by altering parameters adiabatically in a closed-loop fashion, a geometric phase122 may be induced, so-called as the amount of imparted phase is determined by half the solid area bounded by the transformations in the parameter space. Applied to light, one may use a change in polarization to induce such a relative phase in the electric field, as illustrated in Fig. 12(a), which is otherwise known as the Pancharatnam–Berry phase (after the authors), applying the concept to the classical123,124 and quantum125 light, respectively.
The concept of geometric phase. (a) Illustration of a diagonally polarized state traversing a series of waveplates that take the state through a closed loop transformation, which is (b) depicted on the Poincaré sphere. (c) Depiction of an element with spatially varying optical indices of refraction that may be considered a series of HWPs with varying optical axes. (d) Illustration of associated paths taken by different rays of a light beam traversing an element with four arbitrary HWPs with varying fast axis rotational angles.
The concept of geometric phase. (a) Illustration of a diagonally polarized state traversing a series of waveplates that take the state through a closed loop transformation, which is (b) depicted on the Poincaré sphere. (c) Depiction of an element with spatially varying optical indices of refraction that may be considered a series of HWPs with varying optical axes. (d) Illustration of associated paths taken by different rays of a light beam traversing an element with four arbitrary HWPs with varying fast axis rotational angles.
To see this effect, we may consider the polarization parameter space, which is illustrated on the polarization equivalent of the Bloch sphere in Fig. 12(b), known as a Poincaré sphere. Here, linear polarizations are placed along the equator and circular polarizations at the poles. The “closed loop” evolution of an input photon or light beam can then be followed by placing a series of waveplates, as illustrated in (a). When diagonally polarized light is passed through a half-waveplate (HWP) orientated at 22.5°, it becomes horizontally polarized. On the Poincare sphere (b), this is seen by the change in position from |D⟩ to |H⟩ through path A. Passing the light through a quarter-waveplate (QWP) at 45° then alters the state to |R⟩, i.e., following path B and with another QWP at 90°, the |R⟩ state is changed back to |D⟩ through following path C and a closed loop has been formed on the Poincare sphere. The light at this stage is not just diagonally polarized but also has a supplementary phase of ϕG = eiΩ/2, where Ω is the area enclosed by the evolution.
We may now consider applying this physical principle from the time (a) to the spatial126 (c) domain. Here, instead of evolution through waveplates in time, we consider evolving small waveplate elements [α(r, ϕ)] across the transverse spatial coordinates, (r, ϕ), of an optic as depicted in (c). Here, if a change in polarization occurs at points across the transverse spatial plane of the incoming beam, an associated spatially varying geometric phase is also generated. This is illustrated in (d), showing how geometric phase may be mapped between right-circularly polarized (RCP) waves traversing four arbitrary points [denoted α0(r, ϕ) to α3(r, ϕ)] on an optic (c), which has effective HWP elements with different orientations at each transverse coordinate. With the polarization change occurring spatially, the “closed path” is formed by the difference in the path between the other elements, resulting in the geometric phase being both relative and spatially varying.126,127 One may, therefore, engineer the relative optical axis orientations to generate any number of variable geometric phase acquisitions that may be used to manipulate the spatial mode of the overall beam.
Here, ϕ refers to the efficiency of QP, which is tied to the retardance of the slow and fast axes in the element. The cosine term alongside the identity matrix indicates that this portion of the incident beam remains unaffected where no polarization or phase changes occur. The second matrix term refers to the action of the QP whereby an azimuthally-varying geometric phase is imparted as described by Eq. (37). It may thus be seen that when φ = π (HWP retardance), the QP is 100% efficient, with the first matrix term falling away.129,130 Q-plates meeting this condition are referred to as “tuned.”
This effective twisting of the light beam produced by the geometric phase has additional implications in the physical interpretation, whereby the CP polarization may also be seen in terms of spin angular momentum (SAM). Here, when RCP is incident on the QP, an OAM of 2qℏ per photon is generated, and the flip in CP corresponds to a flip in SAM from 1ℏ per photon to −1ℏ. It is well known that the transference of SAM and OAM can occur between light and certain matter.130 Here, SAM interaction occurs in optically anisotropic media such as birefringent material and OAM in transparent inhomogeneous, isotropic media.128 The combination of a thin birefringent (liquid crystal) plate with an inhomogeneous optical axis in the QP subsequently results in the element coupling these two forms of angular momentum such that flipping in the SAM may be seen to generate OAM, making the QP a spin-to-orbital angular momentum converter (STOC) where the symmetry of the optical axis patterning effects the conversion values.130
QPs have been instrumental in realizing practical applications of quantum structured light that interface the spin and orbital components of photons for quantum communication in free-space,21,131,132 underwater,133 between satellites,134 and through optical fiber,135,136 for quantum memories,132 computing,115,137–139 quantum metrology,140 and for engineering novel quantum entangled states such as entangled vectorial fields,141 hybrid entangled,116,142,143 and hyper-entangled144,145 quantum channels. In most of these applications, the QP is used as the main component for encoding quantum information in quantum structured light.
These devices thus allow one to arbitrarily map between polarizations and orbital angular momentum, allowing one full control of the full angular momentum (J) domain and, as such, have been denoted J-Plates. Furthermore, the versatility of engineering the material properties allows one to extend past the OAM domain and map polarization to the pixels domain (images) such that one polarization excites an image like a cat and the other a dog,155 or, conversely, different OAM states exciting different holographic images,156,157 and the HWP dependence changed to that of a QWP158 and structuring the total angular momentum as it propagates.159 Good reviews tracking the burgeoning progress of these and other metasurface-based devices can be found in Refs. 146 and 160. With such capabilities, these metasurface devices have also found possibilities in quantum applications. A good review considering this is Ref. 161, where applications ranging from state creation to manipulation and weak measurements are considered.
2. Refractive mode sorters
Therefore, we have explored detection by the single generation of conjugate projections in order to detect OAM. Alternatively, the application of geometric transformation concepts such as coordinate transformations through the use of optical systems can exact a desired transformation in light.162 This is the basic principle behind mode sorting and, therefore, detecting OAM. Here, the technique takes advantage of the circular geometry associated with OAM so that a geometrical mapping translates circular to rectangular geometry,163 as illustrated in Fig. 13(a). The resultant phase distribution “unwrapping” causes OAM to be transformed into transverse momentum with a linear phase gradient,120 as demonstrated in the (u, v) coordinate space of the figure.
Mode sorting with geometrical phase transformation. The (a) illustration of a conformal mapping that “unwraps” an OAM = 2 mode to a transverse phase gradient and the (b) color map illustration of the phase gradient resulting from the OAM geometric transformation for l = −1 (left) and l = 2 (right). (c) Depiction of the sorting action performed by a Fourier transforming lens after the phase correction element.
Mode sorting with geometrical phase transformation. The (a) illustration of a conformal mapping that “unwraps” an OAM = 2 mode to a transverse phase gradient and the (b) color map illustration of the phase gradient resulting from the OAM geometric transformation for l = −1 (left) and l = 2 (right). (c) Depiction of the sorting action performed by a Fourier transforming lens after the phase correction element.
The resultant phase distribution “unwrapping” causes the OAM to become transverse momentum with a now linear phase gradient of across the beam length.118,119 As the “unwrapped” mode contains a phase gradient limited to the length, d, all OAM modes result in a transverse phase gradient that is integer multiples of each other, as shown in Fig. 13(b).
Moreover, the intensity of the spot indicates the “amount” of any OAM mode present. The mode sorter technique employed with refractive elements allows for the efficient detection of a large range of OAM modes and the associated weightings, enabling the detection of low intensity sources in comparison to other techniques such as SLM projections. For instance, sorting as many as 50 states was demonstrated.119
It may be noted that, as a result of the transformative action, such devices are more sensitive to alignment than devices such as SLMs or QPs. They do, however, offer the advantage of deterministic detection, as the transformation always sends the photons to a detectable state. This is in contrast to single-outcome projective measurements such as those performed by the SLM, where the photons that do not collapse into the mode the SLM is interrogating for are discarded, making the detection probabilistic. Accordingly, this detection has been used from detecting vector vortex states deterministically117 to high-dimensional quantum cryptography,36 efficient detection for a classical quantum walk resonator,139 and interfacing between the path and OAM entanglement.164
Finally, we note that there is a wide range of specialized optics and extensions thereof, such as g-plates, proceeding from liquid crystal patterning like QPs to allow for topological quantum simulations165 and linear photonic gears.166 Mode sorters have been extended to include radial orders167 and other spatial modes such as Hermite–Gaussian modes168 and have taken on more complex forms with multiple conversion planes169 even extending into fibers.170 Metasurfaces, furthermore, have a very wide reach from multi-wavelength lenses171–173 to nonlinear engineering.174–176 Here, we selected a few special optical elements that have made a basic impact on structured quantum light.
V. HIGH DIMENSIONAL STATE CHARACTERIZATION
A. 2D Bell curves and Bell violations
In most quantum entanglement experiments, it is common practice to demonstrate a violation of the Clauser-Horne-Shimony-Holt (CHSH) Bell inequality.177 The violation of the inequality confirms non-locality in quantum experiments. Following the initial demonstration with the polarization states of photons,44 today the Bell-inequality violation is used in a myriad of experiments as a characterization tool. Here we show how it can be performed on an OAM basis. We direct the reader to Refs. 24 and 23 for further reading.
Bell-inequality violation. (a) An experiment for demonstrating the Bell inequality violation using structured light. Here, SPDC photons are generated from nonlinear crystal (NC) and propagated to spatial light modulators (SLM), where spatial projections onto the states |θA(B)⟩ are performed. θA(B) corresponds to the physical rotation of the encoded hologram. Correlation measurements between photons A and B for the subspaces (b) ℓ = 1 and (c) ℓ = 2. Using these correlations, the Bell parameter can be evaluated.
Bell-inequality violation. (a) An experiment for demonstrating the Bell inequality violation using structured light. Here, SPDC photons are generated from nonlinear crystal (NC) and propagated to spatial light modulators (SLM), where spatial projections onto the states |θA(B)⟩ are performed. θA(B) corresponds to the physical rotation of the encoded hologram. Correlation measurements between photons A and B for the subspaces (b) ℓ = 1 and (c) ℓ = 2. Using these correlations, the Bell parameter can be evaluated.
While our demonstration above shows that the Bell-inequality test can be used to confirm entanglement in multiple subspaces of a high dimensional entangled state, it does not confirm the entanglement between all the subspaces. In order to confirm that a higher dimensional state possesses entanglement over all the dimensions, generalized Bell inequalities have to be employed. One such version was introduced by Collins et al.,180 sometimes called the CGLMP inequality, and has been instrumental in demonstrating high dimensional Bell violations ranging from OAM qutrits181,182 and reaching up to 12 dimensions25 of entangled OAM states. Furthermore, an intriguing aspect of the CGLMP inequality is that higher dimensional states can obtain violations above the Tsirelson bound.
B. Quantum state tomography
The concept of tomography relies on the idea of projecting a quantum state onto observable basis states and measuring the probability that the particle is in the state. One then works backward in order to determine what state would result in the outcomes measured. Tutorial references covering this topic may be found in Refs. 23 and 183 with a more brief overview tailored to this work detailed here. The concept is illustrated in Fig. 15(a). Here the projective measurements equate to using a light source to project the shadow of the object of interest onto a plane such as those normal to the x-, y-, and z-axes shown in the top panel of Fig. 15(a). The shapes and dimensions of the projected profiles then allow one to reconstruct what the object was as indicated in the bottom panel of Fig. 15(a). It should be noted that here only 3 projective measurements were used, which assumes a fairly simple and symmetric object. For more complex objects or measurements that introduce uncertainties, a larger number of projections onto different sets of planes can allow for a more accurate reconstruction.
State reconstruction. (a) Concept of quantum state tomography, where projections are made into states (“planes”) that reveal the composition of the state (“shape”). The projections can be subsequently combined to reproduce the state. Qubit state (d = 2), (b) tomographic projections onto the ℓ = ±1 subspace, and corresponding (c) reconstructed state. For two particle qubits, up to 36 measurements are performed. In higher dimensions, more projections are needed. (d) Tomography of a d = 3 entangled state where the projections are performed on smaller qubit spaces at a time. (e) Corresponding reconstructed density matrix for the qutrit entangled state.
State reconstruction. (a) Concept of quantum state tomography, where projections are made into states (“planes”) that reveal the composition of the state (“shape”). The projections can be subsequently combined to reproduce the state. Qubit state (d = 2), (b) tomographic projections onto the ℓ = ±1 subspace, and corresponding (c) reconstructed state. For two particle qubits, up to 36 measurements are performed. In higher dimensions, more projections are needed. (d) Tomography of a d = 3 entangled state where the projections are performed on smaller qubit spaces at a time. (e) Corresponding reconstructed density matrix for the qutrit entangled state.
Quantum mechanically, we can carry out this reconstructive concept by performing various operations in order to manipulate the state and thus characterize it. To do so, we apply a variety of projections on many copies of the quantum system, and with the subsequent measurements, information on the quantum state being interrogated is built up, analogous to the projections of the object shown in Fig. 15(a). The question now arises: what are the optimal projections necessary in order to accurately determine the quantum state?
Several types of approaches to such tomographic measurements have been put forward that allow one to extract the information needed to reconstruct the state. Here, these range from generalized Bell tests25,181,182 to using mutually unbiased bases26,28,184 or incorporating self-guided approaches,185,186 with each approach having certain merits. One may also ask, how many measurements are enough to accurately reconstruct the state using the chosen method? The answer to this lies in the uncertainties in the measurements being made. Initially, one may consider the projections in the concept figure. Here, only three projective measurements are actually necessary in order to find the object. These would then form a tomographically complete set. If there is any uncertainty or “blurriness” in the projection, the actual size and perhaps the fine structural details of the outlines may be in question. To improve this, one may consider making additional projections on the object where the planes are rotated to some degree. While these projections may not provide additional information, they will allow the reconstructed object to be checked against them and adjusted as necessary. This results in the “blurriness” or uncertainty being reduced. This then forms an overcomplete set of measurements but allows a greater degree of accuracy in the reconstruction.187,188
The aforementioned method can be extended to higher dimensions (d > 2) by taking the higher dimensional state and projecting it onto smaller qubit spaces.26 The localized measurements for each particle are then spread across combinations of two-dimensional subspaces within the d-dimensional system. This naturally results in far more measurements as each two-dimensional subspace combination in the d-dimensions of each particle needs to be measured along with all the different MUBs therein. In Fig. 15(d), we show the measurements for |ℓj⟩ ∈ {|−1⟩, |0⟩, |1⟩} basis states, where projections are performed on the OAM eigenstates, as well as the corresponding superpositions, , with the ℓ1 < ℓ2. As such, the number of measurements scale as with d dimensions per particle for a two-particle system, where is a binomial coefficient.
C. Fidelity and purity
To characterize the quality of the generated states, we can use some of the measures that were introduced in Sec. III. For example, we can characterize how similar the measured density matrix is to our desired one by using the state fidelity in Eq. (12), but rewritten here as , where ρ is the measured density matrix and ρT is the target density matrix. The fidelity evaluates to 0 for distinct (non-equivalent) density matrices and 1 for identical density matrices. Here, the density matrices we reconstructed were compared to maximally entangled states in the same dimensions, i.e., |1⟩|−1⟩ + |−1⟩|1⟩ for d = 2 and |1⟩|−1⟩ + |0⟩|0⟩ + |−1⟩|1⟩ for d = 3, corresponding to the density matrices in Figs. 15(c) and 15(d), respectively. The fidelity was measured as f = 0.98 ± 0.01 and f = 0.92 ± 0.01 for our d = 2 and d = 3 dimensional states, respectively, showing that the reconstructed states are similar to the maximally entangled states we desired.
Furthermore, the degree of the mixture in the states can be quantified from Eq. (15) via the linear entropy (SL) or Von Neumann entropy. Accordingly, the purity of the state can be computed from 1 − SL, producing values equal to 1/d for mixed states and 1 for pure states. For our measured density matrices in Figs. 15(c) and 15(d), the linear entropies are SL = 0.01 ± 0.02 and SL = 0.11 ± 0.07 for d = 2 and d = 3, respectively. The corresponding purities are thus 1 − SL = 0.99 ± 0.02 and 1 − SL = 0.89 ± 0.07, showing that the states are similar to maximally entangled states and possess a high level of purity.
D. Toward fast and efficient witness and quantum measures
In the techniques, we have presented so far, we chose the dimensions of the state we wanted to probe (in the case of the Bell measurement), or we needed to reconstruct the full density matrix (tomography) of the system in order to extract the purity or fidelity of the state. However, the state reconstruction can be a tedious process because the measurement complexity can scale to d4 for quantum systems where each photon occupies d dimensions.
The benefit of having access to such measurement techniques is that key information needed for quantum information processing protocols, such as the information capacity of quantum states and the allowed error bounds in secure communication systems, can be estimated quickly without wasting resources and increasing measurement times.
VI. EXAMPLE APPLICATIONS
A. Quantum key distribution
1. High dimensional BB84 protocol
One of the emerging key applications of structured light is in the area of quantum key distribution (QKD). Protocols such as the BB84 (prepare and measure)46 or the E91 (entanglement based)198 protocols were designed to replace computationally difficult problems to maintain secrecy in communication channels. This is because the rise of quantum computing algorithms might compromise traditional key generation techniques that may rely on difficult problems such as factoring.199 QKD is superior in that it is provably secure thanks to the uncertainty principle and no-cloning theorem,200 making the QKD protocols robust against an eavesdropper that is armed with unlimited resources.
Here, we will mainly focus on the prepare measure protocol, where photons are encoded and transmitted by Alice and detected at the receiver by Bob. Following the implementation of these protocols with polarization qubits, it was later realized that polarized photons could only transport 1 bit of information and, therefore, more efforts were focused on generalizing the protocol to higher dimensions. Once the theoretical framework was in place,66 adaptations of the protocol were implemented beyond qubit encoding using alternative degrees of freedom. Several experiments were reported21,35,36,131 that made use of structured light patterns encoded with OAM. We will describe, in a tutorial style, how the prepare measure protocol is executed with quantum structured light.
Using the OAM basis, Alice and Bob can agree on a set of basis modes, |Ψj⟩ ∈ {|ℓ1⟩, |ℓ2⟩, …, |ℓd⟩}, that have d independent states. Subsequently, they select a second set of basis modes |Φj⟩, that are mutually unbiased to the standard basis. The two bases must satisfy the relation |⟨Ψ|Φ⟩|2 = 1/d. For a given d dimensional basis, there are at most d + 1 mutually unbiased bases to choose from.197
Quantum key distribution with structured light. (a) Alice prepares a state on a randomly chosen basis, and Bob can measure the said state on a basis that he chooses randomly as well. (b) The states they choose must be selected from a pair of mutually unbiased bases {|Ψ⟩j}, {|Φ⟩j}, which in this example, we have chosen to be comprised of orthogonal spatial states |Ψ⟩ ∈ {|−1⟩, |0⟩, |1⟩} and superposition states . (c) The overlap matrix shows orthogonality between states on each basis but the non-zero overlap between pairs of states taken from both bases. (d) The key generation scheme starts with Alice preparing a state on a randomly chosen basis and Bob performing a measurement on another randomly chosen basis. They then sift through all the measurements and discard all those in which the preparation and measurement bases were not correlated. The remaining measurements they keep as their key.
Quantum key distribution with structured light. (a) Alice prepares a state on a randomly chosen basis, and Bob can measure the said state on a basis that he chooses randomly as well. (b) The states they choose must be selected from a pair of mutually unbiased bases {|Ψ⟩j}, {|Φ⟩j}, which in this example, we have chosen to be comprised of orthogonal spatial states |Ψ⟩ ∈ {|−1⟩, |0⟩, |1⟩} and superposition states . (c) The overlap matrix shows orthogonality between states on each basis but the non-zero overlap between pairs of states taken from both bases. (d) The key generation scheme starts with Alice preparing a state on a randomly chosen basis and Bob performing a measurement on another randomly chosen basis. They then sift through all the measurements and discard all those in which the preparation and measurement bases were not correlated. The remaining measurements they keep as their key.
2. Security analysis
Secure key rate. Secure key rates (bits per photon) for higher dimensional QKD as a function of the error rates (Q) for several d-dimensional encoding bases.
Secure key rate. Secure key rates (bits per photon) for higher dimensional QKD as a function of the error rates (Q) for several d-dimensional encoding bases.
To realize the full potential of using higher dimensional QKD with spatial modes, several improvements are needed for current implementations. This includes having more efficient photon sources, efficient and fast modulation techniques for spatial mode encoding, and detection methods that are fast, deterministic, lossless, and resilient to perturbations in complex channels (see references focusing on turbulence203 and optical fiber204). One main challenge is improving the measurement step due to increasing measurement complexity for higher dimensional modes. For example, because 2 × d detectors are needed for d dimensional states, filter based techniques, i.e., measuring one mode at a time, are commonly used to reduce costs as well as avoid building complicated setups that require many optical elements. However, this reduces the performance by a factor of 1/d. To overcome this limitation, spatial mode sorting using two refractive optical elements36,117,118,120,205 is being incorporated into the detection step (see the section on special optics, Sec. IV f). These techniques map the desired modes onto spots that can be easily measured with detector arrays. Recently, multi-plane light converters (MPLCs) applied to classical structured light detection169 are showing promise with recent demonstrations extending to the quantum regime.206
B. Teleportation
A building block holding potential for developing quantum technology is quantum teleportation. Having received much attention since the seminal paper by Bennett et al.,207 the protocol derives from the idea of transporting an entity between two distant points without it traveling or existing in the connecting distance. In this technique, the entities in question are states, which have been represented in many different systems from nuclear magnetic resonance208 to atomic ensembles,209–212 solid state systems,213–216 and trapped atoms217–221 along with photons, where degrees of freedom such as time,222–224 path,41,42 polarization,47,225–230 or structured patterns20 have been exploited.
Concept of spatial teleportation. Initially, a pair of entangled photons is shared between Alice (particle C) and Bob (particle B), establishing an entanglement channel. A third particle A containing an unknown or prepared state to be teleported is then mixed with Alice’s particle. Alice then performs a measurement in an entanglement or “Bell” basis of A and C. This Bell projection collapses the entanglement between particles A and B while simultaneously entangling photons A and C as well as destroying the state of photon C. Alice then transmits classical information to Bob about the Bell measurement projection (j), allowing him to correct for any unitary rotation on the state being teleported (Uj) and realizing teleportation. Physical mixing of A and C may be realized through either a beam-splitter (BS) or a non-linear crystal (NC).
Concept of spatial teleportation. Initially, a pair of entangled photons is shared between Alice (particle C) and Bob (particle B), establishing an entanglement channel. A third particle A containing an unknown or prepared state to be teleported is then mixed with Alice’s particle. Alice then performs a measurement in an entanglement or “Bell” basis of A and C. This Bell projection collapses the entanglement between particles A and B while simultaneously entangling photons A and C as well as destroying the state of photon C. Alice then transmits classical information to Bob about the Bell measurement projection (j), allowing him to correct for any unitary rotation on the state being teleported (Uj) and realizing teleportation. Physical mixing of A and C may be realized through either a beam-splitter (BS) or a non-linear crystal (NC).
Before the Bell measurement, photons A and C could be thought of as being in a superposition of every possible entangled combination. For each possible entangled state, Bob’s photon has a form of the carrier’s state, each with a unitary phase variation related to each entangled state that A and C might realize. A measurement of photons A and C on this basis then forces them to realize one of the entangled states. As a result, Bob has the teleported state within the unitary rotation resulting from the projective measurement. To retrieve the state, ρ, Alice then classically communicates the outcome of the Bell detection, j, to Bob such that the unitary correction, Pj, can be applied, and the state has thus been successfully teleported to Bob. The required classical communication then results in excluding the protocol from communication faster than the speed of light. Consequently, by mixing one of the entangled particles with that of a secondary entity carrying an unknown state and projecting them onto an entangled superposition, the unknown state can be conveyed without knowing or measuring it. As there is no direct observation of the state being sent, the superposition, as well as all the correlations, are maintained, preserving both the anonymity and any quantum nature.
As a result, teleportation can be exploited through a wide range of applications. For instance, it is an active component in the development of quantum information science,231,235–237 with the conceptual scheme forming a fundamental step in formal quantum information theories, and its physical process forms a basic building block toward the development of many quantum technologies, allowing information exchange protected against eavesdropping.238 Quantum technologies such as quantum repeaters,239 measurement-based quantum computing240 as well as quantum gate teleportation241 derive from the quantum teleportation protocol, and it may be extended to the idea of a quantum network.242
Significant focus and progress have been given to extending the distance and fidelities243 at which the protocols may be executed across both fiber222,224,225,244,245 and free-space229,230,246 with efforts extending to a low orbit satellite.247 These advances bring forward the practicality of implementing large quantum networks, which can then form a quantum internet.242,248,249 Only recently, however, has progress been made toward physically increasing the dimensionality of the states that can be sent using this protocol. This requires utilizing a basis of maximally entangled states and a unitary operator basis, {Uj}, whereby 250 so that the protocol can be extended beyond 2 dimensions and designed for any finite-dimensional Hilbert space of dimension, d. Here, the protocol remains similar to that of the 2-dimensional case, with Pj changed to Uj and d2 possible outcomes to be communicated classically.
While theoretical protocols and proposals have been suggested,250,251 this has proven difficult to implement physically, not from a lack of available properties to use as high dimensional states but rather from difficulty in mixing the carrier and entangled particle such that high-dimensional Bell projections can be made. Consequently, the limiting factor has been the black box in Fig. 18 to which two emerging solutions have been recently realized.
The first retains the traditional linear mixing tool where the two inputs are incident on different ports of a beam-splitter, and correlation measurements at the outputs allow one to select out a Bell state and thus render teleportation. Here, Refs. 41 and 42 exploit the path degree of freedom to realize three-dimensional teleportation, marking the first demonstrations to break the high-dimensional barrier. Previously, this was not achieved as linear optics is not able to distinguish d > 2 Bell states alone.252 With the introduction of additional ancillary photons to the projection measurements, however, the degeneracy may be broken.252,253 Refs. 41 and 42 subsequently constructed setups to achieve this with a scaling of d − 2 additional single photons and log2(d) − 1 additional pair, respectively. The demonstrations here remained in three dimensions, however, as this quickly becomes resource intensive and practically complex as d increases. It may be further noted that an intrinsic bound on efficiency exists with using a beam-splitter, as not all the Bell states may be detected and thus projected onto. Nevertheless, the schemes represent implementable high-dimensional teleportation of quantum states.
The second approach instead utilizes non-linear principles87,254,255 to mix the carrier and channel photon. Here, instead of the linear element, a non-linear crystal is used in reverse.43,256 Where before, one pump photon produced two entangled daughter photons, now two inputs (the carrier and channel photons) mix in the crystal such that an up-converted photon shares the properties and energy of the inputs. The upconverted photon thus forms a superposition of correlations that can be the projected onto the crystal by measurement and thus realize the necessary Bell projection. This can be understood using OAM as an example, with the projection of the upconverted photon onto the ℓBell = 0 state. From Eq. (23), we know that a pump photon of ℓp = 0 generates ℓB = −ℓC. When using the process in reverse, the same conservation rules apply, such that projection onto ℓBell = 0 results in coincidences only when ℓA = −ℓB = ℓC. It follows then that for the coincidences detected, Bob’s photon must have the same state as A. Projection onto a different ℓBell then only requires a unitary ℓ-related rotation on Bob’s side, as was demonstrated in Ref. 256.
Accordingly, teleportation in dimensions significantly exceeding three was demonstrated.43,256 Interestingly, this approach holds versatility in that the system supports spatial modes as a whole, such that various bases could be used, ranging from OAM states and arbitrary superpositions thereof43 to pixels states256 without adjustment. A point to note here, however, is that due to the low efficiencies associated with the non-linear process, the carrier is forced to be in a bright coherent state, such that enough single photons are present that the probability of upconvertion is large enough. As such, the nature of the carrier is classical until significant improvements in efficiency result, leading to this being called stimulated teleportation. A main consequence is that the teleportation technique is now limited in that correlations cannot be teleported, as this is a feature of quantum carriers.257 Nonetheless, it has facilitated a significant development in that high-dimensional teleportation exceeds three dimensions several times over as well as offering on-demand teleportation across the spatial mode basis with potential toward quantum entangled states with technological development.
C. Quantum imaging
1. Ghost imaging protocol with structured pixel modes
It is known that the spatial correlations between signal and idler photon pairs produced by SPDC can be used within a quantum imaging system.91,258–260 Image information is revealed by the correlations between the signal and idler photons and is not present in the detection of each individual photon. Fundamental to this approach to imaging, widely known as ghost imaging, is that the output plane of the NC in which the down-conversion process occurs is imaged such that each photon (signal and idler) is imaged onto the detector array and the object [O(r)], respectively. The photons that interact with the object (idler) are collected by a “bucket detector,” i.e., a detector with no spatial resolution. This can be a single mode or multi-mode fiber that is coupled to a single pixel detector. All that remains is to measure the spatial information of the idler photon.
Conceptualization of a quantum imaging. (a) One photon from a pair of spatially entangled photons interacts with an object and is collected by a detector without spatial resolution, while the photon that does not interact with the object is spatially resolved. When both photons are detected in coincidence, an image of the object is reconstructed. (b) The image [I(x, y)] is reconstructed as a linear combination of each patterned mask [Pi(x, y)], weighted by the detected coincidences (ci). The top panel shows an example of image reconstruction by the random basis, while the bottom panel is an example of image reconstruction by the Hadamard basis.
Conceptualization of a quantum imaging. (a) One photon from a pair of spatially entangled photons interacts with an object and is collected by a detector without spatial resolution, while the photon that does not interact with the object is spatially resolved. When both photons are detected in coincidence, an image of the object is reconstructed. (b) The image [I(x, y)] is reconstructed as a linear combination of each patterned mask [Pi(x, y)], weighted by the detected coincidences (ci). The top panel shows an example of image reconstruction by the random basis, while the bottom panel is an example of image reconstruction by the Hadamard basis.
For binary random masks, each pixel is randomly assigned a value of either 0 or 1. In this way, the masks contain randomly distributed binary pixels that are either turned on or off.263 In the top panel of Fig. 19(b), we show examples of random patterned mask types used in a typical ghost image reconstruction. A drawback is that a large number of patterns are required (2 N2 or more, where N × N is the number of pixels in the image) to reconstruct an image that is noisy and, therefore, of poor quality.264 Small improvements can be seen if half the pixels in each pattern are activated.265
The mask resolution, independent of the type of mask chosen to spatially resolve the signal photons, will determine the resolution at which the object will be imaged. A higher resolution results in a larger number of basis elements and, therefore, an increased number of masks is needed to reconstruct the image. Increasing the resolution has direct consequences for the reconstruction time; the number of Walsh–Hadamard masks required to form a complete set scales as N2. Specifically, for a complete or general image solution on the Walsh–Hadamard basis, N2 masks are required. Although N2 masks are required for a general image solution, it has been shown that it is possible to acquire image information up to 10× faster with the use of smart algorithms.272,273
Masks are generated as a pre-measurement step, i.e., they are generated prior to starting the experiment and stored in memory or in a local directory. Many programming languages can be used to generate the masks as they are defined earlier. Programming languages often come with native functions, allowing the user quick and easy mask generation. As an example, in the MATLAB programming language, by calling the function Hadamard(N), a Hadamard matrix of order N is generated, where N defines the number of pixels in the mask. Similarly, to generate a random mask, one would use rand(N,N) in MATLAB.
2. Resolution limits for ghost imaging with SPDC photons
3. Structured light in quantum imaging
The quantum imaging approaches we discussed in this tutorial used structured light modes in the form of patterned pixels encoded on SLMs as part of the measurements. Interestingly, spiral phases that are reminiscent of OAM modes, as well as binary phase steps reminiscent of HG modes, have also been incorporated to improve the performance of phase contrast images in quantum ghost imaging experiments with single pixel scanning.22 Recently, it has also been demonstrated that the necessary information for phase retrieval is naturally embedded in the correlation measurements formed from traditional patterned pixels, as detailed earlier.275
While these projective approaches that make use of structured pixels are cost effective, quantum imaging with single photon sensitive cameras has recently taken center stage (see Refs. 276 and 277), making it easier to execute quantum imaging experiments. Recent approaches involve using quantum sources (mainly SPDC photons) to illuminate objects and measure the spatially and temporally modulated photons using single photon cameras.278–280 In addition, these cameras have also been used for performing fundamental tests of quantum mechanics (spatial Bell-inequality tests,281,282 spatial quantum interference283) and for characterizing quantum structured light (spatial-momentum entanglement witnessing,284 OAM entanglement state tomography285). Entanglement sources beyond single degrees of freedom, i.e., hyper-entanglement being the simultaneous entanglement of multiple degrees of freedom (polarization-spatial, in this case), have been used for quantum holography.286
On the other hand, interaction free (see early papers on this topic287–289) imaging approaches are becoming prevalent, where single photon interference is used as the main tool for imaging transparent objects.290 Alternatively, image formation can be achieved through quantum interference,291 i.e., using the Hong–Ou–Mandel interference effect. While these techniques make use of linear optical elements to achieve the interference, the authors in Ref. 292 used multiple nonlinear crystals, inducing coherence between pump and signal photons that share a common idler mode, allowing the transfer of information from signal to idler through SPDC293,294 (see Ref. 295 for a detailed tutorial). Here, the degrees of freedom that are manipulated are the transverse spatial and wavelength/frequency of photons, demonstrating superb control of quantum information in an imaging experiment by exploiting nonlinear optics.
VII. CONCLUDING REMARKS
In this tutorial, we have outlined the basic theory behind realizing high-dimensional quantum states on the basis of spatial modes. We have explained with practical examples how to get started with the experiment, how to prepare and measure quantum states, and finally, how to use this as a resource in quantum information processing and imaging. Rather than a comprehensive monograph, it is a quick guide that serves as a useful starting point for both students and new researchers in the field. As such, we have used common “vanilla” forms of experiments as a means to convey the message, for example, using SPDC as the source of entangled photons and OAM as the basis for measurement. We stress that the state-of-the-art is far more versatile and encourage the interested reader to explore more widely for an advanced treatment beyond this tutorial.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Isaac Nape: Writing – original draft (lead); Writing – review & editing (lead). Bereneice Sephton: Writing – original draft (equal); Writing – review & editing (equal). Pedro Ornelas: Writing – original draft (equal); Writing – review & editing (supporting). Chane Moodley: Writing – original draft (supporting); Writing – review & editing (supporting). Andrew Forbes: Writing – original draft (lead); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.