Skip to main content

ORIGINAL RESEARCH article

Front. Phys., 04 September 2023
Sec. Optics and Photonics

Photonic quantum chromodynamics

  • Center for Exploratory Research Laboratory, Research and Development Group, Hitachi, Ltd., Tokyo, Japan

Twisted photons with finite orbital angular momentum have a distinct mode profile with topological charge at the center of the mode while propagating in a certain direction. Each mode with different topological charges of m is orthogonal, in the sense that the overlap integral vanishes among modes with different values of m. Here, we theoretically consider a superposition state among three different modes with left and right vortices and a Gaussian mode without a vortex. These three states are considered to be assigned to different quantum states; thus, we employed the su(3) Lie algebra and the associated SU(3) Lie group to classify the photonic states. We calculated expectation values of eight generators of the su(3) Lie algebra, which should be observable, since the generators are Hermite matrices. We proposed to call these parameters Gell-Mann parameters, named after the theoretical physicist Murray Gell-Mann, who established quantum chromodynamics (QCD) for quarks. The Gell-Mann parameters are represented on the eight-dimensional hypersphere with its radius fixed due to the conservation law of the Casimir operator. Thus, we discussed a possibility of exploring photonic QCD in experiments and classified SU(3) states to embed the parameters in SO(6) and SO(5).

1 Introduction

The Lie algebra and Lie group [16] were developed mathematically much earlier than the discoveries of quantum mechanics [5, 710]. The theory formulates general principles on how to classify various matrices with complex numbers (C) and provides deep insights into the topological structure underlying matrix calculations. The theory covers quite wide areas and thus applicable to many fields in quantum physics, including various two-level systems, described by the special unitary group of two dimensions, SU(2), to understand polarization [8, 9, 1127]. Historically, however, the powerful mathematical features are not widely recognized for SU(2) states since it is not so much complicated to deal with, even if we do not employ the knowledge of Lie algebra. The situation completely changed once Murray Gell-Mann identified the underlying symmetries for composite elementary particles of baryons and mesons, establishing the quantum chromodynamics (QCD) [4, 5, 10, 2831]. The Lie algebra is now an indispensable tool in physics on elementary particles. Here, we propose to introduce the framework of the Lie algebra and Lie group to photonics, especially for exploring the photonic analog of QCD by utilizing photonic orbital angular momentum (Figures 1A–D).

FIGURE 1
www.frontiersin.org

FIGURE 1. Concept of photonic QCD. Trajectories of a constant phase for photons with (A) left vortex, (B) no vortex, and (C) right vortex are schematically shown. These three states have different topological charges, characterized by orbital angular momentum, and thus, they are orthogonal to each other. We assign three different colors for different charges, red, green, and blue, for these states, as SU(3) states. We consider (D) an arbitrary superposition state among states with different topological charges for variable amplitudes and phases.

Photons are elementary particles of light [32, 33], which have both spin [11, 13, 14] angular momentum and orbital angular momentum [23, 3451] as internal degrees of freedom. The nature of spin is known as polarization [11, 13, 14, 52], which is widely used in sunglasses, liquid-crystal displays, and digital coherent communications [5356], while orbital angular momentum is used in optical tweezers, laser-patterning, and quantum optics [5765]. Previously, spin angular momentum and orbital angular momentum of photons were considered to be impossible for splitting into two independent degrees of freedom [66] in a free space [41, 42, 67, 68]. However, it was recently reported that these degrees of freedom could be successfully separated in a proper gauge invariant way by plane wave expansions [69]. We also confirmed that both spin angular momentum and orbital angular momentum are well-defined quantum observables for photons in a waveguide and a free space as far as the propagation mode is sufficiently confined in the core [26]. Therefore, we can measure quantum mechanical averages of the angular momentum, which could be macroscopic values for coherent photons.

2 Summary of the Lie algebra

First, we summarize theoretically minimum knowledge on fundamental properties of the su(3) Lie algebra in order to make our discussions self-contained and clarify our notations. This study would help photonic researchers, who are not familiar with the su(3) Lie algebra, understand the idea to treat three orthogonal quantum states in an equal fashion. It is far from a comprehensive summary, such that interested readers should refer to excellent textbooks [15, 7]. Those who are familiar with the su(3) algebra may skip this section.

The Lie algebra and Lie group were mathematically developed as early as the 1870s, without specific applications in physics [14, 6]. The first serious applications in physics was found in elementary physics, leading to the discoveries of quarks [5]. Obviously, the su(3) Lie algebra and more general representation theories are robust, and they will be applicable to various cases. On the other hand, here, we focus on applications in photonics, and we use this example to review fundamental characteristics of the su(3) Lie algebra. Thus, we lose the generality in our construction of the logic, but it is easier to understand the concept, and applications in higher dimensions will be straightforward.

2.1 Generators of the su(3) algebra

We consider three orthogonal quantum states, namely, left- and right-twisted states and the no-vortex state, which are described in the Hilbert space with three complex numbers, C3. We allow arbitrary mixing of these three states, realized by the superposition principle, and the wavefunction could be considered to be normalized to 1 or to the fixed number of photons, N, such that the radius of the complex sphere is fixed. Consequently, the number of freedom is 2 × 3–1 = 5, and the Hilbert space is equivalent to the sphere of five dimensions, S5. In order to describe arbitrary rotational operations of the wavefunction in the Hilbert space, we need complex matrices of 3 × 3 for the SU(3) Lie group, which is realized by the exponential mapping form of the su(3) Lie algebra. The SU(3) forms a group, whose determinant must be unity, which corresponds to the traceless condition for the Lie algebra. Therefore, we need 3 × 3–1 = 8 bases, defined by Gell-Mann [4, 5, 10, 2831] as

λ̂1=010100000,(1)
λ̂2=0i0i00000,(2)
λ̂3=100010000,(3)
λ̂4=001000100,(4)
λ̂5=00i000i00,(5)
λ̂6=000001010,(6)
λ̂7=00000i0i0,(7)
λ̂8=13100010002,(8)

which are all Hermite matrices, λ̂i=λ̂i (i = 1, …, 8), implying that their expectation values must be real and observable. We can also use the bases, êi=λ̂i/2, reflecting the underlying su(2) symmetry between two orthogonal states. The bases satisfy the normalization relationship for the trace, Trλ̂iλ̂j=2δij, where δij is the Kronecker delta function.

2.2 Commutation relationship

The commutation relationship is obtained by the straightforward calculation of basis matrices, which is derived as:

λ̂i,λ̂j=2ikCijkλ̂k,(9)

where the structure constants, Cijk, are listed in Table 1. Cijk is an asymmetric tensor, such that odd permutation of indices changes its sign. Most commutation relationships involve only one term in the summation on the right-hand side of the equation, similar to the su(2) commutation relationship for spin. On the other hand, we must account for two terms involved in equations

λ̂4,λ̂5=2i12λ̂3+32λ̂8,(10)
λ̂6,λ̂7=2i12λ̂3+32λ̂8.(11)

TABLE 1
www.frontiersin.org

TABLE 1. Structure constant of the commutation relationship, λ̂i,λ̂j=2ikCijkλ̂k, in the su(3) Lie algebra.

We also confirm that there are two mutually commutable operators,

λ̂1,λ̂8=λ̂2,λ̂8=λ̂3,λ̂8=0(12)

while we see

λ̂1,λ̂20,λ̂1,λ̂30,λ̂2,λ̂30.(13)

Therefore, the rank-2 character of the su(3) Lie algebra is confirmed. It is also evident that λ̂3 and λ̂8 are already diagonalized in our representation for the basis.

2.3 Basis operators for the su(2) algebra within the su(3) algebra

We consider three orthogonal states for the su(3) Lie algebra, and we select two states among three available states. There are three ways to choose two states, and each of the pairs of states will form the su(2) Lie algebra.

For example, if we choose the first and second states, corresponding to the left- and right-twisted states, respectively, we use bases

ê1t=12λ̂1,(14)
ê2t=12λ̂2,(15)
ê3t=12λ̂3,(16)

for describing the SU(2) states since they are equivalent to Pauli matrices,

σ̂1=0110,(17)
σ̂2=0ii0,(18)
σ̂3=1001,(19)

if we neglect the third quantum state for the no-vortex state. These operators, êi(t), were originally used for describing isospin for quarks [5]. For our applications, they will be useful to describe the rotation between the left- and right-twisted states. The rotation corresponds to mixing the left- and right-twisted states, which are described in the Poincaré sphere for vortices.

The introduction of the coupling between the left- and right-twisted states corresponds to allowing the direct SU(2) coupling between states with = −1 and = +1, as shown in Figure 2B. In standard quantum mechanics for orbital angular momentum [5, 8, 31], this coupling is not considered since it changes to the topological charge of ± 2, such that the ladder operations of SU(2) cannot directly transfer states to each other. On the other hand, we allow this coupling for photons simply by mixing states with certain amplitudes and phases [23, 25, 35, 47, 48, 50, 51, 70], which allows us to extend photonic states to obtain an SU(3) symmetry.

FIGURE 2
www.frontiersin.org

FIGURE 2. SU(3) states with photonic orbital angular momentum. (A) Fundamental multiplet of the su(3) Lie algebra. Fundamental basis states of |ψ1⟩, |ψ2⟩, and |ψ3⟩ are shown on the (t3, t8) plane, characterized by their quantum numbers. t3 is known as isospin for quarks. The states are shown by points, given by the eigenvalues, which are separated by the same distance and form an equilateral triangle as their topology, implying the states are treated in an equal footing in the su(3) Lie algebra. We can use u3, v3, or hypercharge of y = 2 (u3 + v3)/3 instead of t8, but only two vectors are required to span the t3t8 plane due to the rank-2 character of the su(3) Lie algebra. (B) Bending of the quantization axis of SU(2) to form SU(3) states. The quantum number (3) of orbital angular momentum along the quantization axis is usually characterized by SU(2) states, as shown in the upper diagram. By allowing SU(2) rotation between left- and right-twisted states, we effectively bend the 3 to realize the superposition state. By combining superposition states with the no-vortex state, we mix the three orthogonal states to realize SU(3) states.

Another pair of states is made of the right-twisted state and the no-vortex state, whose bases are

ê1u=12λ̂6,(20)
ê2u=12λ̂7,(21)
ê3u=1212λ̂3+32λ̂8,(22)
=12000010001.(23)

Here, it is noteworthy that we can define a new vector operator of ê3(u), for example, from λ̂3 and λ̂8, since they are basis vector operators in the su(3) Lie algebra, which forms a vector space. We cannot simply add components in the SU(3) Lie group since the SU(3) Lie group is not a vector space. We observe that ê3(u) is normalized to be similar to ê3(t), such that it is useful to consider SU(2) rotations by ê1(u), ê2(u), and ê3(u).

Similarly, we consider the pair of states consisting of left-vortex state and the no-vortex state, whose bases are

ê1v=12λ̂4,(24)
ê2v=12λ̂5,(25)
ê3v=1212λ̂3+32λ̂8,(26)
=12100000001.(27)

These su(2) commutation relationships are summarized as

ê1x,ê2x=iê3x,(28)

where x = t, u, or v. We used lower-case letters (x = t, u, or v) for operators describing a single quanta such as a quark or a photon, and we used upper-case letters for coherent states of photons (X = T, U, or V) under Bose–Einstein condensation, where a macroscopic number, N, of photons are occupying the same state, as discussed further in this paper.

The su(3) algebra does not contain a non-trivial invariant group. For example, we see that

ê1t,ê4=14λ̂1,λ̂4=14iλ̂7=12ê7,(29)

such that the internal SU(2) groups are connected by commutation relationships, and the su(2) Lie algebra is not closed.

2.4 Ladder operators

We utilize the Cartan–Dynkin formulation [5] for describing the SU(3) states. In the formalism, we consider to fix the quantization axis, rather than isotropic to all directions in the su(2) Lie algebra, and consider ladder operators for increasing and decreasing the quantum number along the quantization axis [5, 9]. More specifically, we define

t̂±=12λ̂1±iλ̂2,(30)
t̂3=12λ̂3,(31)

where t̂3 stands for the operator of the z component of the rotationally symmetric su(2) operator t̂=(t̂1,t̂2,t̂3) and t̂+ and t̂ represent the rising and the lowering operators, respectively, to increase and decrease the quantum number for t̂3. We use t̂3 and t̂± instead of êi(t) (i = 1, 2, 3), and their commutation relationships become

t̂3,t̂±=±t̂±,(32)
t̂+,t̂=2t̂3.(33)

For applications in isospin, t̂3 provides the fixed isospin value (t3) for each elementary particle, such as a proton (t3 = 1/2) and a neutron (t3 = −1/2), a deuterium (D, t3 = 0), and a tritium (T, t3 = 1/2). In elementary particle physics, a superposition state between a proton and a neutron, for example, is not realized due to the superselection rule [5] since the superposition state between different charged states is prohibited. On the other hand, for applications in vortices, we can safely consider the superposition state between the left- and right-twisted states [35, 63], such that we can consider arbitrary mixing of left- and right-twisted states with an arbitrary phase between them. The ladder operators t̂± correspond to changing the topological charge at the center of the vortices for changing its orbital angular momentum from the left to the right circulation or vice versa.

Similarly, we consider the rising and lowering ladder operators, û+ and û, respectively, for the superposition state between the right-twisted and no-vortex states, and the z component of the rotationally symmetric su(2) operator û=(û1,û2,û3):

û±=12λ̂6±iλ̂7,(34)
û3=14λ̂3+3λ̂8,(35)

whose commutation relationships become

û3,û±=±û±,(36)
û+,û=2û3.(37)

For the mixing of the left-twisted and no-vortex states, the corresponding ladder operators, v̂+ and v̂, and the z component of the rotationally symmetric su(2) operator v̂=(v̂1,v̂2,v̂3) become

v̂±=12λ̂4±iλ̂5,(38)
v̂3=14λ̂3+3λ̂8,(39)

whose commutation relationships become

v̂3,v̂±=±v̂±,(40)
v̂+,v̂=2v̂3.(41)

Here, we defined nine operators for ladders and the quantization components of the three sets of su(2) operators (t̂,û,v̂), while only eight bases are required for the su(3) algebra due to the traceless requirement. Consequently, we obtained one identity,

v̂3=û3+t̂3,(42)

which must be met for all states. This means that only two quantum numbers are independently chosen, regardless of apparent three sets of SU(2) states, which is, in fact, consistent with the rank-2 nature of the su(3) Lie algebra.

2.5 Hypercharge and topological charge

As discussed previously, we select two quantum operators from three operators, t̂3, û3, and v̂3, for describing the SU(3) quantum states. If we choose t̂3, we can choose û3 or v̂3. Alternatively, we can consider the superposition state, made of both û3 andv̂3, whose z component becomes

λ̂8=23û3+v̂3.(43)

Equivalently, we define the hypercharge operator [5] as

ŷ=13λ̂8,(44)
=23û3+v̂3=43û3+23t̂3=43v̂323t̂3,(45)

which was indispensable to understand quarks, their 2-body (3-body) compounds of meson, and the 3-body compounds of baryons. They are commutative with the other three operators,

ŷ,t̂3=ŷ,û3=ŷ,v̂3=0,(46)

meaning that hypercharge could be the simultaneous quantum number with the other parameter. Therefore, we expect that the eigenstate is labeled by the quantum numbers, t3, u3, and v3 with y = (u3 + v3)/3, to satisfy

ŷt3,u3,v3=yt3,u3,v3=yt3,y.(47)

We also confirm the identity

ŷ,t̂±=23û3,t̂±+23v̂3,t̂±=0,(48)

which ensures that

ŷt̂±t3,y=yt̂±t3,y,(49)

meaning that the application of ladder operations by t̂± preserves the hypercharge, y.

On the other hand, we find that

ŷ,û±=23û3+v̂3,û±=±û±,(50)

which leads to

ŷû±t3,y=y±1û±t3,y,(51)

which means û+ increments y and û decrements y, respectively. We also confirm the same rule for v̂± as

ŷ,v̂±=23û3+v̂3,v̂±=±v̂±,(52)

which corresponds to

ŷv̂±t3,y=y±1v̂±t3,y.(53)

Finally, we obtain the fundamental multiplets (Figures 2A, B), given by three states,

|ψ1=t3=12,t8=1332=100,(54)
|ψ2=t3=12,t8=1332=010,(55)
|ψ3=t3=0,t8=2332=001.(56)

An arbitrary quantum state can be generated by mixing these three states using the superposition principle: multiplying complex numbers to fundamental ket states and adding up. In a standard matrix formulation of quantum mechanics, a general state is given by a row of three complex numbers.

For quarks, there exists an identity relationship between hypercharge and charge, q, as

q=t3+12y.(57)

Therefore, one can use charge instead of hypercharge for an alternative quantum number.

For our applications to photonic orbital angular momentum, we consider superposition states between left- and right-twisted states and no-vortex state. We use the orbital angular momentum along the quantization axis, z, which is the direction of the propagation, as the first quantum number, instead of the isospin of t3. For the second quantum number, instead of hypercharge, we choose the topological charge, defined by

qt=y+23,(58)

which becomes 0 for the no-vortex state and 1 for both left- and right-twisted states. The topological charge corresponds to the winding number of the mode at the core, propagating along a certain z direction. It is also linked to the magnitude of photonic orbital angular momentum. In this paper, we only consider vortices with a winding number of 1 or 0, but it will be straightforward to extend our discussions to higher-order states.

2.6 Casimir operators

There are other conservative properties in the su(3) Lie algebra. We define a Casimir operator as

Ĉ1=14i=1nλ̂i2=i=1nêi2.(59)

We calculate the commutation relationship as follows:

Ĉ1,λ̂j=14i=1nλ̂i2,λ̂j,(60)
=14i=1nλ̂i2λ̂jλ̂jλ̂i2,(61)
=14i=1nλ̂iλ̂iλ̂jλ̂jλ̂iλ̂i,(62)
=14i=1nλ̂iλ̂i,λ̂jλ̂j,λ̂iλ̂i,(63)
=2i4ikCijkλ̂iλ̂kCjikλ̂kλ̂i,(64)
=0,(65)

where we changed the dummy indices in the last line as ikCjikλ̂kλ̂i=kiCjkiλ̂iλ̂k=ikCijkλ̂iλ̂k. Therefore, the Casimir operator, Ĉ1, obtains the simultaneous eigenstate with the rank-2 states for λ̂i. In fact, we observe, from direct calculations, that

Ĉ1=1+130001+130001+13=43,(66)

which means that Ĉ1 is constant for SU(3) states. Here, it is obvious that we abbreviated the unit matrix of 3 × 3, 13, multiplied with 43=4313, for simplicity.

There is another Casimir operator, defined by

Ĉ2=ijkDijkt̂it̂jt̂k=18ijkDijkλ̂iλ̂jλ̂k,(67)

where Dijk is a symmetric tensor as defined in the anti-commutation relationship given as follows. We observe that Ĉ2 is also constant in the su(3) Lie algebra, such that the commutation relationship

Ĉ2,λ̂i=0,(68)

vanishes.

2.7 Anti-commutation relation

We also obtain the anti-commutation relationship as

λ̂i,λ̂j=43δij+2k=18Dijkλ̂k,(69)

which is equivalent to

êi,êj=13δij+k=18Dijkêk,(70)

where 43δij should be considered as 43δij13, as before. The symmetric tensor, Dijk, is shown in Table 2.

TABLE 2
www.frontiersin.org

TABLE 2. Structure constant of the anti-commutation relationship, λ̂i,λ̂j=4δij/3+2kDijkλ̂k, in the su(3) Lie algebra.

By multiplying λ̂k with the anti-commutation relationship, we obtain

λ̂i,λ̂jλ̂k=43δijλ̂k+2Dijkλ̂kλ̂k.(71)

We take the trace of the equation, while using Tr(λ̂i)=0 and Tr(λ̂iλ̂j)=2δij, and we obtain

Dijk=14Trλ̂i,λ̂jλ̂k.(72)

Similarly, we also obtain

Cijk=14iTrλ̂i,λ̂jλ̂k(73)

from the commutation relationship.

Finally, we show that

Ĉ2=Ĉ12Ĉ1116=109.(74)

To prove the identity, we use

λ̂i,λ̂j=λ̂iλ̂j+λ̂jλ̂i(75)

and

λ̂i,λ̂j=λ̂iλ̂jλ̂jλ̂i.(76)

By adding these equations, we obtain

λ̂i,λ̂j+λ̂i,λ̂j=2λ̂iλ̂j,(77)

which becomes

43δij+2Dijkλ̂k+2iCijkλ̂k=2λ̂iλ̂j(78)

from commutation and anti-commutation relationships. Then, we multiply a factor of λ̂iλ̂j and sum up to obtain

ijλ̂i2λ̂j2=23iλ̂i2+ijkDijkλ̂iλ̂jλ̂kiijkCijkλ̂iλ̂jλ̂k,(79)

where the last term becomes

ijkCijkλ̂iλ̂jλ̂k=12ijkCijkλ̂iλ̂jλ̂k+Cikjλ̂iλ̂kλ̂j=12ijkCijkλ̂iλ̂jλ̂kCijkλ̂iλ̂kλ̂j=12ijkCijkλ̂iλ̂jλ̂kλ̂kλ̂j=12ijkCijkλ̂iλ̂j,λ̂k=iijkCijkCjklλ̂iλ̂l=iijkCjkiCjklλ̂iλ̂l=3iiλ̂i2,(80)

which leads to

ijλ̂i2λ̂j2=ijkDijkλ̂iλ̂jλ̂k+23+3iλ̂i2.(81)

Therefore, we obtain

Ĉ2=Ĉ12Ĉ1116=109,(82)

which is in fact constant under the su(3) Lie algebra.

3 SU(3) state for twisted modes

3.1 Gell-Mann hypersphere

Now, we discuss how to classify the superposition state between left- and right-twisted states and no-vortex state using the su(3) Lie algebra and the SU(3) Lie group. We assume Laguerre–Gauss modes with a topological charge of qt = 1 for both left- and right-twisted states [23, 34, 35, 39, 40, 4751, 63, 65, 7173] for simplicity.

Here, our main idea is to assign the three states of twisted modes and no-vortex mode to orthogonal states of the SU(3) states (Figures 1, 2). The most important part of the twisted modes for orbital angular momentum is its azimuthal (ϕ) dependence [34]; i.e., the wavefunction of the ray with orbital angular momentum of m is given by

ϕ|m=eimϕ,(83)

which is orthogonal to each other for states with different charges of m in a sense,

m|m=02πdϕ2πei(mmϕ=δm,m.(84)

This means that the modes with different orbital angular momentum could be treated as orthogonal quantum mechanical states. For our consideration and notation [2427, 63, 70, 7476], we assign left- and right-twisted states as |L⟩ = |1⟩ = |ψ1⟩ and |R⟩ = | − 1⟩ = |ψ2⟩, respectively, and the no-vortex Gaussian state as |O⟩ = |0⟩ = |ψ3⟩. These states are also considered to have different topological charge, such as red, blue, and green (Figure 1).

We also assume that all modes have the same polarization state, such that our SU(3) state is polarized. Then, we consider the polarization degree of freedom, which comes from the SU(2) spin of photons [8, 9, 1121, 2427], such that we explore the photonic states with the SU(2) × SU(3) symmetry.

We consider a coherent ray of photons emitted from a laser source [1719, 24, 63, 70, 75], such that a macroscopic number of photons per second, N, pass through the cross section of the ray. We use upper-case letters to describe macroscopic observables and expectation values, such as photonic orbital angular momentum [23, 34, 35, 39, 40, 4751, 63, 65, 7173],

L̂i=N̂i=Nλ̂i,(85)

where = h/(2π) is the Dirac constant, defined by the Plank constant (h), divided by 2π, while lower-case letters are used for a single-quantum operator or a normalized parameter, such as a normalized orbital angular momentum operator,

̂i=λ̂i,(86)

for i = 1, 2, and 3.

There is a difference of factor of 2 in the definition between the orbital angular momentum operator ̂i and the isospin operator of t̂3, but it would be more appropriate to use ̂i for photonic vortices since the orbital angular momentum is quantized in the unit of [2426, 34, 35, 63].

First, let us review the SU(2) coupling between left- and right-twisted states [35, 63]. For this, we consider the following state:

|θl,ϕl=eiϕl2cosθl2e+iϕl2sinθl20,(87)

where the amplitudes of left- and right-vortex states are controlled by the polar angle of θl and the phase is defined by ϕl. We can realize this state using an exponential map from the su(3) Lie algebra to the SU(3) Lie group:

D̂2θl=expiλ̂2θl2,(88)
=cosθl2sinθl20sinθl2cosθl20000,(89)

which is a phase-shifter with its fast axis rotated for π/4 from the horizontal axis [24], together with another exponential map of

D̂3ϕl=expiλ̂3ϕl2,(90)
=expiϕl2000expiϕl20000,(91)

which is a rotator. We apply these operators to a unit vector, |ψ1⟩, to confirm the SU(2) state,

|θl,ϕl=D̂3ϕlD̂2θl|ψ1,(92)

made of left- and right-twisted states.

By calculating a standard quantum mechanical average from |θl, ϕl⟩, such that

i=λi=̂i=λ̂i=θl,ϕl|λ̂i|θl,ϕl(93)

for i = 1, 2, and 3, respectively, we obtain

λ1=sinθlcosϕl,(94)
λ2=sinθlsinϕl,(95)
λ3=cosθl.(96)

Thus, the SU(2) states between left- and right-twisted states could be shown on the Poincaré sphere for orbital angular momentum [35, 39, 40, 63, 72]. Usually, the rotational symmetry and the corresponding orbital angular momentum are considered by the SU(2) symmetry since the states between |L⟩ = |1⟩ and |R⟩ = | − 1⟩ cannot be transferred by the change in Δm = ±1, and instead, Δm = ±2 is required. This could be achieved by using a spiral phase plate [37] with a topological charge of m = 2. Alternatively, it is possible to create a superposition state between |L⟩ = |1⟩ and |R⟩ = | − 1⟩, and SU(2) states could be realized by controlling the amplitudes and phases [35, 63]. For our considerations in the SU(3) states, this corresponds to bending the quantization axis, ̂3, for allowing three states to couple among each other (Figures 2A, B).

Next, we consider coupling between the no-vortex state and left- or right-vortex states. This corresponds to changing the hypercharge and topological charge. We call these SU(2) couplings hyperspin since they exhibit spin-like SU(2) behaviors, yet they are different from spin. For elementary particles, such as quarks, states with different charged particles cannot be realized at all due to the superselection rule, such that composite particles, such as a neutron and a proton, cannot be in their superposition state [5]. However, for coherent photons, we consider a superposition state among different topologically charged states, such that we can mix the no-vortex state and twisted state at an arbitrary ratio in amplitudes with a certain definite phase. Topologically, the vortex is well known to be equivalent to a shape of a doughnut, which cannot be continuously changed to a shape of a ball. Our challenge could be considered to realize a superposition state between a doughnut and a ball, which is impossible classically, while we would have a chance since photons are elementary particles with a wave character allowing a superposition state of orthogonal states.

Here, we consider the hyperspin coupling, which means that we explore mixing between twisted states and no-vortex state. In order to achieve this, the easiest option is to follow the previous approach of the SU(2) state between left and right vortices. We simply need to change from λ̂2/2=ê2(t) and λ̂3/2=ê3(t) to λ̂5/2=ê2(v) and ê3(v), respectively, and we define

D̂2vθy=expiê2vθy,(97)
=cosθy20sinθy2000sinθy20cosθy2,(98)

and

D̂3vϕy=expiê3vϕy,(99)
=expiϕy20000000expiϕy2.(100)

We obtain a general SU(3) state,

|θl,ϕl;θy,ϕy=D̂3ϕlD̂2θlD̂3vϕyD̂2vθy|ψ1=eiϕy2eiϕl2cosθl2cosθy2eiϕy2e+iϕl2sinθl2cosθy2e+iϕy2sinθy2.(101)

Finally, we can calculate the expectation values for all generators of the su(3) Lie algebra, which becomes a vector in an eight-dimensional space, given by

λ=λ1λ2λ3λ4λ5λ6λ7λ8=sinθlcosϕlcos2θy2sinθlsinϕlcos2θy2cosθlcos2θy2cosϕy+ϕl2sinθycosθl2sinϕy+ϕl2sinθycosθl2cosϕyϕl2sinθysinθl2sinϕyϕl2sinθysinθl236+32cosθy.(102)

An arbitrary state of SU(3) is characterized by this vector, which is similar to the Stokes parameters [11] on the Poincaré sphere [12]. The higher-dimensional vector of λ satisfies the norm conservation

i=18λi2=43,(103)

upon rotations in the eight-dimensional space, which is guaranteed from the constant Casimir operator of Ĉ1=4/3, as shown previously. Therefore, an SU(3) state is represented as a point on the hypersphere with the radius of

i=18λi2=23.(104)

We propose this hypersphere as the Gell-Mann hypersphere, named after Gell-Mann who found the SU(3) symmetry of baryons and mesons, leading to the discovery of quarks [2830]. In fact, we have the eightfold way [2830] to allow the SU(3) superposition state by changing the amplitudes and the phases of the wavefunction. We can attribute color charge of red, green, and blue to three fundamental states of |ψ1⟩, |ψ2⟩, and |ψ3⟩, respectively, similar to QCD [4, 10, 2830]. In QCD for quarks, only certain sets of multiplets, such as baryons and mesons, are observed as stable bound states of quarks, due to the spontaneous symmetry breaking of the universe [7781]. In our photonic QCD, on the other hand, we can discuss an arbitrary superposition state by mixing three orthogonal states of left- and right-vortices and no-twisted rays. Therefore, we can discuss the SU(3) state before the symmetry is broken; in other words, the symmetry can be recovered without injecting additional energies to the system, similar to the Nambu–Goldstone bosons [7781]. This corresponds to the rotation of the hyperspin of λ in the eight-dimensional Gell-Mann space by using eight generators of rotation λ̂i (i = 1, ⋯, 8) to change the amplitudes and phases. In experiments, this is achieved by using rotators and phase-shifters of SU(2) [2427, 63, 70, 7476] since we can realize arbitrary rotations of the SU(3) state by using three sets of SU(2) rotations, as shown previously.

Among the eight Gell-Mann parameters, λ̂i, two are especially important since the su(3) Lie algebra is of rank-2 nature. One of them is 3 = λ3, which determines the average orbital angular momentum along the direction of propagation, z. The other important parameter is

y3=13λ8=16+12cosθy,(105)

which determines the average hypercharge. We confirm the expected maximum and minimum hypercharge of maxy3=1/3 and miny3=2/3, respectively. Hypercharge is simply converted to the topological charge, qt = y3 + 2/3, and we confirm maxqt=1 and minqt=0 as expected maximum and minimum hypercharge for vortices and no-twisted state, respectively.

The Gell-Mann parameters are composed of eight real parameters, and the vector, λ, has a unit length, |λ|=1. Therefore, the rotation of the vector λ is achieved by the special orthogonal group of eight dimensions, SO(8). The corresponding generators of the so(8) Lie algebra are adjoint representations of the su(3) generators (λ̂i, i = 1, …, 8), which become the structure constants of Cijk (i, j, k = 1, …, 8).

3.2 Hyperspin with the left/right vortex

The Gell-Mann hypersphere contains all practical information about the SU(3) states in terms of amplitudes and phases. Unfortunately, it is impossible to recognize the eight-dimensional hypersphere in the three-dimensional space and time. In the previous sub-section, we showed that the coupling between left and right vortices could be represented by the Poincaré sphere for the twisted photons [35, 63], which corresponds to the coupling controlled by the su(2) generators of ê1(t), ê2(t), and ê3(t). Here, we consider other su(2) generators and discuss how hyperspin is represented in a similar way to the Poincaré sphere.

First, we consider the coupling between the left-twisted state and no-vortex state. This corresponds to the limit (θl, ϕl) → (0, 0), and the Gell-Mann parameters become

λ=00121+cosθysinθycosϕysinθysinϕy0036+32cosθy.(106)

In this case, the parameter λ3 can take a value between 1 and 0 since the left vortex has a topological charge of 1 due to λ3̂|L=|L, while the no-vortex state, (|O⟩), does not have a topological charge, as λ3̂|O=0. The superposition state is characterized to be the non-zero average of λ3, and if the amount of the right-vortex component is less than that of the left-vortex component, λ3 becomes positive. This corresponds to the net left circulation of orbital angular momentum. The other Gell-Mann parameters are given by θy and ϕy. In the limit of the zero right-vortex component, it is convenient to consider the average of the su(2)-generating vector,

v=v1,v2,v3=v̂,(107)
=λ4/2,λ5/2,v3,(108)
=12sinθycosϕysinθysinϕycosθy,(109)

which corresponds to introducing v3=(λ3+3λ8)/4, instead of λ8 or tq, since only two parameters are independent among (t3, u3, v3) due to the rank-2 character of the su(3) Lie algebra.

Similarly, we also check the coupling between the right-vortex state and the no-vortex state, which corresponds to take the limit of (θl, ϕl) → (π, 0), and we obtain the Gell-Mann parameters,

λ=00121+cosθy00sinθycosϕysinθysinϕy36+32cosθy.(110)

In this case, the sign of λ3 changed compared with the coupling with the left vortex since we assigned a negative sign for λ3̂|R=|R to the right vortex, which is observed from the observer side against the light coming to the detector [24, 63]. Therefore, λ3 can take a value between −1 and 0, which corresponds to the average right circulation of orbital angular momentum. For the right circulation, it is useful to calculate the average of the su(2)-generating vector, û as

u=u1,u2,u3=û,(111)
=λ6/2,λ7/2,u3,(112)
=12sinθycosϕysinθysinϕycosθy,(113)

which becomes the same formula for v, when only one chirality (i.e., left or right vortex) is involved. In fact, the parameters θy and ϕy account for the relative phase and the amplitudes, respectively, between the state with |m| = 1 and the state with m = 0 without including the difference in chiralities. Here, u3 is introduced by u3=(λ3+3λ8)/4, and it satisfies the conservation law of v3u3 = t3.

Consequently, we obtained three vectors, (t, u, v), where t = (1, 2, 3)/2 is obtained from the average orbital angular momentum. Each vector of t, u, or v is three-dimensional, such that they are represented by the Poincaré spheres. However, care must be taken when dealing with the radiuses of the Poincaré spheres since they depend on the relative amplitudes determined by θl and θy. This comes from the mutual dependence among three sets of the su(2) Lie algebra since the su(3) Lie algebra does not contain the non-trivial invariant group, as confirmed previously. As a result, we obtained three mutually dependent spheres, which have 3 × 3 = 9 parameters, with one identity of v3 = u3 + t3. For visualization purposes, the three Poincaré spheres with variable radiuses might be practically more useful for humans living in three spatial dimensions, rather than an eight-dimensional Gell-Mann hypersphere of the constant radius of 2/3, whose surface is equivalent to a seven-dimensional spherical surface of S7, given by real numbers, with a fixed radius in the eight-dimensional space.

3.3 Hyperspin embedded in SO(6)

Gell-Mann parameters in SO(8) are useful for understanding the coupling among |L⟩, |R⟩, and |O⟩. However, we can easily recognize that the generators of the su(3) Lie algebra cannot span the whole hypersurface of SO(8). For example, parameters λi (i = 1, …, 7), except for λ8, cannot take values above 1 or below −1, while the radius of 2/3 is larger than 1. This clearly shows that a point like (2/3,0,,0) cannot be covered at all, such that SO(8) is much larger than the parameter space required to represent the photonic states, composed of three orthogonal states.

Then, let us consider the number of freedom required for mixing |L⟩, |R⟩, and |O⟩. In general, we should consider the variable density of photons since the radius of the Poincaré sphere depends on the output power of the ray [8, 9, 1121, 2427, 63, 70, 75]. Then, photons in the coherent state are represented by one complex number per orthogonal degree of freedom for the component of the wavefunction. We consider this for a fixed polarization state while we have three orthogonal states for vortices, and therefore, we have six degrees of freedom (Table 3).

TABLE 3
www.frontiersin.org

TABLE 3. Degrees of freedom for photons with three orthogonal states.

These six degrees of freedom are attributed to corresponding physical parameters (Table 3). One degree of freedom is assigned to the power density of the ray, and another is used for the global U(1) phase, which will not play a role in the expectation values of the Gell-Mann hypersphere. Two degrees of freedom are required for describing the superposition state for orbital angular momentum, which is shown in the Poincaré sphere with variable radiuses (Figure 3A). Therefore, the remaining two parameters should be assigned to hyperspin to account for the mixing of |L⟩ and/or |R⟩ with |O⟩. This picture is consistent with the wavefunction of |θl, ϕl; θy, ϕy⟩, where θy and ϕy account for hyperspin. On the other hand, we used eight Gell-Mann parameters for describing the superposition state from the expectation values. All eight parameters are required to understand the full rotational ways on the Gell-Mann hypersphere; however, fewer parameters are required to scan the full wavefunction over the expected Hilbert space of S5. Here, we try to reduce the number of Gell-Mann parameters to embed hyperspin in SO(6). The aim is to represent the hyperspin

y1=12sinθycosϕy,(114)
y2=12sinθysinϕy,(115)
y3=16+12cosθy,(116)

as shown on the Poincaré sphere (Figure 3B), which should be enough for representing θy and ϕy topologically.

FIGURE 3
www.frontiersin.org

FIGURE 3. Renormalization of Gell-Mann parameters. The eight-dimensional Gell-Mann hypersphere was reduced to two Poincaré spheres for (A) photonic orbital angular momentum and (B) hyperspin. The radius of the Poincaré sphere for orbital angular momentum is cos2 (θy/2), while it is 1/2 for hyperspin. The maximum and minimum y3 correspond to a hypercharge of 1/3 and −2/3, respectively, which are equivalent to a topological charge of 1 (pure vortex of |L⟩ or |R⟩) and 0 (no vortex, |O⟩).

A hint is found in Gell-Mann parameters, λ4, ⋯, λ7, which has the magnitude

λ42+λ52+λ62+λ72=sin2θy(117)

upon changing the other parameters θl, ϕl, and ϕy. Therefore, we can eliminate λ6 and λ7 by renormalizing the operators for the su(3) Lie algebra.

By inspecting λ4, ⋯, λ7, we realize that the phases of ϕy and ϕl are coupled in a mixed form. If we convert ϕy ± ϕl/2 → ϕy, the rest of parameters are easily converted upon rotations. This could be achieved if we remember the rotation matrices of

Rθ=cosθsinθsinθcosθ(118)

from a group to satisfy the associative requirement

Rϕy+ϕl2=Rϕl2Rϕy.(119)

Then, we obtain

cosϕy+ϕl2sinϕy+ϕl2=Rϕl2cosϕysinϕy,(120)

whose reverse relationship becomes

cosϕysinϕy=Rϕl2cosϕy+ϕl2sinϕy+ϕl2.(121)

Using these formulas, we define

λ4λ5=Rϕl2λ4λ5,(122)
=cosϕl2sinϕl2sinϕl2cosϕl2λ4λ5,(123)
=cosϕysinθycosθl2sinϕysinθycosθl2,(124)

and

λ6λ7=Rϕl2λ6λ7,(125)
=cosϕl2sinϕl2sinϕl2cosϕl2λ6λ7,(126)
=cosϕysinθysinθl2sinϕysinθysinθl2.(127)

Then, we successfully converted ϕy ± ϕl/2 → ϕy, as intended. Finally, we can rotate between λ4 and λ6 to eliminate λ6 by defining

λ4λ6=Rθl2λ4λ6,(128)
=cosϕysinθy0.(129)

Similarly, we define

λ5λ7=Rθl2λ5λ7,(130)
=sinϕysinθy0.(131)

In order to obtain these expectation values for Gell-Mann parameters, we should renormalize the original basis operators of the su(3) Lie algebra to define

λ̂4=cosθl2λ̂4+sinθl2λ̂6,(132)
=cosθl2cosϕl2λ̂4+sinϕl2λ̂5+sinθl2cosϕl2λ̂6sinϕl2λ̂7,(133)
=00eiϕl2cosθl200eiϕl2sinθl2eiϕl2cosθl2eiϕl2sinθl20(134)

and

λ̂5=cosθl2λ̂5+sinθl2λ̂7,(135)
=cosθl2sinϕl2λ̂4+cosϕl2λ̂5+sinθl2sinϕl2λ̂6+cosϕl2λ̂7,(136)
=00ieiϕl2cosθl200ieiϕl2sinθl2ieiϕl2cosθl2ieiϕl2sinθl20.(137)

If we use these operators, the Gell-Mann parameters become

λ=sinθlcosϕlcos2θy2sinθlsinϕlcos2θy2cosθlcos2θy2cosϕysinθysinϕysinθy0036+32cosθy,(138)

such that we successfully removed λ6 and λ7. These parameters are equivalent to using photonic orbital angular momentum

=sinθlcosϕlcos2θy2sinθlsinϕlcos2θy2cosθlcos2θy2(139)

and hyperspin

y=12sinθycosϕy12sinθysinϕy16+12cosθy,(140)

which can be shown on two Poincaré spheres with the radiuses of cos2 (θy/2) and 1/2, respectively, instead of the original three spheres. This is consistent with the four degrees of freedom for orbital angular momentum and hyperspin, as confirmed previously (Table 3), and it is also expected from the rank-2 nature of the su(3) Lie algebra, which requires only two sets of the su(3) Lie algebra among three sets of (t̂,û,v̂). In practice, we do not know the angles of θl and ϕl, a priori, such that the angles are obtained from expectation values or experimental results.

Finally, we successfully embedded Gell-Mann parameters in SO(6) to renormalize

S=S1S2S3S4S5S6=λ1λ2λ3λ4λ5λ8=sinθlcosϕlcos2θy2sinθlsinϕlcos2θy2cosθlcos2θy2cosϕysinθysinϕysinθy36+32cosθy,(141)

which satisfies the conservation law of the norm,

i=16Si2=43,(142)

which was derived from the constant Casimir operator of Ĉ1.

3.4 Alternative coherent states

We used D̂2(v)θy and D̂3(v)ϕy to define an arbitrary state, but we can define an alternative coherent state using original bases of the su(3) Lie algebra by the expression

D̂8ϕy=expiλ̂8ϕy2,(143)
=expi13ϕy2000expi13ϕy2000expi23ϕy2(144)

and D̂5θy=D̂2(v)θy, as

|θl,ϕl;θy,ϕy=D̂8ϕyD̂3ϕlD̂2θlD̂5θy|ψ1=ei13ϕy2eiϕl2cosθl2cosθy2ei13ϕy2e+iϕl2sinθl2cosθy2e+i23ϕy2sinθy2.(145)

Using this coherent state, we obtain the Gell-Mann parameters as expectation values:

λ=sinθlcosϕlcos2θy2sinθlsinϕlcos2θy2cosθlcos2θy2cos32ϕy+ϕl2sinθycosθl2sin32ϕy+ϕl2sinθycosθl2cos32ϕyϕl2sinθysinθl2sin32ϕyϕl2sinθysinθl236+32cosθy.(146)

We embedded Gell-Mann parameters into SO(6) for this coherent state as shown previously. To achieve such a conversion, we need to transfer 3ϕy/2±ϕl/23ϕy/2, which appeared in λ4, ⋯, λ7, by confirming

cos32ϕy+ϕl2sin32ϕy+ϕl2=Rϕl2cos32ϕysin32ϕy,(147)

whose inverse becomes

cos32ϕysin32ϕy=Rϕl2cos32ϕy+ϕl2sin32ϕy+ϕl2.(148)

The rest of the calculations are exactly the same as those shown in the previous subsection. We can use the same renormalized operators of λ4̂λ5̂, while we remove λ6̂=0 and λ7̂=0. Then, the renormalized Gell-Mann parameters become

λ=sinθlcosϕlcos2θy2sinθlsinϕlcos2θy2cosθlcos2θy2sinθycos32ϕysinθysin32ϕy0036+32cosθy,(149)

which keep remain unchanged, while hyperspin becomes

y=12sinθycos32ϕy12sinθysin32ϕy16+12cosθy.(150)

This equation just corresponds to the change in the azimuthal angle, ϕy3ϕy/2, in the Poincaré sphere shown in Figure 3B. Consequently, we embedded Gell-Mann parameters in SO(6) as

S=S1S2S3S4S5S6=λ1λ2λ3λ4λ5λ8=sinθlcosϕlcos2θy2sinθlsinϕlcos2θy2cosθlcos2θy2sinθycos32ϕysinθysin32ϕy36+32cosθy,(151)

which also keeps the norm

i=16Si2=43,(152)

upon arbitrary rotations in the six-dimensional space in SO(6). In practical experiments, however, it is more complex, if we set up a rotator for D̂8ϕy, since three waves are involved rather than two waves. In conventional optical experiments, various splitters and combiners are prepared for two waves such that it is much easier to rely on SU(2) rotations, including D̂3(v)ϕy and D̂3(u)ϕy=expiê3(u)ϕy, such that we do not have to use the original bases of λ̂i for SU(3) states.

4 Embedding in SO(5)

For the complete description of the eightfold way to rotate the SU(3) states, Gell-Mann parameters in SO(8) are more useful for understanding the differences in phases and amplitudes among |L⟩, |R⟩, and |O⟩. On the other hand, SO(8) is larger to show the nature of the wavefunction, made of three complex numbers (C3) with its norm conserved to cover S5 in the Hilbert space.

We could successfully reduce the dimension of Gell-Mann parameters from SO(8) to SO(6) or SO(3) × SO(3) to represent SU(3) states, in terms of orbital angular momentum and hyperspin, as expectation values. On the other hand, we have only four parameters (θl, ϕl, θy, and ϕy), such that we can reduce one more dimension to represent S4 in SO(5).

Before proceeding further, we review the relationship between SU(2) and SO(3) for describing spin states or polarization states for photons [8, 9, 1121, 2427]. For polarization, we have two orthogonal states, such that a ray of coherent photons are described by SU(2) states, which require two complex numbers (C2). The SU(2) wavefunction was normalized for a fixed power density, such that one degree of freedom disappeared and the wavefunction covered S3 in the Hilbert space. In fact, according to the fundamental theorem of homomorphism [16], SU(2)/SU(1)SU(2)/U(1)S3, which means that the SU(2) wavefunction is equivalent to S3, except for the global phase factor of U(1). On the other hand, it is also well known that SU(2)/S0SU(2)/Z2SO(3), where S0={1,1} and Z2={0,1}. This means that if we neglect the impact of the global phase factor, such as those expected from the geometrical Pancharatnam–Berry phases [82, 83] in closed loops, the expectation values of SU(2) states are indicated on the sphere, represented by the SO(3) group. Consequently, the original topology of the wavefunction on S3 is reduced to the Poincaré sphere of S2 in expectation values.

Similarly, in SU(3), the fundamental theorem of homomorphism [15] leads to SU(3)/SU(2)S5. This means that we obtain a degree of freedom of SU(2) symmetry within SU(3) states, which maintains the states essentially equivalent to S5, as confirmed from the identity of v3 = u3 + t3 to allow two arbitrary choices of SU(2) states from three sets of SU(2) bases, (t̂,û,v̂). Similar to the SU(2) states, one of the degrees of freedom in S5 would be obtained from the global phase, such that we can represent the expectation values on S4 in SO(5).

However, we could not establish surjective mapping from SO(6) to SO(5) purely upon rotations using our bases of the su(3) Lie algebra because the expectation values of λi (i = 1, …, 7) cannot be larger than 1, while we needed to renormalize λ8 to combine with λ4 and λ5. Then, we focused on the conservation relationships of

λ12+λ22+λ32=cos2θy2,(153)
λ42+λ52+λ82=43cos2θy2(154)

and considered the following non-surjective mapping from SO(6) to SO(5) while we renormalize:

S=S1S2S3S4S5=λ1λ2λ3λ4λ5,(155)
=sinθlcosϕlcos2θy2sinθlsinϕlcos2θy2cosθlcos2θy243cos2θy2cosϕysinθy43cos2θy2sinϕysinθy,(156)

which preserve

λ42+λ52=43cos2θy2.(157)

We also confirm that the renormalized Gell-Mann parameters conserve the norm

i=15Si2=43,(158)

which is consistent with the constant Casimir operator. Consequently, expectation values are embedded on a compact Gell-Mann hypersphere of S4 in SO(5).

5 Discussion

5.1 SU(2) × SU(3) and higher-dimensional systems

However, we assumed that the ray is polarized such that the polarization state is fixed. We can control the polarization state using a phase-shifter and a rotator. We recently proposed a Poincaré rotator, which allows an arbitrary rotation of the polarization state by realizing SU(2) rotations in a combination of half- and quarter-wave plates and phase-shifters [63, 70, 75, 76]. If we use the Poincaré rotator for the ray with SU(3) states of vortices under certain polarization, we can realize SU(2) × SU(3), since spin angular momentum and orbital angular momentum are different quantum observables, such that a general state is made of a direct product state for spin and orbital angular momentum. We can also realize a state created by a sum of these states with different spin and orbital angular momentum states. For example, if we realize the SU(2) state of left and right vortices and assign horizontally and vertically polarized states, respectively, we can realize both singlet and triplet states by controlling the phase among two different many-body states.

Recently, the relationships between topology and polarization are being extensively studied in various forms of structured lights [6, 49, 76, 8488]. Three-dimensional polarization states [8486] are novel structured lights to realize knots and links in intensity profiles. It is also exciting that skyrmions were realized by combining spin and orbital angular momentum of photons [89, 90]. Our results suggest that photons have a higher order SU(N) symmetry by allowing various superposition states among orthogonal basis states of spin and orbital angular momentum.

5.2 Cavity QCD and photonic mesons

It is well established that a photonic crystal is an excellent test bed to explore a cavity quantum electro-dynamics (QED) in an artificial environment [91]. Here, we consider an analog to a cavity QED as a cavity QCD. We construct a one-dimensional cavity, for example, a Fabry–Perot interferometer, where |L⟩, |R⟩, and |O⟩ states are realized. The ray propagates in the cavity along z and is reflected back to propagate along the opposite direction of −z. The chiralities of spin and orbital angular momentum are reversed upon reflections [1521, 2427], such that the state along −z would be a conjugate state to the state along z. Consequently, we can construct multiplets, similar to mesons, made of quarks and anti-quarks [4, 5, 9, 10]. For quarks, an individual quark is very difficult to observe in experiments due to the strong confinements in composite materials of mesons and baryons. On the other hand, we expect an opposite behavior since photons trapped inside the cavity are difficult to observe as is, while photons escaping from the cavity are observed and analyzed using detectors. This corresponds to observing an individual quark, which is a ray of photons propagating at either z or −z. It is quite hard to observe the composite meson analog, which is realized inside the cavity, and it would be difficult to allocate detectors to observe photons propagating in the opposite directions at the same time, which would require a transparent detector. However, it is not essential to observe within the cavity since we can examine the state inside the cavity from the photons escaping from both ends. The cavity QCD experiments will allows us to explore SU(3) and SU(2) × SU(3) multiplets in a standard photonic experimental setup. If we distinguish each polarization state with different orbital angular momentum states as an individual orthogonal state, we can also explore SU(6) states, for example, and it will also be possible to investigate how symmetry breaking from SU(6) to SU(2) × SU(3) affects the photonic states by observing the corresponding expectation values of generators of rotations in a higher-dimensional space. Another remarkable difference in the proposed photonic systems with quarks is quantum statistics; quarks are fermions, and photons are bosons. Our analysis is quite primitive, such that some of our ideas could be applicable to fermionic systems. However, coherent photons out of lasers are quite easy to treat due to technological advances, while a macroscopic number of photons are coherently degenerate, which would be ideal for experiments that require coherent interference. As mentioned previously, phases and amplitudes of a wavefunction determine the crucial Gell-Mann parameters, similar to Stokes parameters for polarization. Polarization is a macroscopic manifestation of the nature of spin for photons, represented on the Poincaré sphere. A similar argument will hold for orbital angular momentum of coherent photons, and the Gell-Mann hypersphere can play a similar role in clarifying the SU(3) states for photons.

5.3 Correlation between SU(n) and SO(n2 − 1)

Finally, we discuss the relationship between SU(n) wavefunctions and expectation values in SO(n2 − 1). It is well known that the SU(2) wavefunction for spin is related to spin average values in SO(3), and therefore, the rotation in SU(2) is linked to the corresponding rotation in SO(3). This fact is also explained by the relationship SU(2)/Z2SO(3), claiming that the SU(2) is the twofold coverage of SO(3). In this paper, we discussed the relationship between SU(3) and SO(8). More generally, we show that a quantum mechanical average of an generator in SU(n) is related to a rotation in SO(n2 − 1) using an adjoint representation of the su(n) Lie algebra.

We assume that a generator of rotation in SU(n) is X̂a and the commutation relationship is [X̂a,X̂b]=icfabcX̂c [5]. In the aforementioned example of SU(3), this corresponds to X̂a=λ̂a. We consider that an initial SU(n) state of |I⟩ will be rotated by an exponential map of exp(iX̂aθ) with an angle of θ to be the final state:

|F=eiX̂aθ|I.(159)

Then, we consider how an average expectation value of X̂b in the initial state X̂bI=I|X̂b|I is transferred to the final state:

X̂bF=F|X̂b|F,(160)
=I|eiX̂aθX̂beiX̂aθ|I,(161)
I|1+iX̂aθX̂b1iX̂aθ|I+Oθ2,(162)
I|X̂b+iθX̂a,X̂b|I+Oθ2,(163)
δbccfabcθX̂cI+Oθ2,(164)
ceF̂aθbcX̂cI+Oθ2,(165)

where we assumed that θ is infinitesimally small and considered only the first order in the expansion, and F̂a is an adjoint operator, whose matrix element becomes (F̂a)bc=fabc, which is a matrix of (n2 − 1) × (n2 − 1). Therefore, the rotation of the wavefunction in SU(n) becomes the rotation of the corresponding expectation value in SO(n2 − 1), as expected.

We also checked its validity in the second order of θ as

Oθ2=θ22I|X̂a2X̂b|I+θ2I|X̂aX̂bX̂a|Iθ22I|X̂b2X̂a|I,(166)
=θ22I|X̂a2X̂b2X̂aX̂bX̂a+X̂b2X̂a|I=θ22I|X̂a2X̂b2X̂aX̂aX̂bicfabcX̂c+X̂a2X̂bicfabcX̂aX̂c+X̂cX̂a|I=θ22I|2icfabcX̂aX̂c|I,(167)
=iθ22cI|fabcX̂aX̂c+fcbaX̂cX̂a|I=iθ22cfabcI|X̂a,X̂c|I,(168)
=θ22cdfabcfacdI|X̂d|I,(169)
=θ22cF̂a2bcI|X̂c|I,(170)

and therefore, the aforementioned formula is also valid in the second order. Actually, this is a reflection of the differentiability of the Lie group, which was originally called an infinitesimal group. Once a formula is derived in an infinitesimal small value, it is straightforward to extend it to the finite value. In our case, we can repeat the infinitesimal amount of rotation with an angle of θ/N, while we can repeat N times, and we take the limit N as

X̂bF=limNc1F̂aθNbcNX̂cI,(171)
=ceF̂aθbcX̂cI.(172)

Therefore, we proved that the quantum mechanical rotation of the wavefunction in SU(N), which is given by Cn on S(n1) upon the normalization, will rotate the expectation value of the generator, which is a vector of Rn21 in SO(n2 − 1), using the adjoint operator of the su(3) Lie algebra.

6 Conclusion

In this study, we proposed to use photonic orbital angular momentum for exploring the SU(3) states as a photonic analog of QCD. We showed that the eight-dimensional Gell-Mann hypersphere in SO(8) characterizes the SU(3) state, made of left- and right-twisted photons and no-twisted photons. There are several ways to visualize the Gell-Mann hypersphere, and we calculated expectation values for the orbital angular momentum and defined hyperspin to represent the coupling between twisted and no-twisted states, which could be shown on two Poincaré spheres or one hypersphere in SO(6) or SO(5). We believe that the proposed superposition state of photons is useful for exploring photonic many-body states to gain some insights into the nature of the symmetries in photonic states.

Data availability statement

The original contributions presented in the study are included in the article/supplementary material; further inquiries can be directed to the corresponding author.

Author contributions

The author confirms being the sole contributor of this work and has approved it for publication.

Funding

This work was supported by JSPS KAKENHI (grant number: JP 18K19958).

Acknowledgments

The author thanks Prof I. Tomita for continuous discussions and encouragements.

Conflict of interest

Author SS is employed by Hitachi, Ltd.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

1. Stubhaug A. The mathematician sophus Lie - it was the audacity of my thinking. Berlin: Springer-Verlag (2002).

Google Scholar

2. Fulton W, Harris J. Representation theory: A first course. New York: Springer (2004).

Google Scholar

3. Hall BC. Lie groups, Lie algebras, and representations; an elementary introduction. Switzerland: Springer (2003).

Google Scholar

4. Pfeifer W. The Lie Algebras su(N) An Introduction. Berlin: Springer Basel AG (2003).

Google Scholar

5. Georgi H. Lie algebras in particle physics: From isospin to unified theories (Frontiers in physics). Massachusetts: Westview Press (1999).

Google Scholar

6. Cisowski C, Götte JB, Franke-Arnold S. Colloquium: Geometric phases of light: Insights from fiber bundle theory. Rev Mod Phys (2022) 94:031001. doi:10.1103/RevModPhys.94.0310015

CrossRef Full Text | Google Scholar

7. Dirac PAM. The principle of quantum mechanics. Oxford: Oxford University Press (1930).

Google Scholar

8. Baym G. Lectures on quantum mechanics. New York: Westview Press (1969).

Google Scholar

9. Sakurai JJ, Napolitano JJ. Modern quantum mechanics. Edinburgh: Pearson (2014).

Google Scholar

10. Weinberg S. The quantum theory of fields: Foundations volume 1. Cambridge: Cambridge University Press (2005).

Google Scholar

11. Stokes GG. On the composition and resolution of streams of polarized light from different sources. Trans Cambridge Phil Soc (1851) 9:399–416. doi:10.1017/CBO9780511702266.010

CrossRef Full Text | Google Scholar

12. Poincaré JH. Théorie mathématique de la lumière, Tome 2. Paris: Georges Carré (1892). Available at: https://gallica.bnf.fr/ark:/12148/bpt6k5462651m Accessed August 23, 2023.

Google Scholar

13. Jones RC. A new calculus for the treatment of optical systems i. description and discussion of the calculus. J Opt Soc Am (1941) 31:488–93. doi:10.1364/JOSA.31.000488

CrossRef Full Text | Google Scholar

14. Fano U. A Stokes-parameter technique for the treatment of polarization in quantum mechanics. Phy Rev (1954) 93:121–3. doi:10.1103/PhysRev.93.121

CrossRef Full Text | Google Scholar

15. Born M, Wolf E. Principles of optics. Cambridge: Cambridge University Press (1999). doi:10.1017/9781108769914

CrossRef Full Text | Google Scholar

16. Jackson JD. Classical electrodynamics. New York: John Wiley and Sons (1999).

Google Scholar

17. Yariv Y, Yeh P. Photonics: Optical electronics in modern communications. Oxford: Oxford University Press (1997).

Google Scholar

18. Gil JJ, Ossikovski R. Polarized light and the mueller matrix approach. London: CRC Press (2016). doi:10.1201/b19711

CrossRef Full Text | Google Scholar

19. Goldstein DH. Polarized light. London: CRC Press (2011). doi:10.1201/b10436

CrossRef Full Text | Google Scholar

20. Hecht E. Optics. Essex: Pearson Education (2017).

Google Scholar

21. Pedrotti FL, Pedrotti LM, Pedrotti LS. Introduction to optics. New York: Pearson Education (2007).

Google Scholar

22. Spreeuw BJC. A classical analogy of entanglement. Found Phys (1998) 28:361–74. doi:10.1023/A:1018703709245

CrossRef Full Text | Google Scholar

23. Shen Y. Rays, waves, su(2) symmetry and geometry: Toolkits for structured light. J Opt (2021) 23:124004. doi:10.1088/2040-8986/ac3676

CrossRef Full Text | Google Scholar

24. Saito S. Spin of photons: Nature of polarisation (2023). arXiv 2303 17112. doi:10.48550/arXiv.2303.17112

CrossRef Full Text | Google Scholar

25. Saito S. Quantum commutation relationship for photonic orbital angular momentum. Front Phys Sec Opt Photon (2023) 11. doi:10.3389/fphy.2023.1225346

CrossRef Full Text | Google Scholar

26. Saito S. Spin and orbital angular momentum of coherent photons in a waveguide. Front Phys (2023) 11:1225360. doi:10.3389/fphy.2023.1225360

CrossRef Full Text | Google Scholar

27. Saito S. Dirac equation for photons: Origin of polarisation (2023). arXiv 2303 18196. doi:10.48550/arXiv.2303.18196

CrossRef Full Text | Google Scholar

28. Gell-Mann M. The eightfold way: A theory of strong interaction symmetry (1961). doi:10.2172/4008239

CrossRef Full Text | Google Scholar

29. Gell-Mann M. A schematic model of baryons and mesons. Phys Lett (1964) 8:214–5. doi:10.1016/S0031-9163(64)92001-3

CrossRef Full Text | Google Scholar

30. Ne’eman Y. Derivation of strong interactions from a gauge invariance. Nuc Phys (1961) 26:222–9. doi:10.1016/0029-5582(61)90134-1

CrossRef Full Text | Google Scholar

31. Sakurai JJ. Advanced quantum mechanics. New York: Addison-Wesley Publishing Company (1967).

Google Scholar

32. Plank M. On the theory of the energy distribution law of the normal spectrum. Verhandl Dtsch Phys Ges (1900) 2:237–45. doi:10.1016/B978-0-08-012102-4.50013-9

CrossRef Full Text | Google Scholar

33. Einstein A. Concerning an heuristic point of view toward the emission and transformation of light. Ann Phys (1905) 17:132. Available at: https://einsteinpapers.press.princeton.edu/papers Accessed August 23, 2023.

Google Scholar

34. Allen L, Beijersbergen MW, Spreeuw RJC, Woerdman JP. Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes. Phys Rev A (1992) 45:8185–9. doi:10.1103/PhysRevA.45.8185

PubMed Abstract | CrossRef Full Text | Google Scholar

35. Padgett MJ, Courtial J. Poincaré-sphere equivalent for light beams containing orbital angular momentum. Opt Lett (1999) 24:430–2. doi:10.1364/OL.24.000430

PubMed Abstract | CrossRef Full Text | Google Scholar

36. Allen L, Padgett MJ. The poynting vector in Laguerre-Gaussian beams and the interpretation of their angular momentum density. Opt Comm (2000) 184:67–71. doi:10.1016/S0030-4018(00)00960-3

CrossRef Full Text | Google Scholar

37. Golub MA, Shimshi L, Davidson N, Friesem AA. Mode-matched phase diffractive optical element for detecting laser modes with spiral phases. Appl Opt (2007) 46:7823–8. doi:10.1364/AO.46.007823

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Bliokh K. Geometrodynamics of polarized light: Berry phase and spin Hall effect in a gradient-index medium. J Opt A: Pure Appl Opt (2009) 11:094009. doi:10.1088/1464-4258/11/9/094009

CrossRef Full Text | Google Scholar

39. Holleczek A, Aiello A, Gabriel C, Marquardt C, Leuchs G. Classical and quantum properties of cylindrically polarized states of light. Opt Exp (2011) 19:9714–36. doi:10.1364/OE.19.009714

PubMed Abstract | CrossRef Full Text | Google Scholar

40. Milione G, Sztul HI, Nolan DA, Alfano RR. Higher-order poincaré sphere, Stokes parameters, and the angular momentum of light. Phys Rev Lett (2011) 107:053601. doi:10.1103/PhysRevLett.107.053601

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Bliokh KY, Rodríguez-Fortuño FJ, Nori F, Zayats AV. Spin-orbit interactions of light. Nat Photon (2015) 9:796–808. doi:10.1038/NPHOTON.2015.201

CrossRef Full Text | Google Scholar

42. Barnett SM, Allen L, Cameron RP, Gilson CR, Padgett MJ, Speirits FC, et al. On the natures of the spin and orbital parts of optical angular momentum. J Opt (2016) 18:064004. doi:10.1088/2040-8978/18/6/064004

CrossRef Full Text | Google Scholar

43. Barnett SM, Babiker M, Padgett MJ. Optical orbital angular momentum. Phil Trans R Soc A (2016) 375:20150444. doi:10.1098/rsta.2015.0444

PubMed Abstract | CrossRef Full Text | Google Scholar

44. Bliokh KY, Bekshaev AY, Nori F. Optical momentum and angular momentum in complex media: From the abraham-minkowski debate to unusual properties of surface plasmon-polaritons. New J Phys (2017) 19:123014. doi:10.1088/1367-2630/aa8913

CrossRef Full Text | Google Scholar

45. Moreau PA, Toninelli E, Gregory T, Aspden RS, Morris PA, Padgett MJ. Imaging Bell-type nonlocal behavior. Sci Adv (2019) 5:eaaw2563. doi:10.1126/sciadv.aaw2563

PubMed Abstract | CrossRef Full Text | Google Scholar

46. Shen Y, Yang X, Naidoo D, Fu X, Forbes A. Structured ray-wave vector vortex beams in multiple degrees of freedom from a laser. Optica (2020) 7:820–31. doi:10.1364/OPTICA.382994

CrossRef Full Text | Google Scholar

47. Shen Y, Wang Z, Fu X, Naidoo D, Forbes A. Su(2) poincar’e sphere: A generalised representation for multidimensional structured light. Phys Rev A (2020) 102:031501. doi:10.1103/PhysRevA.102.031501

CrossRef Full Text | Google Scholar

48. Shen Y, Nape I, Yang X, Fu X, Gong M, Naidoo D, et al. Creation and control of high-dimensional multi-particle classically entangled light. Light Sci Appl (2021) 10:50. doi:10.1038/s41377-021-00493-x

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Zdagkas A, McDonnell C, Deng J, Shen Y, Li G, Ellenbogen T, et al. Observation of toroidal pulses of light. Nat Photon (2022) 16:523–8. doi:10.1038/s41566-022-01028-5

CrossRef Full Text | Google Scholar

50. Shen Y, Wang X, Xie Z, Min C, Fu X, Liu Q, et al. Optical vortices 30 years on: Oam manipulation from topological charge to multiple singularities. Light Sci Appl (2019) 8:90. doi:10.1038/s41377-019-0194-2

PubMed Abstract | CrossRef Full Text | Google Scholar

51. He C, Shen Y, Forbes A. Towards higher-dimensional structured light. Light Sci Appl (2022) 11:205. doi:10.1038/s41377-022-00897-3

PubMed Abstract | CrossRef Full Text | Google Scholar

52. Sotto M, Tomita I, Debnath K, Saito S. Polarization rotation and mode splitting in photonic crystal line-defect waveguides. Front Phys (2018) 6:85. doi:10.3389/fphy.2018.00085

CrossRef Full Text | Google Scholar

53. Bull JD, Jaeger NAF, Kato H, Fairburn M, Reid A, Ghanipour P. 40 GHz electro-optic polarization modulator for fiber optic communications systems. Proc Spie, Photon North (Spie) (2004) 5577:133–43. doi:10.1117/12.567640

CrossRef Full Text | Google Scholar

54. Goi K, Kusaka H, Oka A, Ogawa K, Liow TY, Tu X, et al. 128-Gb/s DP-QPSK using low-loss monolithic silicon IQ modulator integrated with partial-rib polarization rotator. In: Optical fiber communication conference (OFC). San Francisco: Optica Publishing Group (2014). p. W1I–2. doi:10.1364/OFC.2014.W1I.2

CrossRef Full Text | Google Scholar

55. Doerr CR. Silicon photonic integration in telecommunications. Front Phys (2015) 3:37. doi:10.3389/fphy.2015.00037

CrossRef Full Text | Google Scholar

56. Kikuchi K. Fundamentals of coherent optical fiber communications. J Light Technol (2016) 34:157–79. doi:10.1109/JLT.2015.2463719

CrossRef Full Text | Google Scholar

57. Guan B, Scott RP, Qin C, Fontaine NK, Su T, Ferrari C, et al. Free-space coherent optical communication with orbital angular, momentum multiplexing/demultiplexing using a hybrid 3d photonic integrated circuit. Opt Exp (2013) 22:145–56. doi:10.1364/OE.22.000145

PubMed Abstract | CrossRef Full Text | Google Scholar

58. Sotto M, Debnath K, Khokhar AZ, Tomita I, Thomson D, Saito S. Anomalous zero-group-velocity photonic bonding states with local chirality. J Opt Soc Am B (2018) 35:2356–63. doi:10.1364/JOSAB.35.002356

CrossRef Full Text | Google Scholar

59. Sotto M, Debnath K, Tomita I, Saito S. Spin-orbit coupling of light in photonic crystal waveguides. Phys Rev A (2019) 99:053845. doi:10.1103/PhysRevA.99.053845

CrossRef Full Text | Google Scholar

60. Al-Attili AZ, Kako S, Husain MK, Gardes FY, Higashitarumizu N, Iwamoto S, et al. Whispering gallery mode resonances from ge micro-disks on suspended beams. Front Mat (2015) 2:43. doi:10.3389/fmats.2015.00043

CrossRef Full Text | Google Scholar

61. Devlin RC, Ambrosio A, Rubin NA, Mueller JPB, Capasso F. Arbitrary spin-to-orbital angular momentum conversion of light. Science (2018) 358:896–901. doi:10.1126/science.aao5392

PubMed Abstract | CrossRef Full Text | Google Scholar

62. Saito S, Tomita I, Sotto M, Debnath K, Byers J, Al-Attili AZ, et al. Si photonic waveguides with broken symmetries: Applications from modulators to quantum simulations. Jpn J Appl Phys (2020) 59:SO0801. doi:10.35848/1347-4065/ab85ad

CrossRef Full Text | Google Scholar

63. Saito S. Poincaré rotator for vortexed photons. Front Phys (2021) 9:646228. doi:10.3389/fphy.2021.646228

CrossRef Full Text | Google Scholar

64. Angelsky OV, Bekshaev AY, Dragan GS, Maksimyak PP, Zenkova CY, Zheng J. Structured light control and diagnostics using optical crystals. Front Phys (2021) 9:715045. doi:10.3389/fphy.2021.715045

CrossRef Full Text | Google Scholar

65. Andrews DL. Symmetry and quantum features in optical vortices. Symmetry (2021) 13:1368. doi:10.3390/sym.13081368

CrossRef Full Text | Google Scholar

66. v Enk SJ, Nienhuis G. Commutation rules and eigenvalues of spin and orbital angular momentum of radiation fields. J Mod Opt (1994) 41:963–77. doi:10.1080/09500349414550911

CrossRef Full Text | Google Scholar

67. Chen XS, Lü XF, Sun WM, Wang F, Goldman T. Spin and orbital angular momentum in gauge theories: Nucleon spin structure and multipole radiation revisited. Phys Rev Lett (2008) 100:232002. doi:10.1103/PhysRevLett.100.232002

PubMed Abstract | CrossRef Full Text | Google Scholar

68. Leader E, Lorcé C. The angular momentum controversy: What’s it all about and does it matter? Phys Rep (2014) 541:163–248. doi:10.1016/j.physrep.2014.02.010

CrossRef Full Text | Google Scholar

69. Yang LP, Khosravi F, Jacob Z. Quantum field theory for spin operator of the photon. Phys Rev Res (2022) 4:023165. doi:10.1103/PhysRevResearch.4.023165

CrossRef Full Text | Google Scholar

70. Saito S. SU(2) symmetry of coherent photons and application to poincaré rotator. Front Phys (2023) 11:1225419. doi:10.3389/fphy.2023.1225419

CrossRef Full Text | Google Scholar

71. Naidoo D, Roux FS, Dudley A, Litvin I, Piccirillo B, Marrucci L, et al. Controlled generation of higher-order poincaré sphere beams from a laser. Nat Photon (2016) 10:327–32. doi:10.1038/NPHOTON.2016.37

CrossRef Full Text | Google Scholar

72. Liu Z, Liu Y, Ke Y, Liu Y, Shu W, Luo H, et al. Generation of arbitrary vector vortex beams on hybrid-order poincaré sphere. Photon Res (2017) 5:15–21. doi:10.1364/PRJ.5.000015

CrossRef Full Text | Google Scholar

73. Erhard M, Fickler R, Krenn M, Zeilinger A. Twisted photons: New quantum perspectives in high dimensions. Light: Sci Appl (2018) 7. doi:10.1038/lsa.2017.146

PubMed Abstract | CrossRef Full Text | Google Scholar

74. Saito S. Special theory of relativity for a graded index fibre. Front Phys (2023) 11:1225387. doi:10.3389/fphy.2023.1225387

CrossRef Full Text | Google Scholar

75. Saito S. Macroscopic single-qubit operation for coherent photons (2023). arXiv 2304 00013. doi:10.48550/arXiv.2304.00013

CrossRef Full Text | Google Scholar

76. Saito S. Topological polarisation states (2023). arXiv 2304.00014. doi:10.48550/arXiv.2304.00014

CrossRef Full Text | Google Scholar

77. Nambu Y. Quasi-particles and gauge invariance in the theory of superconductivity. Phys Rev (1960) 117:648–63. doi:10.1103/PhysRev.117.648

CrossRef Full Text | Google Scholar

78. Goldstone J, Salam A, Weinberg S. Broken symmetries. Phy Rev (1962) 127:965–70. doi:10.1103/PhysRev.127.965

CrossRef Full Text | Google Scholar

79. Higgs PW. Broken symmetries, massless particles and gauge fields. Phys Lett (1962) 12:132–3. doi:10.1103/PhysRevLett.13.508

CrossRef Full Text | Google Scholar

80. Anderson PW. Random-phase approximation in the theory of superconductivity. Phys Rev (1958) 112:1900–16. doi:10.1103/PhysRev.112.1900

CrossRef Full Text | Google Scholar

81. Schrieffer JR. Theory of superconductivity. Boca Raton: CRC Press (1971). doi:10.1201/9780429495700

CrossRef Full Text | Google Scholar

82. Pancharatnam S. Generalized theory of interference, and its applications. Proc Indian Acad Sci Sect A (1956) XLIV:398–417. doi:10.1007/BF03046050

CrossRef Full Text | Google Scholar

83. Berry MV. Quantual phase factors accompanying adiabatic changes. Proc R Sco Lond A (1984) 392:45–57. doi:10.1098/rspa.1984.0023

CrossRef Full Text | Google Scholar

84. Bauer T, Neugebauer M, Leuchs G, Banzer P. Optical polarization Möbius strips and points and purely transverse spin density. Phys Rev Lett (2016) 117:013601. doi:10.1103/PhysRevLett.117.013601

PubMed Abstract | CrossRef Full Text | Google Scholar

85. Kuznetsov NY, Grigoriev KS, Vladimirova YV, Makarov VA. Three-dimensional structure of polarization singularities of a light field near a dielectric spherical nanoparticle. Opt Exp (2020) 28:27293–9. doi:10.1364/OE.398602

CrossRef Full Text | Google Scholar

86. Intaravanne Y, Wang R, Ahmed H, Ming Y, Zheng Y, Zhou ZK, et al. Color-selective three-dimensional polarization structures. Light Sci Appl (2022) 11:302. doi:10.1038/s41377-022-00961-y

PubMed Abstract | CrossRef Full Text | Google Scholar

87. Forbes A, d Oliveira M, Dennis MR. Structured light. Nat Photon (2021) 15:253–62. doi:10.1038/s41566-021-00780-4

CrossRef Full Text | Google Scholar

88. Nape I, Sephton B, Ornelas P, Moodley C, Forbes A. Quantum structured light in high dimensions. APL Photon (2023) 8:051101. doi:10.1063/5.0138224

CrossRef Full Text | Google Scholar

89. Beckley AM, Brown TG, Alonso MA. Full poincaré beams. Optica (2010) 10:10777–85. doi:10.1364/OE.18.010777

CrossRef Full Text | Google Scholar

90. Shen Y, Martinez EC, Rosales-Guzmán C. Generation of optical Skyrmions with tunable toplogical textures. ACS Photon (2022) 9:296–303. doi:10.1021/acsphotonics.1c01703

CrossRef Full Text | Google Scholar

91. Joannopoulos JD, Johnson SG, Winn JN, Meade RD. Photonic crystals: Molding the flow og light. New York: Princeton Univ. Press (2008).

Google Scholar

Keywords: Gell-Mann hypersphere, SU(3), orbital angular momentum, coherent state, Photonic QCD, optical vortex, topological colour charge, Lie algebra

Citation: Saito S (2023) Photonic quantum chromodynamics. Front. Phys. 11:1225488. doi: 10.3389/fphy.2023.1225488

Received: 19 May 2023; Accepted: 14 August 2023;
Published: 04 September 2023.

Edited by:

Hao Jiang, Huazhong University of Science and Technology, China

Reviewed by:

Yijie Shen, University of Southampton, United Kingdom
Bernhard Johan Hoenders, University of Groningen, Netherlands

Copyright © 2023 Saito. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Shinichi Saito, shinichi.saito.qt@hitachi.com

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.