Skip to main content
Log in

The Fock space of loopy spin networks for quantum gravity

  • Research Article
  • Published:
General Relativity and Gravitation Aims and scope Submit manuscript

Abstract

In the context of the coarse-graining of loop quantum gravity, we introduce loopy and tagged spin networks, which generalize the standard spin network states to account explicitly for non-trivial curvature and torsion. Both structures relax the closure constraints imposed at the spin network vertices. While tagged spin networks merely carry an extra spin at every vertex encoding the overall closure defect, loopy spin networks allow for an arbitrary number of loops attached to each vertex. These little loops can be interpreted as local excitations of the quantum gravitational field and we discuss the statistics to endow them with. The resulting Fock space of loopy spin networks realizes new truncation of loop quantum gravity, allowing to formulate its graph-changing dynamics on a fixed background graph plus local degrees of freedom attached to the graph nodes. This provides a framework for re-introducing a non-trivial background quantum geometry around which we would study the effective dynamics of perturbations. We study how to implement the dynamics of topological BF theory in this framework. We realize the projection on flat connections through holonomy constraints and we pay special attention to their often overlooked non-trivial flat solutions defined by higher derivatives of the \(\delta \)-distribution.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20

Similar content being viewed by others

Notes

  1. The definition of cylindrical consistency can be generalized and extended to a mapping or identification between two wave-functions living on a given graph and a refinement of it. Formulated as such, it becomes equivalent to the choice of a coarse-graining procedure for the quantum states of geometry. This logic has been used to construct new Hilbert spaces for loop quantum states describing the excited states of geometry above non-trivial vacua [24].

  2. The Ashtekar–Barbero connection is only a space connection defined on the canonical hypersurface and is not generically the pull-back of a space–time connection [4448], except in the case of the self-dual and anti-self dual connections given by the purely imaginary choice of Immirzi paramater \(\beta =\pm i\).

  3. In this definition of the wave-functions and cylindrical equivalence, we only deal with only actual edges of the graphs \(\Gamma \) and \(\widetilde{\Gamma }\). So-called “virtual” links or edges, which are used to unfold vertices and define spin basis for the intertwiners, do not enter into play at this stage.

  4. Another approach is to define spin networks made of intertwiners directly interpretable as dual to polyhedron in a curved space. These has been developed in the framework of spin networks for loop quantum gravity with a non-vanishing cosmological constant and is based on a quantum deformation of the \(\mathrm {SU}(2)\) gauge group [5056]. However, it is not yet clear how, if possible, to depart from a homogeneous curvature and glue pieces carrying a different curvature, thus obtaining actual spin networks with variable curvature. One possible link with our present framework would be to show that these curved and quantum-deformed intertwiners can be obtained in a continuum limit as a vertex with an infinite number of little loops creating a constant homogeneous curvature excitation (for instance, triangulating a hyperbolic tetrahedron with finer and finer tetrahedra which can be considered as flat in an infinite refinement limit).

  5. This is different from the more usual approach of using a tree (with “virtual links”) at each vertex to label a spin basis for the intertwiners living at that vertex. In that usual context, a change of tree leads to another set of basis states of the space Hilbert space of intertwiners. Here, in our new context, changing the tree \(\mathcal {T}_v\) explores another sector of the Hilbert space of states. This is natural from the coarse-graining perspective, since that unfolding tree \(\mathcal {T}_v\) was actually part of the finer graph on which the state was initially defined.

  6. We insist that the unfolding tree is about remembering the initial graph. It is not about introducing virtual links in order to define a spin basis for the intertwiners. In the previous folded spin network state context, unfolding trees would label orthogonal subspaces of the full Hilbert space, while in the present loopy spin network scenario, those orthogonal subspaces are identified and the unfolding tree data is forgotten. In both cases, one is always free to introduce virtual links as an auxiliary tool to define suitable basis of intertwiners, but those virtual links never enter the definition of any Hilbert space of spin networks.

  7. The background lattice can then be adapted to the studied models. We could choose a regular lattice or a much simpler graph, such as a flower with a single vertex and an arbitrary number of little loops. Such simple graphs could reveal useful in the study of highly symmetric problems as is the case in cosmology or in the study of Einstein–Rosen waves [59, 60].

  8. In order to identify a complete set of operators acting on the Hilbert space \({\mathcal {H}}^{\mathrm {loopy}}\), we should further consider multi-loops holonomy operators or grasping operators or deformation operators such as \(\mathrm {U}(N)\) operators [61], but in the first exploration we propose, in this paper, we decide to focus on the single-loop holonomy operator.

  9. Once the dynamics of BF theory is properly implemented and well under control in a certain framework, one usually use it as a starting point for imposing the true gravity dynamics, with local degrees of freedom, relying on the reformulation of general relativity as a BF theory with constraints. This is for instance the logic behind the construction of spinfoam models for a quantum gravity path integral [2123].

  10. Indeed, since we already proved that the \(\delta \)-distribution is the only gauge-invariant solution to the holonomy constraint, all the derivative solutions can not be gauge-invariant. We can also prove this directly. To get a solution invariant under conjugation, we need to contract the derivative indices together. Now a fundamental theorem on rotational invariants states that all \(\mathrm {SO}(3)\)-invariant polynomial of n 3d-vectors \(\vec {v}_{i=1 \ldots n}\) are generated by scalar products \(\vec {v}_{i}\cdot \vec {v}_{j}\) and triple products \(\vec {v}_{i}\cdot (\vec {v}_{j}\wedge \vec {v}_{k})\). We simply have to check that the Laplacian and triple-grasping of the \(\delta \)-distribution are not solutions of the holonomy constraints:

    $$\begin{aligned}&\int f (\chi _{\frac{1}{2}}-2)\partial _{a}\partial _{a}\delta = \Delta \chi _{\frac{1}{2}}({\mathbb {I}})\,f({\mathbb {I}}) = \frac{-3}{2}\,f({\mathbb {I}})\ne 0, \end{aligned}$$
    $$\begin{aligned}&\quad \int f (\chi _{\frac{1}{2}}-2)\epsilon ^{abc}\partial _{a}\partial _{b}\partial _{c}\delta = \frac{-i^{3}}{2^{3}}\epsilon ^{abc}\chi _{\frac{1}{2}}(\sigma _{c}\sigma _{b}\sigma _{a})\,f({\mathbb {I}}) = \frac{3}{2}\,f({\mathbb {I}})\ne 0. \end{aligned}$$
  11. As an example, we can see how \(\Delta \) and \(\widetilde{\Delta }\) differ through their action on the coupled character \(\chi (h_{1}h_{2})\):

    $$\begin{aligned} \Delta _{1}\chi _{\frac{1}{2}}(h_{1}h_{2})= & {} -\frac{1}{4}\chi _{\frac{1}{2}}(h_{1}\sigma _{a}\sigma _{a}h_{2})=-\frac{3}{4}\chi _{\frac{1}{2}}(h_{1}h_{2}), \quad \widetilde{\Delta }_{1}\chi _{\frac{1}{2}}(h_{1}h_{2})=-\frac{1}{4}\chi _{\frac{1}{2}}(\sigma _{a}h_{1}\sigma _{a}h_{2}) \\= & {} -\frac{1}{4}\,\big (2\chi _{\frac{1}{2}}(h_{1})\chi _{\frac{1}{2}}(h_{2})-\chi _{\frac{1}{2}}(h_{1}h_{2})\big ), \end{aligned}$$

    which are of course equal at \(h_{1}={\mathbb {I}}\).

  12. As an example, we have in mind the recursion relation satisfied by the 6j symbol, which is understood to be the expression of the action of the holonomy operator on the flat state on the tetrahedron graph [6264]. Our results suggest that the double and triple graspings on the 6j symbol might be other solutions to this recursion relation. That would be specially interesting since the triply grasped 6j symbols is understood to be the first order correction of the q-deformed 6j-symbol [65].

  13. If we had more solutions to the constraints, for example if we only impose the holonomy constraints without the Laplacian constraints, we would propose to render the distributions normalizable by modulating them by a damping operator \(\exp (-\tau \Delta )\) (heat kernel) or similar. Expanding all the functions and distributions in Fourier modes, all the scalar products would become finite and we could then compute expectation values of operators on the resulting physical Hilbert space, for instance:

    $$\begin{aligned} \int \delta e^{-\tau \Delta }\delta =\sum _{j}(2j+1)^{2}e^{-\tau j(j+1)} <+\infty ,\quad \int \delta e^{-\tau \Delta }\partial _{a}\delta =i\sum _{j}(2j+1)^{2}e^{-\tau j(j+1)}\chi _{j}(J_{a})=0. \end{aligned}$$

    But we haven’t investigated this line of research further.

  14. We would use an extension of the finite symmetry groups \(S_{n}\) to the group of permutations of integers which only act non-trivially on a finite subset:

    $$\begin{aligned} S_\infty = \{f : \mathbb {N} \rightarrow \mathbb {N}, f \text { bijective and }\exists n \in \mathbb {N}, \forall m>n, f(m)=m\}. \end{aligned}$$

    We would use the canonical action of \(S_\infty \) on the Hilbert spaces of loopy spin networks \({\mathcal {H}}_{E}\) with finite number of loops:

    $$\begin{aligned} \sigma : {\mathcal {H}}_{E} \rightarrow {\mathcal {H}}_{\sigma (E)}, \quad (\sigma \triangleright f)(\{h_{e_i}\}) = f(\{h_{\sigma ^{-1}(e_i)}\}). \end{aligned}$$

    This action is compatible with the cylindrical consistency conditions and naturally extends to the projective limit. However, requiring invariance of states \(|\Psi \rangle \in {\mathcal {H}}^{\mathrm {loopy}}\) under permutations, \(\forall \sigma \in S_\infty ,~\sigma \triangleright |\Psi \rangle = |\Psi \rangle \), only provides non-normalizable states. This forces us to work on the dual space to define symmetrized states and creates unnecessary technicalities for our present purpose.

  15. This leads us to conjecture a set of complete holonomy constraints for BF theory. Considering all the multi-loop holonomy operators for arbitrary spins acting on the Fock space of loopy intertwiners \({\mathcal {H}}^{\mathrm {sym}}\):

    $$\begin{aligned} \forall j\in \frac{{\mathbb {N}}^{*}}{2},\quad \forall n\in {\mathbb {N}}^{*},\quad \widehat{\chi }^{(n)}_{j}\,|\phi \rangle =\,(2j+1)\,|\phi \rangle , \end{aligned}$$

    then the only solution to all these constraints is the flat state \(\phi =\delta \). We have checked this conjecture up to the three-loop component of the state, \(N=3\), but we haven’t gone further. This would require explicitly and carefully defining the multi-loop holonomy operators. We should also take special care of working with legitimate states, controlling the convergence/divergence of the series in j and N to ensure that the states are distributions.

  16. The simplest way to proceed is, as in the gauge fixing procedure, to choose a maximal tree on \(\Gamma \), then to associate one loop to each edge which does not belong to the tree. This ensures that each of those loops contains one edge that no other loop contain and are thus independent.

  17. This is the point where we choose not to use the completion and just have an isomorphism of pre-Hilbertian spaces in order to have the existence of \(F_f\).

References

  1. Rovelli, C.: Quantum Gravity. Cambridge Monographs on Mathematical Physics. Cambridge University Press, Cambridge (2007)

    Google Scholar 

  2. Thiemann, T.: Modern Canonical Quantum General Relativity. Cambridge Monographs on Mathematical Physics. Cambridge University Press, Cambridge (2007)

    Book  Google Scholar 

  3. Gambini, R., Pullin, J.: A First Course in Loop Quantum Gravity. Oxford University Press, Oxford (2011)

    Book  MATH  Google Scholar 

  4. Ashtekar, A.: New variables for classical and quantum gravity. Phys. Rev. Lett. 57, 2244–2247 (1986)

    Article  ADS  MathSciNet  Google Scholar 

  5. Barbero G, J.F.: Real Ashtekar variables for Lorentzian signature space times. Phys. Rev. D 51, 5507–5510 (1995). arXiv:gr-qc/9410014

    Article  ADS  MathSciNet  Google Scholar 

  6. Immirzi, G.: Real and complex connections for canonical gravity. Class. Quantum Gravity 14, L177–L181 (1997). arXiv:gr-qc/9612030

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. Thiemann, T.: Anomaly—free formulation of nonperturbative, four-dimensional Lorentzian quantum gravity. Phys. Lett. B 380, 257–264 (1996). arXiv:gr-qc/9606088

    Article  ADS  MathSciNet  MATH  Google Scholar 

  8. Thiemann, T.: Quantum spin dynamics (QSD). Class. Quantum Gravity 15, 839–873 (1998). arXiv:gr-qc/9606089

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Thiemann, T.: Quantum spin dynamics (QSD). 2. Class. Quantum Gravity 15, 875–905 (1998). arXiv:gr-qc/9606090

    Article  ADS  MathSciNet  MATH  Google Scholar 

  10. Thiemann, T.: The Phoenix project: master constraint program for loop quantum gravity. Class. Quantum Gravity 23, 2211–2248 (2006). arXiv:gr-qc/0305080

    Article  ADS  MathSciNet  MATH  Google Scholar 

  11. Alesci, E.: Regularized Hamiltonians and spinfoams. J. Phys. Conf. Ser. 360, 012041 (2012). arXiv:1110.6150

    Article  ADS  Google Scholar 

  12. Alesci, E., Assanioussi, M., Lewandowski, J., Mkinen, I.: Hamiltonian operator for loop quantum gravity coupled to a scalar field. Phys. Rev. D 91(12), 124067 (2015). arXiv:1504.02068

    Article  ADS  MathSciNet  Google Scholar 

  13. Assanioussi, M., Lewandowski, J., Mkinen, I.: New scalar constraint operator for loop quantum gravity. Phys. Rev. D 92(4), 044042 (2015). arXiv:1506.00299

    Article  ADS  MathSciNet  Google Scholar 

  14. Bonzom, V., Laddha, A.: Lessons from toy-models for the dynamics of loop quantum gravity. SIGMA 8, 009 (2012). arXiv:1110.2157

    MathSciNet  MATH  Google Scholar 

  15. Engle, J., Livine, E., Pereira, R., Rovelli, C.: LQG vertex with finite Immirzi parameter. Nucl. Phys. B 799, 136–149 (2008). arXiv:0711.0146

    Article  ADS  MathSciNet  MATH  Google Scholar 

  16. Ben Geloun, J., Gurau, R., Rivasseau, V.: EPRL/FK group field theory. Europhys. Lett. 92, 60008 (2010). arXiv:1008.0354

    Article  ADS  Google Scholar 

  17. Vidotto, F., Rovelli, C.: Covariant Loop Quantum Gravity. Cambridge Monographs on Mathematical Physics. Cambridge University Press, Cambridge (2014)

    Google Scholar 

  18. Dupuis, M., Livine, E.R.: Holomorphic simplicity constraints for 4D spinfoam models. Class. Quantum Gravity 28, 215022 (2011). arXiv:1104.3683

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. Speziale, S., Wieland, W.M.: The twistorial structure of loop-gravity transition amplitudes. Phys. Rev. D 86, 124023 (2012). arXiv:1207.6348

    Article  ADS  Google Scholar 

  20. Wieland, W.M.: Hamiltonian spinfoam gravity. Class. Quantum Gravity 31, 025002 (2014). arXiv:1301.5859

    Article  ADS  MATH  Google Scholar 

  21. Livine, E.R.: The spinfoam framework for quantum gravity. PhD Thesis, Lyon, IPN (2010). arXiv:1101.5061

  22. Perez, A.: The spin foam approach to quantum gravity. Living Rev. Rel. 16, 3 (2013). arXiv:1205.2019

    MATH  Google Scholar 

  23. Bianchi, E., Hellmann, F.: The construction of spin foam vertex amplitudes. SIGMA 9, 008 (2013). arXiv:1207.4596

    MathSciNet  MATH  Google Scholar 

  24. Koslowski, T., Sahlmann, H.: Loop quantum gravity vacuum with nondegenerate geometry. SIGMA 8, 026 (2012). arXiv:1109.4688

    MathSciNet  MATH  Google Scholar 

  25. Dittrich, B., Geiller, M.: A new vacuum for loop quantum gravity. Class. Quantum Gravity 32(11), 112001 (2015). arXiv:1401.6441

    Article  ADS  MathSciNet  MATH  Google Scholar 

  26. Bahr, B., Dittrich, B., Geiller, M.: A new realization of quantum geometry. arXiv:1506.08571

  27. Oriti, D.: Group field theory as the 2nd quantization of loop quantum gravity. arXiv:1310.7786

  28. Rivasseau, V.: Quantum gravity and renormalization: the tensor track. AIP Conf. Proc. 1444, 18–29 (2011). arXiv:1112.5104

    ADS  Google Scholar 

  29. Rivasseau, V.: The tensor track, III. Fortsch. Phys. 62, 81–107 (2014). arXiv:1311.1461

    Article  ADS  MathSciNet  MATH  Google Scholar 

  30. Carrozza, S.: Tensorial methods and renormalization in group field theories. PhD Thesis, Orsay, LPT (2013). arXiv:1310.3736

  31. Carrozza, S.: Group field theory in dimension \(4-\epsilon \). Phys. Rev. D 91(6), 065023 (2015). arXiv:1411.5385

    Article  ADS  MathSciNet  Google Scholar 

  32. Livine, E.R., Terno, D.R.: Reconstructing quantum geometry from quantum information: area renormalisation, coarse-graining and entanglement on spin networks. arXiv:gr-qc/0603008

  33. Livine, E.R.: Deformation operators of spin networks and coarse-graining. arXiv:1310.3362

  34. Ashtekar, A., Lewandowski, J.: Projective techniques and functional integration for Gauge theories. J. Math. Phys. 36, 2170–2191 (1995). arXiv:gr-qc/9411046

    Article  ADS  MathSciNet  MATH  Google Scholar 

  35. Ashtekar, A., Lewandowski, J.: Differential geometry on the space of connections via graphs and projective limits. J. Geom. Phys. 17, 191–230 (1995). arXiv:hep-th/9412073

    Article  ADS  MathSciNet  MATH  Google Scholar 

  36. Ashtekar, A., Lewandowski, J.: Representation theory of analytic holonomy C* algebras. arXiv:gr-qc/9311010

  37. Rovelli, C., Smolin, L.: Discreteness of area and volume in quantum gravity. Nucl. Phys. B 442, 593–622 (1995). arXiv:gr-qc/9411005

    Article  ADS  MathSciNet  MATH  Google Scholar 

  38. Rovelli, C., Smolin, L.: Spin networks and quantum gravity. Phys. Rev. D 52, 5743–5759 (1995). arXiv:gr-qc/9505006

    Article  ADS  MathSciNet  Google Scholar 

  39. Freidel, L., Speziale, S.: Twisted geometries: a geometric parametrisation of SU(2) phase space. Phys. Rev. D 82, 084040 (2010). arXiv:1001.2748

    Article  ADS  Google Scholar 

  40. Dupuis, M., Ryan, J.P., Speziale, S.: Discrete gravity models and loop quantum gravity: a short review. SIGMA 8, 052 (2012). arXiv:1204.5394

    MathSciNet  MATH  Google Scholar 

  41. Knizhnik, V., Polyakov, A.M., Zamolodchikov, A.: Fractal structure of 2D quantum gravity. Mod. Phys. Lett. A 3, 819 (1988)

    Article  ADS  MathSciNet  Google Scholar 

  42. Freidel, L., Livine, E.R.: Spin networks for noncompact groups. J. Math. Phys. 44, 1322–1356 (2003). arXiv:hep-th/0205268

    Article  ADS  MathSciNet  MATH  Google Scholar 

  43. Holst, S.: Barbero’s Hamiltonian derived from a generalized Hilbert–Palatini action. Phys. Rev. D 53, 5966–5969 (1996). arXiv:gr-qc/9511026

    Article  ADS  MathSciNet  Google Scholar 

  44. Samuel, J.: Is Barbero’s Hamiltonian formulation a gauge theory of Lorentzian gravity? Class. Quantum Gravity 17, L141–L148 (2000). arXiv:gr-qc/0005095

    Article  ADS  MathSciNet  MATH  Google Scholar 

  45. Alexandrov, S.: On choice of connection in loop quantum gravity. Phys. Rev. D 65, 024011 (2002). arXiv:gr-qc/0107071

    Article  ADS  MathSciNet  Google Scholar 

  46. Geiller, M., Lachieze-Rey, M., Noui, K., Sardelli, F.: A Lorentz-covariant connection for canonical gravity. SIGMA 7, 083 (2011). arXiv:1103.4057

    MathSciNet  MATH  Google Scholar 

  47. Geiller, M., Lachieze-Rey, M., Noui, K.: A new look at Lorentz-covariant loop quantum gravity. Phys. Rev. D 84, 044002 (2011). arXiv:1105.4194

    Article  ADS  MATH  Google Scholar 

  48. Charles, C., Livine, E.R.: Ashtekar–Barbero holonomy on the hyperboloid: Immirzi parameter as a cut-off for quantum gravity. arXiv:1507.00851

  49. Freidel, L., Ziprick, J.: Spinning geometry = Twisted geometry. Class. Quantum Gravity 31(4), 045007 (2014). arXiv:1308.0040

    Article  ADS  MathSciNet  MATH  Google Scholar 

  50. Dupuis, M., Girelli, F.: Quantum hyperbolic geometry in loop quantum gravity with cosmological constant. Phys. Rev. D 87(12), 121502 (2013). arXiv:1307.5461

    Article  ADS  Google Scholar 

  51. Bonzom, V., Dupuis, M., Girelli, F., Livine, E.R.: Deformed phase space for 3d loop gravity and hyperbolic discrete geometries. arXiv:1402.2323

  52. Dupuis, M., Girelli, F., Livine, E.R.: Deformed spinor networks for loop gravity: towards hyperbolic twisted geometries. Gen. Relativ. Gravit. 46(11), 1802 (2014). arXiv:1403.7482

    Article  ADS  MathSciNet  MATH  Google Scholar 

  53. Charles, C., Livine, E.R.: Closure constraints for hyperbolic tetrahedra. Class. Quantum Gravity 32(13), 135003 (2015). arXiv:1501.00855

    Article  ADS  MathSciNet  MATH  Google Scholar 

  54. Haggard, H.M., Han, M., Kamiński, W., Riello, A.: SL(2, C) Chern–Simons theory, a non-planar graph operator, and 4D loop quantum gravity with a cosmological constant: semiclassical geometry, Nucl. Phys. B900, 1–79 (2015). arXiv:1412.7546

    Article  ADS  MATH  Google Scholar 

  55. Haggard, H.M., Han, M., Riello, A.: Encoding curved tetrahedra in face holonomies: a phase space of shapes from group-valued moment maps. arXiv:1506.03053

  56. Haggard, H.M., Han, M., Kamiński, W., Riello, A.: Four-dimensional quantum gravity with a cosmological constant from three-dimensional holomorphic blocks. Phys. Lett. B 752, 258–262 (2016). arXiv:1509.00458

    Article  ADS  MathSciNet  Google Scholar 

  57. Pithis, A.G., Ruiz Euler, H.-C.: Anyonic statistics and large horizon diffeomorphisms for loop quantum gravity black holes. Phys. Rev. D 91, 064053 (2015). arXiv:1402.2274

    Article  ADS  MathSciNet  Google Scholar 

  58. Yang, J., Ma, Y.: Quasi-local energy in loop quantum gravity. Phys. Rev. D 80, 084027 (2009). arXiv:0812.3554

    Article  ADS  MathSciNet  Google Scholar 

  59. Korotkin, D., Samtleben, H.: Canonical quantization of cylindrical gravitational waves with two polarizations. Phys. Rev. Lett. 80, 14–17 (1998). arXiv:gr-qc/9705013

    Article  ADS  MathSciNet  MATH  Google Scholar 

  60. Ashtekar, A., Bicak, J., Schmidt, B.G.: Behavior of Einstein–Rosen waves at null infinity. Phys. Rev. D 55, 687–694 (1997). arXiv:gr-qc/9608041

    Article  ADS  MathSciNet  Google Scholar 

  61. Borja, E.F., Freidel, L., Garay, I., Livine, E.R.: U(N) tools for loop quantum gravity: the return of the spinor. Class. Quantum Gravity 28, 055005 (2011). arXiv:1010.5451

    Article  ADS  MathSciNet  MATH  Google Scholar 

  62. Bonzom, V., Livine, E.R., Speziale, S.: Recurrence relations for spin foam vertices. Class. Quantum Gravity 27, 125002 (2010). arXiv:0911.2204

    Article  ADS  MathSciNet  MATH  Google Scholar 

  63. Bonzom, V., Freidel, L.: The Hamiltonian constraint in 3d Riemannian loop quantum gravity. Class. Quantum Gravity 28, 195006 (2011). arXiv:1101.3524

    Article  ADS  MathSciNet  MATH  Google Scholar 

  64. Bonzom, V., Dupuis, M., Girelli, F.: Towards the Turaev-Viro amplitudes from a Hamiltonian constraint. Phys. Rev. D 90(10), 104038 (2014). arXiv:1403.7121

    Article  ADS  Google Scholar 

  65. Freidel, L., Krasnov, K.: Discrete space–time volume for three-dimensional BF theory and quantum gravity. Class. Quantum Gravity 16, 351–362 (1999). arXiv:hep-th/9804185

    Article  ADS  MathSciNet  MATH  Google Scholar 

  66. Bonzom, V., Livine, E.R.: A new hamiltonian for the topological BF phase with spinor networks. J. Math. Phys. 53, 072201 (2012). arXiv:1110.3272

    Article  ADS  MathSciNet  MATH  Google Scholar 

  67. Bonzom, V., Dittrich, B.: Dirac’s discrete hypersurface deformation algebras. Class. Quantum Gravity 30, 205013 (2013). arXiv:1304.5983

    Article  ADS  MathSciNet  MATH  Google Scholar 

  68. Feller, A., Livine, E.R.: Ising spin network states for loop quantum gravity: a toy model for phase transitions. Class. Quantum Gravity 33(6), 065005 (2016). arXiv:1509.05297

    Article  ADS  MathSciNet  MATH  Google Scholar 

  69. Alesci, E., Cianfrani, F.: Quantum-reduced loop gravity: cosmology. Phys. Rev. D 87(8), 083521 (2013). arXiv:1301.2245

    Article  ADS  MathSciNet  MATH  Google Scholar 

  70. Noui, K., Perez, A.: Three-dimensional loop quantum gravity: physical scalar product and spin foam models. Class. Quantum Gravity 22, 1739 (2005). doi:10.1088/0264-9381/22/9/017. arXiv:gr-qc/0402110

  71. Livine, E.R.: Projected spin networks for Lorentz connection: linking spin foams and loop gravity. Class. Quantum Gravity 19, 5525–5542 (2002). arXiv:gr-qc/0207084

    Article  ADS  MathSciNet  MATH  Google Scholar 

  72. Dupuis, M., Livine, E.R.: Lifting SU(2) spin networks to projected spin networks. Phys. Rev. D 82, 064044 (2010). arXiv:1008.4093

    Article  ADS  Google Scholar 

Download references

Acknowledgments

C.C. would like to thank Michel Fruchart and Dimitri Cobb for their keen insights and useful discussions with them.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Etera R. Livine.

Appendices

Appendix 1: Projective limits of loopy spin networks

The general framework of a projective family and the projective limit can be found in [34], where it is applied to define the kinematical Hilbert space of spin network states for loop quantum gravity. Here we apply these definitions to loopy spin networks, in order to define superposition states of potentially an infinite number of little loops. To this purpose, we focus on the flower graph, with a single vertex and an arbitrary number of loops attached to that central node.

In order to define precisely this idea of varying number of loops, we start with wave-functions over a finite number of loops and define a nesting, that is describe how to include a set of loops inside a larger one. We will identity the set of all potential loops with the set of integers. Finite sets of loops are defined as finite subsets of integers. Loops are labeled by the integers and are a priori distinguishable. For instance, a wave-function with the support on the loop number 2 and a wave-function on the loop number 277 are not the same though they both are one-loop states and depend on only one variable, as illustrated in Fig. 21.

Fig. 21
figure 21

We distinguish the different potential loops and therefore consider the resulting wave-functions as different even when they excite the same number of loops

Mathematically, we consider the set of all finite subsets of integers \(\mathcal {P}_{<\infty }(\mathbb {N})\). To each subset \(E \in \mathcal {P}_{<\infty }(\mathbb {N})\), we associate the set \(\mathrm {SU}(2)^{E}\) of colorings of the corresponding loops by \(\mathrm {SU}(2)\) group elements. Then wave-functions on E are gauge-invariant functions over \(\mathrm {SU}(2)^{E}\):

$$\begin{aligned} \{\Psi : \mathrm {SU}(2)^E \rightarrow \mathbb {C}~:~ \forall g \in \mathrm {SU}(2),~ \Psi (\{gh_{\ell }g^{-1}\}_{\ell \in E}) = \Psi (\{h_{\ell }\}_{\ell \in E})\}. \end{aligned}$$
(137)

Defining the scalar product using the Haar measure over \(\mathrm {SU}(2)^{E}\), the Hilbert space \({\mathcal {H}}_{E}\) of quantum states on the loopy spin network defined by the subset E of loops is the \(L^{2}(\mathrm {SU}(2)^{E}/\text {Ad}\mathrm {SU}(2))\).

The space of loops is equipped with a partial directed order given by the inclusion of subsets of integers. The partial directed order encodes how different subsets are nested within one another: a wave-function over the loop number 2 and a wave-function over the loop number 277 are different but they are both embedded in the larger class of wave-functions which depend on both loop number 2 and loop number 277 as illustrated in Fig. 22.

Fig. 22
figure 22

Though different, two functions over two different loops can be embedded in a larger space of functions depending on several loops by identifying them with functions with trivial dependency on some loops

This partial ordering by inclusion of subsets induces a projective structure on the loop colorings by group elements. We define a projector \(p_{EE'}\) defined for every pair of subsets \((E,E')\) such that \(E \subseteq E'\) by:

This projector is simply the canonical restriction from the larger subset \(E'\) to the smaller subset E. These projectors satisfy a key transitivity property:

$$\begin{aligned} \forall E,E',E'', \quad E\subset E'\subset E''\, \Longrightarrow p_{E'E''} \circ p_{EE'} = p_{EE''}, \end{aligned}$$
(138)

so that the couple of sets \((\mathrm {SU}(2)^{E},p_{EE'})_{E,E' \in \mathcal {P}_{<\infty }(\mathbb {N})}\) form what is called a projective family. The projective limit \(\overline{\mathrm {SU}(2)}\) is then defined by:

$$\begin{aligned} \overline{\mathrm {SU}(2)} = \Big \{ (g_{E})_{E \in \mathcal {P}} \in \times _{E\in \mathcal {P}} \mathrm {SU}(2)^{E} ~:~ E \subseteq E' \Rightarrow p_{EE'} g_{E'} = g_{E} \Big \} \end{aligned}$$
(139)

Intuitively, this corresponds to collections of colorings on all possible subsets of loops which are compatible with each other with respect to the inclusion. Therefore, these compatibility conditions between all finite samplings of the collection, as illustrated in Fig. 23, is the precise implementation of the notion of a coloring of an infinite number of loops.

Fig. 23
figure 23

The projective limit is made of collections of colorings of finite subsets equipped with compatibility conditions: the coloring of a finite subset is the projection of the coloring of any larger subset (colour figure online)

We translate the projective structure to the space of wave-functions. The projectors \(p_{EE'}\) for \(E\subset E'\) turn into injections \(I_{EE'}\equiv p^{*}_{EE'}\) defined by their pull-backs:

$$\begin{aligned} I_{EE'} : {\mathcal {H}}_{E}\rightarrow & {} {\mathcal {H}}_{E'} \nonumber \\ f\mapsto & {} p^{*}_{EE'}f : p^{*}_{EE'}f \big (\{g_{\ell }\}_{\ell \in E'}\big )=f \big (\{\{g_{\ell }\}_{\ell \in E}\}\big ), \end{aligned}$$
(140)

where \(p^{*}_{EE'}f\) trivially depends on group elements living on loops of \(E'\) which do not belong to the smaller set E. The compatibility conditions translates into an equivalence relation:

$$\begin{aligned} f_{E_1} \sim f_{E_2} \Leftrightarrow \forall E_3\in \mathcal {P},~E_1 \subseteq E_3,~E_2\subseteq E_3,~p^*_{E_1 E_3} f_{E_1} = p^*_{E_2E_3} f_{E_2} \end{aligned}$$
(141)

This allows to define wave-functions on the projective limit of the loop colorings \(\overline{\mathrm {SU}(2)}\) and give a precise sense to functions over an infinite number of loops:

$$\begin{aligned} \mathcal {H}=\left( \bigcup _{E\in \mathcal {P}} \mathcal {H}_{E}\right) /\sim \end{aligned}$$
(142)

In order to make this projective limit less abstract and easier to handle, we use another representation. For every equivalence class of wave-functions in the projective limit \({\mathcal {H}}\), let us remove all the trivial dependency and pick its representant based on the smallest subset of loops. So, in practice, we define spaces of “proper states”, i.e. wave-functions that have no trivial dependency:

$$\begin{aligned} \mathcal {H}_{E}^0 = \left\{ f \in \mathcal {H}_{E} : \forall \ell \in E,\,\int _{\mathrm {SU}(2)}\mathrm {d}h_{\ell }\,f=0 \right\} . \end{aligned}$$
(143)

This is the space of functions really defined on the subset E, with an actual dependance on each loop and no constant term. The integral condition removes all the spin-0 components of the wave-functions. First, we show that an arbitrary wave-function over the subset E of loops can be fully decomposed into proper states with support on all the subsets of E:

Lemma 6.1

The following isomorphism holds as a pre-Hilbertian space isomorphism:

$$\begin{aligned} \forall E \in \mathcal {P}_{<\infty }(\mathbb {N}), \quad \mathcal {H}_{E} \simeq \bigoplus _{F \in \mathcal {P}(E)} \mathcal {H}_{F}^0, \end{aligned}$$
(144)

where the direct sum is over all subsets \(F\subset E\). This isomorphism is realized through the projections \(f_{F}=P_{E,F}f\) acting on wave-functions \(f\in {\mathcal {H}}_{E}\):

$$\begin{aligned} f_{F}\big (\{h_{\ell }\}_{\ell \in F} \big ) \!=\! \sum _{\widetilde{F}\subset F} (-1)^{\#\widetilde{F}} \int \prod _{\ell \in E{\setminus } F}\mathrm {d}g_{\ell } \prod _{\ell \in \widetilde{F}}\mathrm {d}k_{\ell }\, f\big (\{h_{\ell }\}_{\ell \in F{\setminus } \widetilde{F}}, \{k_{\ell }\}_{\ell \in \widetilde{F}}, \{g_{\ell }\}_{\ell \in E{\setminus } F} \big ). \end{aligned}$$
(145)

Its inverse is the re-summation of the projections:

$$\begin{aligned} f=\sum _{F\subset E} f_{F}. \end{aligned}$$
(146)

Proof

We proceed in two steps. First we check that each projection \(f_{F}\) is a proper state,

$$\begin{aligned} \forall \ell \in F,\quad \int \mathrm {d}h_{\ell }\,f_{F}=0, \end{aligned}$$

and that re-summing these projections \(\sum _{F\subset E} f_{F}\) yields f. Second, we check that the integral condition, ensuring that there is no spin-0 mode, also implies that the subspaces \(\mathcal {H}_{F}^0\) are pairwise orthogonal, which concludes the proof. \(\square \)

This decomposition generalizes to the projective limit:

Proposition 6.2

The following isomorphism holds as a pre-Hilbertian space isomorphism:

$$\begin{aligned} \mathcal {H} \simeq \bigoplus _{E \in \mathcal {P}_{<\infty }(\mathbb {N})} \mathcal {H}_{E}^0. \end{aligned}$$
(147)

Proof

If \((f_E)_{E \in \mathcal {P}_{<\infty }(\mathbb {N})}\) is in \(\bigoplus _{E \in \mathcal {P}_{<\infty }(\mathbb {N})} \mathcal {H}_{E}^0\), we define the set of subsets on which the state f does not vanish:

$$\begin{aligned} C_f = \{ E \in \mathcal {P}_{<\infty }(\mathbb {N}) : f_E \ne 0 \}. \end{aligned}$$
(148)

By definition of the direct sum, \(C_f\) is finite, so we can define the finite subset \(F = \cup _{E \in C_f} E\) and the re-summation map:

$$\begin{aligned} \phi : \bigoplus _{E \in \mathcal {P}_{<\infty }(\mathbb {N})} \mathcal {H}_{E}^0\rightarrow & {} \mathcal {H} \nonumber \\ (f_E)_{E \in \mathcal {P}_{<\infty }(\mathbb {N})}\mapsto & {} \left[ \sum _{E \in C_f} f_E\right] \end{aligned}$$
(149)

where the brackets refer to the equivalence class of the function. This map is obviously linear and we now look for a definition of its inverse. So let us consider a state in \(\mathcal {H}\), that is an equivalence class s. We define the set of subsets of loops on which it has support:

$$\begin{aligned} D_s = \{ E \in \mathcal {P}_{<\infty }(\mathbb {N}) : \exists f \in s,~f \in \mathcal {H}_{E} \} \end{aligned}$$
(150)

Then we consider the smallest set in \(D_s\), which can be definedFootnote 17 as the intersection \(F_s = \bigcap _{E \in D_s} E\). In a sense, this is the minimal support of the state s. We choose a representative \(f^{s}\) of the equivalence class s in \(F_{s}\). It is actually unique by definition of the equivalence relation. Then we consider the decomposition in proper states of \(f^{s}\) over all subsets F of \(F_{s}\) and define:

$$\begin{aligned} \psi : \mathcal {H}\rightarrow & {} \bigoplus _{E \in \mathcal {P}_{<\infty }(\mathbb {N})} \mathcal {H}_{E}^0 \nonumber \\ s\mapsto & {} \sum _{F\subset F_{s}}P_{F_{s},F} f^{s} \end{aligned}$$
(151)

It is direct to check that it is indeed the inverse of \(\phi \). \(\square \)

This decomposition into proper states is very useful to visualize the space: each wave-function can be decomposed into a sum of wave-functions over a finite number of loops but with no trivial dependency. This gives a precise meaning to superpositions of numbers of loops.

Appendix 2: Holonomy constraint on \(\mathrm {SU}(2)\)

1.1 Distributions on \(\mathrm {SU}(2)\) and conjugation-invariant solutions

We would like to impose the holonomy constraints for BF theory which read for a single group element:

$$\begin{aligned} \forall g \in \mathrm {SU}(2), \quad \widehat{\chi }\varphi \,(g)= \chi _\frac{1}{2}(g) \varphi (g) = 2 ,\varphi (g). \end{aligned}$$
(152)

If we stay in the strict framework of the Hilbert space \(L^{2}(\mathrm {SU}(2))\), no square integrable function actually provides such an eigenvector for \(\widehat{\chi }\) and we should solve this equation in the dual space. As is standard in quantum mechanics, the natural framework for solving the equation is a rigged-Hilbert space (or Gelfand triple), that is a triplet: \(\mathcal {S} \subset \mathcal {H} \subset \mathcal {S}^*\). The space \(\mathcal {H}\) is the Hilbert space. The smaller space \(\mathcal {S}\) is provided with a stronger topology than the induced one and can thought of as the test function space, while its dual \(\mathcal {S}^{*}\) is the space of continuous linear forms on \({\mathcal {S}}\) and defines the distribution space. The major property of \(\mathcal {S}\) is to be small enough for the algebra of observables to be defined over it. Then the operator algebra can be naturally extended on \(\mathcal {S}^{*}\) and thus on \(\mathcal {H}\). For instance, an operator A defined on \(\mathcal {S}\) acts on a (dual) state \(\varphi \) be in \(\mathcal {S}^*\) as:

$$\begin{aligned} \forall f \in \mathcal {S},~A\varphi (f) = \varphi (A^\dagger f). \end{aligned}$$
(153)

So let us be explicit for functions over \(\mathrm {SU}(2)\). The Hilbert space \(\mathcal {H}\) is the space of square-integrable functions. The space \(\mathcal {S}\) is usually chosen to be the Schwarz space so that canonical position and momentum operator can be defined. Here, the rapid fall-off condition is not needed since we are dealing with a compact group, but we keep the smoothness requirement:

$$\begin{aligned} \mathcal {S} = \{\psi \in \mathcal {H} / \psi \in \mathcal {C}^\infty \} \end{aligned}$$
(154)

Regarding the topology, the space \(\mathcal {S}\) is naturally endowed with the convergence on every \(\mathcal {C}^k\) space. More precisely the space of \(\mathcal {C}^k\) functions is equipped with the following norm:

$$\begin{aligned} \Vert f\Vert _{\mathcal {C}^k} = \sup _{0 \le i \le k} \Vert \partial _{a_1,\ldots ,a_i} f\Vert _\infty \end{aligned}$$
(155)

This norm has two nice properties. First, differentiation is continuous from \(\mathcal {C}^k\) to \(\mathcal {C}^{k-1}\). Second, the topology induced by the norms are finer as k goes to infinity. So the limit topology on \(\mathcal {S} = \bigcap _{k \in \mathbb {N}} \mathcal {C}^k\)goes as follows: a sequence of functions \(f_{n\in {\mathbb {N}}}\) in \(\mathcal {S}\) admits 0 as its limit if the sup-norm of all its derivatives \(\Vert \partial _\alpha f_{n}\Vert _{\infty }\) go to 0 for arbitrary multi-index \(\alpha \). This is topology is naturally finer than all the \(\mathcal {C}^k\) topologies and the differentiation is still continuous. Provided with this topology, \(\mathcal {S}\) is a Frechet space: it is complete and metrizable (though no norm is defined). Note that, although all the \(\mathcal {C}^k\) are Banach spaces, their descending intersection \(\bigcap _{k \in \mathbb {N}} \mathcal {C}^k\) is not.

Things are usually clearer and more explicit in the Fourier decomposition. Let us consider the Fourier decomposition of a function over \(\mathrm {SU}(2)\) on the Wigner matrices:

$$\begin{aligned} f(g)=\sum _{j,a,b}f^{j}_{ab}D^{j}_{ab}. \end{aligned}$$

By the Fourier convergence theorem, smoothness actually translates into a rapid fall-off of the Fourier coefficients:

$$\begin{aligned} f\in {\mathcal {S}}\quad \Longleftrightarrow \quad \forall K\in {\mathbb {N}}, \quad \sum _{j} |f^{j}| d_j^K < +\infty , \end{aligned}$$
(156)

where \(d_{j}=(2j+1)\) is the dimension of the spin-j representation and \(|f^{j}|\) can equally be the sup-norm or the square-norm of the matrix \(f^{j}_{ab}\). This also means that the Fourier coefficients of a distribution cannot diverge faster than polynomially:

$$\begin{aligned} \varphi \in {\mathcal {S}}^{*} \quad \Longleftrightarrow \quad \exists K, \quad \sum _{j} |\varphi ^{j}| d_j^{-K} < +\infty . \end{aligned}$$
(157)

The strong topology on \({\mathcal {S}}\) means that a sequence of smooth functions \(f_{n}\) converges to 0 in \({\mathcal {S}}\) if and only if all the K-power sums go to 0:

$$\begin{aligned} \lim _{n\rightarrow \infty }f_{n}=0 \quad \Longleftrightarrow \quad \forall K\in {\mathbb {N}}, \quad \sum _{j} |f^{j}_{n}| d_j^K \quad \underset{n\rightarrow \infty }{\longrightarrow }0. \end{aligned}$$
(158)

This ensures that the evaluations of a distribution \(\varphi \) will also converge \(\varphi (f_{n})\rightarrow 0\).

Let us apply this to the holonomy constraints for functions invariant under conjugation on \(\mathrm {SU}(2)\). In this case, all functions decompose on the characters,

$$\begin{aligned} \varphi (g)=\sum _{j\in \frac{{\mathbb {N}}}{2}}\varphi _{j}\chi _{j}(g), \end{aligned}$$

and the eigenvalue problem \(\chi (g)\varphi (g)=2\varphi (g)\) translates into a recursion relation on the Fourier coefficients:

$$\begin{aligned} \rho \varphi _{0}=\varphi _{\frac{1}{2}},\quad \rho \varphi _{j\ge \frac{1}{2}}=\varphi _{j-\frac{1}{2}}+\varphi _{j+\frac{1}{2}}. \end{aligned}$$
(159)

Once we fix the initial condition \(\varphi _{0}\), this recursive equation has a solution for every complex value \(\rho \in {\mathbb {C}}\), but this does not systematically defines a solution state, in \(L^{2}\) or a distribution. The solution to the recursion is given in terms of the two solutions \(\mu _{\pm }\) of the quadratic equation \(\mu ^{2}-\rho \mu +1=0\):

$$\begin{aligned} \forall j\in \frac{{\mathbb {N}}}{2}, \quad \varphi _{j}=\frac{\mu _{+}^{2j+1}-\mu _{-}^{2j+1}}{\mu _{+}-\mu _{-}}. \end{aligned}$$
(160)

For \(\rho =2\), the discriminant \((\rho ^{2}-4)\) vanishes and this ansatz fails leads: instead of the power law, we get a linear growth \(f_{j}=(2j+1)\), which leads back to the \(\delta \)-distribution peaked on the identity. For real values \(|\rho |< 2\), the discriminant is negative and we get an oscillatory solution. Mapping \(\rho \) to an angle \(\theta \) defined as \(\rho =2\cos \theta \), the two solutions are \(\mu _{\pm }=\exp (\pm i \theta )\) and the Fourier coefficients are \(\varphi _{j}=\sin (2j+1)\theta \,/\sin \theta \,=\chi _{j}(\theta )\), which gives a \(\delta \)-distribution fixing the class angle of the group element g to \(\theta \). For \(|\rho |>2\), the positive discriminant will leads to exponentially divergent coefficients \(\varphi _{j}\) and do not define a proper distribution.

1.2 The holonomy constraint as a recursion relation

Let us write the holonomy constraint equation \(\widehat{\chi }\,\varphi =2\,\varphi \) for distributions on \(\mathrm {SU}(2)\) in the Fourier decomposition. We decompose the distribution \(\varphi \) on the Wigner matrices:

$$\begin{aligned} \varphi (h)=\sum _{j,m,n} (2j+1)\varphi ^{j}_{nm}\,D^{j}_{mn}(h), \end{aligned}$$

with the magnetic moment indices \(-j\le m,n\le +j\). We know from spin recoupling the action of a spin-\(\frac{1}{2}\) Wigner matrix on an arbitrary spin (for example calculating the corresponding Clebsh–Gordan coefficients using the Schwinger representation of the \({\mathfrak {su}}(2)\) Lie algebra in terms of a pair of harmonic oscillators):

$$\begin{aligned} \forall A,B= & {} \pm , \quad D^{\frac{1}{2}}_{\frac{A}{2},\frac{B}{2}}(h)\,D^{j}_{m,n}(h) \nonumber \\= & {} \frac{1}{2j+1}\Bigg [AB\sqrt{(j+Am+1)(j+Bn+1)}\,D^{j+\frac{1}{2}}_{m+\frac{A}{2},n+\frac{B}{2}}(h) \nonumber \\&+\, \sqrt{(j-Am)(j-Bn)}\,D^{j-\frac{1}{2}}_{m+\frac{A}{2},n+\frac{B}{2}}(h) \Bigg ], \end{aligned}$$
(161)

with the obvious action on the trivial spin \(j=0\) mode. The action of the spin-\(\frac{1}{2}\) character is obtained by adding the two operators for \(A=B=\pm \):

$$\begin{aligned} \chi _{\frac{1}{2}}(h) = D^{\frac{1}{2}}_{+\frac{1}{2},+\frac{1}{2}}(h)+D^{\frac{1}{2}}_{-\frac{1}{2},-\frac{1}{2}}(h). \end{aligned}$$

We can then translate the functional equation \(\chi _{\frac{1}{2}}\,\varphi =2\,\varphi \) into a recursion relation on the Fourier coefficients \(\varphi ^{j}_{nm}\):

$$\begin{aligned} 2\varphi ^{0}= & {} \varphi ^{\frac{1}{2}}_{-\frac{1}{2},-\frac{1}{2}} + \varphi ^{\frac{1}{2}}_{+\frac{1}{2},+\frac{1}{2}}, \nonumber \\ 2(2j+1)\varphi ^{j}_{n,m}= & {} \varphi ^{j-\frac{1}{2}}_{n-\frac{1}{2},m-\frac{1}{2}}\sqrt{(j+m)(j+n)} + \varphi ^{j+\frac{1}{2}}_{n-\frac{1}{2},m-\frac{1}{2}}\sqrt{(j-m+1)(j-n+1)} \nonumber \\&+\, \varphi ^{j-\frac{1}{2}}_{n+\frac{1}{2},m+\frac{1}{2}}\sqrt{(j-m)(j-n)} \nonumber \\&+\, \varphi ^{j+\frac{1}{2}}_{n+\frac{1}{2},m+\frac{1}{2}}\sqrt{(j+m+1)(j+n+1)}. \end{aligned}$$
(162)

We easily check that \(\varphi ^{j}_{nm}=\delta _{nm}=D^{j}_{nm}({\mathbb {I}})\) is a solution to these recursion relations, which corresponds to the \(\delta \)-distribution on the group \(\varphi (h)=\delta (h)\). But straightforward computations also tell us that \(\varphi ^{j}_{nm}=D^{j}_{nm}(J_{a})\) for \(a=z,\pm \) are also solutions, or explicitly:

$$\begin{aligned} \varphi ^{j}_{nm}= \left| \begin{array}{l} D^{j}_{nm}(J_{z})=m\delta _{n,m}\\ D^{j}_{nm}(J_{+})=\sqrt{(j+m+1)(j-m)}\delta _{n,m+1}\\ D^{j}_{nm}(J_{-})=\sqrt{(j-m+1)(j+m)}\delta _{n,m-1} \end{array}\right. \end{aligned}$$
(163)

In general, we can show that for every vector \(\vec {x}\in {\mathbb {R}}^{3}\), the coefficients \(\varphi ^{j}_{nm}=D^{j}_{nm}(\vec {x}\cdot \vec {J})\) solve the recursion equation. Indeed, one differentiates \(\chi _{\frac{1}{2}}(h)D^{j}_{nm}(h)\) and evaluates it at the identity,

$$\begin{aligned} \left. -i\partial _{x}\chi _{\frac{1}{2}}(h)D^{j}_{nm}(h)\right| _{{\mathbb {I}}} = \chi _{\frac{1}{2}}({\mathbb {I}})D^{j}_{nm}(\vec {x}\cdot \vec {J})+\chi _{\frac{1}{2}}(\vec {x}\cdot \vec {J})D^{j}_{nm}({\mathbb {I}}) =2D^{j}_{nm}(\vec {x}\cdot \vec {J}). \end{aligned}$$
(164)

Then one applies the recoupling relations given above to decompose \(\chi _{\frac{1}{2}}D^{j}_{nm}\) onto the Wigner matrices for spins \(j\pm \frac{1}{2}\) in order to prove that \(\varphi ^{j}_{nm}=D^{j}_{nm}(\vec {x}\cdot \vec {J})\) do satisfy the required recursion relation. These coefficients actually correspond to the first derivative of the \(\delta \)-distribution, \(\varphi (h)=\partial _{x}\delta (h)\).

In order to classify all the solutions to these recursion relations, we need to determine the initial data required to implement the recursion. The problem is that the size of the matrices \(\varphi ^{j}\) increases with the mode j. If we start with fixing \(\varphi ^{0}\), then this determines only one component out of four of the next matrix \(\varphi ^{\frac{1}{2}}\), thus leaving three new initial conditions to be freely chosen. The \(4=(1+3)\) solutions corresponding to those required initial conditions are \(\delta \) and the \(\partial _{a}\delta \)’s.

Moving up to the spin \(j=1\) components, we have four recursion relations determining the matrix elements \(\varphi ^{1}\) in terms of the \(\varphi ^{\frac{1}{2}}\) matrix (indeed one relation per matrix elements of \(\varphi ^{\frac{1}{2}}\)). This leaves us with \(5=(9-4)\) undetermined matrix elements, which we need to specify as initial conditions. Choosing vanishing \(\varphi ^{0}\) and \(\varphi ^{\frac{1}{2}}\) initial conditions, these five new initial conditions correspond to the five linearly-independent second-order-derivative solutions of the holonomy constraint: \(\partial _{a}\partial _{b}\) and \(\partial _{a}\partial _{a}-\partial _{b}\partial _{b}\) for \(a\ne b\). So in total, we need to specify 9 initial conditions up to the spin 1 modes. And this will inexorably grow as we explore higher and higher spins, with \(4j+1=[(2j+1)^{2}-(2j)^{2}]\) new initial conditions for the matrix \(\varphi ^{j}\), thus reproducing the infinite dimensional space of solutions of the holonomy constraint, with the tower of higher and higher derivatives.

1.3 Laplacian constraint, double recursion and flatness equations

We introduce another constraint supplementing the holonomy constraint in order to truly impose flatness and get the \(\delta \)-distribution as unique solution: we impose the Laplacian constraint \((\widetilde{\Delta }-\Delta )=0\), where \(\Delta =\partial ^{L}_{a}\partial ^{L}_{a}=\partial ^{R}_{a}\partial ^{R}_{a}\) is the usual Laplacian operator and \(\widetilde{\Delta }=\partial ^{L}_{a}\partial ^{R}_{a}\) mixes the right and left derivations. These two operators do not change the spin j and act rather simply on the Wigner matrices:

$$\begin{aligned} \Delta D^{j}_{mn}(h)=-D^{j}_{mn}(hJ_{a}J_{a})=-j(j+1)D^{j}_{mn}(h) , \quad \Delta D^{j}_{mn}(h)=-D^{j}_{mn}(J_{a}hJ_{a}). \end{aligned}$$
(165)

Using the explicit action of the three \({\mathfrak {su}}(2)\) generators, and applying the Cauchy–Schwarz inequality to bound the sums, we can check that the operator \((\widetilde{\Delta }-\Delta )\) is positive. We can also translate the Laplacian constraint \(\Delta \varphi =\widetilde{\Delta }\varphi \) into equations on the Fourier coefficient matrices \(\varphi ^{j}\):

$$\begin{aligned} j(j+1)\varphi ^{j}_{nm}= & {} nm\,\varphi ^{j}_{nm} + \frac{1}{2}\varphi ^{j}_{n-1,m-1}\sqrt{(j+n)(j-n+1)(j+m)(j-m+1)}\nonumber \\&+\, \frac{1}{2}\varphi ^{j}_{n+1,m+1}\sqrt{(j-n)(j+n+1)(j-m)(j+m+1)}. \end{aligned}$$
(166)

This is a recursion at fixed spin j on the matrix elements of each \(\varphi ^{j}\) independently. It works at fixed \((n-m)\), that is along the diagonals of the matrix, determining the matrix elements from, say, the highest weight components:

$$\begin{aligned} \begin{array}{ll} \varphi ^{j,j}&{}\rightarrow \,\varphi ^{j-1,j-1}\rightarrow \varphi ^{j-2,j-2}\rightarrow \dots \\ \varphi ^{j,j-1}&{}\rightarrow \, \varphi ^{j-1,j-2}\rightarrow \varphi ^{j-2,j-3}\rightarrow \dots \\ \varphi ^{j,j-2}&{}\rightarrow \,\varphi ^{j-1,j-3}\rightarrow \varphi ^{j-2,j-4}\rightarrow \dots \end{array} \end{aligned}$$

Putting this constraint with the holonomy constraint, we get a double recursion structure. The holonomy constraint realizes a recursion on the spin j, determining the matrix \(\varphi ^{j}\) from the lower spin matrices, while the Laplacian constraint implements a recursion on the magnetic moment m within each matrix \(\varphi ^{j}\):

$$\begin{aligned} \varphi ^{0} \rightarrow \left( \begin{array}{cc}\searrow &{} \searrow \\ \searrow &{} \searrow \end{array} \right) \rightarrow \left( \begin{array}{ccc}\searrow &{} \searrow &{} \searrow \\ \searrow &{} \searrow &{} \searrow \\ \searrow &{} \searrow &{} \searrow \end{array} \right) \rightarrow \left( \begin{array}{cccc}\searrow &{} \searrow &{} \searrow &{} \searrow \\ \searrow &{} \searrow &{} \searrow &{} \searrow \\ \searrow &{} \searrow &{} \searrow &{} \searrow \\ \searrow &{} \searrow &{} \searrow &{} \searrow \end{array} \right) \rightarrow \dots \end{aligned}$$

This allows to solve the problem of the infinite initial conditions needed for the holonomy constraint. We easily check that \(\varphi _{nm}=\delta _{nm}\) is a solution:

$$\begin{aligned} j(j+1)=m^{2}+\frac{1}{2}(j+m)(j-m+1)+\frac{1}{2}(j-m)(j+m+1). \end{aligned}$$

But then, we completely solve the constraint and show that it implies that the function is invariant under conjugation:

Proposition 6.3

Let us consider the Laplacian constraint \((\Delta -\widetilde{\Delta })\,\varphi =0\) translated to the Fourier decomposition \(\varphi (h)=\sum _{j,m,n}(2j+1){\mathrm {Tr}}\,\varphi ^{j}D^{j}(h)\). Then the each of the Fourier coefficient matrix \(\varphi ^{j}\) at fixed spin j is proportional to the identity. This means that \(\varphi \) is invariant under conjugation.

Proof

Let us fix \(j\ge \frac{1}{2}\). The spin-0 component \(\varphi ^{0}\) is unconstrained and left free. The recursion relation (166) allows to start with an element \(\varphi _{j,j-M}\) with \(0\le M \le 2j\) and to determine all of the following components along the corresponding diagonal, \(\varphi _{j-N,j-M-N}\) for \(0\le N\le (2j-M)\). One actually gets a relation in terms of combinatorial factors:

$$\begin{aligned} \varphi _{j-N,j-M-N}= \varphi _{j,j-M}\, \sqrt{\frac{(M+N)!}{M!N!}\,\frac{(2j)!(2j-M-N)!}{(2j-M)!(2j-N)!}}. \end{aligned}$$
(167)

In particular, one obtains for the other end of the diagonal:

$$\begin{aligned} \varphi _{-j+M,-j} = \varphi _{j,j-M}\,\frac{(2j)!}{M!(2j-M)!}. \end{aligned}$$
(168)

The trick is that the recursion relation (166) is symmetric under the exchange \((n,m)\leftrightarrow (-m,-n)\): we start now from the other end of the same diagonal and work our way back to the initial top element. Therefore the previous equality holds but in the opposite way:

$$\begin{aligned} \varphi _{j,j-M}= & {} \varphi _{-j+M,-j}\,\frac{(2j)!}{M!(2j-M)!} = \varphi _{j,j-M}\left( \frac{(2j)!}{M!(2j-M)!}\right) ^{2}, \nonumber \\&\text {thus either} \quad \frac{(2j)!}{M!(2j-M)!}=1 \quad \text {or}\quad \varphi _{j,j-M}=0. \end{aligned}$$
(169)

In the special case \(M=0\), along the principal diagonal, the recursion relation simplifies and reads \(\varphi ^{j,j}=\varphi ^{j-1,j-1}=\varphi ^{j-2,j-2}=\dots \). For all the other cases \(M\ne 0\), the matrix elements must vanish. This concludes the proof that the matrix \(\varphi ^{j}\) must be proportional to the identity. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Charles, C., Livine, E.R. The Fock space of loopy spin networks for quantum gravity. Gen Relativ Gravit 48, 113 (2016). https://doi.org/10.1007/s10714-016-2107-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10714-016-2107-5

Keywords

Navigation