3.1. The SU(2) × U(1) × SU(2) “Fano Sub-Group” Symmetry and X-states
As mentioned above, an interesting sub-group, or sub-algebra, of the 15-generator SU(4) is provided by a subset of seven of them which plays a role in many physical systems. To identify them,
Table 1 shows that each of the 15 operators
can serve as a non-trivial center U(1), since it commutes with six others. Take as an example the operator
, or
, which we will use as a running example in later sections—we emphasize, however, that any of fifteen choices can serve. Its six companions in such a sub-group are
; that is,
. For charged spin-1/2 particles in an external magnetic field along the
z-axis, the Hamiltonian, with scalar couplings
and what are termed cross-coherences
, is an example of a physical situation with this sub-symmetry [
7]. It is realized in the CNOT quantum logic gate constructed out of two Josephson junctions [
56]. Another example is to take
, or
, as a center. Now, the other six are
or
; or, more compactly,
. That is, all three Pauli matrices of the first spin, along with their multiplication by (any) one of the matrices of the second, such as
, clearly provides six matrices that commute with this choice of center
. These two examples differ, however, in their quantum entanglement properties.
While they do not, as they stand, split into two sets of three that mutually commute with each other, the linear combinations
—that is, the triplet
and similar triplet with minus signs—indeed provide two sets of triplets that obey SU(2) commutation relations within themselves, while each member commutes with all three of the other set. The operators
behave like projection operators. The presence of a non-trivial center, along with the trivial unit center, leads immediately to such projection operators and a division of the space into two separate ones. These are termed, generically,
P and
Q; with
. For the purposes of unitary integration, since only commutation relations enter, the evolution
does indeed simplify into two independent factors, as in Equation (
4), which may be termed pseudo-spin SU(2)s [
7]. The prefix “pseudo-” is invoked because each member of the triplet no longer squares to unity, as with Pauli spinors, but into the overall commuting objects
, involving both centers. Note that, in more general contexts beyond our currently considered multiple qubits, the presence of a non-trivial center that squares to unity leads to a similar decomposition into orthogonal
P and
Q spaces. While all fifteen
lead to such a separation into two complementary projected spaces within a set of seven generators, only the nine involving both spins can accommodate quantum entanglement, as will be discussed further below. Explicitly, when
is the U(1) center, the two mutually commuting SU(2) triplets are
and
. They square to
, and again behave similarly to pseudo-spins, but cannot be written in the same factorized form of ⊗ of the two spins (as at the beginning of this paragraph) for the center
(or for any choice of center with only one of the spins).
The above discussion for operators and generators of a sub-group symmetry of SU(4) applies also to the states of a two-qubit system. In a matrix representation, they are again equivalent to 4 × 4 Hermitian matrices. The general
density matrix of pure or mixed states is characterized by 15 parameters, three real ones along the diagonal and six complex off-diagonal elements in a Hermitian matrix. It was natural in the original heuristic definition [
9] to denote those with only two non-zero off-diagonal entries, namely those on the anti-diagonal, as
X-states, from visual appearance, as in Equation (
3). There are now seven parameters in all, and this indeed provides an instance of the previous paragraph’s SU(2) × U(1) × SU(2) sub-group symmetry [
7]. Depending on the center U(1), the density matrix may or may not look like the letter X, but this symmetry perspective shows their commonality [
10]. In addition, under operations by members of that same set of seven operators, the X character of the physical system is preserved. This proves very convenient in many physical applications, by reducing the number of parameters and operators that we are required to handle. This accounts for the popularity of discussing the
X-states of a two-qubit system. It also points toward a natural extension; to higher multiples of qubits and of higher-dimensional qudits. We will take this up in
Section 3.4. We note that both the Lie-algebraic aspect—that the seven operators close under commutation—and their Clifford-algebraic structure—that they close under multiplication—are important [
10]. A mathematical description of the occurrence of such sets, due to Clifford groups, is in [
57,
58].
Analytical handling is also simplified, reducing to merely evaluating traces. This is because all the
are traceless and square to unity. With any such subset of seven
out of the
, the density matrix that remains invariant under their operations can be rendered as a linear superposition of them,
in analogy to that for a single spin,
. The seven real coefficients
in the sum in Equation (
6) parametrize
X-states, and are given by Tr[
]. Eigenvalues, and entanglement or other correlation properties, can be expressed compactly in terms of these coefficients [
10]. Furthermore, the triplet structure of the Lie–Clifford algebra is most conveniently captured geometrically, in
Figure 5 [
10,
55,
59,
60].
This figure, a beautifully symmetric pattern in its own right, occurs in projective geometry as the “Fano Plane” [
22], where it is described as the finite projective geometry PG(2,2). Arranging the seven operators at the vertices, the mid-points of the sides, and in the center of an equilateral triangle, the seven lines shown (including the inscribed circle) each pass through three points, providing the multiplication rule for those
. The center U(1) element occupies the center of the triangle. On the three un-arrowed median lines, all three operators mutually commute, so that the product of two gives the third, regardless of order. On the four arrowed lines, the operators mutually anticommute, so that the product of two gives
times the third, with plus (minus) signs along (against) the line of the arrow [
10,
55]. They may be termed “cyclic”, to contrast with “commuting” (also called isotropic in [
61] and orthogonal in [
62]) for the other set of three medians. The central element commutes, of course, with all six of the others. Each of those six has one ’conjugate’ element with which it commutes, and four with which it anticommutes. All of this can be read off by merely glancing at
Figure 5, which provides simple rules for their manipulation when calculating entanglement and discord [
10]. Indeed, this figure may be regarded as a direct extension of the “
cycle”, familiar from the multiplication or commutation of the Pauli operators for a single qubit. (Additionally, in vector product and quaternion multiplication rules.) It seems natural to call this
Figure 5 the “Fano triangle”, after its Italian geometer originator Gino Fano, and the sub-group symmetry of two qubits as the “Fano sub-group”, in addition to its designation as the Fano Plane of finite projective geometry [
22].
Besides the labels of the
that are shown,
Figure 5 also displays a binary and a quaternionic labelling to be discussed in subsequent sub-sections. For the purpose of this later discussion of quaternionic groups, note the placement of the
cyclic triplet on four lines, with one minus sign on the three edges and three minus signs for the circle; the cyclicity arrow is used in opposite senses between them. The extension to octonions, that have seven independent square roots of
and all seven lines arrowed, will be taken up at the end of
Section 3. In addition, the finite projective geometry PG( 2, 2), of seven points and seven lines, with a complete duality between these points and lines, differs interestingly from the finite Euclidean geometry EG(2, 2) of four points and six lines—this is obtained by dropping the midpoints in the diagram and terminating the median lines at the center. In finite geometries, only the points matter, and not the continuous lines connecting them. Furthermore, while two points define a line in Euclidean geometry, three do so in projective geometries. Another perspective is that the midpoints in
Figure 5 are points “at infinity”, just as the circle is a line at infinity in Euclidean terms, but projective geometry makes no distinction between points at infinity and “regular” points at finite location.
In Dirac language, a sub-set of seven operators in a Fano sub-group —such as, for example, the one at the beginning of this section with
as the center—are one A, three V, and three T of three indices different to the ones chosen in A. Other possibilities are one V plus three each of T and A, or P plus the six T. Now, classify the Dirac matrices into five groups: three
, three
, three
, three
(cyclic), and the three singletons
,
,
. In terms of these five classes, the 15 Fano sub-group sets are
,
. In each set of seven, the first entry, separated by a semi-colon, is the commuting U(1) element. A glance at the sets shows the involvement of the five classes in natural symmetric patterns. Interestingly, the division of 15 Dirac gamma matrices into five classes, four triplets or vector quantities and three scalar ones, parallels a geometric discussion where twelve are numbered numerically and three with alphabets a, b, and c [
62], or an analogous division among the generators of the group symmetry of the hydrogen atom [
23]. These connections between widely disparate problems may deserve further exploration.
While each
acting as a center produces 15 different
X-states, they differ in terms of their quantum entanglement which rests on the cross-correlation between the 1–2 and 3–4 sub-spaces of each spin of the two-qubit system in the canonical basis. When the
is a single spin operator in
Table 1, it does not mix these two spaces, and the projection operators provided by such a center do not describe entanglement. As an example, neither the
nor the
diagonal operators, with (1, 1, −1, −1) and (1, −1, 1, −1) entries along the diagonal, respectively, has entanglement, whereas for
with (1, −1, −1, 1),
X-states may display entanglement for certain values of the parameters in the density matrix. The first of the three diagonal operators acts as a unit operator of opposite sign in the 1–2 and 3–4 spaces of the two qubits; the second, similarly, within 1–3 and 2–4, which are the spaces of same spin orientation—up or down. It is the third, with center
and grouping 1–4 and 2–3, that pairs a qubit with another of the opposite spin. This simultaneous involvement of both particle and spin seems necessary for quantum entanglement. There is a striking correspondence to Dirac theory, where the lower 3 and 4 components of negative-energy electron states are reinterpreted as positive-energy positron states, with a similar spin-flip involved; the 4 interpreted as up spin and the 3 as down spin of the positron (Section 3.10 of [
24]). Thus, charge conjugation in that context is the analog of the entanglement of two qubits.
Such a sub-division of the 15 into 6 + 9, with single and double spin centers—the former always separable while the latter may admit entanglement—has also been discussed in detail from a finite geometric perspective [
58]. Fifteen different Fano planes are listed in their Appendix A. A particular type of geometric hyperplane, called a perp-set, is identified as a symplectic polar space of rank 2 and order 2, called W(3, 2). Depending on a unique quadric
of this space, that involves only non-trivial Pauli matrices, and whether the perp-sets intersect that quadric tangentially or transversally, one can distinguish the groups of 9 and 6, respectively. An
X-state set, such as
from the beginning of this section, is described in that language as one vector,
, that is orthogonal to the other six [
61] (in place of commuting in Lie-algebraic language). There will be further discussion in
Section 4, but note that the simpler perspective is provided by spin/qubit language in terms of the nature of the center, whether it is a single or double spin operator. In terms of Dirac matrices, it is the nine
as centers that exhibit entanglement, not the other six of the
and the singletons enumerated above.
3.2. The SO(5) “Desargues Sub-Group” Symmetry
Identifying sub-group symmetries other than the Fano sub-group of the previous section proceeds, again, through the commutator
Table 1 and by picking subsets that close under commutation as triplets. As mentioned, such a closed sub-algebra is all that is required for efficient construction of the evolution operator
. Thus,
is such a set. Atomic and molecular four-level systems often have Hamiltonians that involve only ten parameters, because of the dipole selection rules for transitions between the four states. As a result, two parameters characterize energy positions along the diagonal, as in the case of two identical qubits when they share the same energy separation, and four complex off-diagonal dipole couplings display such a sub-group symmetry. Together then, ten real parameters define such a system [
37]. It is the symmetry SO(5) of five-dimensional rotations. (Actually, it is the double covering group Spin(5), just as SU(2) is such a cover of SO(3); however, the distinction is unimportant for most of our discussion.)
As in the previous SU(2) × U(1) × SU(2) Fano sub-group, there are many such SO(5) that can be identified in
Table 1. Indeed, the above set of ten operators, when compared with a similar set in
Section 3.1, has the first six common, while the previous center of that Fano sub-group has been removed and replaced with the last four. This points to a systematic way of picking out the SO(5) examples, just as before for SU(2) × U(1) × SU(2). Again, for every
in
Table 1, pick the six other zeroes in that row or column and supplement by four others as required to close the sub-algebra. In terms of Dirac matrices, the above mentioned set of ten are four of the
, with indices 1, 2, 4, and 5, and their pairwise combinations. That is, three each of V, A, and T plus P. There is no involvement of the
. On the other hand, an alternative set of ten
, with the same initial six, but a different set of four (to replace the
element) is V+T with no involvement of
or any pseudoscalar aspect. Yet another example is A+T, in the language of Dirac matrices.
The pleasing geometric figure of an equilateral triangle with an inscribed circle, as well as seven line triplets of seven operators (
Figure 5), provided a rendering of the Fano sub-group in
Section 3.1. Similarly, the well-known Desargues diagram of projective geometry [
63,
64], placing ten points on ten lines, produces a rendering of the SO(5) sub-group which may, therefore, be called the “Desargues sub-group”. Various renderings are in [
55,
59], as well as in [
65], whose
Figure 5 refers to it as the “Petersen” graph, dual to a five-point “ovoid”. These objects are to be discussed further in
Section 4. Yet another geometric alternative, that follows the previous paragraph’s prescription of dropping the center in the set of seven and adding four others, is to remove the center and add a new vertex off the plane of the triangle, along with edges to the other three already-extant vertices and their corresponding three mid-points. This produces the next order simplex after the 2-simplex triangle, namely the 3-simplex tetrahedron, which represents the SO(5) Desargues sub-group. The ten lines are the six edges and four face circles of a tetrahedron, which will be shown and discussed in
Section 3.3.
We turn now to the evolution
, as per Equation (
4); for such an SO(5), the Hamiltonian with
is most naturally chosen as
, so that all handling is of 2 × 2 block matrices. For the ten-parameter
H, a convenient representation [
37] is
, where the ten arbitrarily time-dependent coefficients
form a
antisymmetric real matrix, keeping with the aspect of five-dimensional rotations. (We will use
and
with summation over repeated indices.) As noted, several quantum optics and multiphoton problems of four levels, driven by time-dependent electric fields, have such a Hamiltonian. It has also been considered extensively in coherent population transfer for many molecular and solid state systems [
66]. Casting this Hamiltonian in the form of Equation (
5), we have
The 2 × 2 block matrix
in
Figure 4 obeys a matrix Riccati equation, and the four entries (
) can be chosen as real:
;
. The equation for
takes the form [
37]
(As an alternative,
and
can also be rendered in terms of quaternions
). We can now construct a five-dimensional unit vector
out of the four real
z,
The nonlinear Equation (
8) in
z, takes on a simple, linear Bloch-like form,
which is the obvious analog of the Bloch equation of a single spin involving the cross product; that is, its higher-dimensional antisymmetric counterpart for rotations in five dimensions.
Solving this provides the first two factors in
Figure 4, as well as the effective Hamiltonian for the two diagonal blocks of the third factor, which may in turn be analyzed as the spiked Bloch sphere of a single SU(2) [
37]. As in the single-spin case, one can perform an inverse stereographic projection, now from the four-dimensional plane
to the four-sphere
. It provides a higher-dimensional polarization vector for describing two spin problems such as these. Hamiltonians possessing Spin(5) symmetry are, therefore, described by the geometrical picture of one
and two
spheres, along with two phases, as shown in
Figure 6, and the unitary evolution operator depicted as in
Figure 4. The former base manifold is similar, albeit of higher dimension than that of a single qubit, while the fiber is now a six-dimensional object and not a single phase. This can be pictured as two spiked Bloch spheres sitting on each point of the base manifold. Although a much larger, ten-dimensional, object than in the single-spin case, it is nevertheless an elegant and accessible generalization of
Figure 2 [
37].
3.3. SU(3) Sub-Groups and the Complete SU(4) Hamiltonian Involving all Fifteen Operators
Other sub-group symmetries of SU(4) include SU(3) with eight parameters; these can be thought of as two independent energy parameters along the Hamiltonian’s diagonal and three complex off-diagonal couplings. A general three-level system, embedded into four with the fourth level completely uncoupled, constitutes a trivial example of such an su(3) sub-algebra. However, less trivial examples can also occur. The
now contains two non-zero complex
z for a total of four parameters. The description [
53] of this four-dimensional manifold, as well as the remaining SU(2) and a U(1) phase, parallel the discussion of the general SU(4) that we now take up. Their manifolds are, however, more complex than spheres and one-dimensional fibers.
Moving beyond sub-groups to the full SU(4), consider an arbitrary
Hamiltonian with its entire complement of 15 operators/matrices. This
H is obtained by adding five additional terms to the previous Spin(5) Hamiltonian in
Section 3.2:
. This corresponds to the energy levels being arbitrarily positioned, as they would be in a general four-level (not necessarily two-qubit) system, and the two other couplings restored. Correspondingly, Equation (
7) acquires an additional term,
, in the diagonal
. Meanwhile, in
, the
are replaced by
. Thus, the full SU(4) amounts to a simple modification to the previously considered Spin(5), executed by adding a term, which is proportional to the unit operator, to the diagonal blocks and making the four
complex;
is absorbed as their imaginary parts [
37]. A full 6 × 6 antisymmetric collection of generators (see Appendix B of [
37]) may then be viewed as the symmetry SO(6) to be discussed further below.
Combined with the tetrahedron introduced above for the ten-parameter SO(5), the full fifteen-parameter SU(4) completes that figure by adding the four face centers and the body center. In turn, there are now 35 triplet lines that have Lie–Clifford algebra. Besides the previous 6 edges and 4 face circles, there now are 12 medians, 4 altitudes, 3 lines that link the body center to two midpoints, and 6 that link two face centers to a midpoint. The full tetrahedron is shown in
Figure 7 [
55]. A similar description is provided in [
67]. It is difficult to display all of the 35 lines, but
Figure 8 of [
62] makes a good attempt. Each of the four faces of the tetrahedron is now a Fano subgroup. A less immediately visual one is formed by the six midpoints and body centre of the tetrahedron; the seven lines now being the facial circles and the three lines connecting opposite midpoints to the body centre. The various Desargues sub-groups of 10 points and 10 triplet lines (6 edges and 4 face circles, as noted) with their labelling, and the ovoid consisting of five complementary points, are shown in [
67].
In constructing
, now for the full SU(4), the Riccati Equation (
8), now for complex
z, becomes
Just as the structure of Equation (
8) suggests that
and
, with suitable normalization, define a five-dimensional unit vector
in Equation (
9), we now have the same for a set of six complex quantities
. The nonlinear Riccati equation for the four complex
in Equation (
11) becomes a linear Bloch-like equation, as before, but now in six dimensions,
Once again, the
obey a first-order equation with an antisymmetric matrix which describes rotations in six dimensions. Since the 15
are real, the real and imaginary parts of the six
each obey a six-dimensional rotational transformation. The geometrical picture is of a Grassmannian manifold [
37], with details as follows. The six complex
obey three constraints, and thus describe a nine-dimensional Stiefel manifold St(6, 2, R) with SU(4)/[SU(2) × SU(2)] symmetry. This differs in a phase parameter from an eight-dimensional Grassmannian manifold Gr(4, 2, C), according to St(6, 2, R) ≃ Gr(4, 2, C) × U(1). Such a Gr manifold describes the four complex
. An alternative view, in terms of a five-sphere S
that has two orthogonal six-dimensional unit vectors radiating from its origin, each of which rotated within that sphere, is given in [
37]. Yet another description is to use what are called Plücker coordinates—six complex parameters which identify complex hyperplanes of Gr(4, 2, C), and are again discussed in [
37]. They have been further discussed for pure states of three qubits, which is also a system with 15 real parameters, and used [
68] for transforming between the so-called W and GHZ (Greenberger–Horne–Zeilinger) states that are familiar from quantum information theory [
69]. Such a transformation has also been discussed using a SU(2) × SU(2) sub-group symmetry [
70].
The occurrence of five- and six-dimensional antisymmetric equations in Equations (
10) and (
12) that are simple generalizations of the familiar vector Bloch equation for a single qubit, reflects the isomorphism between the groups SU(4) and SO(6) (more accurately, its covering group Spin(6): SU(4): SO(6) ∼ SU(4)
). They suggest a mapping between the 15 generators of these groups, as given in Equation (B1) of [
37]. That mapping also extends to the full set of operators that describe the non-relativistic hydrogen atom and its transitions in quantum mechanics [
23]. Interestingly, this correspondence between SU and SO symmetries is only present in the single- and two-qubit problems; it does not hold for any higher number of qubits. This fact rests on a curiosity in number theory called the Ramanujan–Nagell theorem, which states that the Diophantine equation which relates squares of integers and integer powers of two,
, has solutions for only five values of integer
n and
k [
71].
3.4. Larger Numbers of Qubits and Their X-States
Recognizing the symmetry group and structure behind
X-states permits simple generalization to a larger number of qubits. For this purpose, when we step back from two-qubits to a single qubit, any 2 × 2 density matrix looks necessarily like the letter X! It has, of course, SU(2) symmetry. The Fano sub-group symmetry of two-qubit
X-states, SU(2) × U(1) × SU(2), may be regarded as repeating the previous one-qubit SU(2), along with attaching the center U(1) in between. This view also fits into a 4 × 4 density matrix, interpreted as two 2 × 2 ones of 1–4 and 2–3 spaces in the canonical basis, with a mutual phase between these spaces. This specific breakdown into 1–4 and 2–3, as in Equation (
3) and the example in
Figure 5, as opposed to other 2 × 2 breakdowns, will be discussed further below. The generalization to a higher number of
q qubits is immediate. At each step, two copies of the previous step, with an added U(1) in between, provides the corresponding symmetry and set of
X-states. Thus, for a system of three qubits, the SU(2) × U(1) × SU(2) × U(1) × SU(2) × U(1) × SU(2) group of 15 generators, a sub-group of the full SU(2
= 8) of 63 generators, is the sub-group symmetry of the relevant three-qubit
X-states. This corresponds to seven real diagonal and four complex anti-diagonal elements in an
matrix. For any
q, the full symmetry group is SU(
with an explosively large number (
) of generators; however, the smaller
set provides the
X-states and their operators. These
q-qubit
X-states constitute the finite projective geometry PG(
q, 2) and, for
, can be geometrically represented by the same 3-simplex tetrahedron (in
Figure 7) that was used for all two-qubit generators. When compared to general mixed states, pure states additionally form a subset with fewer parameters; that number coincides with the
X-state values above, so that, again, the same figures can be used to represent them.
X-states of a two-qubit system entail seven parameters, whereas those with three qubits have fifteen; this latter value is also the number of generators or independent parameters of a full two-qubit system, suggesting an interesting nesting of projective geometries. PG(2, 2) of seven sits within PG(3, 2) of fifteen, with an additional eight members that may be seen as the vertices of a cube (
Figure 8). This triangle-plus-cube provides an alternative to the tetrahedron in
Figure 7, representing all fifteen operators and 35 triplet Lie–Clifford lines of the full SU(4), as discussed in
Section 3.3. The Fano triangle’s seven vertices and seven triplet lines are supplemented by the eight vertices and 28 triplet lines (12 edges, 12 medians of faces, 4 body diagonals) of the Clifford cube, as shown in
Figure 8. Such a cube has been used by computer scientists to represent a three-color (RGB) imaging scheme [
72]. Generalizing to higher
q, the PG(
q, 2) of
X-states has
points running through the sequence
, each with the previous number of lines, triangles, tetrahedrons, etc., as supplemented by the hypercube of
dimensions with
vertices: 4, 8, 16, 32,
…. On the other hand, the total number of parameters for a density matrix of
q qubits is the sequence
of PG(2
q-1,2), which is the number of all generators of SU(
) [
73,
74]. A binary labelling in the next sub-section provides another convenient addition to these algebraic and geometric perspectives of such sequences; for example, see Equation (
13). The identification of Clifford algebras with PG(
n, 2) over GF(2), the Galois field of order 2, for
has also figured in the mathematics literature, although from a different approach. This was found in [
75,
76], and independently by another group [
77,
78], who have also investigated the connection of these projective geometries to qubits.
3.6. A Quaternionic Correspondence
Hamilton’s quaternions, a four-dimensional division algebra, has long been regarded as an alternative to Pauli matrices for describing a single qubit with quantum spin-1/2 [
17]. Similar correspondences for multiple qubits also bring out group-theoretic links between the discrete/finite groups of multiple quaternions and the continuous SU(
) groups and generators of the qubit systems. As noted in the Introduction, quaternions are natural for geometric algebra, and Klein himself gives a nice description of three-dimensional rotations in terms of them [
79]—see [
80] for a pedagogical treatment. Maxwell too had advocated their use, although he himself wrote out his equations for electromagnetism in component form. However, subsequent developments in physics went in a different direction [
59,
64]. Vectors, and scalar and cross products of them, became the standard; geometric algebra and quaternions would not have separated a single product in this way. This would have had the merit of permitting division, which is not defined for two vectors in arbitrary directions. Many advantages would have accrued [
16,
79]. Interestingly, the consideration of SU symmetries in qubit systems, as discussed in this review, naturally highlighted connections [
59] to finite projective geometries and geometric algebra, along with correspondences between continuous Lie groups and discrete finite groups. We will now discuss this. In addition, in
Section 4, we will consider their development from the point of view of the purely geometric approach that is not motivated by quantum information. We will bring both views together on a common platform.
A variant of the more-economical binary labelling at the end of
Section 3.5, which fits better with the geometric diagrams of simplexes and a consistent build-up of their labels, is the one adopted for the Fano Plane in
Figure 5, the Clifford cube in
Figure 8, and the tetrahedron in
Figure 7. Consistently using round brackets for this binary (to distinguish from the earlier one with square brackets), start with the basic triplet of quaternions
or the Pauli matrices
. With the correspondence
of the two triplets that obey the same cyclic multiplication rules, the 1-simplex of a single qubit or quaternion is a line of three points. With all seven lines of
Figure 5 completely equivalent, including the inscribed circle, any of them can be the starting one. Choose the right edge, labelling the three points with a 2-binary of (01), (10) and (11), as shown. Again, for convenience of generalization, place
as the mid point, in assonance with the spin language convention that puts
Z as the diagonal object in quantum physics. The minus sign is for later consistent generalization of every further simplex; these are, again, obtained by introducing a step into a new dimension, with a new
k-like square root of minus 1 and its partner
-like generalization of the mid-point of a line to a facial, space/body, etc., center. They will always carry a string of 1’s in the round bracket binary label. For these reasons, the correspondence to quaternions is
.
Next, the basic quaternion group Q
[
81,
82] of the set
, with its Cayley table shown in
Table 3, can be set in correspondence with the Fano sub-group of seven operators and both placed on the Fano Plane’s equilateral triangle, as shown in
Figure 5. This step to the 2-simplex introduces a new vertex, which may be denoted
k and connected to the previous three points. It joins with the previous endpoints
, to provide the vertices of the Fano triangle. The center of the group,
, is the geometric center, and two new mid-points, “conjugate” negatives of the previous vertices
, arise at this step. Note the four cyclic lines (three edges and circle), shown arrowed, and three commuting median lines. As stated before, all lines are equivalent in a finite projective geometry. The circulation of the arrows is counter-clockwise around the edges and clockwise in the inscribed circle. Correspondingly, the extension from 2- to 3-binary labelling proceeds by adding an initial 0 to the points of the 1-simplex, and calling the new vertex (100). The other points then acquire labels by binary addition, with the commuting center as (111). In this manner, the Cayley table of Q
is placed on the Fano Plane.
Such a 3-binary
with round brackets labels the points in
Figure 5 on purely geometric grounds:
as the right edge,
as the left edge, and
as the base edge. The ascribing of quaternion
to the points is, to some extent, arbitrary—all points and lines are equivalent, and are related by simple geometric transformations, such as rotations in the plane. However, it is natural to place
and, equivalently (000)/(111), as the center of the triangle. Pairs of opposite signs then stand on opposite ends of the medians, which are conjugates under binary addition. As stated, it proves convenient for what follows to standardize the new vertex as
k-like when proceeding to generalize to higher dimensions or a larger number of qubits. Another perspective, provided by geometric and Clifford algebra, is that 1 is a scalar and
a vector, while
, formed out of antisymmetric pairwise products, is a bi-vector, and
is a pseudoscalar. Together, they are placed on the Fano Plane in
Figure 5.
With the equivalence of multiplication rules between quaternions and Pauli matrix generators of SU(2), a similar placement can be made of 1-qubit generators with
and
labelling points in
Figure 5. Such an assignment also occurs in Figure 1 of [
77]. However, a sign is irrelevant when dealing with generators of a continuous group. Instead, 2-qubit generators of the Fano sub-group provide an appropriate correspondence and natural identification with the 7-generator Fano sub-group and
X-states of
Section 3.1. Again, since any
can serve as a center, the example shown in
Figure 5 is for
as the center; the remaining six are commuting pairs at opposite ends of the medians. However, this match to 2-qubit generators is not to Q
as such, but to another closely related order-8 “co-quaternion” group that is isomorphic to the dihedral group D
. This will be discussed further at the end of this sub-section. Note that the change in sign of a quaternion has, as its counterpart, the duality exchange
.
The same quaternion and spin generator labelling in
Figure 5 is shown in
Table 2, and it must again be noted that it is for the specific example chosen—other centers and choices of the seven generators yield other correspondences between quaternions and generators. As noted, the example chosen corresponds to a physical set up of two spins in a magnetic field along the
z-axis, in addition to the four operators of magnetic interactions in the orthogonal
plane. Because any
can serve as a center (and is thus placed as the geometric center), their square bracket and double spinor names cannot be universally related 1:1 to the geometric round bracket labels [
55]. Furthermore, both sets of 4-binary can be rendered in base ten to run as a single number, so that points in
Figure 7 and
Figure 8 may be labelled from 1 to 15 as in [
67], and the 2-qubit generators as in [
55]; however, we have retained the
names due to previous usage in [
7,
8,
55].
The next step is to consider two independent quaternions (each commutes with all of the other set), which can be denoted by lower and upper case
and
, along with the unit element and all bilinear products. Taking all sixteen with plus/minus signs forms the 32-element finite group Q
. A half-way step is to include just one of the upper case elements, say
K, to acquire a group of order 16. This element could also be regarded as an ordinary complex square root of unity, so that we might call this the complex quaternion group. It has previously been referred to as a“bi-quaternion”, a term that Hamilton himself seems to have introduced [
81,
82]. (Bi-quaternion could more properly have been kept for the full order-32 group with all multiplicative combinations of two independent quaternions.) Its Cayley table is shown in
Table 4, and its 15 elements have been placed in 1:1 correspondence with the Clifford cube and tetrahedron of
Figure 7 and
Figure 8. As a group, it is C
Q
or
. (The pair
would also provide all 16 elements, but does not form a C
group; the pair
achieves this.)
In terms of the extension from the seven point/line triangle/2-simplex to this 8-vertex/28-line cube or 15-point/35-line tetrahedron/3-simplex, the previous 3-binary is extended to a 4-binary
and, geometrically, a new vertex (1000) or
K, in a new dimension is introduced. In the cube, that vertex is connected to the previous seven
as mid-points and extended to vertices
whereas in the tetrahedron, the seven lines from vertex to the base Fano triangle introduce seven new midpoints—three of them are face centers and one is a body center. Opposite vertices of the cube are of opposite sign:
and the center is the
of quaternions and the
generator of the example chosen. With the quaternions placed at midpoints of the edges at the lower corner, their negatives stand at the midpoints of the orthogonal face in keeping with their bi-vector nature noted earlier. In the tetrahedron, the new vertex
K introduces edge midpoints
, face centers the same triplet with minus signs, and its conjugate
is the body center, with
and
the face center of the Fano triangle remaining in the base.
Table 2 brings together all the alternative renderings of the 15 generators of the two-qubit system in terms of
, Pauli matrices, Dirac gamma matrices, binary, and quaternionic labels. Note the consistent pattern of each subsequent simplex having a
-like center with a string of 1’s as its binary representation.
The 4-binary has a natural language in terms of space-time
in physics, but could equally be rendered as alternatives, including four colors [
72] or four acoustic notes, to describe similar constructs of 15 basic objects; it has even been used [
55] in the context of a well-known combinatorics problem to be described below. In
Table 2, the association of this round-bracket 4-binary with the bi-quaternions is fixed. That is, the latter read off the former with the simple 1:1 association introduced in an above paragraph:
, i.e.,
, etc. and conjugates such as
. An initial 1 brings in
:
, etc. Multiplication of quaternions corresponds to binary addition. However, since any
denoted by its spinor and square bracket binary can be chosen as the center, there is no one-to-one link of them, the correspondence shown being for the specific choice of
as center and
.
There are other order-8 sub-groups of the full order-16 complex quaternion group. One is the set,
, forming the co-quaternion group D
with a Cayley table as shown in
Table 4. In a standard minimal notation [
82] for D
, it can be rendered in terms of two parameters as
. With the same labelling as in
Table 2, the set of seven generators are
with
as center. Correspondingly, the physical system is now of two spins in a magnetic field in the
y-direction, with four magnetic coupling operators in the orthogonal
plane. Another is the set of eight elements
which is
or C
C
. It corresponds, for the example in
Table 2, to the set
with
as center; that is, two spins in a magnetic field in the
x-direction with coupling terms in the orthogonal
plane. Differing only in a renaming of the magnetic field in terms of the
directions, they share the same entanglement and other physics. All these sets show the same
duality noted above. The full set of sixteen elements of a complex quaternion, or the generators of SU(4) in
Table 2, have a Cayley table shown in
Table 5. They may be viewed as direct products of C
with the order-8 groups, whether Q
or D
.
As noted, the quaternion labels placed on the vertices in
Figure 5 are arbitrary, given the natural geometric symmetries of the triangle such as rotations through multiples of
. The independent placement of the qubit generators is also arbitrary, as is their correspondence to the quaternions. Given that any of the 15
can serve as a center and define the
X-states of the Fano sub-group, there are as many choices. However, entanglement properties differ; only the nine
involving both qubits accommodate quantum entanglement. Their set of seven generators has two single spin generators, the
and
, and five two-spin operators. All square to unity. For the six single-spin centers, a similar set is composed of the three single operators of the other spin and the three products of them with the center. With these generators, a multiplicative sign or constant, such as
, is irrelevant and, as observed, the
satisfy all the multiplication rules of quaternions—including that they square to
. Turning to the order-8 groups, the quaternion group Q
has six
along the diagonal in its Cayley
Table 3. Meanwhile, the co-quaternion group in
Table 4, that is isomorphic to the dihedral D
, has two, and the C
C
has four (the upper left 8 × 8 block of
Table 5). Thus, while all three of them can be associated with
X-states, it is the co-quaternion which best matches the entangled class in having two of the seven square to −1. Those could be set to match in the set with center
; then, the two
and
, with
factors, would square to −1.
Thus, the choice made in
Figure 5,
Figure 7 and
Figure 8, and
Table 2 puts
as the two
and
, with a corresponding co-quaternion
. This set is trivially different from the one displayed in
Table 4. For this choice, all rows of
Table 2 can be retained unchanged, especially the binary and qubit generators; however, in the last row we must multiply entries involving
by
K with
. In
Figure 5,
Figure 7 and
Figure 8, we similarly have to multiply
by
K, and leave the other points unchanged. On the other hand, had we chosen the quaternion
, the entries in
Table 2 could be left unchanged, but this would correspond to the
X-state set
with
as center. This would be at a slight conflict with the convention in physics of choosing
z as the magnetic field direction or quantization axis. As stated, with no unique correspondence between the three labelling systems, binary, quaternion, and qubit generators, it is partly convention and partly aesthetics dictating the choice made in
Table 2 and in our figures. The choice made in
Table 2 is to tie the square-bracket 4-binary to qubit generators and the round bracket to bi-quaternions, with
as the center of the former matched to the
center of quaternions.
An interesting connection can also be established with octonions, the only other division algebra besides reals, complex numbers, and quaternions [
64]. With seven independent square roots of −1, they can also be laid on the seven points of
Figure 5 but with crucial differences. With a cyclic triplet
replacing
in order, and the seventh square root
s placed at the center, these seven independent imaginaries
have seven cyclic lines on the triangle
. All lines are now arrowed with the same circulation sense for the edges and the circle, unlike in
Figure 5 which has them opposite, and medians are now also arrowed from midpoint to vertex. This turns out to be crucial, since octonionic multiplication is no longer associative, which was the case with the previous three division algebra. This means that opposite circulation is a necessary feature of the quaternions in
Figure 5, in contrast to octonions. Interestingly, the seven quaternionic triplets can also be depicted on the cube in
Figure 8, with the unit element at the lower left corner,
on the three connected vertices to it,
at their opposite corners, and
s at the body center. An alternative placement of the seven at the seven corners of the cube is in [
64].
This process of counting how many −1 occur along the diagonal of a Cayley table, that arises naturally in our discussion with qubit generators, also has a bearing on further extension. There are no more division algebras beyond octonions to place on all points after the 2-simplex of
Figure 5. In the 3-simplex tetrahedron of
Figure 7,
, eight of them square to −1. Fifteen square roots of −1, placed at each of the points, represent what are called “sedenions” [
83]. Products of pairs of them as 35 triplets can be specified and, as with octonions, the products are not commutative or associative. That loss of associativity in multiplication precludes, of course, matrix representation. Physics has seen little use of octonions or sedenions. (See, however, [
84,
85]. In addition, octonions do allow what is termed “limited associativity” [
64]). However, from the 35 triplets we have discussed for the tetrahedron in either quaternionic or spinor language and a Cayley table such as
Table 5, one can build a Cayley table for sedenions as in [
83]. Indeed, a more symmetric arrangement than in [
83] is to group the 35 triplets into seven columns of 5 rows each, with all 15 elements occurring once and only once in each column for the Kirkman arrangement of schoolgirls, as discussed below in
Section 4. Note that a proper Cayley table requires the 15 square roots with both plus/minus signs, along with
, and is a group of order 32. As an alternative to the Cayley–Dickson construction in [
83], the higher
q-qubit simplexes provide another route to constructing these hypercomplex numbers, including sedenions and beyond, and associating them with finite projective geometries.
The correspondence to quaternions and higher string binaries for more than two qubits proceeds naturally. Each further qubit in a
q-qubit sequence introduces a new initial entry of 1 in the string with a new independent
K-like entry and new vertex in the next simplex—the previous simplex’s points are assigned an initial 0.
appears with
as the new simplex’s body center, just as the mid-point, face center, tetrahedron’s body center, did for 1-, 2-, 3-simplex of 1-, 2-, 3-qubit systems, respectively. This gives a geometric realization in simplexes of
X-states of
q-qubits. A somewhat different approach was adopted in [
86], which followed the Cayley-Dickson
-dimensional algebra with imaginary units
encoded in PG(
, 2), and triads of points with
as lines. Binomial configurations
are then identified with octonions for
, sedenions for
, and higher
–nions. Again, the PG(
, 2) is a
-dimensional projective space over Galois field GF(2), as noted at the end of
Section 3.4, and the
are isomorphic to Grassmannian G
. While [
78] noted that there is no “neat picture” for these higher 2
-nions, the
N-binary string and simplex schemes discussed above provide a unified picture of all of them and associate with the
X-states (or, equivalently, pure states) of
N-qubits. Additionally, in correspondence to quaternions, one new independent
K-like imaginary unit is added at each step, in order to acquire the simplex of the next-highest dimension.