Skip to main content
Log in

Recovering the QNEC from the ANEC

  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

We study the relative entropy in QFT comparing the vacuum state to a special family of purifications determined by an input state and constructed using relative modular flow. We use this to prove a conjecture by Wall that relates the shape derivative of relative entropy to a variational expression over the averaged null energy (ANE) of possible purifications. This variational expression can be used to easily prove the quantum null energy condition (QNEC). We formulate Wall’s conjecture as a theorem pertaining to operator algebras satisfying the properties of a half-sided modular inclusion, with the additional assumption that the input state has finite averaged null energy. We also give a new derivation of the strong superadditivity property of relative entropy in this context. We speculate about possible connections to the recent methods used to strengthen monotonicity of relative entropy with recovery maps.

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

Similar content being viewed by others

Notes

  1. Actually we only need that there is at least one state \(u' \psi \) with finite ANE.

  2. Despite being a significant result about modular hamiltonians, a rigorous mathematical proof of this formula does not yet exist in the literature. We we do not actually use this formula in our derivations and proof of the QNEC as formulated using relative entropy and half sided modular inclusions.

  3. Our conventions are not standard. The state labels on the relative modular operators are switched. We follow the conventions in [6] where the relative entropy and relative modular operators are labelled in the same way. Our labelling on the Connes cocycle are standard.

  4. From the algebraic point of view \(P\ge 0\) follows more directly from properties of modular Hamiltonians under inclusion [6].

  5. This should not be confused with the boosted states discussed in [50, 51]. These states are more singular since they involve modular flow with only half the \(\psi \)-modular Hamiltonian. The Connes cocycle is one way to deal with issues related to divergences that arise in that case, and some of the resulting physics is related to that discussed in [50, 51]. In particular we expect the bulk description of these states, in the context of AdS/CFT, to be the same for the part of the bulk spacetime that is (bulk) causally separated from the boundary entangling surface.

  6. Wall wrote down this conjecture in a different form involving entanglement entropy and the half integrated ANE. It is essentially equivalent to what is stated here, although the original form involves quantities that are not obviously well defined in algebraic quantum field theory. He also for the most part had in mind 2d QFTs. The original conjecture came from arguing that there was no other quantity that he could imagine except \(- \partial S\) that satisfies all the properties of the right hand side of (62).

  7. For a CFT a more general discussion is possible, but we do not consider this here.

  8. Recall that strong continuity of a vector uses the Hilbert space norm and weak continuity demands that the inner product with any fixed vector is continuous. A weakly continuous vector valued holomorphic function can be shown to be strongly continuous via the Cauchy integral formula.

  9. This works for a domain twice the size of \(T_I\), however the necessary bound below, as far as we are aware, does not extend beyond \(T_I\).

  10. The multi-dimensional edge of the wedge theorem is overkill here, we can apply the one dimensional edge of the wedge theorem along \(\eta \)-strips at fixed s or equivalently s-strips at fixed \(\eta \). The result is a function \(g(s,\eta )\) that is holomorphic in one of the variables when the other one is held fixed, which by Hartog’s theorem is holomorphic in both variables.

  11. The proof of this follows by considering the topological invariant that counts the zeros enclosed in some subset \(\varGamma \subset C\): \(2\pi i N_\epsilon = \int _{\partial \varGamma } ds \left( y_\epsilon '(s)/y_\epsilon (s)\right) \) of a holomorphic function. If the limiting function vanishes at some discrete point inside \(\varGamma \) then for small enough \(\varGamma \) we have, by uniform convergence applied to the integral, \(\lim _{\epsilon \rightarrow 0} N_\epsilon =1\). This contradicts the original assumption which gives \(N_\epsilon = 0\) for all \(\epsilon \).

  12. We should again take an arbitrary sequence \(\epsilon _n\) converging to zero and show that for any such sequence we have convergence to the same limit function. This follows the same logic as above so we do not repeat it.

  13. To show this set \(\lambda \zeta =t \) and consider \(\left| (1-e^{i \zeta \lambda })/(i\zeta \lambda )\right| = \left| (1-e^{i t})/t \right| \) which we have to show is bounded by 1. Now for real t we have: \(\left| (1-e^{it})/t\right| = 2 |\sin (t/2)/t| \le 1\). This later inequality extends, via the Phragmén-Lindelöf principle, into the t-upper half plane since \((1-e^{it})/it\) is holomorphic and bounded in the upper half plane (UHP). There is probably a much easier way to show this.

  14. The order of an entire function f(z) is \(\limsup _{r \rightarrow \infty } (\log \log {\max _{|z|=r} |f(z)|)/\log r}\).

  15. This is the well known fact that an entire function can be determined by its zeros up to an overall exponential of a polynomial who’s degree is the same as the order of the entire function.

  16. Actually Araki showed this for states \({\widehat{\varOmega }}, {\widehat{\psi }},{\widehat{\psi }}_n\) that are representatives of \(\varOmega ,\psi ,\psi _n\) in a natural positive cone associated to \({\mathcal {A}}_C'\) and some cyclic separating vector. Taking this vector to be \(\varOmega \) and using the fact that \(\varDelta '_{\varOmega |{\widehat{\psi }}_n} = \varDelta '_{\varOmega |\psi _n}\), where \({\widehat{\psi }}_n\) is related to \(\psi _n\) by some unitary in \({\mathcal {A}}_C\), we have the desired continuity.

  17. See Corollary 3, where we don’t actually need entire states to prove the existence of the following limit at real \(s,\eta \) given \(P_s\) is finite

References

  1. Bousso, R., Fisher, Z., Koeller, J., Leichenauer, S., Wall, A.C.: Proof of the quantum null energy condition. Phys. Rev. D 93, 024017 (2016). arXiv:1509.02542

    ADS  MathSciNet  Google Scholar 

  2. Bousso, R., Fisher, Z., Leichenauer, S., Wall, A.C.: Quantum focusing conjecture. Phys. Rev. D 93, 064044 (2016)

    ADS  MathSciNet  Google Scholar 

  3. Bekenstein, J.D.: Black holes and entropy. Phys. Rev. D 7, 2333 (1973)

    ADS  MathSciNet  MATH  Google Scholar 

  4. Casini, H.: Relative entropy and the Bekenstein bound. Class. Quantum Grav. 25, 205021 (2008). arXiv:0804.2182

    ADS  MathSciNet  MATH  Google Scholar 

  5. Wall, A.C.: Proof of the generalized second law for rapidly changing fields and arbitrary horizon slices. Phys. Rev. D 85, 104049 (2012)

    ADS  Google Scholar 

  6. Witten, E.: APS Medal for Exceptional Achievement in Research: invited article on entanglement properties of quantum field theory. Rev. Mod. Phys. 90, 045003 (2018). arXiv:1803.04993

    ADS  Google Scholar 

  7. Balakrishnan, S., Faulkner, T., Khandker, Z.U., Wang, H.: A General Proof of the Quantum Null Energy Condition, arXiv:1706.09432

  8. Holzhey, C., Larsen, F., Wilczek, F.: Geometric and renormalized entropy in conformal field theory. Nucl. Phys. B 424, 443 (1994). arXiv:hep-th/9403108

    ADS  MathSciNet  MATH  Google Scholar 

  9. Calabrese, P., Cardy, J.L.: Entanglement entropy and quantum field theory. J. Stat. Mech. 0406, P06002 (2004). arXiv:hep-th/0405152

    MATH  Google Scholar 

  10. Lewkowycz, A., Maldacena, J.: Generalized gravitational entropy. JHEP 08, 090 (2013). arXiv:1304.4926

    ADS  MathSciNet  MATH  Google Scholar 

  11. Hollands, S., Sanders, K.: Entanglement Measures and Their Properties in Quantum Field Theory. arXiv:1702.04924

  12. Longo, R.: Entropy Distribution of Localised States. arXiv:1809.03358

  13. Longo, R., Xu, F.: Relative entropy in CFT. Adv. Math. 337, 139 (2018). arXiv:1712.07283

    MathSciNet  MATH  Google Scholar 

  14. Xu, F.: Some Results on Relative Entropy in Quantum Field Theory. arXiv:1810.10642

  15. Xu, F.: On Relative Entropy and Global Index. arXiv:1812.01119

  16. Kang, M.J., Kolchmeyer, D.K.: Holographic Relative Entropy in Infinite-Dimensional Hilbert Spaces. arXiv:1811.05482

  17. Wall, A.C.: Lower bound on the energy density in classical and quantum field theories. Phys. Rev. Lett. 118, 151601 (2017). arXiv:1701.03196

    ADS  MathSciNet  Google Scholar 

  18. Klinkhammer, G.: Averaged energy conditions for free scalar fields in flat space-times. Phys. Rev. D 43, 2542 (1991)

    ADS  MathSciNet  Google Scholar 

  19. Kelly, W.R., Wall, A.C.: Holographic proof of the averaged null energy condition. Phys. Rev. D 90, 106003 (2014). arXiv:1408.3566

    ADS  Google Scholar 

  20. Faulkner, T., Leigh, R.G., Parrikar, O., Wang, H.: Modular Hamiltonians for deformed half-spaces and the averaged null energy condition. JHEP 09, 038 (2016). arXiv:1605.08072

    ADS  MathSciNet  MATH  Google Scholar 

  21. Hartman, T., Kundu, S., Tajdini, A.: Averaged null energy condition from causality. JHEP 07, 066 (2017). arXiv:1610.05308

    ADS  MathSciNet  MATH  Google Scholar 

  22. Kravchuk, P., Simmons-Duffin, D.: Light-ray operators in conformal field theory. JHEP 11, 102 (2018). arXiv:1805.00098

    ADS  MathSciNet  MATH  Google Scholar 

  23. Borchers, H.-J.: The CPT-theorem in two-dimensional theories of local observables. Commun. Math. Phys. 143, 315 (1992)

    ADS  MathSciNet  MATH  Google Scholar 

  24. Wiesbrock, H.-W.: Half-sided modular inclusions of von-Neumann-algebras. Commun. Math. Phys. 157, 83 (1993)

    ADS  MathSciNet  MATH  Google Scholar 

  25. Borchers, H.: Half-sided modular inclusion and the construction of the poincaré group. Commun. Math. Phys. 179, 703 (1996)

    ADS  MATH  Google Scholar 

  26. Araki, H., Zsidó, L.: Extension of the structure theorem of borchers and its application to half-sided modular inclusions. Rev. Math. Phys. 17, 491 (2005)

    MathSciNet  MATH  Google Scholar 

  27. Maldacena, J., Shenker, S.H., Stanford, D.: A bound on chaos. JHEP 08, 106 (2016). arXiv:1503.01409

    ADS  MathSciNet  MATH  Google Scholar 

  28. Hartman, T., Jain, S., Kundu, S.: Causality constraints in conformal field theory. JHEP 05, 099 (2016). arXiv:1509.00014

    ADS  Google Scholar 

  29. Fawzi, O., Renner, R.: Quantum conditional mutual information and approximate Markov chains. Commun. Math. Phys. 340, 575 (2015)

    ADS  MathSciNet  MATH  Google Scholar 

  30. Wilde, M.M.: Recoverability in quantum information theory. Proc. R. Soc. A 471, 20150338 (2015)

    ADS  MathSciNet  MATH  Google Scholar 

  31. Junge, M., Renner, R., Sutter, D., Wilde, M.M., Winter, A.: Universal recovery maps and approximate sufficiency of quantum relative entropy. In: Annales Henri Poincaré, vol. 19, pp. 2955–2978, Springer, Berlin (2018)

  32. Swingle, B., Wang, Y.: Recovery Map for Fermionic Gaussian Channels. arXiv preprint arXiv:1811.04956 (2018)

  33. Haag, R.: Local Quantum Physics- Fields Particles, Algebras. Springer, Berlin (1996)

    MATH  Google Scholar 

  34. Witten, E.: Black Holes, Singularity Theorems, and All That. https://static.ias.edu/pitp/2018/sites/pitp/files/gr_lectures_edited.pdf (2018)

  35. Reeh, H., Schlieder, S.: Bemerkungen zur unitäräquivalenz von lorentzinvarianten feldern. Il Nuovo Cimento (1955–1965) 22, 1051 (1961)

    ADS  MATH  Google Scholar 

  36. Bisognano, J.J., Wichmann, E.H.: On the duality condition for quantum fields. J. Math. Phys. 17, 303 (1976)

    ADS  MathSciNet  Google Scholar 

  37. Koeller, J., Leichenauer, S., Levine, A., Shahbazi-Moghaddam, A.: Local modular Hamiltonians from the quantum null energy condition. Physi. Rev. D 97, 065011 (2018)

    ADS  MathSciNet  Google Scholar 

  38. Casini, H., Teste, E., Torroba, G.: Modular Hamiltonians on the null plane and the Markov property of the vacuum state. J. Phys. A 50, 364001 (2017). arXiv:1703.10656

    MathSciNet  MATH  Google Scholar 

  39. Borchers, H.-J.: On the use of modular groups in quantum field theory. In: Annales de l’Institut Henri Poincare-A Physique Theorique, vol. 63, pp. 331–382, Paris: Gauthier-Villars, c1983–c1999 (1995)

  40. Buchholz, D., D’Antoni, C., Longo, R.: Nuclear maps and modular structures. I. General properties. J. Funct. Anal. 88, 233 (1990)

    MathSciNet  MATH  Google Scholar 

  41. Jefferson, R.: Comments on Black Hole Interiors and Modular Inclusions. arXiv:1811.08900

  42. Araki, H.: Relative entropy of states of von Neumann algebras. Publ. Res. Inst. Math. Sci. 11, 809 (1976)

    MathSciNet  MATH  Google Scholar 

  43. Araki, H.: Relative entropy for states of von Neumann algebras II. Publ. Res. Inst. Math. Sci. 13, 173 (1977)

    MathSciNet  MATH  Google Scholar 

  44. Araki, H., Masuda, T.: Positive cones and lp-spaces for von Neumann algebras. Publ. Res. Inst. Math. Sci. 18, 759 (1982)

    MathSciNet  MATH  Google Scholar 

  45. Uhlmann, A.: Relative entropy and the Wigner–Yanase–Dyson–Lieb concavity in an interpolation theory. Commun. Math. Phys. 54, 21 (1977)

    ADS  MathSciNet  MATH  Google Scholar 

  46. Ohya, M., Petz, D.: Quantum Entropy and Its Use. Springer, Berlin (2004)

    MATH  Google Scholar 

  47. Lieb, E.H., Ruskai, M.B.: Proof of the strong subadditivity of quantum-mechanical entropy. J. Math. Phys. 14, 1938 (1973)

    ADS  MathSciNet  Google Scholar 

  48. Bousso, R., Casini, H., Fisher, Z., Maldacena, J.: Entropy on a null surface for interacting quantum field theories and the bousso bound. Phys. Rev. D 91, 084030 (2015)

    ADS  MathSciNet  Google Scholar 

  49. Blanco, D.D., Casini, H.: Localization of negative energy and the bekenstein bound. Phys. Rev. Lett. 111, 221601 (2013)

    ADS  Google Scholar 

  50. Jafferis, D.L., Suh, S.J.: The gravity duals of modular hamiltonians. JHEP 09, 068 (2016). arXiv:1412.8465

    ADS  MathSciNet  MATH  Google Scholar 

  51. Faulkner, T., Li, M., Wang, H.: A Modular Toolkit for Bulk Reconstruction. arXiv:1806.10560

  52. Shenker, S.H., Stanford, D.: Black holes and the butterfly effect. JHEP 03, 067 (2014). arXiv:1306.0622

    ADS  MathSciNet  MATH  Google Scholar 

  53. Caron-Huot, S.: Analyticity in spin in conformal theories. JHEP 09, 078 (2017). arXiv:1703.00278

    ADS  MathSciNet  MATH  Google Scholar 

  54. Araki, H.: Some properties of modular conjugation operator of von neumann algebras and a non-commutative radon-nikodym theorem with a chain rule. Pac. J. Math. 50, 309 (1974)

    MathSciNet  MATH  Google Scholar 

  55. Araki, H.: Relative hamiltonian for faithful normal states of a von Neumann algebra. Publ. Res. Inst. Math. Sci. 9, 165 (1973)

    MathSciNet  MATH  Google Scholar 

  56. Witten, E.: Notes on Some Entanglement Properties of Quantum Field Theory. arXiv:1803.04993

  57. Schiff, J.L.: Normal Families. Springer, Berlin (2013)

    MATH  Google Scholar 

  58. Lashkari, N.: Constraining Quantum Fields using Modular Theory. arXiv:1810.09306

  59. Petz, D.: Sufficient subalgebras and the relative entropy of states of a von neumann algebra. Commun. Math. Phys. 105, 123 (1986)

    ADS  MathSciNet  MATH  Google Scholar 

  60. Petz, D.: Sufficiency of channels over von neumann algebras. Q. J. Math. 39, 97 (1988)

    MathSciNet  MATH  Google Scholar 

  61. Accardi, L., Cecchini, C.: Conditional expectations in von neumann algebras and a theorem of takesaki. J. Funct. Anal. 45, 245 (1982)

    MathSciNet  MATH  Google Scholar 

  62. Uhlmann, A.: The “transition probability” in the state space of \(\text{ a }^*\)-algebra. Rep. Math. Phys. 9, 273 (1976)

    ADS  MathSciNet  MATH  Google Scholar 

  63. Kosaki, H.: Applications of the complex interpolation method to a von neumann algebra: non-commutative lp-spaces. J. Funct. Anal. 56, 29 (1984)

    MathSciNet  MATH  Google Scholar 

  64. Haagerup, U.: Lp-spaces associated with an arbitrary von Neumann algebra. In: Algebres d’opérateurs et leurs applications en physique mathématique (Proc. Colloq., Marseille, 1977), vol. 274, pp. 175–184 (1979)

  65. Leichenauer, S., Levine, A., Shahbazi-Moghaddam, A.: Energy density from second shape variations of the von Neumann entropy. Phys. Rev. D 98, 086013 (2018). arXiv:1802.02584

    ADS  MathSciNet  Google Scholar 

  66. Balakrishnan, S., Chandrasekaran, V., Faulkner, T., Levine, A., Shahbazi-Moghaddam, A.: Entropy Variations and Light Ray Operators from Replica Defects. arXiv:1906.08274

  67. Hofman, D.M., Maldacena, J.: Conformal collider physics: energy and charge correlations. JHEP 05, 012 (2008). arXiv:0803.1467

    ADS  Google Scholar 

  68. Callebaut, N., Verlinde, H.: Entanglement Dynamics in 2D CFT with Boundary: Entropic origin of JT gravity and Schwarzian QM. arXiv:1808.05583

  69. Cordova, C., Shao, S.-H.: Light-Ray Operators and the BMS Algebra. arXiv:1810.05706

  70. Engelhardt, N., Wall, A.C.: Coarse Graining Holographic Black Holes. arXiv:1806.01281

  71. Engelhardt, N., Wall, A.C.: Decoding the apparent horizon: a coarse-grained holographic entropy. Phys. Rev. Lett. 121, 211301 (2018). arXiv:1706.02038

    ADS  Google Scholar 

  72. Neuenfeld, D., Saraswat, K., Van Raamsdonk, M.: Positive gravitational subsystem energies from CFT cone relative entropies. JHEP 06, 050 (2018). arXiv:1802.01585

    ADS  MathSciNet  MATH  Google Scholar 

  73. Casini, H., Teste, E., Torroba, G.: All the entropies on the light-cone. JHEP 05, 005 (2018). arXiv:1802.04278

    ADS  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We especially thank Raphael Bousso, Ven Chandrasekaran, Netta Engelhardt, Ben Freivogel, Marius Junge, Nima Lashkari, Juan Maldacena, Arvin Shahbazi-Moghaddam for discussions related to this work. This work was supported by the DOE: Award Number DE-SC0019517.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Thomas Faulkner.

Additional information

Communicated by Y. Kawahigashi.

Publisher's Note

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

Relative Modular Operator

Relative Modular Operator

In this appendix we collect various formula related to the relative modular operator. We are particularly interested in defining these objects when the vector states are not necessarily cyclic and separating. These considerations are standard and can be found in the Appendix of [44]. We warn the reader that we have a different convention for labeling our Sand\(\varDelta \)relative modular operators - the state labels are switched. This convention was used in [56] and we stick with this.

Take \(\psi \) to not be cyclic and separating. This means that there could be some \(\alpha \in {\mathcal {A}}\) such that \(\alpha | \psi \rangle =0\) (not separating) and also that \({\mathcal {A}} | \psi \rangle \) may generate a proper subspace of \({\mathcal {H}}\) instead of the full Hilbert space (not cyclic). To describe this situation we define support projections as the minimal projectors that satisfy:

$$\begin{aligned} s^{\mathcal {A}}(\psi ) | \psi \rangle =&| \psi \rangle \qquad s^{\mathcal {A}}(\psi ) \in {\mathcal {A}} \end{aligned}$$
(276)
$$\begin{aligned} s^\mathcal {A'}(\psi ) | \psi \rangle =&| \psi \rangle \qquad s^\mathcal {A'}(\psi ) \in {\mathcal {A}}'. \end{aligned}$$
(277)

An equivalent definition follows from finding the projector onto the following subspaces:

$$\begin{aligned} \left[ {\mathcal {A}}' | \psi \rangle \right]&= \pi (\psi ){\mathcal {H}} \subset {\mathcal {H}} \end{aligned}$$
(278)
$$\begin{aligned} \left[ {\mathcal {A}} | \psi \rangle \right]&= \pi '(\psi ){\mathcal {H}} \subset {\mathcal {H}}. \end{aligned}$$
(279)

These are seen to be equivalent as follows. Firstly the \(\pi (\psi )\) commutes with \({\mathcal {A}}'\) since for arbitrary state \(|\phi _i\rangle \in {\mathcal {H}}\) and for all \(\alpha ' \in {\mathcal {A}}'\):

$$\begin{aligned} \langle \phi _1 | \left[ \pi _{\mathcal {A}}(\psi ), \alpha ' \right] | \phi _2 \rangle&= \langle \chi _1 | \alpha ' | \phi _2 \rangle - \langle \phi _1 | \alpha ' | \chi _2 \rangle \qquad&\left( | \chi _i \rangle = \pi (\psi ) | \phi _i \rangle \right) \end{aligned}$$
(280)
$$\begin{aligned}&= \langle \chi _1 | \alpha ' | \chi _2 \rangle - \langle \chi _1 | \alpha ' | \chi _2 \rangle = 0 \qquad&\left( \alpha ' | \chi _i \rangle = \pi (\psi ) \alpha ' | \chi _i \rangle \right) \end{aligned}$$
(281)

so \(\pi \) is in \({\mathcal {A}}\). Secondly it is the minimal such projector leaving \(\psi \) invariant since if it were not there would be another projector \(\pi _2 \in {\mathcal {A}}\) with \(\pi _2 {\mathcal {H}} \subset \pi (\psi ) {\mathcal {H}}\) which also leaves invariant the subspace:

$$\begin{aligned} \left[ A' | \psi \rangle \right] = \left[ A' \pi _2 | \psi \rangle \right] = \pi _2 \left[ A' | \psi \rangle \right] \end{aligned}$$
(282)

such that \(\pi _2 {\mathcal {H}} \subset \pi (\psi ) {\mathcal {H}} = \pi _2 \pi (\psi ) {\mathcal {H}} \subset \pi _2 {\mathcal {H}}\) implying that \(\pi _2 = \pi (\psi )\). Thus \(\pi = s^{\mathcal {A}}\) and similarly for the commutant.

Note that if \(\pi (\psi )\) is not the unit operator, then \(\psi \) is not cyclic for \({\mathcal {A}}'\) and \((1-\pi (\psi ))\) annihilates \(\psi \) which means that \(\psi \) is not separating for \({\mathcal {A}}\). That is the lack of either of these two properties exchange under \({\mathcal {A}} \leftrightarrow {\mathcal {A}}'\).

We now move to the modular operators. We will consider two state \(\psi ,\phi \) neither of which needs to be cyclic and separating. We start with the definition of the Tomita operators:

$$\begin{aligned} S_{\psi |\phi }\left( \alpha | \psi \rangle + | \chi ' \rangle \right)&= \pi (\psi ) \alpha ^\dagger | \phi \rangle \qquad \forall \,\, \chi ' \in (1-\pi '(\psi )) {\mathcal {H}} \end{aligned}$$
(283)
$$\begin{aligned} S_{\phi |\psi } \left( \alpha | \phi \rangle + |\xi '\rangle \right)&= \pi (\phi ) \alpha ^\dagger | \psi \rangle \qquad \forall \, \, \xi ' \in (1-\pi '(\phi )) {\mathcal {H}} \end{aligned}$$
(284)

for \(\alpha \in {\mathcal {A}}\). Note that, for the first equation above, if both \(\alpha | \psi \rangle =0\) and \(| \chi ' \rangle =0\) vanish then \(\alpha [ {\mathcal {A}}' | \psi \rangle ] = 0 \implies \pi (\psi ) \alpha ^\dagger = 0\) so 0 is mapped to 0 as is necessary for a linear operator. The Tomita operators are closable as defined and we will use the same symbol for the closure as the original operator. The support of these operators is:

$$\begin{aligned} \mathrm{supp} ( S_{\psi |\phi }, \, S_{\phi |\psi }^\dagger ) = \pi '(\psi ) \pi (\phi ) {\mathcal {H}} \qquad \mathrm{supp} ( S_{\phi |\psi }, \, S_{\psi |\phi }^\dagger ) = \pi (\psi ) \pi '(\phi ) {\mathcal {H}}. \end{aligned}$$
(285)

Applying the definitions twice we have:

$$\begin{aligned} S_{\psi |\phi } S_{\phi |\psi }&= \pi (\psi ) \pi '(\phi ) \end{aligned}$$
(286)
$$\begin{aligned} S_{\phi |\psi } S_{\psi |\phi }&= \pi '(\psi ) \pi (\phi ). \end{aligned}$$
(287)

For the commutant algebra we have:

$$\begin{aligned} S_{\psi |\phi }'\left( \alpha ' | \psi \rangle + | \chi \rangle \right)&= \pi '(\psi ) (\alpha ')^\dagger | \phi \rangle \qquad \forall \, \, \chi \in (1-\pi (\psi )) {\mathcal {H}} \end{aligned}$$
(288)
$$\begin{aligned} S_{\phi |\psi }' ( \alpha ' | \phi \rangle + | \xi \rangle )&= \pi '(\phi ) (\alpha ')^\dagger | \psi \rangle \qquad \forall \, \, \chi \in (1-\pi (\phi )) {\mathcal {H}} \end{aligned}$$
(289)

for \(\alpha ' \in {\mathcal {A}}'\) with support that is complementary to (285). And similar equations hold for the commutant as in (286). Now consider:

$$\begin{aligned}&\left( \beta ' | \psi \rangle + | \chi \rangle , S_{\psi |\phi }\left( \alpha | \psi \rangle + | \chi ' \rangle \right) \right) = \left( \beta ' | \psi \rangle + | \chi \rangle , \pi (\psi ) \alpha ^\dagger | \phi \rangle \right) \end{aligned}$$
(290)
$$\begin{aligned}&\quad = \left( \beta ' | \psi \rangle , \alpha ^\dagger | \phi \rangle \right) = \left( \alpha | \psi \rangle , (\beta ')^\dagger | \phi \rangle \right) = \left( \alpha | \psi \rangle + | \chi '\rangle , \pi '(\psi ) (\beta ')^\dagger | \phi \rangle \right) \end{aligned}$$
(291)
$$\begin{aligned}&\quad = \left( \alpha | \psi \rangle + | \chi '\rangle , S_{\psi |\phi }'\left( \beta ' | \psi \rangle + | \chi \rangle \right) \right) \end{aligned}$$
(292)

which means that (because the above states are dense on the appropriate support)

$$\begin{aligned} S_{\psi |\phi }' = S_{\psi |\phi }^\dagger \,, \qquad S_{\phi |\psi }' = S_{\phi |\psi }^\dagger \end{aligned}$$
(293)

where the later equation follows a similar analysis.

We move now to the relative modular operators. Consider the positive self adjoint operators:

$$\begin{aligned} \varDelta _{\psi |\phi } = S_{\psi |\phi }^\dagger S_{\psi |\phi } \qquad \varDelta _{\phi |\psi } = S_{\phi |\psi }^\dagger S_{\phi |\psi } \end{aligned}$$
(294)

with support

$$\begin{aligned} \mathrm{supp} ( \varDelta _{\psi |\phi } ) = \pi '(\psi ) \pi (\phi ) {\mathcal {H}} \qquad \mathrm{supp} ( \varDelta _{\phi |\psi } ) = \pi (\psi ) \pi '(\phi ) {\mathcal {H}}. \end{aligned}$$
(295)

For the commutant algebra we learn that:

$$\begin{aligned} \varDelta _{\psi |\phi }' =&(S_{\psi |\phi }')^\dagger S_{\psi |\phi }' = S_{\psi |\phi } S_{\psi |\phi }^\dagger \quad \implies \quad \varDelta _{\psi |\phi }' \varDelta _{\phi |\psi } = \pi (\psi ) \pi '(\phi ) \end{aligned}$$
(296)
$$\begin{aligned} \varDelta _{\phi |\psi }' =&S_{\phi |\psi } S_{\phi |\psi }^\dagger \quad \implies \quad \varDelta _{\phi |\psi }' \varDelta _{\psi |\phi } = \pi '(\psi ) \pi (\phi ). \end{aligned}$$
(297)

We can define powers of the modular operators \(\varDelta _{\psi |\phi }^z\) etc. to be zero when acting on \((1-\pi '(\psi ) \pi (\phi )) {\mathcal {H}}\) and to be the usual power when acting on the support of \(\varDelta _{\psi |\phi }\) which for example means that \(\varDelta _{\psi |\phi }^0 =\pi '(\psi ) \pi (\phi ) \). So for example we have:

$$\begin{aligned} (\varDelta _{\psi |\phi }')^{z} \varDelta _{\phi |\psi }^{-z} = \pi (\psi ) \pi '(\phi ). \end{aligned}$$
(298)

Furthermore we can apply polar decompositions to the Tomita operators, where the anti-linear part is not unitary, but rather a partial anti-linear isometry with the support and range of the Tomita operators:

$$\begin{aligned} S_{\psi |\phi } = J_{\psi |\phi } \varDelta _{\psi |\phi }^{1/2} \qquad \qquad \mathrm{etc} \end{aligned}$$
(299)

where:

$$\begin{aligned} J_{\psi |\phi }^\dagger J_{\psi |\phi }&= \pi '(\psi ) \pi (\phi ) \qquad J_{\psi |\phi } J_{\psi |\phi }^\dagger = \pi (\psi ) \pi '(\phi ) \end{aligned}$$
(300)
$$\begin{aligned} J_{\phi |\psi }^\dagger J_{\phi |\psi }&= \pi (\psi )\pi '(\phi ) \qquad J_{\phi |\psi } J_{\phi |\psi }^\dagger = \pi '(\psi ) \pi (\phi ) \qquad \qquad \mathrm{etc} \end{aligned}$$
(301)

(with appropriate support and range.) Plugging back into the Tomita operators and (294) we have:

$$\begin{aligned} J_{\psi |\phi } \varDelta ^{1/2}_{\psi |\phi } J_{\phi |\psi } \varDelta ^{1/2}_{\phi |\psi } = \pi (\psi ) \pi '(\phi ) \quad \implies \quad J_{\psi |\phi } J_{\phi |\psi } ( J_{\phi |\psi }^\dagger \varDelta _{\psi |\phi }^{1/2} J_{\phi |\psi } ) = \varDelta _{\phi |\psi }^{-1/2} \end{aligned}$$
(302)

which by the uniqueness of the polar decomposition implies that:

$$\begin{aligned} J_{\psi |\phi } J_{\phi |\psi } = \pi (\psi ) \pi '(\phi ) \qquad J_{\psi |\phi } = J_{\phi |\psi }^\dagger \qquad J_{\psi |\phi } \varDelta _{\psi |\phi }^{1/2} J_{\phi |\psi } = \varDelta _{\phi |\psi }^{-1/2} \end{aligned}$$
(303)

where the last equation implies that:

$$\begin{aligned} J_{\psi |\phi } \varDelta _{\psi |\phi }^{is} = \varDelta _{\phi |\psi }^{is} J_{\psi |\phi } \end{aligned}$$
(304)

by the anti-linearity of J.

For the complement we use (293) to derive:

$$\begin{aligned} J_{\psi |\phi }' = J_{\phi |\psi } \qquad (\varDelta _{\psi |\phi }')^z = \varDelta _{\phi |\psi }^{-z}. \end{aligned}$$
(305)

In order to understand relative modular flow and cocycles we have to apply the Connes \(2\times 2\) or \(3\times 3\) trick which we present here in a vector language. Consider the Hilbert space:

$$\begin{aligned} {\mathcal {H}}_{\mathrm{tot}} = {\mathcal {H}}_L \otimes {\mathcal {H}}_R \otimes {\mathcal {H}}_{QFT} \end{aligned}$$
(306)

where \({\mathcal {H}}_{L,R}\) are both simple n-dimensional qunit Hilbert spaces with basis \(|i \rangle ; i = 0,1, \ldots , n-1\). In this Hilbert space we consider the state:

$$\begin{aligned} | \varPsi \rangle = \sum _{i=1}^{n} \frac{1}{\sqrt{n}} | i_L i_R \rangle \otimes | \phi _i \rangle \end{aligned}$$
(307)

where \(| \phi _i \rangle \) are vector states in the QFT Hilbert space that need not be cyclic and separating. We will actually consider the state as living in the subspace projected by the support projectors of the states \(\phi _i\):

$$\begin{aligned} \widetilde{{\mathcal {H}}} = \left[ | i_L j_R \rangle \otimes \pi '(\phi _i) \pi (\phi _j) {\mathcal {H}}_{QFT}: i,j = 0,1, \ldots n-1 \right] \subset {\mathcal {H}}_{\mathrm{tot}} \end{aligned}$$
(308)

where the square brackets means the linear span. We now consider the algebra of operators acting on this new Hilbert space:

$$\begin{aligned} \gamma \in \mathbb {A}\,: \qquad \gamma = \sum _{ij} 1_L \otimes \left( | i \rangle \langle j | \right) _R \otimes c_{ij} \qquad c_{ij} \in \pi (\phi _i) {\mathcal {A}}\, \pi (\phi _j). \end{aligned}$$
(309)

The commutant is:

$$\begin{aligned} \gamma ' \in \mathbb {A}'\,: \qquad \gamma ' = \sum _{ij} \left( | i \rangle \langle j | \right) _L \otimes 1_R \otimes c'_{ij} \qquad c_{ij}' \in \pi '(\phi _i) {\mathcal {A}}'\, \pi '(\phi _j) \end{aligned}$$
(310)

and \(\varPhi \) is cyclic and separating for these algebras after we project to the subspace \(\tilde{{\mathcal {H}}}\). The generalized Tomita operator is:

$$\begin{aligned} \mathbb {S} \gamma | \varPsi \rangle = \gamma ^\dagger | \varPsi \rangle \end{aligned}$$
(311)

from which one finds:

$$\begin{aligned} \mathbb {S} = \sum _{ij} | j_L i_R \rangle \langle i_L j_R | \otimes {\tilde{S}}_{i|j} \,, \qquad {\tilde{S}}_{i|j} c_{ji} | \phi _i \rangle = c_{ji}^\dagger | \phi _j \rangle . \end{aligned}$$
(312)

Note that this later Tomita operator acts between Hilbert spaces:

$$\begin{aligned} {\tilde{S}}_{i|j}\, : \pi '(\phi _i) \pi (\phi _j) {\mathcal {H}}_{QFT} \rightarrow \pi '(\phi _j) \pi (\phi _i) {\mathcal {H}}_{QFT}. \end{aligned}$$
(313)

We can relate this to our original definition of the Tomita operators by passing back to the original unprojected Hilbert spaces setting \(c_{ji} = \pi (\phi _j) \alpha \pi (\phi _i)\):

$$\begin{aligned} {\tilde{S}}_{i|j} \pi (\phi _j) \alpha | \phi _i \rangle = \pi (\phi _i) \alpha ^\dagger | \phi _j \rangle . \end{aligned}$$
(314)

We can lose the \(\pi (\phi _j)\) on the left hand side since the orthogonal part is killed on the right hand side anyway. Or in other words we can extend the definition \({\tilde{S}}_{i|j}\) to the larger Hilbert space consistent with the right hand side by dropping this projector and also demanding:

$$\begin{aligned}&S_{i|j} \alpha | \phi _i \rangle = \pi (\phi _i) \alpha ^\dagger | \phi _j \rangle \end{aligned}$$
(315)
$$\begin{aligned}&S_{i|j} (1- \pi '(\phi _i) ) {\mathcal {H}}_{QFT} = 0 \end{aligned}$$
(316)

which is the same definition we gave for \(i=\psi \) and \(j=\phi \) in (283).

The modular operator for the \(n \times n\) Hilbert space is:

$$\begin{aligned} \varvec{\Delta } = \mathbb {S}^\dagger \mathbb {S} = \sum _{ij} | i_L j_R \rangle \langle i_L j_R | \otimes {\tilde{\varDelta }}_{i|j}\,, \qquad {\tilde{\varDelta }}_{i|j} = {\tilde{S}}_{i|j}^\dagger {\tilde{S}}_{i|j} \end{aligned}$$
(317)

and the modular conjugation operator is:

$$\begin{aligned} \mathbb {J} = \sum _{ij} | j_L i_R \rangle \langle i_L j_R | \otimes {\tilde{J}}_{i|j} \qquad {\tilde{S}}_{i|j} = {\tilde{J}}_{i|j} {\tilde{\varDelta }}_{i|j}^{1/2} \end{aligned}$$
(318)

where \(\mathbb {J}^2 = 1,\, \mathbb { J} \varvec{\Delta } \mathbb {J} = \varvec{\Delta }^{-1}\). The projected modular operators act between the following Hilbert spaces:

$$\begin{aligned} {\tilde{\varDelta }}_{i|j}\, : \pi '(\phi _i) \pi (\phi _j) {\mathcal {H}}_{QFT}&\rightarrow \pi '(\phi _i) \pi (\phi _j) {\mathcal {H}}_{QFT} \end{aligned}$$
(319)
$$\begin{aligned} {\tilde{J}}_{i|j}\, : \pi '(\phi _i) \pi (\phi _j) {\mathcal {H}}_{QFT}&\rightarrow \pi '(\phi _j) \pi (\phi _i) {\mathcal {H}}_{QFT} \end{aligned}$$
(320)

and can be extended to \({\mathcal {H}}_{QFT}\) as we did with the Tomita operators.

We now apply the results of Tomita-Takesaki theory to these new modular operators. That is we know that \(\mathbb {A} = \varvec{\Delta }^{is} \mathbb {A} \varvec{\Delta }^{-is}\) and \(\mathbb {A}' = \mathbb {J} \mathbb {A} \mathbb {J}\). Computing:

$$\begin{aligned} \varvec{\Delta }^{is} \gamma \varvec{\Delta }^{-is}&= \sum _k \sum _{ij} ( | k \rangle \langle k | )_L \otimes ( | i \rangle \langle j | )_R \otimes {\tilde{\varDelta }}_{k|i}^{is} c_{ij} {\tilde{\varDelta }}_{k|j}^{-is} \end{aligned}$$
(321)
$$\begin{aligned} \mathbb {J} \gamma \mathbb {J}&= \sum _k \sum _{ij} ( | i \rangle \langle j |)_L \otimes (| k \rangle \langle k |)_R \otimes {\tilde{J}}_{k | i } c_{ij} {\tilde{J}}_{j | k}. \end{aligned}$$
(322)

This is only consistent with the form of the algebra given in (309) if we have:

$$\begin{aligned} {\tilde{\varDelta }}_{k|i}^{is} c_{ij} {\tilde{\varDelta }}_{k|j}^{-is} \, \in \pi (\phi _i) {\mathcal {A}} \pi (\phi _j)\,, \qquad c_{ij} \in \pi (\phi _i) {\mathcal {A}} \pi (\phi _j) \end{aligned}$$
(323)

and the flowed operator is the same operator for all k. Extending these statements to the full Hilbert space by setting the flowed and conjugated operators to zero away from the support we find that:

$$\begin{aligned} \varDelta _{k|i}^{is} \alpha \varDelta _{k|j}^{-is} \, \in {\mathcal {A}} \pi '(\phi _k)\,, \qquad \alpha \in {\mathcal {A}} \end{aligned}$$
(324)

where we are forced to add \(\pi '(\phi _k)\) so that it vanishes away from this support. For example this is consistent with \(s \rightarrow 0\) where we find \(\pi (\phi _i) \alpha \pi (\phi _j) \pi '(\phi _k)\). Note that we can pick \(\alpha \) in \({\mathcal {A}}\) rather than the projected algebra since the modular operators above anyway apply this projection. Note the flowed operator is now marginally not independent of k due to \(\pi '(\phi _k)\).

If we set \(i=j\) this defines the standard modular automorphism group but now for a non-cyclic and separating vector:

$$\begin{aligned} \varDelta _{i}^{is} \alpha \varDelta _i^{-is} = \sigma _s^{\phi _i}(\alpha ) \pi '(\phi _i) \,, \qquad \sigma _s^{\phi _i}(\alpha ) = \varDelta _{\varOmega |i}^{is} \alpha \varDelta _{\varOmega |i}^{-is} \end{aligned}$$
(325)

where in the later equation we have used a cyclic and separating vector \(\varOmega \) to define this flow.

For \(\alpha = 1\) we can define the operator in \({\mathcal {A}}\) that is independent of k as the cocycle:

$$\begin{aligned} \varDelta _{k|i}^{is} \varDelta _{k|j}^{-is} \equiv (D \phi _i : D\phi _j )_s \pi '(\phi _k) \,, \qquad (D \phi _i : D\phi _j )_s \in {\mathcal {A}} \end{aligned}$$
(326)

which can be extracted by picking \(\phi _k\) to be a cyclic and separating vector. The cocycle satisfies:

$$\begin{aligned} (D \phi _i : D\phi _j)^\dagger _s&= (D \phi _j : D\phi _i)_s \end{aligned}$$
(327)
$$\begin{aligned} (D \phi _i : D\phi _j)^\dagger _s (D \phi _i : D\phi _j)_s&= \sigma _s^{\phi _i}( \pi (\phi _j)) \end{aligned}$$
(328)
$$\begin{aligned} (D \phi _i : D\phi _j)_s (D \phi _i : D\phi _j)^\dagger _s&= \sigma _s^{\phi _j}(\pi (\phi _i)) \end{aligned}$$
(329)

where the right hand side of the later two equations are projection operators if \([ \pi (\phi _i), \pi (\phi _j)] =0\) which means that for states with commuting support projectors the cocycle is a partial isometry.

There is the following relation on triples of the cocycle:

$$\begin{aligned} (D \phi _1 : D\phi _2)_s (D\phi _2:D\phi _3)_s = (D\phi _1:D\phi _3)_s \qquad \mathrm{if} \,\,\,\, \begin{matrix} \pi (\phi _1)\pi (\phi _2) = \pi (\phi _1) \,\, \mathrm{or} \\ \pi (\phi _2)\pi (\phi _3) = \pi (\phi _3) \end{matrix} \end{aligned}$$
(330)

where the later conditions on the projectors can be guaranteed by demanding:

$$\begin{aligned} \pi (\phi _1){\mathcal {H}} \subset \pi (\phi _2){\mathcal {H}} \,\, \mathrm{or} \,\, \pi (\phi _3){\mathcal {H}} \subset \pi (\phi _2){\mathcal {H}}. \end{aligned}$$
(331)

Similary for the modular conjugation operators we have:

$$\begin{aligned} {\tilde{J}}_{k|i} c_{ij} {\tilde{J}}_{j|k} \, \in \pi '(\phi _i) {\mathcal {A}}' \pi '(\phi _j)\,, \qquad c_{ij} \in \pi (\phi _i) {\mathcal {A}} \pi (\phi _j) \end{aligned}$$
(332)

which extends to the \({\mathcal {H}}_{QFT}\) in the usual way with:

$$\begin{aligned} J_{k|i} \alpha J_{j|k} \in {\mathcal {A}}' \pi (\phi _k)\,, \qquad \alpha \in {\mathcal {A}} \end{aligned}$$
(333)

and where apart from the projector \(\pi (\phi _k)\) this is the same operator independent of k. We define the non relative modular conjugation action as:

$$\begin{aligned} J_i \alpha J_i = j^{\phi _i}(\alpha ) \pi (\phi _i) \,,\qquad j^{\phi _i}(\alpha ) = J_{\varOmega |i} \alpha J_{i|\varOmega }. \end{aligned}$$
(334)

The equivalent of the cocycles are the following linear operators:

$$\begin{aligned} J_{k|i} J_{j|k} = \varTheta _{i|j}' \pi (\phi _k)\,, \qquad \varTheta _{i|j}' \in {\mathcal {A}}' \end{aligned}$$
(335)

which satisfies:

$$\begin{aligned} (\varTheta '_{i|j})^\dagger = (\varTheta '_{j|i}) \,, \quad (\varTheta '_{i|j})^\dagger (\varTheta '_{i|j}) = j^{\phi _j}(\pi (\phi _i)) \,, \quad (\varTheta '_{i|j}) (\varTheta '_{i|j})^\dagger = j^{\phi _i}(\pi (\phi _j)) \end{aligned}$$
(336)

where this is then a partial isometry if \([\pi (\phi _i),\pi (\phi _j)] =0\).

For example if we specify that \(\phi _i = \varOmega \), cyclic and separating, and \(\phi _j = \psi \) we find that:

$$\begin{aligned} J_\varOmega J_{\psi |\varOmega } = \varTheta '_{\varOmega |\psi }\,, \qquad (\varTheta '_{\varOmega |\psi })^\dagger (\varTheta '_{\varOmega |\psi }) = \pi '(\psi )\,, \qquad (\varTheta '_{\varOmega |\psi }) (\varTheta '_{\varOmega |\psi })^\dagger = J_{\varOmega } \pi (\psi ) J_\varOmega \end{aligned}$$
(337)

where the support of this operator is: \(\mathrm{supp}(\varTheta _{\varOmega |\psi }') = \pi '(\psi ) {\mathcal {H}}\) and \(\mathrm{supp}(\varTheta _{\varOmega |\psi }')^\dagger =J_\varOmega \pi (\psi ) J_{\varOmega } {\mathcal {H}}\). We also have the useful relations:

$$\begin{aligned} J_{\psi |\varOmega } = J_\varOmega \varTheta '_{\varOmega |\psi } \qquad J_{\varOmega |\psi } = ( \varTheta '_{\varOmega |\psi })^\dagger J_\varOmega \qquad J_\psi = (\varTheta '_{\varOmega |\psi })^\dagger J_\varOmega \varTheta '_{\varOmega |\psi }. \end{aligned}$$
(338)

There is also a triple relation:

$$\begin{aligned} \varTheta '_{1|2} \varTheta '_{2|3} = \varTheta '_{1|3} \qquad \mathrm{if} \,\,\,\, \begin{matrix} \pi (\phi _1)\pi (\phi _2) = \pi (\phi _1) \,\, \mathrm{or} \\ \pi (\phi _2)\pi (\phi _3) = \pi (\phi _3) \end{matrix} \end{aligned}$$
(339)

where again this later condition can be achieved only under the conditions specified for the support projectors.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ceyhan, F., Faulkner, T. Recovering the QNEC from the ANEC. Commun. Math. Phys. 377, 999–1045 (2020). https://doi.org/10.1007/s00220-020-03751-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-020-03751-y

Navigation