Skip to main content
Log in

Quantum Entanglement: An Analysis via the Orthogonality Relation

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

In the literature there has been evidence that a kind of relational structure called a quantum Kripke frame captures the essential characteristics of the orthogonality relation between pure states of quantum systems, and thus is a good qualitative mathematical model of quantum systems. This paper adds another piece of evidence by providing a tensor-product construction of two finite-dimensional quantum Kripke frames. We prove that this construction is exactly the qualitative counterpart of the tensor-product construction of two finite-dimensional Hilbert spaces over the complex numbers, and thus show that composition of quantum systems, especially the phenomenon of quantum entanglement, can be modelled in the framework of quantum Kripke frames. The assumptions used in our construction hint that we need complex numbers in quantum theory. Moreover, for this construction, we give a new and interesting characterization of linear maps of trace 0 in terms of the orthogonality relation.

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.

Similar content being viewed by others

Data Availability

Data sharing not applicable to this article as no datasets were generated or analysed during the current study.

Notes

  1. Here \({\mathcal {L}} ({\mathcal {H}})\) denotes the lattice formed by the closed linear subspaces of a Hilbert space \({\mathcal {H}}\).

  2. In this paper, we assume that every vector space V, including a Hilbert space, is of dimension at least 1, i.e. \(V \ne \{ {\textbf{0}} \}\).

  3. For simplicity, we henceforth omit all ‘\(\cdot \)’ for the scalar multiplication in a vector space.

  4. In the following, for simplicity, we omit all ‘\(\cdot \)’ for the multiplication in a ring.

  5. An Hermitian form is defined to be conjugate in the second argument, while an inner product is usually defined to be conjugate in the first argument. This difference is just a matter of notation.

  6. Here \(G[\Sigma _B]\) denotes the image of \(\Sigma _B\) under the partial function \(G: \Sigma _B \dashrightarrow \Sigma _A\).

  7. Please remember that the following condition (Ad)\(_F\) is not a property of a Kripke frame or a partial function; instead it is a property of an (ordered) pair of two elements in two Kripke frames, in this case, \((t_A,t_B)\).

  8. In the following, for simplicity, we omit all subscripts of function symbols and constant symbols of division rings. We also omit the dot for multiplication in a division ring.

  9. Here we use superscripts as indices, for subscriptes have been used to distinguish between objects related to \(V_A\) and those related to \(V_B\).

  10. A Hamel basis of a vector space V is an independent set \(B \subseteq V\) such that every vector in V is a (finite) linear combination of vectors in B.

  11. An Hermitian form \(\Phi \) is anisotropic if, for each \({\textbf{v}} \in V\), \(\Phi ( {\textbf{v}}, {\textbf{v}} ) = 0\) implies \({\textbf{v}} = {\textbf{0}}\). Every orthomodular Hermitian form is anisotropic. A proof can be found below Definition 1.2 on page 206 of [45].

  12. Since \({\mathcal {K}}\) is a field, the fraction notation like \(\frac{a}{b}\) is legitimate and intuitive.

  13. A geometry is a simple matroid.

  14. A map \(f: V_1 \rightarrow V_2\) is additive, if \(f (x+y) = f(x) + f(y)\) for any \(x,y \in V_1\).

  15. For a vector space V, its algebraic dual \(V^*\) is the vector space of the linear functionals on V. For a projective geometry \((G, {\star })\), its dual geometry is the ordered pair \((G^*, {\star }^*)\) such that \(G^*\) is the set of hyperplanes of \((G,{\star })\) and, for any \(A, B \in G^*\), \(A \star ^* B\) is the set of hyperplanes each of which includes \(A \cap B\).

  16. For a vector space V, \(V^\bullet \) denotes the set of non-zero vectors in V.

  17. In [39] each projective space is assumed to have at least two lines (the sentences in a box on Page 7).

References

  1. Engesser, K., Gabbay, D., Lehmann, D.: Handbook of quantum logic and quantum structures: quantum structures. Elsevier Science B.V, Amsterdam (2007)

    MATH  Google Scholar 

  2. Engesser, K., Gabbay, D., Lehmann, D.: Handbook of quantum logic and quantum structures: quantum logic. Elsevier Science B.V, Amsterdam (2009)

    MATH  Google Scholar 

  3. Birkhoff, G., von Neumann, J.: The logic of quantum mechanics. Ann. Math. 37, 823–843 (1936)

    MathSciNet  MATH  Google Scholar 

  4. Mackey, G.: Mathematical foundations of quantum mechanics. W.A.Benjamin, INC., New York (1963)

    MATH  Google Scholar 

  5. Jauch, J.: Foundations of quantum mechanics. Addison-Wesley Publishing Company, Boston (1968)

    MATH  Google Scholar 

  6. Piron, C.: Foundations of quantum physics. W.A. Benjamin Inc., New York (1976)

    MATH  Google Scholar 

  7. Aerts, D.: Quantum axiomatics. In: Engesser, K., Gabbay, D.M., Lehmann, D. (eds.) Handbook of quantum logic and quantum structures: quantum logic, pp. 79–126. Elsevier B.V, Amsterdam (2009)

    MATH  Google Scholar 

  8. Mielnik, B.: Geometry of quantum states. Commun. Math. Phys. 9(1), 55–80 (1968)

    ADS  MathSciNet  MATH  Google Scholar 

  9. Zabey, P.: Reconstruction theorems in quantum mechanics. Found. Phys. 5(2), 323–342 (1975)

    ADS  MathSciNet  Google Scholar 

  10. Dalla Chiara, M., Giuntini, R., Greechie, R.: Reasoning in quantum theory: sharp and unsharp quantum logics trends in logic. Kluwer Acadamic Press, Dordrecht (2004)

    MATH  Google Scholar 

  11. Baltag, A., Smets, S.: Complete axiomatizations for quantum actions. Int. J. Theor. Phys. 44(12), 2267–2282 (2005)

    MathSciNet  MATH  Google Scholar 

  12. Döring, A., Isham, C.: A topos foundation for theories of physics: I. formal languages for physics. J. Math. Phys. (2008). https://doi.org/10.1063/1.2883740

    Article  MathSciNet  MATH  Google Scholar 

  13. Döring, A., Isham, C.: A topos foundation for theories of physics: II. daseinisation and the liberation of quantum theory. J. Math. Phys. (2008). https://doi.org/10.1063/1.2883742

    Article  MathSciNet  MATH  Google Scholar 

  14. Döring, A., Isham, C.: A topos foundation for theories of physics: III the representation of physical quantities with arrows. J. Math. Phys. (2008). https://doi.org/10.1063/1.2883742

    Article  MathSciNet  MATH  Google Scholar 

  15. Döring, A., Isham, C.: A topos foundation for theories of physics: IV. categories of systems. J. Math. Phys. (2008). https://doi.org/10.1063/1.2883742

    Article  MathSciNet  MATH  Google Scholar 

  16. Flori, C.: A first course in topos quantum theory. Springer, Berlin (2013)

    MATH  Google Scholar 

  17. Flori, C.: A second course in Topos quantum theory. Springer, Berlin (2018)

    MATH  Google Scholar 

  18. Holik, F., Sáenz, M., Plastino, A.: A discussion on the origin of quantum probabilities. Ann. Phys. 340(1), 293–310 (2014)

    ADS  MathSciNet  MATH  Google Scholar 

  19. Abramsky, S., Coecke, B.: A categorical semantics of quantum protocols. In: Proceedings of the 19th IEEE Conference on Logic in Computer Science (LiCS’04), pp. 415–425. IEEE Press, Los Alamitos (2004)

  20. Coecke, B., Kissinger, A.: Picturing quantum processes: a first course in quantum theory and diagrammatic reasoning. Cambridge University Press, Cambridge (2017)

    MATH  Google Scholar 

  21. Heunen, C., Vicary, J.: Categories for quantum theory: an introduction. Cambridge University Press, Cambridge (2019)

    MATH  Google Scholar 

  22. D’Ariano, G., Chiribella, G., Perinotti, P.: Quantum theory from first principles: an informational approach. Cambridge University Press, Cambridge (2017)

    MATH  Google Scholar 

  23. Foulis, D., Randall, C.: Lexicographic orthogonality. J. Comb. Theor. Series A 11(2), 157–162 (1971)

    MathSciNet  MATH  Google Scholar 

  24. Dishkant, H.: Semantics of the minimal logic of quantum mechanics. Studia Logica 30(1), 23–30 (1972)

    MathSciNet  MATH  Google Scholar 

  25. Goldblatt, R.: Semantic analysis of orthologic. J. Philos. Logic 3, 19–35 (1974)

    MathSciNet  MATH  Google Scholar 

  26. Zhong, S.: Quantum states: an analysis via the orthogonality relation. Synthese 199, 15015–15042 (2021)

    MathSciNet  Google Scholar 

  27. Zhong, S.: A formal state-property duality in quantum logic. Stud. Logic 10(2), 112–133 (2017)

    Google Scholar 

  28. Moore, D.: On state spaces and property lattices. Stud. Hist. Phil. Mod. Phys. 30(1), 61–83 (1999)

    MathSciNet  MATH  Google Scholar 

  29. Randall, C., Foulis, D.: Tensor products of quantum logics do not exist. Notices Amer. Math. Soc. 26, 6 (1979)

    Google Scholar 

  30. Foulis, D., Bennett, M.: Tensor products of orthoalgebras. Order 10(3), 271–282 (1993)

    MathSciNet  MATH  Google Scholar 

  31. Aerts, D., Daubechies, I.: Physical justification for using the tensor product to describe two quantum systems as one joint system. Helvetica Physica Acta 51, 661–675 (1978)

    MathSciNet  Google Scholar 

  32. Aerts, D.: Construction of the tensor product for the lattices of properties of physical entities. J. Math. Phys. 25(5), 1434–1441 (1984)

    ADS  MathSciNet  Google Scholar 

  33. Pulmannová, S.: Tensor product of quantum logics. J. Math. Phys. 26(1), 1–5 (1985)

    ADS  MathSciNet  MATH  Google Scholar 

  34. Baltag, A., Smets, S.: Correlated information: a logic for multi-partite quantum systems. Electron. Notes Theor. Comput. Sci. 270(2), 3–14 (2011)

    MATH  Google Scholar 

  35. Renou, M.O., Trillo, D., Weilenmann, M., et al.: Quantum theory based on real numbers can be experimentally falsified. Nature 600, 625–629 (2021)

    ADS  Google Scholar 

  36. Chen, M.-C., Wang, C., Liu, F.-M., Wang, J.-W., Ying, C., Shang, Z.-X., Wu, Y., Gong, M., Deng, H., Liang, F.-T., Zhang, Q., Peng, C.-Z., Zhu, X., Cabello, A., Lu, C.-Y., Pan, J.-W.: Ruling out real-valued standard formalism of quantum theory. Phys. Rev. Lett. 128, 040403 (2022)

    ADS  Google Scholar 

  37. Li, Z.-D., Mao, Y.-L., Weilenmann, M., Tavakoli, A., Chen, H., Feng, L., Yang, S.-J., Renou, M.-O., Trillo, D., Le, T.P., Gisin, N., Acín, A., Navascués, M., Wang, Z., Fan, J.: Testing real quantum theory in an optical quantum network. Phys. Rev. Lett. 128, 040402 (2022)

    ADS  Google Scholar 

  38. Faure, C., Frölicher, A.: Modern projective geometry. Springer, The Netherlands (2000)

    MATH  Google Scholar 

  39. Beutelspacher, A., Rosenbaum, U.: Projective geometry: from foundations to applications. Cambridge University Press, Cambridge (1998)

    MATH  Google Scholar 

  40. Stubbe, I., van Steirteghem, B.: Propositional systems, Hilbert lattices and generalized Hilbert spaces. In: Engesser, K., Gabbay, D., Lehmann, D. (eds.) Handbook of quantum logic and quantum structures: quantum structures, pp. 477–523. Elsevier B.V, The Netherlands (2007)

    MATH  Google Scholar 

  41. Zhong, S.: Correspondence between Kripke frames and projective geometries. Studia Logica 106, 167–189 (2018)

    MathSciNet  MATH  Google Scholar 

  42. Kadison, R., Ringrose, J.: Fundamentals of the theory of operator algebras, elementary theory. American Mathematical Society, New York (1997)

    MATH  Google Scholar 

  43. Zhang, C., Feng, Y., Ying, M.: Unambiguous discrimination of mixed quantum states. Phys. Lett. A 353(4), 300–306 (2006)

    ADS  Google Scholar 

  44. Hessenberg, G.: Beweis des Desarguesschen Satzes aus dem Pascalschen. Math. Ann. 61(2), 161–172 (1905)

    MathSciNet  MATH  Google Scholar 

  45. Holland, S.: Orthomodularity in infinite dimensions: a theorem of M. Solèr. Bull. Amer. Math. Soc. 32, 205–234 (1995)

    MathSciNet  MATH  Google Scholar 

  46. Solèr, M.: Characterization of Hilbert spaces with orthomodularity spaces. Commun. Algebra 23, 219–243 (1995)

    MATH  Google Scholar 

  47. Wigner, E.: Group theory and its application to the quantum mechanics of atomic spectra. Expanded and improved edn. Academic Press Inc., New York (1959)

    Google Scholar 

  48. Zhong, S.: Orthogonality and quantum geometry: Towards a relational reconstruction of quantum theory. PhD thesis, University of Amsterdam (2015). http://www.illc.uva.nl/Research/Publications/Dissertations/DS-2015-03.text.pdf

  49. Zhong, S.: On characterizing linear maps of trace \(0\). Preprint on ResearchGate, https://doi.org/10.13140/RG.2.2.11138.58568/1

Download references

Acknowledgements

Most technical results in Sect. 3 are already in my PhD thesis [48] finished at the Institute for Logic, Language and Computation (ILLC), University of Amsterdam, in 2015, while the characterization of quasi-linear maps with the same accompanying field isomorphism and that of linear maps of trace 0 in this paper use completely different and more clever ideas than those in [48]. I am very grateful to my supervisors, Prof. Johan van Benthem, Dr. Alexandru Baltag and Prof. Sonja Smets, for their supervision and the detailed comments on my thesis which are invaluable in the writing of this paper. I am also very grateful to the members in my PhD committee, Dr. Nick Bezhanishvili, Prof. Robert Goldblatt, Prof. John Harding, Prof. Yde Venema, Prof. Ronald de Wolf and Prof. Mingsheng Ying, for their comments on my thesis which are very helpful in writing this paper. The results were also presented in the Delta Workshop 9 (2018) and the Quantum & Beyond Workshop (2020), and I am very grateful to the audience for the very helpful discussion. The improvement of the technical results in [48], reported in [49], has been obtained when I used the Research and Writing Room in Guangzhou Library in the summer of 2018. I’m very grateful to Guangzhou Library for providing the great environment and facilities for my research. Last but not the least, my PhD research was funded by China Scholarship Council (CSC), and the writing of this paper is supported by the National Social Science Fund of China (No. 20CZX048).

Funding

The writing of this paper is supported by the National Social Science Fund of China (No. 20CZX048).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shengyang Zhong.

Additional information

Publisher's Note

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

Appendices

Appendix A

In this appendix, we are going to prove Proposition 2.17 which is crucial to the dimension theory of quantum Kripke frames.

We start from reviewing three results about projective geometry.

The first one is about a way of constructing the linear closure of a set.

Proposition A.1

Let \({\mathfrak {G}} = ( G, {\star } )\) be a projective geometry. For each \(A\subseteq G\), define a sequence \(\{ A_i \}_{i \in {\mathbb {N}}}\) of subsets of G as follows:

  • \(A_0=A\);

  • \(A_{n+1}=\bigcup \{ a\star b\mid a,b\in A_n \}\).

Then \({\mathcal {C}}(A)=\bigcup _{i\in {\mathbb {N}}} A_i\).

Proof

First we prove by induction that \(A_i\subseteq {\mathcal {C}}(A)\), for every \(i\in {\mathbb {N}}\).

Base Step: \(i=0\). By the definition of linear closures \(A_0=A\subseteq {\mathcal {C}}(A)\).

Induction Step: \(i=n+1\). Let \(c\in A_{n+1}\) be arbitrary. By the definition of \(A_{n+1}\) there are \(a,b\in A_n\) such that \(c\in a\star b\). By the induction hypothesis \(a,b\in A_n \subseteq {\mathcal {C}}(A)\), so \(c\in a\star b\subseteq {\mathcal {C}}(A)\) since \({\mathcal {C}}(A)\) is a subspace.

This finishes the proof by induction. Therefore, \(\bigcup _{i\in {\mathbb {N}}} A_i\subseteq {\mathcal {C}}(A)\).

Second we prove that \(\bigcup _{i\in {\mathbb {N}}} A_i\) is a subspace including A, and thus \({\mathcal {C}}(A)\subseteq \bigcup _{i\in {\mathbb {N}}} A_i\). By definition \(A=A_0\subseteq \bigcup _{i\in {\mathbb {N}}} A_i\). Now let \(a,b\in \bigcup _{i\in {\mathbb {N}}} A_i\) be arbitrary. Then there are \(n,n'\in {\mathbb {N}}\) such that \(a\in A_n\) and \(b\in A_{n'}\). Note that by definition \(A_i\subseteq A_{i+1}\), for every \(i\in {\mathbb {N}}\). Hence \(a,b\in A_m\), where \(m=max\{ n,n'\}\). Therefore, \(a\star b\subseteq A_{m+1}\subseteq \bigcup _{i\in {\mathbb {N}}} A_i\). As a result, \(\bigcup _{i\in {\mathbb {N}}} A_i\) is a subspace. \(\square \)

The second one is a very important and useful result in projective geometry called the projective law.

Theorem A.2

(Corollary 2.4.5 in [38]) Let \({\mathfrak {G}} = ( G, {\star } )\) be a projective geometry. For any non-empty sets \(A,B\subseteq G\),

$$\begin{aligned} {\mathcal {C}}(A\cup B)=\bigcup \{ a\star b\mid a\in {\mathcal {C}}(A),\ b\in {\mathcal {C}}(B)\}. \end{aligned}$$

The third one is a corollary of the projective law to be used later.

Corollary A.3

Let \({\mathfrak {G}} = ( G, {\star } )\) be a projective geometry. For any \(a\in G\) and \(A\subseteq G\), \({\mathcal {C}}(\{ a\}\cup {\mathcal {C}} (A))={\mathcal {C}}(\{ a\} \cup A)\).

Proof

By definition \({\mathcal {C}}(A)\) is a subspace, and thus \({\mathcal {C}}({\mathcal {C}}(A)) = {\mathcal {C}}(A)\). Using the projective law,

$$\begin{aligned} {\mathcal {C}}(\{ a\} \cup A)&= \bigcup \{ c\star d\mid c\in {\mathcal {C}}(\{ a\}),\ d\in {\mathcal {C}}(A) \} \\&= \bigcup \{ c\star d\mid c\in {\mathcal {C}}(\{ a\}),\ d\in {\mathcal {C}}({\mathcal {C}}(A)) \} \\&= {\mathcal {C}}( \{ a \} \cup {\mathcal {C}}(A)) \end{aligned}$$

\(\square \)

Now we are going to prove that in a quantum Kripke frame, for any \(n\in {\mathbb {N}}\) and \(s_1,...,s_n\in \Sigma \), \({\mathcal {C}}(\{ s_1,...,s_n\})={\sim }{\sim }\{ s_1,...,s_n\}\). Please remind that, according to Theorem 2.16, in a quantum Kripke frame \({\mathfrak {F}} = (\Sigma , {\rightarrow })\), for any \(s, t \in \Sigma \), \({\sim }{\sim }\{ s, t \}\) is the line \(s \star t\) in the projective geometry corresponding to \({\mathfrak {F}}\).

Lemma A.4

In a quantum Kripke frame \({\mathfrak {F}} = ( \Sigma , {\rightarrow })\), if \(P\subseteq \Sigma \) is closed, it is a subspace of \({\textbf{G}} ({\mathfrak {F}})\).

Proof

Assume that \(s,t\in P\). Then \(\{ s,t\}\subseteq P\). Applying 2 of Lemma 2.15 twice, one can obtain \({\sim }{\sim }\{ s,t\}\subseteq {\sim }{\sim }P\). Since P is closed, \({\sim }{\sim }\{ s,t\}\subseteq {\sim }{\sim }P=P\). \(\square \)

Lemma A.5

In a quantum Kripke frame \({\mathfrak {F}} = ( \Sigma , {\rightarrow })\), \({\sim }Q={\sim }{\mathcal {C}}(Q)\), for every \(Q\subseteq \Sigma \).

Proof

By definition \(Q\subseteq {\mathcal {C}}(Q)\), so \({\sim }{\mathcal {C}}(Q)\subseteq {\sim }Q\) by 2 of Lemma 2.15. It remains to show that \({\sim }Q\subseteq {\sim }{\mathcal {C}}(Q)\).

We define a sequence of sets \(\{ Q_i \}_{ i \in {\mathbb {N}}}\) in the same way as in Proposition A.1. Then by the proposition \({\mathcal {C}}(Q)=\bigcup _{i\in {\mathbb {N}}} Q_i\). It is easy to see from the definition that \({\sim }{\mathcal {C}}(Q)={\sim }\bigcup _{i\in {\mathbb {N}}} Q_i=\bigcap _{i\in {\mathbb {N}}} {\sim }Q_i\). We prove \({\sim }Q \subseteq \bigcap _{i\in {\mathbb {N}}} {\sim }Q_i = {\sim }{\mathcal {C}}(Q)\) by showing that \({\sim }Q\subseteq {\sim }Q_i\), for every \(i\in {\mathbb {N}}\). Use induction on i.

Base Step: \(i=0\). \({\sim }Q\subseteq {\sim }Q={\sim }Q_0\) obviously holds.

Induction Step: \(i=n+1\). Let \(s\in {\sim }Q\) and \(t\in Q_{n+1}\) be arbitrary. By definition there are \(u,v\in Q_n\) such that \(t\in {\sim }{\sim }\{ u, v \}\). By IH \(s\in {\sim }Q\subseteq {\sim }Q_n\). Hence \(s\not \rightarrow u\) and \(s\not \rightarrow v\), i.e. \(s\in {\sim }\{ u,v\}\). Since \(t\in {\sim }{\sim }\{ u, v \}\), \(s\not \rightarrow t\). Since t is arbitrary, \(s\in {\sim }Q_{n+1}\). Therefore, \({\sim }Q\subseteq {\sim }Q_{n+1}\). \(\square \)

The following lemma suggests a way to get a bigger closed set from a smaller one using linear closures. It is a special case of Proposition 14.2.5 in [38]. Since a direct proof is not long, we give it here to avoid introducing general terminologies.

Lemma A.6

In a quantum Kripke frame \({\mathfrak {F}} = ( \Sigma , {\rightarrow })\), let \(s\in \Sigma \) and \(P\subseteq \Sigma \) be closed. Then \({\mathcal {C}}(\{ s\} \cup P)\) is closed.

Proof

If \(s \in P\), then \({\mathcal {C}} (\{ s \} \cup P) = {\mathcal {C}} (P) = P\) is closed by Lemma A.4 and the definition of subspaces. It remains to show the case when \(s \not \in P\). Since P is closed, \(s \not \in {\sim }{\sim }P\), so there is a \(u \in {\sim }P\) such that \(s \rightarrow u\).

By Lemma 2.15\({\mathcal {C}}(\{ s\} \cup P) \subseteq {\sim }{\sim }{\mathcal {C}}(\{ s\} \cup P)\). It remains to show that \({\sim }{\sim }{\mathcal {C}}(\{ s\} \cup P) \subseteq {\mathcal {C}}(\{ s\} \cup P)\). By Lemma A.5 it suffices to show that \({\sim }{\sim }(\{ s\} \cup P) \subseteq {\mathcal {C}}(\{ s\} \cup P)\).

Let \(w \in {\sim }{\sim }(\{ s\} \cup P)\) be arbitrary. If \(w = s\), then \(w \in {\mathcal {C}}(\{ s\} \cup P )\); so it remains to deal with the case when \(w \ne s\). By Lemma 4.11 in [41] there is a \(v \in {\sim }{\sim }\{ w, s \}\) such that \(u \not \rightarrow v\). Since \(s \rightarrow u\) and \(u \not \rightarrow v\), \(s \ne v\). Hence by Lemma 4.12 in [41] \(w \in {\sim }{\sim }\{ s, v \}\). To show that \(w \in {\mathcal {C}} ( \{ s \} \cup P )\) and thus finish the proof, it remains to show that \(v \in P = {\sim }{\sim }P\).

Let \(x \in {\sim }P\) be arbitrary. When \(x = u\), then \(v \not \rightarrow u\) follows from the construction of v. When \(x \ne u\), by Lemma 4.11 in [41] there is a \(y \in {\sim }{\sim }\{ x, u \}\) such that \(y \not \rightarrow s\). Since \(x, u \in {\sim }P\), \(y \in {\sim }{\sim }\{ x, u \} \subseteq {\sim }{\sim }{\sim }P = {\sim }P\). Together with \(y \not \rightarrow s\), we have \(y \in {\sim }\{ s \} \cap {\sim }P = {\sim }( \{ s \} \cup P )\). Hence \(w \not \rightarrow y\). Since \(v \in {\sim }{\sim }\{ w, s \}\), \(v \not \rightarrow y\). Since \(s \rightarrow u\) and \(y \not \rightarrow s\), \(y \ne u\). Hence by Lemma 4.12 in [41] \(x \in {\sim }{\sim }\{ y, u \}\). Since \(v \not \rightarrow u\) and \(v \not \rightarrow y\), \(v \in {\sim }\{ u, y \}\), so \(x \not \rightarrow v\). Since x is arbitrary, \(v \in {\sim }{\sim }P = P\). \(\square \)

Finally we are ready to prove Proposition 2.17.

Proposition A.7

(Proposition 2.17) In a quantum Kripke frame \({\mathfrak {F}} = ( \Sigma , {\rightarrow })\), for any \(n\in {\mathbb {N}}\) and \(s_1,...,s_n\in \Sigma \), \({\mathcal {C}}(\{ s_1,...,s_n\})={\sim }{\sim }\{ s_1,...,s_n\}\).

Proof

We prove by induction that, for every \(n\in {\mathbb {N}}\), \({\mathcal {C}}(\{ s_1,...,s_n \})\) is closed.

Base Step: \(n=0\). By convention \(\{ s_1,...,s_n \} = \emptyset \), so \({\mathcal {C}} (\emptyset ) = \emptyset \) is closed by Lemma 2.15.

Induction Step: \(n=k+1\). By the induction hypothesis \({\mathcal {C}}(\{ s_1,...,s_k\})\) is closed. Then by Lemma A.6\({\mathcal {C}}(\{ s_{k+1}\} \cup {\mathcal {C}}(\{ s_1,...,s_k\}))\) is closed. By Corollary A.3

$$\begin{aligned} {\mathcal {C}}(\{ s_{k+1}\} \cup {\mathcal {C}}(\{ s_1,...,s_k\}))={\mathcal {C}}(\{ s_{k+1}\} \cup \{ s_1,...,s_k\})={\mathcal {C}}(\{ s_1,...,s_k,s_{k+1}\}) \end{aligned}$$

Hence \({\mathcal {C}}(\{ s_1,...,s_k,s_{k+1}\})\) is closed. This finishes the proof by induction.

Now by Lemma A.5\({\sim }{\sim }\{ s_1,...,s_n\}={\sim }{\sim }{\mathcal {C}} (\{ s_1,...,s_n\} )\). As \({\mathcal {C}} (\{ s_1,...,s_n\} )\) is closed, \({\sim }{\sim }\{ s_1,...,s_n\}={\mathcal {C}} (\{ s_1,...,s_n\} )\). \(\square \)

Appendix B

In this appendix, we list the results in projective geometry in the literature which are used in this paper. The following results with the notations are from [38], and the footnotes are added by me.

Proposition 2.3.3 (Page 34) For every subset A of a projective geometry G there exists a smallest subspace \({\mathcal {C}} (A)\) containing A and the so obtained operator \({\mathcal {C}}\) is a closure operator on G.

Proposition 3.1.13 (Page 59) Let G be a projective geometry and \({\mathcal {C}}\) the closure operator associated to it. Then the set G together with the closure operator \({\mathcal {C}}\) is a geometry.Footnote 13

Lemma 6.3.4 (Page 137) Let \(V_1\), \(V_2\) be vector spaces over \(K_1\), \(K_2\) respectively, \(f: V_1 \rightarrow V_2\) and \(h: V_1 \rightarrow V_2\) two additive mapsFootnote 14 such that \(h(x) \in K_2 \cdot f(x)\) for all \(x \in V_1\). We suppose that \(f(V_1)\) contains at least two linearly independent vectors. Then there exists a unique \(\mu \in K_2\) such that \(h(x) = \mu \cdot f(x)\) for all \(x \in V_1\).

Theorem 10.3.1 (Page 243) For a partial map \(g: G \dashrightarrow G'\) between arguesian geometries the following conditions are equivalent:

  1. (1)

    if homogeneous coordinates \(u: {\mathcal {P}} V \rightarrow G\) and \(u': {\mathcal {P}} V' \rightarrow G'\) are given, there exists a semilinear map \(f: V \rightarrow V'\) such that \(g = u' \circ {\mathcal {P}} f \circ u^ {-1}\),

  2. (2)

    g is described by a semilinear map in some homogeneous coordinates,

  3. (3)

    g is the composite of two non-degenerate morphisms,

  4. (4)

    g is the composite of finitely many non-degenerate morphisms.

Definition 10.3.2 (Page 243) A morphism \(g: G \dashrightarrow G'\) between arguesian geometries is called arguesian if it satisfies the equivalent conditions of the preceding theorem.

Proposition 11.2.5 (Page 259) For any vector space V one has a natural isomorphism \(\Theta _V: {\mathcal {P}} (V^*) \rightarrow ({\mathcal {P}} V)^*\).Footnote 15 It is induced by the map \(\Omega : (V^*)^\bullet \rightarrow ({\mathcal {P}}V)^*, \varphi \mapsto {\mathcal {P}}(\text {ker}\ \varphi )\).Footnote 16

Definition 13.4.1(Page 310) A dualized (projective) geometry is a projective geometry G together with a subspace \(\Gamma \subseteq G^*\) of the dual geometry satisfying \(\bigcap \Gamma = \emptyset \).

Example 13.4.2 (Page 310) Let V be a dualized vector space, i.e. a vector space with a vector subspace \(V' \subseteq V^*\) of the algebraic dual which is separating. This means that for every \(x \ne 0\) there exists \(l \in V'\) such that \(l (x) = 1\). Then \({\mathcal {P}} (V)\) together with the subspace \(\Theta _V ({\mathcal {P}}(V'))\) (cf. 11.2.5) is a dualized geometry.

Proposition 13.5.3 (Page 316) Let \(V_1\) and \(V_2\) be dualized vector spaces. Then for every non-degenerate continuous homomorphism \(g: {\mathcal {P}} (V_1) \dashrightarrow {\mathcal {P}} (V_2)\) there exists a continuous quasilinear map \(f: V_1 \rightarrow V_2\) such that \(g = {\mathcal {P}} f\).

Proposition 14.1.4 (Page 324) Let G be an orthogeometry. Then G together with the set \(\Gamma := \{a^\bot \mid a \in G \}\) is a dualized geometry.

Definition 14.4.1 (Page 334) An adjunction between two orthogeometries \(G_1\) and \(G_2\) consists of two partial maps \(g_1: G_1 \dashrightarrow G_2\) and \(g_2: G_2 \dashrightarrow G_1\) satisfying

  1. 1)

    \(\text {Ker}\ g_1 = (\text {Im}\ g_2)^\bot \),

  2. 2)

    \(\text {Ker}\ g_2 = (\text {Im}\ g_1)^\bot \),

  3. 3)

    for all \(a \in \text {Dom}\ g_1\) and \(b \in \text {Dom}\ g_2\) one has \(g_1 a \perp b\) iff \(a \perp g_2 b\).

Proposition 14.4.4 (Page 335) For a partial map \(g_1: G_1 \dashrightarrow G_2\) between orthogeometries the following conditions are equivalent:

  1. (1)

    there exists a partial map \(g_2: G_2 \dashrightarrow G_1\) such that \((g_1, g_2)\) is an adjunction,

  2. (2)

    \(g_1\) is a continuous homomorphism.

Lemma 14.4.9 (Page 337) A quasilinear map \(f: V_1 \rightarrow V_2\) is continuous if and only if there exists a quasilinear map \(f^\circ : V_2 \rightarrow V_1\) such that \(\Phi _2 (fx,y) = \tau (\Phi _1 (x,f^\circ y))\) for all \(x \in V_1\) and \(y \in V_2\). The map \(f^\circ \) is unique and called the adjoint of f.

The following results with the notations are from [39].

Definition (Page 62) Let \({\textbf{P}}\) be a projective space.Footnote 17 We say that in \({\textbf{P}}\) the theorem of Pappus holds if any two intersecting lines g and h with \(g \ne h\) satisfy the following condition. If \(A_1, A_2, A_3\) are distinct points on g and \(B_1, B_2, B_3\) are distinct points on h all different from \(g \cap h\) then the points

$$\begin{aligned} Q_{12}:= A_1 B_2 \cap B_1 A_2,\ Q_{23}:= A_2 B_3 \cap B_2 A_3 \text{ and } Q_{31}:= A_3 B_1 \cap B_3 A_1 \end{aligned}$$

lie on a common line.

Theorem 2.2.2 (Page 62) Let V be a vector space over a division ring F. Then the theorem of Pappus holds in \({\textbf{P}} (V)\) if and only if F is commutative (in other words, if F is a field).

Theorem 2.2.3 Hessenberg’s Theorem (Page 65) Let \({\textbf{P}}\) be an arbitrary projective space. If the theorem of Pappus holds in \({\textbf{P}}\) then the theorem of Desargues is also true in \({\textbf{P}}\).

Theorem 3.6.7 (Page 132) The projective collineations of \({\textbf{P}} (V)\) are precisely the products of central collineations.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhong, S. Quantum Entanglement: An Analysis via the Orthogonality Relation. Found Phys 53, 75 (2023). https://doi.org/10.1007/s10701-023-00710-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10701-023-00710-0

Keywords

Navigation