Skip to main content
Log in

A model of spinfoam coupled with an environment

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

Abstract

In this paper, an open quantum system theory for spinfoams is developed. This new formalism aims at deriving an effective Lindblad equation to compute the reduced dynamics of a quantum gravitational field. The system parameters are determined from numerical ab initio calculations, based on the spinfoam formalism. This theoretical proposal is illustrated by means of examples. The decoherence effect can induce the relaxation of the quantum gravitational state toward a sate of a small area. This is analogue to the well-known Purcell relaxation of QED, for which the qubits are replaced by the spin-network representation of the gravitational field. Some thermodynamic properties of these systems are computed, and several issues with the thermal time hypothesis are underlined. Moreover, the results suggest that further approximations can be performed to study reduced dynamics of quantum space-time.

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

References

  1. Gambini, R., Pullin, J., Ashtekar, A.: Loops, Knots, Gauge Theories and Quantum Gravity. Cambridge Monographs on Mathematical Physics. Cambridge University Press, Cambridge (1996). https://doi.org/10.1017/CBO9780511524431

    Book  Google Scholar 

  2. Rovelli, C.: Quantum Gravity. Cambridge University Press, Cambridge (2007)

    MATH  Google Scholar 

  3. Ashtekar, A., Reuter, M., Rovelli, C.: From General Relativity to Quantum Gravity. arXiv:1408.4336 (2014)

  4. Rovelli, C.: Space and Time in Loop Quantum Gravity. arXiv:1802.02382 (2018)

  5. Rovelli, C., Vidotto, F.: Covariant Loop Quantum Gravity: An Elementary Introduction to Quantum Gravity and Spinfoam Theory. Cambridge University Press, Cambridge (2014). https://doi.org/10.1017/CBO9781107706910

    Book  MATH  Google Scholar 

  6. Rovelli, C.: Simple model for quantum general relativity from loop quantum gravity. J. Phys.: Conf. Ser. 314, 012006 (2011). https://doi.org/10.1088/1742-6596/314/1/012006. arXiv:1010.1939

    Article  Google Scholar 

  7. Perez, A.: The Spin-Foam Approach to Quantum Gravity. Living Rev. Rel. 16(1), 3 (2013). https://doi.org/10.12942/lrr-2013-3

    Article  MATH  Google Scholar 

  8. Barrau, A., Grain, J.: Loop quantum gravity and observations. arXiv:1410.1714 [astro-ph, physics:gr-qc] (2014)

  9. Cai, Y.F., Wang, Y.: Testing quantum gravity effects with latest CMB observations. Phys. Lett. B 735, 108–111 (2014). https://doi.org/10.1016/j.physletb.2014.06.019. arXiv:1404.6672

  10. Perez, A.: Black holes in loop quantum gravity. Rep. Prog. Phys. 80(12), 126901 (2017). https://doi.org/10.1088/1361-6633/aa7e14. arXiv:1703.09149

    Article  ADS  MathSciNet  Google Scholar 

  11. Christodoulou, M.: Transition de géométrie en gravité quantique à boucles covariante. Ph.D. Thesis, AIX-MARSEILLE Univeristy. http://www.theses.fr/2017AIXM0273/document (2017)

  12. D’Ambrosio, F., Christodoulou, M., Martin-Dussaud, P., Rovelli, C., Soltani, F.: The End of a Black Hole’s Evaporation—Part I. arXiv:2009.05016 [gr-qc] (2020)

  13. Dona, P., Sarno, G.: Numerical methods for EPRL spin foam transition amplitudes and Lorentzian recoupling theory. Gen. Relativ. Gravit. 50(10), 127 (2018). arXiv:1807.03066

    Article  ADS  MathSciNet  Google Scholar 

  14. Dona, P., Fanizza, M., Sarno, G., Speziale, S.: Numerical Study of the Lorentzian EPRL Spin Foam Amplitude. Phys. Rev. D 100, 106003. arXiv:1903.12624 (2019)

  15. Breuer, H.P., Petruccione, F.: The Theory of Open Quantum Systems. Oxford University Press, Oxford (2007)

    Book  Google Scholar 

  16. Feller, A.: Entanglement and Decoherence in Loop Quantum Gravity. Ph.D. Thesis, ENS de Lyon. https://tel.archives-ouvertes.fr/tel-01650029/document (2017)

  17. Azouit, R., Sarlette, A., Rouchon, P.: Adiabatic Elimination for Open Quantum Systems with Effective Lindblad Master Equations. IEEE 55th Conference on Decision and Control (CDC), Las Vegas, NV, USA, pp. 4559–4565. arXiv:1603.04630 [quant-ph] (2016)

  18. Paetz, T.T.: An Analysis of the ‘Thermal-Time Concept’ of Connes and Rovelli. Ph.D. Thesis, Georg-August-Universität Göttingen. http://www.theorie.physik.uni-goettingen.de/forschung/qft/theses/dipl/Paetz.pdf (2010)

  19. Baez, J.C.: An Introduction to Spin Foam Models of Quantum Gravity and BF Theory. Gausterer H., Pittner L., Grosse H. (eds) Geometry and Quantum Physics. Lecture Notes in Physics, vol 543. Springer, Berlin, Heidelberg. https://doi.org/10.1007/3-540-46552-9_2arXiv:gr-qc/9905087 (1999)

  20. Livine, E.R.: A Short and Subjective Introduction to the Spinfoam Framework for Quantum Gravity. arXiv:1101.5061 (2011)

  21. Engle, J., Livine, E., Pereira, R., Rovelli, C.: LQG vertex with finite Immirzi parameter. Nucl. Phys. B 799(1–2), 136 (2008). https://doi.org/10.1016/j.nuclphysb.2008.02.018. arXiv:0711.0146

    Article  ADS  MathSciNet  MATH  Google Scholar 

  22. Christodoulou, M., Rovelli, C., Speziale, S., Vilensky, I.: Realistic observable in background-free quantum gravity: the Planck-star tunnelling-time. Phys. Rev. D 94(8), 084035 (2016). https://doi.org/10.1103/PhysRevD.94.084035

    Article  ADS  MathSciNet  Google Scholar 

  23. Lindblad, G.: On the generators of quantum stochastic semigroups. Commun. Math. Phys. 48(2), 119 (1976). https://doi.org/10.1007/BF01608499

    Article  ADS  MathSciNet  MATH  Google Scholar 

  24. Oeckl, R.: The Tenth Marcel Grossmann Meeting. pp. 2296–2300 (2006). https://doi.org/10.1142/9789812704030_0321. arXiv:gr-qc/0401087

  25. Banburski, A., Chen, L.Q., Freidel, L., Hnybida, J.: Pachner moves in a 4d Riemannian holomorphic Spin Foam model. Phys. Rev. D 92(12), 124014 (2015). https://doi.org/10.1103/PhysRevD.92.124014. arXiv:1412.8247

    Article  ADS  MathSciNet  Google Scholar 

  26. Chen, L.Q.: Bulk amplitude and degree of divergence in 4d spin foams. Phys. Rev. D 94(10), 104025 (2016). https://doi.org/10.1103/PhysRevD.94.104025. arXiv:1602.01825

    Article  ADS  MathSciNet  Google Scholar 

  27. Dittrich, B., Schnetter, E., Seth, C.J., Steinhaus, S.: Coarse graining flow of spin foam intertwiners. Phys. Rev. D 94(12), 124050 (2016). https://doi.org/10.1103/PhysRevD.94.124050. arXiv:1609.02429

    Article  ADS  Google Scholar 

  28. Feller, A., Livine, E.R.: Surface state decoherence in loop quantum gravity, a first toy model. Class. Quantum Grav. 34(4), 045004 (2017). https://doi.org/10.1088/1361-6382/aa525c. arXiv:1607.00182

    Article  ADS  MathSciNet  MATH  Google Scholar 

  29. Delcamp, C., Dittrich, B.: Towards a phase diagram for spin foams. Class. Quantum Grav. 34(22), 225006 (2017). https://doi.org/10.1088/1361-6382/aa8f24. arXiv:1612.04506

    Article  ADS  MathSciNet  MATH  Google Scholar 

  30. Gentle, A.P.: Regge calculus: a unique tool for numerical relativity. Gen. Relativ. Gravit. 34(10), 1701 (2002). https://doi.org/10.1023/A:1020128425143

    Article  MathSciNet  MATH  Google Scholar 

  31. Christodoulou, M., D’Ambrosio, F.: Characteristic Time Scales for the Geometry Transition of a Black Hole to a White Hole from Spinfoams. arXiv:1801.03027 [gr-qc] (2018)

  32. Garraway, B.M.: Nonperturbative decay of an atomic system in a cavity. Phys. Rev. A 55(3), 2290 (1997). https://doi.org/10.1103/PhysRevA.55.2290

    Article  ADS  MathSciNet  Google Scholar 

  33. Barrett, J.W., Dowdall, R.J., Fairbairn, W.J., Hellmann, F., Pereira, R.: Lorentzian spin foam amplitudes: graphical calculus and asymptotics. Class. Quantum Grav. 27(16), 165009 (2010). https://doi.org/10.1088/0264-9381/27/16/165009. arXiv:0907.2440

    Article  ADS  MathSciNet  MATH  Google Scholar 

  34. Magliaro, E., Perini, C.: Regge gravity from spinfoams. Int. J. Mod. Phys. D 22(02), 1350001 (2013). https://doi.org/10.1142/S0218271813500016. arXiv:1105.0216

    Article  ADS  MathSciNet  MATH  Google Scholar 

  35. Han, M., Zhang, M.: Asymptotics of spinfoam amplitude on simplicial manifold: Lorentzian theory. Class. Quantum Grav. 30(16), 165012 (2013). https://doi.org/10.1088/0264-9381/30/16/165012. arXiv:1109.0499

    Article  ADS  MathSciNet  MATH  Google Scholar 

  36. Livine, E.R., Speziale, S.: New spinfoam vertex for quantum gravity. Phys. Rev. D 76(8), 084028 (2007). https://doi.org/10.1103/PhysRevD.76.084028

    Article  ADS  MathSciNet  Google Scholar 

  37. Bianchi, E., Magliaro, E., Perini, C.: Coherent spin-networks. Phys. Rev. D 82(2), 024012 (2010). https://doi.org/10.1103/PhysRevD.82.024012. arXiv:0912.4054

    Article  ADS  Google Scholar 

  38. Rovelli, C., Vidotto, F.: Small black/white hole stability and dark matter. Universe 4(11), 127 (2018). https://doi.org/10.3390/universe4110127. arXiv:1805.03872

    Article  ADS  Google Scholar 

  39. Gross, M., Haroche, S.: Superradiance: an essay on the theory of collective spontaneous emission. Phys. Rep. 93(5), 301 (1982). https://doi.org/10.1016/0370-1573(82)90102-8

    Article  ADS  Google Scholar 

  40. Santos, J.P., Semião, F.L.: Master equation for dissipative interacting qubits in a common environment. Phys. Rev. A 89(2), 022128 (2014). https://doi.org/10.1103/PhysRevA.89.022128. arXiv:1311.0018

    Article  ADS  Google Scholar 

  41. Bianchi, E.: Entropy of Non-extremal Black Holes from Loop Gravity. arXiv:1204.5122 [gr-qc, physics:hep-th] (2012)

  42. Haroche, S., Raimond, J.M.: Exploring the Quantum: Atoms, Cavities, and Photons. OUP Oxford, Oxford (2006)

    Book  Google Scholar 

  43. Louisell, W.H.: Quantum Statistical Properties of Radiation. Wiley, Hoboken (1973)

    MATH  Google Scholar 

  44. Connes, A., Rovelli, C.: Von Neumann algebra automorphisms and time-thermodynamics relation in generally covariant quantum theories. Class. Quantum Grav. 11(12), 2899 (1994). https://doi.org/10.1088/0264-9381/11/12/007

    Article  ADS  MathSciNet  MATH  Google Scholar 

  45. Rovelli, C.: Statistical mechanics of gravity and the thermodynamical origin of time. Class. Quantum Grav. 10(8), 1549 (1993). https://doi.org/10.1088/0264-9381/10/8/015

    Article  ADS  MathSciNet  MATH  Google Scholar 

  46. Menicucci, N.C., Olson, S.J., Milburn, G.J.: Clocks and Relationalism in the Thermal Time Hypothesis. arXiv:1108.0883 [gr-qc] (2011)

  47. Gemmer, J., Michel, M., Mahler, G.: Quantum Thermodynamics: Emergence of Thermodynamic Behavior Within Composite Quantum Systems, 2nd edn. Lecture Notes in Physics. Springer, Berlin (2009)

    Book  Google Scholar 

  48. Barrett, J.W., Naish-Guzman, I.: The Ponzano–Regge model. Class. Quantum Grav. 26(15), 155014 (2009). https://doi.org/10.1088/0264-9381/26/15/155014. arXiv:0803.3319

    Article  ADS  MATH  Google Scholar 

Download references

Acknowledgements

Thanks to Giorgio Sarno for helpful discussions about the SL2Cfoam code, and to David Viennot for advised comments. Simulations have been executed on computers from the Utinam Institute of the Université de Franche-Comté, supported by the Région de Franche-Comté and Institut des Sciences de l’Univers (INSU).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Quentin Ansel.

Ethics declarations

Conflict of interest

The author declares that he has no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

A Some mathematical definitions

This appendix is devoted to some mathematical definitions, which are omitted in Sect. 2 for conciseness. Further technical details on this subject can be found in Refs. [2, 5, 19, 20].

Definition 4

A spin-network \(\psi \) is a triple \((\varGamma ,\chi , \imath \))consisting of:

  1. 1.

    a 1-dimensional oriented complex \(\varGamma \), represented by a graph whose vertices are points of the space-time manifold and edges are paths connecting the points.

  2. 2.

    a labeling \(\chi \) of each edge e by a unitary irreducible representation \(\chi _e\) of \({\mathfrak {H}}\), where \({\mathfrak {H}} \subseteq {\mathfrak {G}}\) is a subgroup of the gauge group of the theory. Elements of \({\mathfrak {H}}\) associated with edges encode the structure of space at the classical level.

  3. 3.

    a labeling \(\imath \) of each vertex v by an intertwiner

    $$\begin{aligned} \imath _v:\chi _{e_1} \otimes \cdot \cdot \cdot \otimes \chi _{e_n} \rightarrow \chi _{e'_1} \otimes \cdot \cdot \cdot \otimes \chi _{e'm} \end{aligned}$$

    where \(e_1,\ldots ,e_n\) are the edges incoming to v and \(e'_1,\ldots ,e'_m\) are the edges outgoing from v.

It is usual to consider \({\mathfrak {G}} = SL(2,{\mathbb {C}})\) and \({\mathfrak {H}} = SU(2)\) in 3+1D gavity, or \({\mathfrak {G}} = {\mathfrak {H}} = SU(2)\) in 3D Euclidan gravity [5]. For reasons that will become clearer below, it is preferred to substitute the names vertex and edge by node and link.

Definition 5

A spinfoam F is a triple \(({{\tilde{\varGamma }}},{{\tilde{\chi }}},{{\tilde{\imath }}} \))consisting of:

  1. 1.

    a 2-dimensional oriented complex \({{\tilde{\varGamma }}}\).

  2. 2.

    a labeling \({{\tilde{\chi }}}\) of each face f by a unitary irreducible representation \({{\tilde{\chi }}}_e\) of \({\mathfrak {G}}\).

  3. 3.

    a labeling \({{\tilde{\imath }}}\) of each vertex e by an intertwiner

    $$\begin{aligned} {{\tilde{\imath }}}_e:{{\tilde{\chi }}}_{f_1} \otimes \cdot \cdot \cdot \otimes {{\tilde{\chi }}}_{f_n} \rightarrow {{\tilde{\chi }}}_{f'_1} \otimes \cdot \cdot \cdot \otimes {{\tilde{\chi }}}_{f'_m} \end{aligned}$$

    where \(f_1,\ldots ,f_n\) are faces incoming to e and \(f_1',\ldots ,f_m'\) are faces outgoing from e.

Notice that here, vertices are D dimensional objects, edges are D-1 dimensional objects, and faces are D-2 dimensional objects.

Definition 6

Let \(\psi \) be a spin-network \(\psi =(\varGamma ,\chi ,\imath )\). A spinfoam \(F:\emptyset \rightarrow \psi \) is a triple \(({{\tilde{\varGamma }}},{{\tilde{\chi }}}, {{\tilde{\i }}})\) verifying the definition 5, and:

  1. 1.

    for any link l of \(\varGamma \), \(\chi _l = P({{\tilde{\chi }}}_f)\) if f is incoming to l, and \((\chi _l)^* = P({{\tilde{\chi }}}_f)\) otherwise.

  2. 2.

    for any node n of \(\varGamma \), \(\imath _n = P({{\tilde{\imath }}}_e)\).

P is a map from unitary irreducible representations of \({\mathfrak {G}}\) to unitary irreducible representations of \({\mathfrak {H}}\). It is an input of the theory.

This definition gives the transition amplitude from the vacuum to a given spin-network. The evolution from a spin-network to another one is given by:

Definition 7

A spinfoam \(F: \psi \rightarrow \psi '\) is defined by \(F:\emptyset \rightarrow \psi ^* \otimes \psi '\), where \(^*\) denotes the dual of a spin-network.

A spinfoam model is defined with a specific choice of (\({\mathfrak {G}},{\mathfrak {H}},\imath ,P,Z\)).

Proof of proposition 1

This appendix is devoted to the proof of proposition 1.

Proof

We provide here a simple proof with a minimum of details. Further information is given in Reference [17].

We are looking for a master equation \(d_t \rho _e = \mathcal {L}_e \rho = [\mathcal {L}_{e,0} + \epsilon \mathcal {L}_{e,1} + o(\epsilon ^2) ]\rho _e\) that provides the evolution of the system up to an error of \(\epsilon ^2\). To determine this equation, we introduce the map:

$$\begin{aligned} \rho (t) = K( \rho _e(t) )= \sum _{m\ge 0} \epsilon ^m K_m (\rho _e(t)). \end{aligned}$$

Then, by definition we have:

$$\begin{aligned} \frac{d \rho }{dt} = \mathcal {L}_0 K(\rho _e) + \epsilon \mathcal {L}_1 K(\rho _e) = K(\mathcal {L}_e \rho _e). \end{aligned}$$

Using the expansion of K in powers of \(\epsilon \), we deduce that:

$$\begin{aligned} \mathcal {L}_0 K_0(\rho _e) = K_0(\mathcal {L}_{e,0} \rho _e) ~~(m=0), \end{aligned}$$
(23)

and,

$$\begin{aligned} \mathcal {L}_0 K_1(\rho _e) + \mathcal {L}_1 K_0(\rho _e) = K_0(\mathcal {L}_{e,1} \rho _e) ~~(m=1). \end{aligned}$$
(24)

Since we are interested in dynamics restricted in \(\mathcal {D}_0\), we define \(K_0\) as a projector on \(\mathcal {D}_0\): \(K_0 (\rho ) = P_0 \rho P_0 ^\dagger \), with \( P_0 = \sum _{n=1}^{dim \mathcal {H}_0} | n \rangle \langle n |\), with \(\{| n \rangle \}\) a basis of \(\mathcal {H}_0\). From the definition of \(\mathcal {D}_0\), we deuce that Eq. (23) gives \(\mathcal {L}_{e,0}=0\). We also introduce the Kraus map of a solution of the unperturbed system when \(t \rightarrow \infty \): \(U_0\rho (0)=\lim _{t \rightarrow \infty }e^{t\mathcal {L}_0}\rho (0) \equiv \sum _\mu M_\mu \rho (0) M_\mu ^\dagger \), with \(\{ M_\mu \}\) an ensemble of operators such that \(\sum _\mu M_\mu ^\dagger M_\mu = {\mathbb {I}}_{\mathcal {H}} \}\).

To arrive at the desired result, we apply the super-operator \(U_0\) on Eq. (24). Since \(U_0.\mathcal {L}_0 =0\), we have:

$$\begin{aligned} U_0 \mathcal {L}_1 K_0(\rho _e) = U_0 K_0(\mathcal {L}_{e,1} \rho _e) \end{aligned}$$
(25)

Moreover, \(U_0\) leaves \(K_0\) unchanged. Therefore,

$$\begin{aligned}&U_0 \mathcal {L}_1 K_0(\rho _e) = K_0(\mathcal {L}_{e,1} \rho _e) \end{aligned}$$
(26)
$$\begin{aligned}&\quad \Rightarrow ~ K_0(U_0 \mathcal {L}_1 K_0(\rho _e)) = \mathcal {L}_{e,1} \rho _e \end{aligned}$$
(27)

In the second line, we have used \(K_0^2 = K_0\) and the fact that \(\mathcal {L}_{e,1}\) is restricted to \(\mathcal {H}_0\). The structure of the operator \(\mathcal {L}_1\) is by assumptions:

$$\begin{aligned} \mathcal {L}_1 \rho _e \equiv \sum _{n = 1}^{dim \mathcal {H}_0} \left( P_n \rho _e P_n ^\dagger -\frac{1}{2} P_n^\dagger P_n \rho _e - \frac{1}{2} \rho _e P_n^\dagger P_n \right) ~~,~~ P_n = | n \rangle \langle n |. \end{aligned}$$

Then,

$$\begin{aligned} \mathcal {L}_{e,1} \rho _e \equiv \sum _{n,m,\mu } |M_{\mu ,nm}|^2 \left( | n \rangle \langle m | \rho _e | m \rangle \langle n | -\frac{1}{2}| m \rangle \langle n | | n \rangle \langle m | \rho _e - \frac{1}{2} \rho _e | m \rangle \langle n | | n \rangle \langle m | \right) \end{aligned}$$

To finish the proof, we define \(\kappa _{nm} = \sum _\mu |M_{\mu ,nm}|^2\). \(\square \)

Numerical investigation of hypothesis 3

In this section, we present briefly the numerical investigation that leads us to the hypothesis 3. Contrary to examples in the main text, the 3D theory with gauge group \({\mathfrak {G}} = SU(2)\) is used. The transition amplitude is given by [5, 48]

$$\begin{aligned} W_{{{\tilde{\varGamma }}}} (j_l) = \mathcal {N}_{{{\tilde{\varGamma }}}} \sum _{j_f} \prod _f (-1)^ {j_f} (2 j_f +1) \prod _v (-1)^{\sum _{a=1}^6 j_{v,a}} \left\{ \begin{array}{ccc} j_{v,1} &{} j_{v,2} &{} j_{v,3} \\ j_{v,4} &{} j_{v,5} &{} j_{v,6} \end{array} \right\} , \end{aligned}$$
(28)

with \(j_l\) the spin label of the link l, of a boundary spin-network. \(j_f\) is the spin label of the face f of the spinfoam, and v is a vertex of the foam. (va) denotes the face a associated with the vertex v, \(\{\ldots \}\) is Wigner’s 6j-symbol, and \( \mathcal {N}_{{{\tilde{\varGamma }}}}\) is a normalization factor.

In the computation, we consider foams with \(V= 2,3\) or 4 vertices glued one-by-one with at most one edge. The 2-complex is defined without bubbles. The transition amplitude can be written as \(W(\psi _{out},\psi _{in})\), with \(\psi _{in}\) a spin-network formed by the boundary of the firsts vertices, and \(\psi _{out}\) the spin-network formed the boundary of other vertices. We can take symmetric networks, but it is not necessary for our purpose. For each in and out states, we are interested in a reduced number of degrees of freedom, given by a sub-spin-network, made of a single node and its three boundary links. The reduced Hilbert space \(\mathcal {H}_r\) is defined by the formal association of its canonical basis to a set of specific sub-spin-networks.

For example, if we consider a sub-spin-network composed of 1 node and 3 adjacent links with spin numbers (1, 1, 1) and (2, 2, 2), we have: \(\mathcal {H}_r = {\mathbb {C}}^2\), \(e_1 \rightarrow (1,1,1)\) and \(e_2 \rightarrow (2,2,2)\). For the numerical calculation presented below, sub-spin-networks are generated randomly, with dim\((\mathcal {H}_r) = 10\).

The transition amplitude between these sub-spin-networks is calculated using a large number (\(\sim 10^6\)) of different spin-networks generated randomly. This defines a map \(\mathcal {H}_r \otimes \mathcal {H}_r \rightarrow {\mathbb {C}}\) such that \((n,m) \rightarrow W_{nm}\). The goal is to find an approximation of \(W_{nm}\) with a simpler model. Here, we choose for the simplest model, a foam made of two unconnected vertices, such that \(W \propto W_{v_1}(j_l)W_{v_2}(j_l')\). With a minimization algorithm (the NMinimize function of Mathematica), we find a linear superposition of spin-networks such that the transition amplitudes between sub-spin-networks of the simplified system are the closest to the values given by the initial system. To quantify the difference between models, the following cost function is used: \(C= \Vert \mathbf {W}_1/\Vert \mathbf {W}_1 \Vert - \mathbf {W}_2/\Vert \mathbf {W}_2 \Vert \Vert \), where \(\mathbf {W}_i\) is a vector of dimension dim\((\mathcal {H}_r)^2\) whose components are transition amplitudes \(W_{nm}\) given by the model i. Vectors are normalized in the definition of the cost function in order to take into account the normalization factor \(\mathcal {N}_{{{\tilde{\varGamma }}}}\) introduced in Eq. (28).

Fig. 8
figure 8

a Minimum value of C found after numerical minimization as a function of the number of vertices V taken into account in the foam. b Probability density function of the value of C in the case \(V=4\). The minimum value is \(\min [C]=0.0056\) and the maximum value is \(\max [C]=1.9998\)

Figure 8 show the minimum value of C after numerical minimization of the cost function, and the probability density function to find a certain value of C. The second graph is computed with\(10^5\) random spin-networks of the simplified model. A step of 0.1 is used to compute the histogram. We observe that very low values of C are reachable, but these low values are not representative of general values of C (the average value is \(\approx 1.5\)). This emphasizes that a specific design of \(\psi _{bath}\) is required for the approximation. Due to many linear dependence between parameters, and the large number of parameters, the numerical optimization is not a difficult task, the convergence is fast. We also notice that the approximation is better for initial spinfoams of large dimension. This is explained by the fact that, the boundary of two distant vertices are weakly correlated.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ansel, Q. A model of spinfoam coupled with an environment. Gen Relativ Gravit 53, 39 (2021). https://doi.org/10.1007/s10714-021-02811-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10714-021-02811-5

Keywords

Navigation