1. Introduction: The ‘Quantum First’ Program
If our world, as is widely believed, is quantum, then it is, in principle, representable as a single vector evolving in Hilbert space; thus, all spatiotemporal notions must be reducible to this description. This suggests a research program that has been variously called ‘Hilbert Space Fundamentalism’ [
1], ‘Mad Dog Everettianism’ [
2], ‘extreme Occam’s razor’ [
3], or, simply, ‘physics from scratch’ [
4]. Essentially, rather than trying to find quantized versions of classical notions, whether of spacetime or of classical fields, one instead starts with a quantum description, aiming to locate the appropriate dynamics within. We will thus think of it as simply the ‘quantum first’ program (cf. [
5]).
This is a formidable task. Hilbert space itself is exhaustively described by its dimension; a vector in Hilbert space contains virtually no information regarding its physical content—a single-qubit space might model an electron’s spin just as well as a current direction in a Josephson junction; a vector in this space might designate to spin up, down, or any superposition of these.
Yet, despite this apparent paucity of information, important advances have been made. Most significantly, a thriving research program has established surprising links between quantum mechanics and general relativity, outlining the possibility of deriving space and time, as well as their dynamics, from an abstract quantum description. Maldacena and Susskind have proposed the
conjecture [
6], conjecturing that quantum entanglement is equivalent to wormholes in spacetime. (However, it should be noted that, due to the quantum no-communication theorem [
7], these wormholes are necessarily non-traversable, and thus they preserve the locality underlying general relativity.) Indeed, the links appear to be so deep that Susskind has proposed the generalization of
to
, stating that ‘where there is quantum mechanics, there is also gravity’ [
8]. Hence, a quantum system does not merely come equipped with a geometry; rather, that geometry is dynamical, approximating general relativistic spacetimes in a classical limit.
The origin of this idea can be traced to work of Jacobson, who has shown that Einstein’s equations can be recovered from thermodynamics in the presence of black hole horizons [
9] (see also [
10,
11]). These arguments rely essentially on the area-law-scaling of the Bekenstein–Hawking entropy of black holes [
12]. This scaling is also observed, in many cases, for the entanglement entropy of quantum systems—indeed, it has been proposed that we explain black hole entropy as the entanglement entropy of degrees of freedom across the horizon [
13,
14].
Along a different thread, van Raamsdonk has suggested that entanglement may be responsible for spacetime connectivity [
15,
16]. In the AdS/CFT correspondence, the Ryu–Takayanagi formula [
17] relates the entanglement between different patches of the conformal field theory on the boundary of Anti-de Sitter space with the area of a surface partitioning the AdS space; if the entanglement between both goes to zero, then the space partitions into two disconnected regions.
Combining these perspectives, the linearized Einstein equations follow from ‘entanglement thermodynamics’ for perturbations of AdS spacetime [
18,
19]; while the bulk of this work assumes the AdS/CFT correspondence, Jacobson has shown that a derivation of the full Einstein equations is possible without assuming an AdS background, given a condition of entanglement equilibrium [
20].
A way to think about this is in terms of a graph, each node of which is itself an abstract element of quantum information—a qubit, or perhaps more generally, some finite-dimensional Hilbert space—that carries no spatial information itself, with the edges encoding the entanglement structure of the whole system. Geometry, then, must be reconstructed from this basic picture.
This is, essentially, the proposal made by Cao and Carroll [
21] (see also [
22]). Starting from a graph, as described above, they show that the metric structure of spacetime, as well as the linearized Einstein equations, can be obtained using an argument closely related to that of Jacobson.
Intriguingly, there appears a natural bipartitioning of the entanglement degrees of freedom in proposals such as the one in Ref. [
21]. Recall that the area law scaling of entanglement entropy plays a crucial role in recovering spacetime. Such a scaling is typically only observed in a very special class of states, such as the ground states of local gapped many-body systems [
23]. In Ref. [
21], a generalized notion termed ‘redundancy-constrained’ (RC) states is used, which correspond to graphs such that the entanglement entropy can be computed by summing over all the contributions of nearest-neighbor links over a given bipartition of the system into a region,
R, and its complement,
. Explicitly,
where
is the quantum mutual information between nodes
i and
j (see
Figure 1).
One can now consider states that only approximately satisfy this constraint. Then, the entanglement entropy receives a subleading correction
which can be associated with long-range entanglement or excited degrees of freedom over and above the vacuum (see
Figure 2). This subleading correction may then be associated with the entanglement entropy of an emergent gauge theory [
24].
Along similar lines, Cao [
25], following work by Harlow [
26], has proposed that we interpret the splitting of contributions to the entanglement entropy in terms of quantum error-correcting codes (QECCs)—specifically, QECCs protecting against subsystem erasure. There, the information in some state.
. is nonlocally encoded into
in such a way as to be protected against erasure of the subsystem,
. Thus, there exists a decoding unitary on
R:
where
is the generically entangled state of the remaining degrees of freedom that restores the encoded information,
. Thus, the full information encoded on
can be recovered from
R alone. This yields for the entropy of
:
where
can be interpreted as the geometric, area law contribution to the entropy, and
in turn can be interpreted as stemming from the emergent matter fields.
It is here that our investigation begins. If dynamical spacetime emerges from the area law contribution to the entanglement, it seems a natural proposal to investigate the emergence of matter degrees of freedom, together with their symmetries, from the entanglement yielding subleading corrections to the entropy—the ‘logical’, encoded state, , on the above view. However, we stress that this motivation is, at this point, heuristic; the discussion in the sequel is independent of this interpretation.
Our main result is then to show how to use a construction of Dubois–Violette and Todorov [
27,
28], as extended and clarified by Krasnov [
29], to derive the exact symmetry and degrees of freedom of (a generation of) Standard Model fermions from the entanglement properties of a three-qubit state. This proceeds by two routes: first, the internal symmetries of a left-handed half-generation are derived from the transformations leaving a preferred complex structure invariant. Then, via a kind of ‘dimensional reduction’, the three-qubit state yields the spacetime degrees of freedom of a fermion generation, together with the internal symmetries of the right-handed part.
One can think of this in the following way: Standard, complex quantum mechanics can, as noticed long ago by von Weizsäcker [
30] (whose ‘ur-theory’ [
31] might be considered the first example of a ‘quantum first’ approach), be associated with
-dimensional spacetime. In more modern times, it has been shown that the three-dimensionality of position space and the state-space of elementary quantum systems, and the necessity of their correspondence, follow from information–theoretical considerations [
32,
33]; indeed, even the Lorentz transformations can be obtained by considering the relation between different experimenters’ descriptions [
34].
These arguments can be generalized to quantum theory over the quaternions
, leading to a
-dimensional spacetime. Furthermore, extending to the octonions
, we obtain
dimensions [
35] (which are, not coincidentally, precisely the dimensions in which a supersymmetric Lagrangian exists [
36,
37]).
These spacetimes can then be broken down to
dimensions by simply singling out a preferred complex structure—that is, fixing a split of
into
and
into
. It is then the observation of Dubois–Violette, Todorov, and Krasnov [
27,
28,
29] that these splits are left invariant by
for the quaternionic case and
for the octonionic one. This can be compared to a ‘quantum first’ realization of the proposal due to Kaluza [
38] and Klein [
39,
40] that higher-dimensional spacetime theories can lead to gauge fields in lower dimensions, with the origin of the gauge symmetry being a
—respectively,
—action on the two- and three-qubit state space.
The appeal to higher division-algebraic generalizations of quantum theory has a long and distinguished history. Of note, Finkelstein, Jauch, Schiminovich, and Speiser provided the first example of quaternionic quantum theory in 1962/1963 [
41,
42]—the latter of which is remarkable for introducing a mass-generating mechanism for an
-gauge theory by means of a quaternion-imaginary dynamical field, playing much the same role as the Higgs field in the later explanation of electroweak symmetry breaking.
Applications of the octonions to particle physics date back to work by Günaydin and Gürsey, who first proposed using them to model the strong force [
43]. The attractivity of appealing to the (normed) division algebras—
,
, and
—comes from the observation that the unit complex numbers form the group
, i.e., the electromagnetic gauge group; meanwhile, the unit quaternions form
, i.e., the gauge group of the weak interaction. Of course, due to their nonassociative nature, the unit octonions form no group at all—but still, their automorphism group is the exceptional
, which contains
(the gauge group of the strong interaction) as a subgroup—in fact, the subgroup that preserves one octonion imaginary direction and thus singles out a preferred complex structure.
This has led to an alternative, algebraic approach to models of particle physics. Dixon, for instance, focuses on the algebra
(where the factor of
is technically redundant and only included for aesthetic purposes), and uses it to build a model incorporating the complete matter and force content of the Standard Model [
44]. More recently, Furey [
45,
46] has proposed her unified theory of ideals, drawing on actions of the Dixon algebra
on itself. Stoica [
47] has proposed a construction in which the Standard Model emerges from the complex Clifford algebra
, with the color- and electromagnetic symmetry appearing in the same way as in the model of Furey. Furthermore, of note here is the model of Morita [
48,
49,
50], giving a quaternionic version of the theory of electroweak interactions, and an octonionic QCD. For a more thorough review of this ‘algebraic’ approach to the Standard Model, see [
47].
The final piece of the puzzle then is the origin of quaternionic/octonionic quantum theory. Here, we appeal to work by Mosseri and Dandoloff [
51,
52] and Bernevig and Chen [
53]: entangled states of two qubits can be written in terms of quaternionic spinors—more accurately, the state space of a two-qubit system is given by the quaternionic Hopf fibration (just as the state space of a single qubit, the Bloch sphere, is the complex Hopf fibration). Likewise, three-qubit systems can be considered as octonionic spinors, whose state space is given by the octonionic Hopf fibration. This also provides the point of origin for the emergence of the Standard Model gauge group: the equivariance group of the octonionic Hopf fibration is
, whose subgroup respecting the splitting
is just
[
29].
Furthermore, the Hopf maps in each case are entanglement sensitive—disentangling each qubit results in the space decomposing into a product of individual qubit state spaces. The information within separable states is then fully representable within ordinary, empty -dimensional spacetime; but entanglement introduces additional, ‘internal’ degrees of freedom. The heuristic picture we propose then is one where -dimensional spacetime with its dynamics emerges from the area law contribution to the entanglement of a quantum state, while area-law-violating corrections can be viewed as ‘bubbles’ of higher-dimensional spacetime; upon reduction to dimensions, these can be interpreted as singling out a preferred tensor product decomposition, adding matter and gauge degrees of freedom.
The origin of the Standard Model symmetry and its attendant idiosyncrasies has been an active topic of research since its inception. Grand Unified Theories (GUTs) have proposed to embed its gauge group into a larger one, recovering its phenomenology after appropriate symmetry breaking; however, there is a large number of such schemes conceivable, with little guidance for choosing among them. The division algebraic approach reduces this embarrassment of riches somewhat, but even here, finding a unique model has been challenging, and the reason for their appearance is not well understood—we are yet to determine what, if anything, singles them out as relevant to the description of elementary particles.
Our main aim here is to provide a possible physical picture that offers some guidance. In particular, as we will see, the division algebraic description is singled out by the geometry of qubit Hilbert spaces, potentially significantly reducing the search space for theory-building. Furthermore, by building a bridge between the ‘quantum first’ and division algebraic programs, we hope that the insights of each may help stimulate research within the other, in particular by introducing the possibility of emergent matter (and gauge) degrees of freedom, but more generally by providing a unified framework for both.
The remainder of this article is structured as follows: In
Section 2, we introduce the basic template which the subsequent three sections will follow. First, we give a brief introduction to the scenario, starting with a single qubit; then, in
Section 2.1, we present the connection of complex quantum mechanics with
-dimensional spacetime. In the following
Section 2.2, we introduce the complex Hopf fibration, and show its relation to the state space of single qubits. Then, in
Section 2.3, we illustrate the importance of the group
as the complex Hopf map’s equivariance group.
In
Section 3, the construction is generalized to two qubits, and quaternions, showing the connection to
-dimensional spacetime (
Section 3.1), and the quaternionic Hopf fibration (
Section 3.2). In
Section 3.3, then, it is illustrated how the group
emerges from the subgroup of
, leaving a quaternion imaginary direction invariant. In
Section 4, the construction is analogously generalized to the octonions and three qubits,
-dimensional spacetime (
Section 4.1), the octonionic Hopf fibration, its relevance for three-qubit entanglement (
Section 4.2), and, finally, the emergence of
as leaving a choice of complex structure invariant (
Section 4.3).
In
Section 5, we will then speculate about an extension of the model to incorporate a full generation of Standard Model fermions by generalizing to the four-qubit case, which will also allow us to include massive states in our model, and yield a possible natural explanation for the chirality of the weak force. As the presented construction develops a correspondence between gauge fields and few-qubit Hilbert spaces,
Section 6 discusses the possibility of using it for simulations on current quantum computers. Finally, we conclude in
Section 7, and propose some directions of further development, such as the incorporation of a Higgs mechanism by taking account of the nontriviality of the embedding of the fiber in the Hopf fibration.
Outline of the Approach
The following
Section 2,
Section 3 and
Section 4 all follow a similar template, which we outline here for clarity. After recapitulating some elementary details on the mathematical description of the systems under consideration (single-, two- and three-qubit systems, respectively), we rehearse an argument that links their state-spaces to the symmetries of Minkowski spacetime of varying spatial dimension. This takes its leave from what von Weizsäcker has called ‘abstract quantum theory’ [
54]: that is, individual qubits are not considered as spatiotemporally located physical systems (e.g., electron spins)—rather, their purely abstract properties as linked to their mathematical description are the subject of examination. (Von Weizsäcker considered these as ‘elementary questions’ that are subject to empirical investigation.) The symmetries of Minkowski spacetime (i.e., the
-dimensional Lorentz group) then is constructed as automorphisms of the state space in each case.
Then, each system is described in terms of the Hopf fibration. In the single qubit case, this just yields the familiar Bloch sphere picture: the two complex parameters of a single qubit yield four real numbers whose squares sum to one, thus parametrizing a point on the sphere in four dimensions, . Since each qubit has an arbitrary phase, we can ‘neglect’ it to yield a point on the Bloch sphere (which is the space whose automorphisms yield the -dimensional Lorentz transformations). This realizes as with an fiber coming from the phase, which is the complex Hopf fibration. An analogous construction reduces the state space of a two- and three-qubit system to an , respectively, base space.
The interesting point here is that these two latter mappings are both entanglement sensitive: the additional dimensions of the base space characterize the entanglement between one qubit and the rest of the system; disentangling the states thus reduces either base to , the Bloch sphere of the single qubit. We can thus introduce a split in the degrees of freedom of the base, between those belonging to a single qubit, and those characterizing the entanglement to the other qubit(s). This motivates then looking at the transformations leaving this split invariant, i.e., those that do not ‘mix’ the entanglement- and qubit-degrees of freedom. These turn out to be for two qubits, and for three qubits, with in each case the acting on the single qubit in the base, and the remaining transformations on the degrees of freedom characterizing the entanglement.
The question remains of how we can relate this mathematical fact to physical reality. Taking the above construction as it is, we obtain exclusively the
internal symmetry of the Standard Model in the three-qubit case; indeed, any talk of spacetime symmetries could be bracketed entirely. While this is interesting, it remains to be seen how to reconnect to a spacetime description of a physical particle. This project is begun in
Section 5. There, we argue that there is an alternative perspective on this construction, which amounts to taking the degrees of freedom of the base qubit as being acted upon by the spacetime symmetries instead. In the case of the two-qubit state, we argue that this yields the spacetime degrees of freedom of a left- and a right-handed two-spinor, together with a
-gauge symmetry on the right-handed part. This yields a Kaluza–Klein-like construction, where spatial dimensions in a higher-dimensional space yield internal symmetries in the lower dimension.
Thus, each instance of this construction can be either interpreted as giving the internal symmetries of a left half-generation of fermions, or as the spacetime symmetries of both a right- and left-handed fermion, together with the internal symmetries of the right-handed part, lacking the -factor. Putting both together—which amounts, as the number of degrees of freedom is concerned, to adding another qubit—then provides all the right ingredients for the description of a single fermion generation.
4. The Octonionic Case: Three Qubits
The third step in this construction takes us to the octonionic case. Remarkably, this is also the final step: the octonions are the largest division algebra, and the octonionic one is the final Hopf fibration. The next algebra that can be constructed via the Cayley–Dickson process, the sedenions, is no longer a division algebra, as it possesses nontrivial zero divisors; hence, the Hopf map cannot always be defined.
There is thus a certain finality to this case. Having this construction naturally stop here then may go some way towards explaining why it is just this case—and not any one further ‘up the ladder’—that might play a role in nature. More complex states then should not be expected to bring anything fundamentally new to the table—or at least, nothing that yields novel gauge symmetries in the way described here.
We will not give an in-depth introduction to the octonion algebra here; instead, we refer the reader to Baez’ magnificent exposition of their manifold subtle, and sometimes surprising, connections with the broader mathematical landscape [
58]. Thus, it suffices to say, here, that octonions can be generated by taking a pair of quaternions and introducing an additional imaginary unit,
(where we identify
), yielding (following the conventions of [
53])
This gives the quaternionic triples
A salient property of the octonions is their nonassociative nature:
for
and
. Octonions are, however, still
alternative: for any two elements
, it holds that
Octonion conjugation takes
to
with
which again lets us define a norm
and a multiplicative inverse
.
An arbitrary three-qubit state can be written as
with
. The normalization condition entails that a three-qubit state can be represented by a point on the 15-sphere
. We now define the following four quaternions:
which we combine into the octonions
This particular choice avoids introducing an anisotropy on
[
53].
We can now introduce an octonionic analogue of the Pauli matrices
where
i runs from 2 to 8. This allows us to define an octonionic ‘Bloch vector’, whose coordinates are
with
i here running from 1 to 9.
4.1. ‘’ and -Dimensional Spacetime
Despite the nonassociativity of the octonions, the general picture described above is still valid. As before, we can take matrices of the form
to parametrize the light-cone in
-dimensional Minkowski space (recall that such matrices always have a well-defined determinant, Equation (
8)), on which the Lorentz transformations are realized via conjugation by matrices from ‘
’. However, for the octonions—being nonassociative in addition to non-commutative—it is not immediately clear what, exactly, could be meant by this group. Nevertheless, it can be shown that a construction exists that generalizes the complex and quaternionic case, and again allows us to think of (pure) octonionic ‘qubits’ as parametrizing the
-dimensional celestial sphere [
62] (s. a. [
58,
63]).
4.2. Three Qubits and the Octonionic Hopf Fibration
Due to the normalization
, an octonionic qubit defines a point on the sphere
. There again exists a map taking points of the sphere to the extended octonionic plane,
which can then be projected to
by means of
once more with
. As before, this is just the map taking
to the ‘Bloch-vector’ coordinates,
This realizes
as an
-fibration over an
base,
and, as in the previous case, the map reduces to a projection into the complex subspace of
in case the state is separable into single-qubit and two-qubit states. Hence, the base space
contains the (projective) state of one qubit, along with entanglement degrees of freedom, while the fiber contains the state of the leftover two qubits.
4.3. and
As with the previous two sections, there exists a generalization of the connection between the spin-groups and Hopf fibrations to the octonionic case. One can arrive at an analogous description of the group
in terms of matrices of the form given in Equation (
15), with
replaced by
, denoting left-multiplication by octonions. However, we can also again simply note that the octonionic Hopf fibration can be equivalently given as [
64]
Then, one must find the subgroup of
respecting a split
, that is, singling out an octonion imaginary direction. It was shown by Dubois–Violette and Todorov [
28] (see also [
27,
65,
66]), and put into a more explicit form by Krasnov [
29], that this is just the gauge group of the Standard Model,
.
We will not give the full argument here, and refer the reader to the cited literature instead. However, as in the quaternionic case, there is an intuitive way of viewing the group’s action on the base space of the Hopf fibration. Under the split , the coordinates split into two sets, from the singled-out complex subspace, and from the additional octonionic complex directions (the entanglement degrees of freedom). Here, the first set characterizes one qubit, while the second gives the entanglement with the other two. Thus, we must look at the transformations that do not mix these sets.
As before, the triplet
is acted on by rotations from
. We now use the embedding of
into
, given by
where
and
[
29]. With
, this then acts on the six remaining directions, as required.
Finally, from the above embedding, one sees that the
-spinor splits into two representations, with the single factor of
being associated with an
-singlet of
-charge
, and the
associated with a
triplet of
-charge
, both transforming as a doublet under
[
29]. Consequently, the octonionic qubit can be seen to contain precisely one generation of left-handed particles with the quantum numbers of the Standard Model, after being split into a leptonic part
L from the factor
and a quark part
Q from
:
4.4. Summary
We have outlined a correspondence between the symmetries of few-qubit states and the gauge symmetries of the standard model. To help intuition, it is useful to recapitulate the construction in a pictorial way, highlighting the action of the symmetry groups.
In
Figure 3, the iteration of the Hopf fibration, starting with the
containing the three qubit state, is depicted. Arrows indicate the coordinate axes of each space. At each stage, the colors in the fibration’s base indicate the actions of the symmetry group: within the
base, the
symmetry acts on the directions
, while the
permutes the remaining three coordinates. On the next level, the directions
are acted on by elements of
, with
yielding rotations in the subplane spanned by the other two directions.
Figure 4 shows the three qubit state, together with the qubit- and entanglement degrees of freedom, highlighting again the action of each symmetry. The dark gray patch represents
, containing one qubit and the degrees of freedom corresponding to its entanglement with the other two qubits in the base, and the other two qubits in the fiber. The
symmetry acts on the coordinates containing the qubit, with the
acting on the coordinates parametrizing the entanglement. Likewise, the
fiber contains in its base the second qubit, on which again
transformations act, and the entanglement of that qubit to the third one, that is itself subject to a
-action.
4.5. Further Development
As noted, the construction above terminates in a natural way with the octonionic case. However, of course, the universe is not limited to three qubits and their entanglement. Therefore, investigation into the properties of state spaces beyond three qubits seems a likely avenue for further developing the model.
Moving to four qubits, it is then tempting to appeal to the next algebra in the Cayley–Dickson sequence—the sedenions, . Of course, because of the existence of zero divisors, the Hopf map cannot be defined in a straightforward way. Hence, a fully analogous continuation of the sequence above does not seem feasible—which indeed may be an indication of the apparent relevance of this case in nature.
Nevertheless, it has been proposed that the sedenions could be utilized to quantify four-qubit entanglement in a similar way, with a single four-qubit state being a point on
[
67]. Furthermore, the discrete
automorphism of
has been argued to yield a natural realization of the triple-generational structure of the Standard Model [
68,
69,
70], providing essentially three copies of the octonion algebra.
As expected, no essentially new phenomenology (e.g., fifth forces) becomes apparent through enlarging the model; meanwhile, it seems nevertheless possible to capture further relevant structure of the Standard Model, such as most notably the generational structure.
5. Entanglement and Massive Particles
The above has given us a way to interpret the degrees of freedom contained within a single octonionic qubit as the left-handed part of one generation of the fermion content of the Standard Model, transforming under the gauge symmetry in the appropriate way, after singling out a complex direction and splitting into .
However, some ambiguity still exists about the coordinates : if we view the octonionic spinor as spanning the 9-dimensional heavenly sphere, these coordinates should be interpreted as the spatial coordinates upon dimensional reduction by means of singling out a special complex direction; yet, they are acted upon by the part of the Standard Model symmetry.
What comes to our aid is the observation that only half the particle content—the left-handed part—of the Standard Model is subject to the symmetry: the Standard Model is a chiral theory. Hence, if we instead view as spanning the 3-dimensional celestial sphere, only the symmetry remains, and the -spinor models a right-handed generation instead.
It then seems suggestive to move on to , which is indeed large enough to accommodate a full generation of Standard Model fermions.
However, let us first step back and observe that, so far, we have only considered pure states. These, as discussed above, give rise to light-like vectors (in
,
, and
dimensions, respectively). Hence, these can, at best, model massless particles propagating at the speed of light. However, we can appeal here to an observation due to Kiosses [
71] and Morikoshi [
72]: while a pure density matrix will have a Bloch vector of norm 1, a mixed-state density matrix defines a point at the interior of the Bloch sphere—hence, in the correspondence above, a time-like direction within the light-cone, in other words, the propagation of a massive particle.
Let us for the moment consider a single, unnormalized spinor
such that
,
. Its density matrix is
which yields as before
We can find a linearization of this equation of the form
where the repeated index is summed over.
Acting on this from the left with
recovers Equation (
45). This has the form of a Weyl equation in the momentum representation; hence, we have renamed our coordinates
to illustrate this interpretation.
Consider now a general two-qubit state of the form
with
.
Then, let us define the matrices
:
With these, we have again
for
Here, the
are the Bloch vector coordinates of a further (unnormalized) spinor
. With this, the four-spinor
defines a light ray in
dimensions.
We can then write in block–matrix form
If we now introduce
we obtain,
or, after multiplication of the first with
[
71],
where we see that
picks up a
phase factor from the rotation through the circle
by an angle
. This, we recall, is the
of the
-subgroup of
respecting the split
. If we thus want to keep this as a symmetry of the theory, we should require that it remains invariant under such transformations. Consequently, we see that the symmetry preserving the split indeed acts as a gauge transformation
, producing a chiral
gauge theory.
This identification of the
phase with electric charge is also made by Budinich [
73], where the above equation is interpreted as describing the proton and the neutron.
Absorbing the phase factor thus into
, this can be brought into a more standard form by introducing
together with the matrices
in the chiral representation
which finally yields the Dirac equation
where we have for ease of notation relabeled
, and otherwise
. Note that, to recover
, we must act on this from the left with
.
Now, introducing
and the left- and right-chiral projectors
we identify the left- and right-chiral components,
and
. Hence, the entangled two-qubit state can be interpreted as a left- and a right-handed spinor in
dimensions of mass
m, with
. For vanishing entanglement, we again obtain two separate instances of
, parametrizing two points, two lightrays on the celestial sphere in three dimensional space.
Consequently, we have two different interpretations of the two-qubit state: in the interpretation given in
Section 3, we obtain the internal symmetries of a left-handed fermion generation (without color degrees of freedom, i.e.,
), while the interpretation in this section yields the spatial degrees of freedom of a left- and a right-handed fermion, together with the internal
of the right-handed part.
Putting both together, then, yields all the right ingredients for a unified description of a full generation (again, without the color symmetry). As , this means that we need a three-qubit state; however, this is now interpreted in terms of two copies of the construction referring to the quaternionic Hopf fibration.
This then suggests an extension of these observations to the octonionic case. The move from to is equivalent to moving from a three-qubit to a four-qubit state. This gives rise to two copies of the octonionic Hopf fibration—which may be motivated by the fact that there is no fourth fibration; thus, the construction as presented in this paper finds its natural endpoint there.
As discussed in
Section 4.3, the coordinates after singling out a complex direction split into two sets,
and
, in the notation of this section, with the norm
made explicit. These obey
where the right side (giving the ‘mass term’ as in Equation (
54)) is acted on by the
into which the
part of the Standard Model symmetry is embedded, as in Equation (
41). Thus, we see that the octonionic qubit can be interpreted, analogously to the quaternionic one, as the left-handed part of one family of fermions transforming under the full Standard Model symmetry
, or as a right-handed part, together with the energy–momentum four-vectors of the left- and right-handed fermion, transforming under the residual
symmetry. Putting both together, then, seems well poised to yield the description of a full generation of fermions. However, we leave the full formulation of the model for future work.
We note here that, in this form, the model is also closely related to that proposed by Budinich [
73].
7. Conclusions
The present study draws inspiration and arguments from several disparate sources, aiming to integrate them into a coherent whole. These avenues are, as delineated above:
- ∘
The ‘quantum first’ program of deriving physics from scratch, starting from a quantum state in Hilbert space;
- ∘
The Kaluza–Klein-like emergence of forces and matter from dimensional reduction;
- ∘
The connection between spacetimes of dimensions , , and and the division algebras , , , and ;
- ∘
The division–algebraic approach to particle physics;
- ∘
The application of division algebras to entanglement theory.
In summary, the following picture suggests itself. The background of spacetime, together with its dynamics, emerges from large-scale properties of the area law contribution to the total entanglement entropy. Area-law-violating contributions, which can be viewed as due to an ‘encoded’ state in the sense of quantum error correction, on the other hand, yield gauge- and matter-fields. These can be viewed as what one might picturesquely call ‘bubbles’ of higher dimensional spacetime, which, upon reduction to the -dimensional background, yield the familiar gauge symmetries of the Standard Model, along with appropriately transforming matter fields, in a novel realization of the Kaluza–Klein mechanism.
In this vein, we have reviewed how a succession of tensor product spaces of elementary systems gives rise to a succession of constructions based on the Hopf fibration, terminating with the final octonionic case. Using only the postulate of singling out a preferred imaginary direction, motivated by the requirement of representing all higher-order systems within the
-dimensional spacetime associated with the elementary case of the single qubit, we have seen that the largest of these cases, as shown in [
27,
29,
65,
66], gives rise to the exact gauge group of the Standard Model, complete with the correct transformation laws for lepton and quark states (of a single handedness).
The construction can be interpreted in two different ways. In
Section 2,
Section 3 and
Section 4, the internal symmetries of a left-handed half-generation of fermions were constructed, starting from the equivariance group of the Hopf map.
Section 5 then introduces a way to think about the reduction as yielding the spacetime degrees of freedom of a Dirac fermion, together with the gauge symmetry of the right-handed half.
Consequently, by including one last doubling, taking us to four qubits in the Hilbert space , we sketched a possibility to include both left- and right-handed fermion states, as well as extending the description to massive particles, giving an intrinsic explanation of the chirality of the weak interaction as due to the fact that one set of directions for one of the two Hopf fibrations in this case is taken up as spanning the celestial sphere.
Thus, a state of four qubits appears to be a microcosm containing many of the ingredients necessary to build up our universe: we obtain the spacetime and internal symmetries of the right-handed half of one generation of massive Standard Model fermions in -dimensions, and the internal symmetries of the left-handed half. Apart from the possibility of model-building, the correspondence described here thus also yields a natural way of implementing physically relevant dynamics with state-of-the-art quantum computers.
Furthermore, as this construction reaches its natural end-point with the octonionic Hopf fibration, the discovery of additional particles or ‘fifth’ forces would not be easily accommodated; thus, they constitute strong evidence against it.
The above does not pretend to give a full theory. Its intent is merely to point out that many of the ingredients making up the most fundamental theory of particle physics, the Standard Model, seem to crop up in an—at first sight—entirely unrelated setting: namely the structure of entanglement in qubit Hilbert spaces. The characteristics of the Standard Model are often regarded as arbitrary, unsightly, and odd—their independent appearance in two unrelated settings, then, would seem to be doubly odd. Hence, one might venture the more parsimonious hypothesis that there is, in fact, a deep relation between the two areas, whose outline the present account has just begun to trace.
On the basis that this is indeed more than accidental, we have provided a sketch pertaining to how a connection between these fields might be realized, by incorporating ideas from the ‘quantum first’ program. Thus, the present proposal is best understood as attempting to provide a suitable physical background within which the seemingly disconnected strands outlined above fall naturally into place.
The construction as outlined so far, yielding the symmetries of -dimensional Minkowski spacetime and internal symmetries from the symmetries of the state space of three qubits upon a ‘dimensional reduction’ that can be interpreted as singling out a complex direction or a set of degrees of freedom as belonging to the ‘base’ qubit, is mathematically exact. Of course, this has no bearing on whether it is realized in nature; many ideas of great mathematical beauty have failed in the confrontation with experimental data. Thus, the physical relevance of these ideas, if any, remains open to investigation.
Furthermore, much work remains for putting this idea on the solid footing it needs to be fully convincing. For one, there still seems to be an element of choice in the question of which complex structure to preserve; several inequivalent options are discussed in Ref. [
29]. There are, of course, also many more elements of the Standard Model that our present discussion has not touched upon—such as the values of elementary masses/Yukawa couplings and interaction strengths, the mixing angles, and others.
One intriguing problem is the inclusion of a Higgs mechanism. Here, we offer up the following observation: the above discussion implicitly assumed that it is possible to associate a single distinguished imaginary axis with each point of the base manifold of the Hopf fibration. However, the nontriviality of the bundle means that this ultimately cannot be achieved. This is well known in the standard qubit case, where it represents just the impossibility of assigning a consistent phase to every point on the Bloch sphere.
A difficulty of this sort was already encountered by Finkelstein, Jauch, Schiminovich, and Speiser in their original formulation of quaternionic quantum mechanics [
42]. There, it was proposed that the ‘special’ quaternion imaginary direction should be promoted to a dynamical variable, thus introducing an additional scalar degree of freedom into the theory (a move that was generalized to the octonionic case by Casalbouni, Domokos, and Kövesi-Domokos [
80]). Furthermore, it was shown that, due to this addition, the gauge bosons of the resulting
gauge theory acquire mass, meaning that the quaternionic field assumes a Higgs boson-like role. It would be highly interesting to investigate this possibility for the present model.
Another salient question lies in the origin of the tensor product structure we have implicitly assumed in resolving an abstract quantum state into distinct qubits. In ordinary quantum mechanics, systems are largely individuated by their spatial relations—and thus, we assign a tensor product structure according to a split of a system in degrees of freedom:
over here versus
over there. However, in the picture outlined here, we do not have recourse to any a priori spatial structure. What then determines the tensor product structure of the Hilbert space? This is the
quantum factorization problem [
3], or the question of
quantum mereology [
81], which—unless one is willing to settle for considering the tensor product structure as an a priori input into the theory—will need a convincing answer to make the notions outlined so far fully well defined. Indeed, without an answer to this question, there is in general no fact of the matter regarding whether a system is entangled.
Recently, Stoica has argued that there is no unique way to reconstruct
space from the quantum state [
82]. If that is the case, then the present proposal will need to appeal to additional structure. One possibility might be to interpret the proposal in an epistemic way, with a preferred basis being picked out by the actually performed observations. This would put the observer into a central role, opening up a ‘first-person’ account of physics, similar to the proposal of ‘law without law’ by Müller [
83]. On this view, the basis—and, with it, the tensor product structure—would then be a necessary input to the model in order to yield a unique phenomenology.
There are many other intriguing possibilities for future development. For instance, while more complex states do not add more chapters to the story told above, that does not necessarily imply that they can contain nothing but additional copies of the above construction. Further degrees of freedom might exist that, while not having a neat decomposition in terms of ordinary particles as per the present construction, nevertheless add to the total mass–energy content of the universe. As the Standard Model particles can only account for ∼15% of this mass–energy content, one might then conjecture that these additional degrees of freedom have something to do with dark matter.