We show that the standard Heisenberg algebra of quantum mechanics admits a noncommutative differential calculus Ω1 depending on the Hamiltonian p2/2m + V(x), and a flat quantum connection ∇ with torsion such that a previous quantum-geometric formulation of flow along autoparallel curves (or “geodesics”) is exactly Schrödinger’s equation. The connection ∇ preserves a non-symmetric quantum metric given by the canonical symplectic structure lifted to a rank (0, 2) tensor on the extended phase space where we adjoin a time variable. We also apply the same approach to obtain a novel flow generated by the Klein–Gordon operator on Minkowski spacetime with a background electromagnetic field, by formulating quantum “geodesics” on the relativistic Heisenberg algebra with proper time for the external geodesic parameter. Examples include quantum geodesics that look like a relativistic free particle wave packet and a hydrogen-like atom.
I. INTRODUCTION
The idea of geometry extended to a possibly noncommutative “coordinate algebra” A has been extensively developed since the 1980s and now has an accepted role as a plausibly better description of spacetime (i.e., “quantum spacetime”) that includes Planck scale effects. There are various approaches and we will use particularly the constructive approach in Ref. 1 and references therein, based on a chosen differential graded algebra (Ω, d) of “differential forms,” a quantum metric g ∈ Ω1 ⊗ AΩ1 and a quantum Levi–Cività connection ∇: Ω1 → Ω1 ⊗ AΩ1 to build up the “quantum Riemannian geometry.” The ∇ here is a bimodule connection in the sense of Refs. 2 and 3. This approach is complementary to the well-known Connes’ approach where the noncommutative geometry is encoded in a spectral triple4 or abstract Dirac operator as starting point. The two approaches can be compatible and with interesting results where they meet.5,6 The bimodule approach has been used to construct toy models of quantum gravity7–9 and more recently to generate particle masses for scalar fields via a Kaluza-Klein mechanism with noncommutative “extra dimensions.”10
In the present paper, we apply the powerful machinery of quantum Riemannian geometry to the more obvious context of ordinary quantum mechanics and quantum theory. Here the noncommutativity parameter will not be the Planck scale but just the usual ℏ. The noncommutativity inherent in quantum theory has long been one of the motivations for results in operator algebras in general and noncommutative geometry in particular, and the role of the latter in actual quantum systems has already been noted in Connes’ approach, for example to understand the quantum Hall effect.11 The role of the quantum Riemannian geometry formalism,1 however, has not been explored so far in this context but makes sense once we note that the “quantum metric” need not be symmetric, so can be equally applied to objects which classically would be antisymmetric. Indeed, we will be led to a “generalised quantum metric” on an extended phase space with time adjoined and which quantises an antisymmetric tensor related to the symplectic structure. This generalised quantum metric will also be degenerate and ∇, although compatible with it, will have a small amount of torsion, both features relating to the extra time direction. Thus, there are some differences but in the broadest terms we will effectively formulate ordinary quantum mechanics somewhat more in the spirit of gravity, rather than the more well-studied idea of formulating gravity in a quantum manner.
We will make particular use of a notion of “quantum geodesics” with respect to any bimodule connection ∇, as recently introduced and studied in Refs. 12–14. The preliminary Sec. II A provides the algebraic definition particularly of quantum geodesic(s) via the notion of a “geodesic A-B bimodule” E where is the geodesic-time coordinate algebra. One choice of E recovers in the classical case a single classical geodesic in a manifold, while in Sec. II B another choice of E recovers classically a dust of particles with density ρ, where each particle moves along a classical geodesic. The tangent vector to all these particles will be a vector field X obeying an autoparallel “geodesic velocity equation.” The actual particle flows are then given classically by exponentiating the vector field X to a diffeomorphism of the manifold, while the natural way to do this in our algebraic formulation turns out to be a corresponding flow equation not for ρ but for an amplitude ψ, where ρ = |ψ|2. This formalism then makes sense when the coordinate algebra of the manifold is replaced by a noncommutative algebra A, i.e., in noncommutative geometry. For our purposes now, we need to go further and Sec. III introduces a new choice of E in which the “geodesic flow” takes place more generally on a representation space of an algebra A rather than on A itself. We can then apply this in Sec. IV to A the Heisenberg algebra in the Schrödinger representation, allowing us to express the standard Schrödinger equation for a Hamiltonian h = p2/2m + V(x) as a quantum geodesic flow. The new result here is not the flow equation, which is just Schrödinger’s equation, but the noncommutative geometric structures that we find behind it, including a generalised quantum metric G, a compatible vector field X and a bimodule connection ∇. Moreover, while the Schrödinger representation necessarily entails the usual baggage of quantum mechanics, the noncommutative geometric structures themselves make sense at an algebraic level.
Note that this Sec. V is driven by the quantum algebra as a natural relativistic version of Sec. IV, and is very different from previous discussions of proper time in the Klein–Gordon context, such as Ref. 20 where the proper time and rest mass come from a canonically conjugate pair of observables. We illustrate our approach on the easy case of a free particle in 1 + 1 Minkowski space, where we analyse a proper time wave packet centred around an on-shell Klein–Gordon field (Example 5.5), and we also outline a proper time atomic model similar to a hydrogen atom.
Section VI rounds off the paper with a self-contained Poisson-level extended phase space formalism as suggested by our results of Sec. IV at the semiclassical level. This helps to clarify the geometric content of our constructions and also provides the physical meaning of ∇ as infinitesimal data for the quantisation of the differential structure in the same way as a Poisson bracket is usually regarded as the data for the quantisation of the algebra. Some concluding remarks in Sec. VII provide directions for further work.
II. PRELIMINARIES: ALGEBRAIC FORMULATION OF GEODESICS
Here, we give a minimal but self-contained account of the algebraic set up of differentials and connections and the formulation of quantum geodesics introduced in Ref. 12 in terms of A-B bimodule connections.1 A possibly noncommutative unital algebra A equipped with an exterior algebra (ΩA, d) will play the role of a manifold, and expresses a geodesic time parameter t with its classical differential dt. The formalism also allows for more general and possibly noncommutative B and (ΩB, d), but we will only need the classical choice in the present paper. Proposition 2.1 is a general version of the classical case treated in Ref. 12, Proposition 2.2 is essentially in Ref. 12 but reworked for right connections (which is needed to mesh later with conventions in quantum mechanics), while Corollary 2.3 is new.
A. Algebraic set up and the case of a single geodesic
Our first task for an algebraic version is the differential structure. If A is any unital algebra, we define a “differential structure” formally by fixing a bimodule over A of 1-forms. This means a vector space where we can associatively multiply by elements of A from either side and a map sending a “function” to a “differential form” obeying the Leibniz rule d(aa′) = da.a′ + a.da′. In the -algebra case over , we require to also have a -operation for which (a.da′)* = (da′*).a* for all a, a′ ∈ A. One normally demands that is spanned by elements of the form ada′ for a, a′ ∈ A, otherwise one has a generalised differential calculus. Any then extends to an exterior algebra ΩA with product denoted ∧ and exterior derivative increasing degree by 1 and obeying a graded-Leibniz rule and d2 = 0. There is a canonical “maximal prolongation” of any which will actually be sufficient in our examples, but one can also consider quotients of it for (Ω, d). In fact the choice of higher degrees does not directly impact the geodesic theory but is relevant to the torsion and curvature of a connection. We also define left and right vector fields as respectively left and right module maps (i.e., maps which are tensorial in the sense of commuting with the left and right multiplication by A).
The above generalises the monoidal category of right A-A-bimodule connections for any fixed differential algebra A (i.e., an algebra with differential structure). This diagonal case, in a left-handed version, is more familiar in noncommutative geometry.2,3,21 By a linear connection on A, we mean an A-A-bimodule connection with associated braiding (or just ∇ with associated braiding σ when the context is clear). In this case, the covariant derivative associated to a left vector field X will be denoted . We will also adopt an explicit notation ∇ξ = ξ(1) ⊗ Aξ(2) (summation understood), so that ∇X(ξ) = ξ(1)X(ξ(2)) for all ξ ∈ Ω1.
This equation makes sense for any differential algebra. We could also have defined slightly more generally with the same σE, albeit this generalisation is of no particular interest at this level. If A = C∞(M) and is a smooth curve, then in local coordinates, , where γ*(ξi)(t) = ξi(γ(t)) pulls back the coefficients of a 1-form ξ. If we also write for Christoffel symbols , then the algebraic geodesic equation reduces to (2.1), as analysed in Ref. 12. Thus, we have introduced ∇∇(σE) = 0 in generality as the notion of a “geodesic bimodule” and shown that the simplest choice E = B as a noncommutative bimodule, where the left action is defined by a curve γ, reduces to a single classical geodesic in the classical case.
B. Algebraic setting of geodesic velocity fields
In algebraic terms, this is just our universal ∇∇(σE) = 0 equation for a different choice of bimodule, namely now E = A ⊗ B, where A is a differential algebra in the role of the manifold equipped with a linear connection and as above with its classical calculus and trivial linear connection ∇dt = 0. More precisely, for a topological algebra, we can take with the left action by A and right action by B when viewed as subalgebras in the obvious way, but to keep things simple we will give formulae for A ⊗ B.
- ∇∇(σE) is a bimodule map if and only ifwhich is equivalent to
- ∇∇(σE) = 0 if and only if in addition, for all ,
This is a right-handed version of a result for left connections in Ref. 12, but we include a brief proof for completeness. Since and in the obvious way, and since dt is a basis of , the content of the bimodule map σE is a bimodule map which when restricted to implies it is given by a time dependent left vector field Xt on A as stated.
Next, writing e = a ⊗ f(t) we have ∇E(e) = ∇E(a.1 ⊗ f) = a.∇E(1 ⊗ 1.f) + σE(da ⊗ f) = a.∇E(1 ⊗ 1).f + a ⊗ df + σE(da ⊗ 1).f gives the formula for ∇E, for some undetermined ∇E(1 ⊗ 1) = κt ⊗ dt and some σE(da ⊗ 1) = Xt(da) ⊗ dt.
In Proposition 2.2, let ∇ have σ invertible, be finitely generated projective as a left module and be the associated right connection on . In these terms, the corresponding conditions are
;
.
C. Probabilistic geodesic flow and the ∇Ee = 0 equation
In our case of interest, and we consider the B-valued output to define a function of “time” with t* = t. Then d on the left is derivative in the coordinate. Hence, if ∇E preserves ⟨, ⟩ and e obeys ∇Ee = 0 as above for geodesic evolution then . We adopted a more mathematical notation but this is equivalent to the usual bra-ket notion other than the values being in B.
It remains to analyse the content of inner product preservation for our specific E where E = A ⊗ B [or ]. Since A could be noncommutative, instead of a measure we fix a positive linear functional or “vacuum state” and define as above, pointwise at each t so that the result is B-valued. Equivalently, we can suppose we are given ⟨, ⟩ and define , where a is viewed in E as constant in time.
III. QUANTUM MECHANICAL GEODESICS ON HILBERT SPACES
Thus, the data here is an “operator-valued time-dependent vector field” which for each t is a bimodule map in the sense and . We can also think of as an element of the dense subspace .
We now proceed as in Sec. II to take the trivial linear connection ∇dt = 0 acting on and an arbitrary linear right connection ∇ on , and the tensor product connections (2.4) and (2.5) on the domain and codomain of σE.
Similarly, the geodesic flow of Sec. II B is recovered with (or some completion thereof) in the classical case or with the left regular representation (or more precisely a completion thereof) in a potentially noncommutative version. In the classical case, compatibility with the calculus Ω1(M) requires the Hamiltonian to have the form for some time dependent vector field and function κt acting by left multiplication on C∞(M) and the condition in Propostion 3.3 reduces to the velocity field Eq. (2.8). In the noncommutative case with ρ the left regular representation, we have discussed in (3.1) how can arise as the image of a bimodule map Xt. This is not quite as general as Proposition 2.2, where we only assumed a left vector field.
It remains to extend Sec. II C to our more general setting. We suppose that ρt is a -homomorphism for each t. ϕt(a) = ⟨e(t)|a|e(t)⟩ = ⟨e(t)|ρte(t)⟩ or when viewed as a B-valued inner product. The main difference is that now the inner product is assumed in the Hilbert space and not given by a vacuum state ϕ0 or a preferred element 1 ∈ E as was possible before.
In the setting of Proposition 3.3, the inner product is preserved by ∇E if and only if is anti-hermitian.
If we consider Proposition 2.2 with a bimodule map as a special case with as in (3.1), then anti-hermtian essentially reduces to the two conditions in Proposition 2.4.
IV. QUANTUM GEODESICS FOR THE HEISENBERG ALGEBRA
Let A be the Heisenberg algebra and h ∈ A a Hamiltonian as above. There exists a canonical differential structure , generalised quantum metric and a metric compatible connection ∇ on such that the standard Schrödinger equation is realised as a quantum geodesic flow equation ∇Eψ = 0 with respect to a geodesic velocity field .
The proof will be a series of results starting with a class of “almost commutative” centrally extended differential structures17 where the classical commutation relations of differentials on phase space acquire a multiple of a central 1-form θ′.
We are not asserting that ∇ is unique, although we are not aware of any other solutions at least for generic V(x). It is natural in the sense of playing well with central 1-forms in and uniquely characterised by this as follows.
- has 2n central 1-formssuch that X(ωi) = X(ηi) = 0 for all i. Moreover, ∇ is the unique right connection such that
- has a central elementsuch that
For (1), the commutation relations of Proposition 4.2 immediately give that ωi, ηi are central, clearly annihilated by X as stated there. That these are covariantly constant requires and ∇(dpi) = −∇(θ′∂iV) = −θ′ ⊗ d∂iV − (∇θ′)∂iV if ∇ is a right connection. We then assume ∇θ′ = 0 and use (4.2).
To discuss the quantum geometry further, we now need to specify . For every there is a canonical “maximal prolongation” obtained by applying d to the degree 1 relations, and other choices are a quotient.
Let be the quotient of the maximal prolongation of by the additional relations dθ′ = 0 and θ′2 = 0.
The torsion of a right connection is T∇ = ∧∇ + d and comes out as shown. The formula for ∧(G) is immediate from the form of G stated in Proposition 4.4 given that θ′ anticommutes with 1-forms. Note that the torsion and curvature for a right connection are right module maps but not necessarily bimodule ones, which is indeed not the case for T∇ here. It follows in our case that R∇ = (id ⊗ d + ∇ ∧id)∇ = 0 as clearly R∇(ωi) = R∇(ηi) = R∇(θ′) = 0, since ∇ itself vanishes on these.□
We see that G is not quantum-symmetric in the sense of ∧(G) = 0 as needed for a strict quantum metric;1 it is “generalised quantum metric” in the notation there (and is, moreover, degenerate). Likewise, the torsion tensor does not vanish, so ∇ is not a “quantum Levi–Cività connection” in the sense of quantum Riemannian geometry either. Rather, G is if anything antisymmetric with respect to σ (but this depends on ∇) and moreover, one can quotient out θ′ = 0 to work in the unextended calculus on A, in which case has the same form as the canonical symplectic 2-form in the classical case, and G becomes its lift. The geodesic velocity field X, however, does not descend to this quotient, while ∇ = 0 at this quotient level, at least in the Heisenberg case studied here. Thus, the geometric picture is not exactly a quantum version of symplectic geometry either. We return to this in Sec. VI.
A. Example of the harmonic oscillator
V. GEODESIC FLOW FROM THE ELECTROMAGNETIC KLEIN GORDON OPERATOR
Our goal in this section is to extend the quantum geodesic flow of Sec. IV to a relativistic setting with flat spacetime metric η = diag(−1, 1, 1, 1) and an electromagnetic background with gauge potential Aa in place of the potential in the Hamiltonian in Sec. IV. This is done in Sec. V A using the Klein–Gordon operator but it is important to note that we are not proposing this as a way of solving the Klein–Gordon equation itself nor as an alternative to its established role in quantum field theory. Rather, this is something new which, unlike the nonrelativistic version, does not land on an established equation, not least due to the external geodesic time parameter in addition to the spacetime time. As a first look at what we have, Sec. V B discusses how it could nevertheless be visualised in a quantum-mechanics like manner in a laboratory frame and shows what we get in some examples.
Although the flow (5.1) is not something usually considered in physics, we will see that it lends itself to a quantum geodesic formulation. Indeed, that this works out will be a minor miracle in terms of the amount of algebra, which in itself suggests that this is a natural relativistic generalisation of the quantum geodesic flow in Sec. IV.
A. Electromagnetic Heisenberg differential calculus
We next want to choose ∇ on such that the conditions (4.3) for ∇∇(σE) in Proposition 3.3 hold.
The constructions so far are manifestly Lorentz invariant as long as θ′ is taken to transform trivially. We will also have recourse to the following quotient which is adapted to the observer in the chosen inertial frame but which is covariant in that one can make this in any inertial frame.
If Aa is time independent then is central in the Heisenberg algebra and . Moreover, there is a subalgebra with subcalculus of generated by where is central in and . Moreover, ∇ restricts on the generators to a connection on and .
Next, we omit x0 from our algebra as under our assumptions it does not appear in Fab or on the right hand side of any of the commutation relations other than as dx0 = −p0θ′/m. The remaining generators and relations are (5.8)–(5.10) as listed below albeit a closed central generator. Further, ∇ restricts to this subcalculus as any dx0 terms given by ∇ can be rewritten in terms of u by the relations.□
B. Relativistic amplitudes and hydrogen-like atom
Here we consider a possible interpretation or way to visualise the quantum geodesic evolution constructed in Sec. V A in a manner that is a little analogous to a modification of quantum mechanics. This is for comparison purposes to start to get a feel for the content of this flow, given that it is not something usually considered.
Proper time s relativistic wave packet dispersing as it moves down and to the right. Shown are the real and absolute values at c = ℏ = m = 1 and u = 1.1. Images produced by Mathematica.
Proper time s relativistic wave packet dispersing as it moves down and to the right. Shown are the real and absolute values at c = ℏ = m = 1 and u = 1.1. Images produced by Mathematica.
Although our quantum geodesic flow equation ∇EΨ = 0 is not Schrödinger’s equation, its similarity at fixed u means that we can use all the tools and methods of quantum mechanics with s in place of time there and u as a parameter in the Hamiltonian, as in the preceding example. This is also somewhat different from the usual derivation of Schrödinger’s equation as a limit of the KG equation, which involves writing where ct = x0 and ΨKG(t, xi) is slowly varying to recover Schrödinger’s equation for ΨKG with corrections. The minus sign is due to the −+ + + conventions. We do not need to make such slow variation assumptions and in fact we proceed relativistically as far as the flow is concerned and in a choice of laboratory frame as far as the interpretation is concerned. This means that our differences from Schrödinger’s equation are now of a different nature from the usual ones coming from the KG equation, although they share some terms in common.
VI. EXTENDED PHASE SPACE POISSON GEOMETRY
Traditionally in physics, one starts at the Poisson level and then “quantises.” In our case the situation was reversed with the quantum geometry of the Heisenberg algebra in Sec. IV dictated by the algebraic set up. We now semiclassicalise this and similar models to a Poisson level version and present that independently. The first thing we notice is that there is an extra dimension θ′ in the calculus, which is not a problem when ℏ ≠ 0 but which means that we do not have an actual differential calculus when ℏ = 0 as θ′ is still present and not generated by functions and differentials of them. This suggests that to have an honest geometric picture, we should work on where is a symplectic manifold with symplectic connection and symplectic form ωμν in local coordnates (we denote its inverse by ωμν with upper indices for the associated Poisson bivector inverse to it) and corresponds to an external time variable t with θ′ = dt. The latter recognises that noncommutative systems can generate their own time in a way that is not explicable in the classical limit. By a symplectic connection we mean torsion free and preserving the symplectic form. (Such connections always exist but are not unique.)
If ω = dθ then ∧(G) = dΘ where Θ = θ + 2hθ′ for the usual contact form Θ on extended phase space as in Ref. 25. On the other hand, our specific results in this section are not related as far as we can tell to metrics on phase space such as the Jacobi metric in Refs. 26 and 27. Nevertheless, we do make use of a natural (possibly degenerate) classical metric gμν on M induced by the Hamiltonian and we do not exclude the possibility that different approaches to geometry on phase space could be linked in future work.
A. Hamiltonian vector fields as autoparallel on extended phase space
For generic h, X is autoparallel with respect to ∇ if and only if .
We now turn to the classical symplectic form ωμνdxμ ∧dxν and its torsion free symplectic connection . In our extended calculus, we would like to find a related 2-form which is preserved by the extended covariant derivative ∇. Given that we have just added a variable t, it is reasonable to do this by extending the symplectic form by something wedged with dt.
The extended covariant derivative ∇ preserves a 2-form of the form ωαβdxα ∧dxβ + df ∧dt for generic f (derivative not vanishing identically on any open region) and has X autoparallel if and only if , and .
B. Quantum geometry of the extended Heisenberg algebra
- Let with coordinates xμ = xi for μ = i and xμ = pi for μ = i + n, where ∈ {1, …, n}. We take symplectic structure ω with tensor ωμν, associated Poisson bivector ωμν, and symplectic connectionThe Hamiltonian vector field associated to and the possibly degenerate inner product gμν areand zero otherwise.
- We extend to with an additional coordinate x0 = t. The extended connection and its torsion areand otherwise zero, with autoparallel vector field X extended by X0 = 1 (i.e., we add ). The preserved antisymmetric tensor and the 1-form η are
Comparing with (6.4), we see that the calculus corresponds to ∇ at the semiclassical level with (as well as τμ = 0). The first term of the first form of ∇(dxμ) also then agrees with ∇ in (6.6) with a further quantum correction. In this way, the formulae in Sec. IV can be written more geometrically on the extended phase space and the meaning of the connection ∇ with respect to which Schrödinger’s equation is “quantum geodesic flow” emerges as the semiclassical data for the quantum differential calculus. This also suggests how Sec. IV could potentially be extended to other quantisations of symplectic manifolds, though this remains to be done. We have only considered the time-independent theory and it seems likely that the above will extend also to the time-dependent case.
VII. CONCLUDING REMARKS
In Sec. III, we extended the formalism of “quantum geodesics” in noncommutative geometry as introduced in Ref. 12 using A-B-bimodule connections from Ref. 1 to geodesics in representation spaces. We then applied this to ordinary quantum mechanics and showed in Sec. IV that the usual Schrödinger equation can be viewed as a quantum geodesic flow for a certain quantum differential calculus on the quantum algebra of observables (the Heisenberg algebra) acting on wave functions in the Schrödinger representation. The quantum differential calculus here encodes the Hamiltonian much as in GR the Riemannian manifold determines the geodesic flow. This idea that physics has new degrees of freedom in the choice of quantum differential structure has been around for a while now and is particularly evident at the Poisson level.31 Such a freedom was already exploited to encode Newtonian gravity by putting the gravitational potential into the spacetime differential structure;15 our now results in Sec. IV are in the same spirit but now on phase space in ordinary quantum mechanics and not as part of Planck scale physics.
We then proceeded in Sec. V to a novel relativistic version of Sec. IV based on the Klein–Gordon operator minimally coupled to an external field. Even the simplest 1 + 1 dimensional case without external field in Example 5.5 proved interesting, with relativistic proper time wave packets Ψ quantum geodesically flowing with constant velocity v = ⟨Ψ|x|Ψ⟩/⟨Ψ|t|Ψ⟩ in the laboratory frame. The example illustrates well that quantum geodesic flow looks beyond the Klein–Gordon equation itself. Just as an ant moving on an apple has feet on either side of the geodesic which keeps it on the geodesic path, the quantum geodesic wave packet spreads off-shell on either side of a Klein–Gordon solution but on average evolves as expected. We also showed how our quantum geodesic flow in the case of a time-independent background field nevertheless amounts to some kind of proper time Schrödinger-like equation if we analyse the geodesic flow at fixed energy u, allowing the usual tools of quantum mechanics to be adapted to our case. We illustrated this with a hydrogen-like atom of atomic number Z. Section VI concluded with a look at the extended phase space geometry that emerges from our constructions at the semiclassical level.
Clearly, many more examples could be computed and studied using the formalism in this paper, including general (non-static) electromagnetic backgrounds to which the theory already applies. Also, in Sec. IV we focused on time-independent Hamiltonians, but the general theory in Proposition 3.3 does not require this. It would be interesting to look at the time dependent case and the construction of conserved currents. The present formalism also allows the possibility of more general algebras B in place of for the geodesic time variable.
On the theoretical side, the formalism can be extended to study quantum geodesic deviation, where classically one can see the role of Ricci curvature entering. This is not relevant to the immediate setting of the present paper since, at least in Sec. IV, the quantum connection on phase space was flat and preserved the extended quantum symplectic structure (rather than being a quantum Levi–Cività connection). It will be looked at elsewhere as more relevant to quantum spacetime and quantum gravity applications, but we don’t exclude the possibility of quantum mechanical systems where curvature is needed, e.g., with a more general form of Hamiltonian. Another immediate direction for further work would be to extend Sec. V from an electromagnetic background on the representation space to a curved Riemannian background on the latter, i.e., to gravitational backgrounds such as the wave-operator black-hole models in Ref. 17. It could also be of interest to consider quantum geodesic flows using a Dirac operator or spectral triple4 as in Connes’ approach instead of the Klein–Gordon operator.
Finally, on the technical side, the role of θ′ needs to be more fully understood from the point of view of the quantum extended phase space and its reductions. In our case, it arises as an obstruction to the Heisenberg algebra differential calculus, which forces an extra dimension, but we ultimately identified it with the geodesic time interval. However, a very different approach to handle this obstruction is to drop the bimodule associativity condition in the differential structure,28,32 which could also be of interest here.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Edwin Beggs: Conceptualization (equal); Formal analysis (equal); Writing – original draft (equal); Writing – review & editing (equal). Shahn Majid: Conceptualization (equal); Formal analysis (equal); Writing – original draft (equal); Writing – review & editing (equal).
DATA AVAILABILITY
Data sharing is not applicable to this article as no new data were created nor analyzed in this study.