Skip to main content
Log in

Mind–Body Interaction and Modern Physics

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

The idea that mind and body are distinct entities that interact is often claimed to be incompatible with physics. The aim of this paper is to disprove this claim. To this end, we construct a broad mathematical framework that describes theories with mind–body interaction (MBI) as an extension of current physical theories. We employ histories theory, i.e., a formulation of physical theories in which a physical system is described in terms of (i) a set of propositions about possible evolutions of the system and (ii) a probability assignment to such propositions. The notion of dynamics is incorporated into the probability rule. As this formulation emphasises logical and probabilistic concepts, it is ontologically neutral. It can be used to describe mental ‘degrees of freedom’ in addition to physical ones. This results into a mathematical framework for psycho-physical interaction (\(\Psi \Phi\)I formalism). Interestingly, a class of \(\Psi \Phi\)I theories turns out to be compatible with energy conservation.

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

Similar content being viewed by others

Notes

  1. The space of all Lorentzian metrics on the spacetime manifold M is denoted by \(\text{ LRiem }(M)\). The space of spacetime geometries is the quotient manifold \(\text{ LRiem }(M)/\text{Diff }(M)\) where \(\text{ Diff }(M)\) is the infinite-dimensional group of diffeomorphisms on M.

  2. Most of these problems are facets of the so-called problem of time in quantum gravity [18,19,20].

  3. Physical theories are traditionally described in terms of kinematics, dynamics and initial conditions. Kinematics defines the physical variables and the symmetries of a system, and how the former relate to measurable quantities. Dynamics describes how physical variables evolve in time. Initial conditions are necessary in order to obtain unique predictions about particular physical systems. In histories theory, the notion of kinematics is incorporated into the definition of the space of history propositions, while the notions of dynamics and initial conditions are incorporated into the rule of probability assignment.

  4. Note that in this paper we only employ the logical structure of quantum histories, and we do not commit to the decoherent histories interpretation. Our analysis makes sense, even in the context of Copenhagen quantum mechanics—see, the discussion in Sect. 3.2.

  5. Unlike her neuroscientist cousin of the same name, who became famous by Jackson’s thought experiment [41, 42], our Mary has had both RED and GREEN experiences in her past.

  6. This follows from an translation of Locke’s famous argument about an ‘inverted spectrum’ [44]—see also Ref. [45] and references therein—into the language of contemporary physics. Let \(i:\mathcal{C}_q \rightarrow \mathcal{V}_{\Phi }\) be the inclusion map of the set \(\mathcal{C}_{q}\) of propositions about qualia into the set \(\mathcal{V}_{\Phi }\) of history propositions about physical properties. Since qualia are not part of the physical theory, the physical predictions ought to be the same for any inclusion map \(i' = f \circ i :\mathcal{C}_q \rightarrow \mathcal{V}_{\Phi }\), where f is an automorphism of \(\mathcal{C}_q\). Thus, the probability assignment is invariant under the group \(\text{ Aut }(\mathcal{C}_q)\) of automorphisms of \(\mathcal{C}_q\). This means that there is no dynamical reason why the insertion of a red rose is correlated with a RED quale rather than a GREEN quale. As a matter of fact, there is no reason why the insertion of the rose is not correlated with a sound quale, i.e., why Mary does not hear Beethoven’s 9th symphony on seeing the red rose.

  7. Energy is conserved only in stationary spacetimes, i.e., a class of spacetimes characterized by a specific symmetry (the existence of a timelike Killing field). Energy is not conserved in generic spacetimes. For example, energy is not conserved in the expanding universe models that are employed in cosmology. Energy conservation holds approximately at scales much smaller than the Hubble length that characterises the expansion of the universe.

  8. In quantum theory, energy is strictly conserved only for a particular class of initial states (namely, the eigenstates of the Hamiltonian operator). In general, what is conserved (i.e., it is the same at all times) is the probability distribution for the values of energy. This means that two individual quantum systems that have been prepared identically will, in general, be measured with different values of energy. Of course, a measured system is an open system, so one does not expect energy to be preserved during the measurement process. In any case, the statement that the value of energy remains the same during time evolution is not true.

  9. This holds for quantum-classical interactions in which the quantum system backreacts on the classical, i.e., if the interaction is two-way. The backreaction to the classical variable must be expressed in terms of probabilities with respect to a specific basis in the Hilbert space. However, a generic Hamiltonian for the evolution of the quantum system generates superpositions in the specified basis, thus, rendering impossible the formulation of backreaction. Hence, to describe backreaction, we must introduce a mechanism of suppressing such interferences in each time step, which is mathematically equivalent to a dynamical reduction process. For examples of such models, see Refs. [61,62,63]. On the other hand, if the coupling is one-way, and there is no backreaction from the quantum system to the classical, the quantum evolution can be fully unitary.

References

  1. Leibniz, G.W.: Theodicy: Essays on the Goodness of God and the Freedom of Man and the Origin of Evil. English translation by E. M. Huggard. Open Court, La Salle, Illinois (1710/1985)

  2. Bunge, M.: The Mind-Body Problem: A Psychobiological Approach. Pergamon, Oxford (1980)

    Google Scholar 

  3. Fodor, J.: The Mind–Body Problem. In: Warner, R., Szubka, T. (eds.) The Mind–Body Problem. Blackwell, Oxford (1994)

    Google Scholar 

  4. Butterfield, J.: Quantum Curiosities of Psychophysics. In: Cornwell, J. (ed.) Consciousness and Human Identity. Oxford University Press, Oxford (1997)

    Google Scholar 

  5. Collins, R.: Modern physics and the energy-conservation objection to mind–body dualism. Am. Philos. Q. 45, 31 (2008)

    Google Scholar 

  6. Pitts, J.B.: Conservation laws and the philosophy of mind: opening the black box, finding a mirror. Philosophia (2019). https://doi.org/10.1007/s11406-019-00102-7

    Article  Google Scholar 

  7. Gibb, S.: Closure principles and the laws of conservation of energy and momentum. Dialectica 64, 363 (2010)

    Google Scholar 

  8. Irwin, K., Amaral, M., Chester, D.: The self-simulation hypothesis interpretation of quantum mechanics. Entropy 22, 247 (2020)

    ADS  MathSciNet  Google Scholar 

  9. Isham, C.J.: Lectures On Quantum Theory: Mathematical And Structural Foundations. Imperial College Press, London (1995)

    MATH  Google Scholar 

  10. Dirac, P.: Public Lecture in the Indian Science Congress. Baroda (1955)

  11. Chalmers, D.: Facing up to the problem of consciousness. J. Consc. Stud. 2, 200 (1995)

    Google Scholar 

  12. Chalmers, D.: The puzzle of conscious experience. Sci. Am. 273, 80 (1995)

    Google Scholar 

  13. Penrose, R.: Shadows of the Mind. Oxford Univesity Press, Oxford (1994)

    Google Scholar 

  14. Plato: Timaeus and Critias, translated by Waterfield, R. Oxford University Press, Oxford (~360 B. C./2008)

  15. Goldstein, H., Poole, C.P., Safko, J.L.: Chapter 13 in Classical Mechanics, 3rd edn. Addison Wesley, Reading (2001)

    Google Scholar 

  16. Harman, P.M.: Energy, Force and Matter: The Conceptual Development of Nineteenth-Century Physics. Cambridge University Press, Cambridge (1982)

    Google Scholar 

  17. Wheeler, J.A., Ford, K.: Geons. W. W. Norton, Black Holes and Quantum Foam, New York (1998)

    MATH  Google Scholar 

  18. Isham, C.J.: Canonical Quantum Gravity and the Problem of Time. Lecture Notes at Salamanca Summer School (1992)

  19. Kuchar, K.: The problem of time in canonical quantum gravity. In: Ashtekar, A., Stachel, J. (eds.) Conceptual Problems of Quantum Gravity. Birkhäuser, Basel (1991)

    Google Scholar 

  20. Anderson, E.: The Problem of Time: Quantum Mechanics Versus General Relativity. Springer, Berlin (2017)

    MATH  Google Scholar 

  21. Mach, E.: The Science of Mechanics; A Critical and Historical Account of its Development, translated by MCormack, T.H. (1906), Open Court, LaSalle (1883/1960)

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

    MathSciNet  MATH  Google Scholar 

  23. Houtappel, R.M.F., van Dam, H., Wigner, E.P.: The conceptual basis and use of the geometric invariance principles. Rev. Mod. Phys. 37, 595 (1965)

    ADS  MathSciNet  Google Scholar 

  24. Griffiths, R.B.: Consistent Quantum Theory. Cambridge University Press, Cambridge (2003)

    Google Scholar 

  25. Omnés, R.: The Interpretation of Quantum Mechanics. Princeton University Press, Princeton (1994)

    MATH  Google Scholar 

  26. Omnés, R.: Understanding Quantum Mechanics. Princeton University Press, Princeton (1999)

    MATH  Google Scholar 

  27. Gell-Mann, M., Hartle, J.B.: Quantum mechanics in the light of quantum cosmology. In: Zurek, W. (ed.). Complexity, Entropy, and the Physics of Information. Addison Wesley, Reading (1990)

  28. Hartle, J.B.: Spacetime quantum mechanics and the quantum mechanics of spacetime. Gravitation and quantizations. In: Proceedings of the 1992 Les Houches Summer School, ed. by B. Julia and J. Zinn- Justin, Les Houches Summer School Proceedings, Vol. LVII. North Holland, Amsterdam (1995)

  29. Isham, C.J.: Quantum logic and the histories approach to quantum theory. J. Math. Phys. 35, 2157 (1994)

    ADS  MathSciNet  MATH  Google Scholar 

  30. Isham, C.J., Linden, N.: Quantum temporal logic and decoherence functionals in the histories approach to generalised quantum theory. J. Math. Phys. 35, 5452 (1994)

    ADS  MathSciNet  MATH  Google Scholar 

  31. Savvidou, K.: The action operator for continuous-time histories. J. Math. Phys. 40, 5657 (1999)

    ADS  MathSciNet  MATH  Google Scholar 

  32. Savvidou, N.: Space-time symmetries in histories canonical gravity. In: Oriti, D. (ed.) Approaches to Quantum Gravity. Cambridge University Press, Cambridge (2009)

    MATH  Google Scholar 

  33. Anastopoulos, C.: Quantum processes on phase space. Ann. Phys. 303, 275 (2003)

    ADS  MathSciNet  MATH  Google Scholar 

  34. Savvidou, K.: General Relativity Histories Theory I: The Spacetime Character of the Canonical Description. Class. Quant. Grav. 21, 615 (2004)

    ADS  MathSciNet  MATH  Google Scholar 

  35. Savvidou, K.: General Relativity Histories Theory II: Invariance Groups. Class. Quant. Grav. 21, 631 (2004)

    ADS  MathSciNet  MATH  Google Scholar 

  36. Okon, E., Sudarsky, D.: Measurements according to consistent histories. Stud. Hist. Philos. Mod. Phys. 48, 7 (2014)

    MathSciNet  MATH  Google Scholar 

  37. Griffiths, R.B.: Consistent Quantum Measurements. arXiv:1501.04813 (2015)

  38. Okon, E., Sudarsky, D.: The consistent histories formalism and the measurement problem. Stud. Hist. Philos. Mod. Phys. 52, 217 (2015)

    MathSciNet  MATH  Google Scholar 

  39. Anastopoulos, C., Savvidou, K.: Quantum probabilities for time-extended alternatives. J. Math. Phys. 48, 032106 (2007)

    ADS  MathSciNet  MATH  Google Scholar 

  40. Anastopoulos, C., Savvidou, K.: Time of arrival and localization of relativistic particles. J. Math. Phys. 60, 032301 (2019)

    ADS  MathSciNet  MATH  Google Scholar 

  41. Jackson, F.: Epiphenomenal qualia. Philos. Q. 32, 127 (1982)

    Google Scholar 

  42. Jackson, F.: What Mary didn’t know. J. Philos. 83, 291 (1986)

    Google Scholar 

  43. Tye, M.: Qualia, The Stanford Encyclopedia of Philosophy (Winter 2016 Edition), In: Zalta, E.N. (ed.). https://plato.stanford.edu/archives/win2016/entries/qualia/

  44. Locke, J.: Essay Concerning Human Understanding. Oxford University Press, Oxford (1689/1975)

  45. Byrne, A.: Inverted Qualia, The Stanford Encyclopedia of Philosophy (Winter 2016 Edition), In: Zalta, E.N. (ed.). https://plato.stanford.edu/archives/win2016/entries/qualia-inverted/ (2016)

  46. Jung, C.G., Pauli W.: Atom and Archetype: The Pauli/Jung Letters, 1932–1958. In: Meier, C.A., translated by Roscoe, D. (ed.). Princeton University Press, Princeton (2014)

  47. Main, R.: The cultural significance of synchronicity for Jung and Pauli. J. Anal. Psychol. 59, 174 (2014)

    Google Scholar 

  48. Jung, C.G.: Synchronicity: An Acausal Connecting Principle. Routledge, East Sussex (2008)

    Google Scholar 

  49. Beck, F., Eccles, J.: Quantum aspects of brain activity and the role of consciousness. Proc. Natl. Acad. Sci (USA) 89, 11357 (1992)

    ADS  Google Scholar 

  50. Hameroff, S.R., Penrose, R.: Consciousness in the universe: a review of the Orch OR theory. Phys. Life Rev. 11, 39 (2014)

    ADS  Google Scholar 

  51. Gao, S.: A quantum theory of conscioussness. Minds Mach. 18, 39 (2008)

    Google Scholar 

  52. Schwartz, J.M., Stapp, H.P., Bauregard, M.: Quantum physics in neuroscience and psychology: a neurophysical model of mind–brain interaction. Philos. Trans. R. Soc. Lond. B 360, 1309 (2005)

    Google Scholar 

  53. Von Neumann, J.: Mathematical Foundations of Quantum Mechanics. Princeton University Press, Princeton (1955)

    MATH  Google Scholar 

  54. Wigner, E.P.: Remarks on the mind body question. Am. J. Phys. 35, 1169 (1967)

    ADS  Google Scholar 

  55. Bassi, A., Ghirardi, G.C.: Dynamical reduction models. Phys. Rep. 379, 257 (2003)

    ADS  MathSciNet  MATH  Google Scholar 

  56. Karolyhazy, F.: Gravitation and quantum mechanics of macroscopic objects. Nuovo Cim. 52, 390 (1966)

    ADS  Google Scholar 

  57. Diosi, L.: Gravitation and quantum-mechanical localization of macro-objects. Phys. Lett. A 105, 199 (1984)

    ADS  Google Scholar 

  58. Diosi, L.: A universal master equation for the gravitational violation of quantum mechanics. Phys. Lett. 120, 377 (1987)

    Google Scholar 

  59. Penrose, R.: Gravity and state vector reduction. In: Penrose, R., Isham, C.J. (eds.) Quantum Concepts in Space and Time. Clarendon Press, Oxford (1986)

    Google Scholar 

  60. Penrose, R.: On gravity’s role in quantum state reduction. Gen. Rel. Grav. 28, 581 (1996)

    ADS  MathSciNet  MATH  Google Scholar 

  61. Blanchard, Ph., Jadcyk, A.: Strongly coupled quantum and classical systems and Zeno’s effect. Phys. Lett. A 175, 157 (1993)

    ADS  MathSciNet  Google Scholar 

  62. Diosi, L., Halliwell, J.J.: Coupling classical and quantum variables using continuous quantum measurement theory. Phys. Rev. Lett. 81, 2846 (1998)

    ADS  MathSciNet  MATH  Google Scholar 

  63. Diosi, L., Gisin, N., Strunz, W.T.: Quantum approach to coupling classical and quantum dynamics. Phys. Rev. A 61, 022108 (2000)

    ADS  Google Scholar 

  64. Anastopoulos, C.: Particle or Wave: The Evolution of the Concept of Matter in Modern Physics. Princeton University Press, Princeton (2008)

    Google Scholar 

  65. Beltrametti, E.G., Cassinelli, G.: The Logic of Quantum Mechanics. Cambridge University Press, Cambridge (2010)

    MATH  Google Scholar 

  66. Marsden, J.E., Ratiu, T.: Introduction to Mechanics and Symmetry. Springer, Berlin (1999)

    MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Charis Anastopoulos.

Additional information

Publisher's Note

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

Appendices

Appendix 1: The Structure of Histories Theory

In this Appendix, we briefly summarize some features of histories theory, both classical and quantum.

1.1 Classical Histories Theory

Consider a system described by a sample space \(\Gamma\) that is a differentiable manifold. Single-time propositions corresponds to measurable subsets C of \(\Gamma\). Let \(\mathcal{T} = \{t_1, t_2, \ldots , t_n\}\) be a discrete time-set for this system with \(t_1< t_2< \ldots < t_n\). In general, \(\mathcal{T}\) may be any partially ordered set, but here we restrict our considerations to the simplest case.

The path space \(\Pi\) consists of all maps \(\gamma :\mathcal{T} \rightarrow \Gamma\). The set of history propositions \(\mathcal{V}\) consists of all measurable subsets of \(\Pi\). Hence, we can write a history proposition for this system as

$$\alpha = (C_{t_{1}}, C_{t_{2}}, \ldots , C_{t_{n}}),$$
(8)

where \(C_{t_i}\) is a subset of \(\Gamma\) that corresponds to the proposition that the system was found in \(C_{t_i}\) at time \(t_i\). Logical operations are standardly defined in terms of set-theoretic operations between subsets of \(\Pi\).

A probability functional assigns a probability \(\text{ Prob }(\alpha )\) to each history proposition \(\alpha \in \mathcal{V}\), as

$$\begin{aligned} \text{ Prob }(\alpha ) = \int dx_1 \ldots dx_n \chi _{C_{t_1}}(x_1) \chi _{C_{t_2}}(x_2) \ldots \chi _{C_{t_n}}(x_n) p(x_1, t_1; x_2, t_2; \ldots , x_n, t_n), \end{aligned}$$
(9)

where \(p(x_1, t_1; x_2, t_2; \ldots , x_n, t_n)\) is a probability measure on \(\Pi\), and \(\chi _C\) the characteristic function of the set \(C \subset \Gamma\).

The theory of stochastic processes is an example of a classical history theory. Of particular interest are Markovian processes, which describe probabilistic systems without memory. The associate probability measures are of the form

$$\begin{aligned} p(x_1, t_1; x_2, t_2; \ldots , x_n, t_n) = \rho _{t_1}(x_1) g(x_2, t_2|x_1, t_1) g(x_3, t_3|x_2, t_2) \ldots g(x_n, t_n|x_{n-1}, t_{n-1}) , \end{aligned}$$
(10)

where \(\rho _{t_1}\) is a probability density on \(\Gamma\) at the initial moment of time, and \(g(x_2, t_2|x_1, t_1) > 0\) is the transition matrix between times \(t_1\) and \(t_2\). The transition matrix is stochastic, i.e., it satisfies \(\int dx_2 g(x_2, t_2|x_1, t_1) = 1\).

Deterministic processes (without memory) correspond to the degenerate case where the transition matrix is a delta function, i.e.,

$$\begin{aligned} g(x_2, t_2|x_1, t_1) = \delta [x_2 - f_{t_2, t_1}(x_1)], \end{aligned}$$
(11)

where \(f_{t,t'}\) is a diffeomorphism on \(\Gamma\) indexed by t and \(t'\). For classical mechanics and field theory, see, Ref. [32].

1.2 Quantum Histories Theory

In Copenhagen quantum mechanics, propositions correspond to measurement outcomes. Ideal measurements are described in terms of projection-valued measures (PVMs) on a Hilbert space \(\mathcal{H}\), i.e., a family of projectors \({\hat{P}}_{a}\) indexed by a, that satisfy the following properties: (i) mutual exclusion, \({\hat{P}}_a {\hat{P}}_b = \delta _{ab} {\hat{P}}_a\), and (ii) exhaustion \(\sum _a {\hat{P}}_a = {\hat{I}}\).

A sequence of measurements at times \(t_1, t_2, \ldots , t_n\) corresponds to a sequence of PVMs \({\hat{P}}_{a_i, t_i}\) indexed by the time-parameter. A history of measurement outcomes is a sequence of projectors

$$\begin{aligned} \alpha := ({\hat{P}}_{a_1, t_1}, {\hat{P}}_{a_2, t_2}, \ldots , {\hat{P}}_{a_n, t_n}). \end{aligned}$$
(12)

Suppose that the system is prepared at the state \({\hat{\rho }}_0\) at \(t = 0\), and that its Hamiltonian is \({\hat{H}}\). The probability associated to \(\alpha\) is

$$\begin{aligned} \text{ Prob }(\alpha ) = Tr\left( {\hat{C}}_{\alpha } {\hat{\rho }}_0 {\hat{C}}_{\alpha }^{\dagger }\right) , \end{aligned}$$
(13)

where

$$\begin{aligned} {\hat{C}}_{\alpha } = {\hat{P}}_{a_n, t_n}(t_n) \ldots {\hat{P}}_{a_2, t_2} {\hat{P}}_{a_1, t_1}(t_1) \end{aligned}$$
(14)

is an operator associated to the history \(\alpha\), defined in terms of the Heisenberg-picture projectors \({\hat{P}}_{a_i, t_i}(t_i) = e^{i{\hat{H}}t_i} {\hat{P}}_{a_i, t_i}e^{-i{\hat{H}}t_i}\).

The consistent/decoherent histories interpretation of quantum mechanics starts with the observation that the sequence \(({\hat{P}}_{a_1, t_1}, {\hat{P}}_{a_2, t_2}, \ldots , {\hat{P}}_{a_n, t_n})\) can be interpreted as referring to propositions about properties of a physical system, and not only to measurement outcomes. Then, it is possible to define logical operations, such as \(\text{ AND }\), \(\text{ OR }\), \(\text{ NOT }\) and so on, between those history propositions, and to define a set \(\mathcal{V}\) if history propositions that is closed under those operations.

The set of history propositions is constructed in terms of the history Hilbert space \(\mathcal{K}_{his}\), defined as the tensor product of single time Hilbert spaces,

$$\begin{aligned} \mathcal{K}_{his} = \otimes _{t \in \mathcal{T}} \mathcal{H}_t, \end{aligned}$$
(15)

where t is an element of the time-set \(\mathcal{T}\), and \(\mathcal{H}\) is a copy of the single-time Hilbert space indexed by t. The history (12) is represented by a projection operator \({\hat{E}} = {\hat{P}}_{a_1, t_1} \otimes {\hat{P}}_{a_2, t_2} \otimes \ldots \otimes {\hat{P}}_{a_n, t_n}\) on \(\mathcal{K}_{his}\). General history propositions are represented by projectors on \(\mathcal{K}_{his}\) that are not factorized. Then, the set \(\mathcal{V}\) of history propositions coincides with the lattice \(\Lambda (\mathcal{K}_{his})\) of projection operators on \(\mathcal{K}_{his}\). Hence, the logical operations on \(\mathcal{V}\) coincide with the lattice operations of \(\Lambda (\mathcal{K}_{his})\) [65].

The incorporation of dynamics into the histories description is rather intricate, and it requires a detailed analysis of the symmetries of the formalism. For continuous time, dynamics are implemented through an action operator that is defined on \(\mathcal{K}_{his}\) [31].

The rule (13) does not define a probability measure on the set of history propositions. Probabilities can only be defined with respect to a given context. In the Copenhagen interpretation, the context is given by the measurement set-up, i.e., the choice of the different PVMs at different moments of time.

In decoherent histories, the context is expressed in terms of the abstract concept of a consistent set. To this end, we define the decoherence functional for a pair of histories \(\alpha\) and \(\beta\) of the form (13), as

$$\begin{aligned} d(\alpha , \beta ) = Tr\left( {\hat{C}}_{\alpha } {\hat{\rho }}_0 {\hat{C}}_{\beta }^{\dagger }\right) . \end{aligned}$$
(16)

The decoherence functional is extended to general history propositions by the requirements of

  • linearity: \(d(\alpha \; \text{ OR }\; \gamma , \beta ) = d(\alpha , \beta ) + d(\gamma , \beta )\), for \(\alpha \; \text{ AND }\; \gamma = \emptyset\), and

  • hermiticity: \(d(\alpha , \beta ) = d(\beta , \alpha )^*\).

Then, d defines a bilinear map on \(\mathcal{V} \times \mathcal{V}\).

Consider a set \(\mathcal{W}\) of history propositions that are mutually exclusive and exhaustive. We call \(\mathcal{W}\) a consistent set, if it satisfies the consistency condition

$$\begin{aligned} \text{ Re } \; d(\alpha , \beta ) = 0, \; \alpha \ne \beta , \end{aligned}$$
(17)

for all \(\alpha , \beta \in \mathcal{W}\). Then, the diagonal elements of the decoherence functional define a probability measure on \(\mathcal{S}\),

$$\begin{aligned} \text{Prob}_{\mathcal{W}}(\alpha) = d(\alpha, \alpha),\,\text{for}\,\text{all}\; \; \alpha \in {\mathcal{W}}. \end{aligned}$$
(18)

Hence, in quantum theory we do not have a complete probability function on \(\mathcal{V}\), but a family of partial probability functions \({\text{Prob}}_{\mathcal{W}}\), each corresponding to a different consistent set \(\mathcal{W}\).

Energy Conservation in \(\Psi \Phi\)I Theories

In this section, we explore the issue of energy conservation in \(\Psi \Phi\)I theories. We consider classical processes, because energy conservation is precisely formulated in classical mechanics. However, some of the results can be appropriately generalised also for quantum systems.

1.1 Energy Conserving Dynamics

We first consider deterministic processes. They describe dynamical systems, i.e., systems with time evolution determined by a set of differential equations. These equations can be interpreted as a flow on a differentiable manifold that plays the role of the state space. For MBI models, the state space is of the form \(\Gamma _{\Phi }\times \Gamma _{\Psi }\), where \(\Gamma _{\Phi }\) contains the physical degrees of freedom and \(\Gamma _{\Psi }\) contains the mental degrees of freedom.

We describe the physical degrees of freedom in terms of classical mechanics, so that energy conservation is well-defined. Let us denote the points of \(\Gamma _{\Phi }\) by \(\xi ^a\), for some discrete index a. Time evolution is given by Hamilton’s equation

$$\begin{aligned} {\dot{\xi }}^a = \mathcal{P}^{ab} \partial _bH, \end{aligned}$$
(19)

where H is the Hamiltonian, i.e., a function on \(\Gamma _{\Phi }\) whose values correspond to the usual notion of energy.

The Poisson tensor \(\mathcal{P}^{ab}\) in Eq. (19) is antisymmetric, non-degenerate and satisfies Jacobi’s identity

$$\begin{aligned} \mathcal{P}^{ad}\partial _d\mathcal{P}^{bc} + \mathcal{P}^{cd}\partial _d\mathcal{P}^{ab} + \mathcal{P}^{bd}\partial _d\mathcal{P}^{ca} = 0. \end{aligned}$$
(20)

The antisymmetry of \(\mathcal{P}\) guarantees the conservation of energy, since

$$\begin{aligned} {\dot{H}} = \partial _aH {\dot{\xi }}^a = \mathcal{P}^{ab} \partial _a H \partial _bH = 0. \end{aligned}$$
(21)

The most general dynamical system on \(\Gamma _{\Phi }\times \Gamma _{\Psi }\), compatible with Eq. (19), is of the form

$$\begin{aligned} {\dot{\xi }}^a= & {} \mathcal{P}^{ab}(\xi ) \partial _bH(\xi ) + G^a(\xi , y) \end{aligned}$$
(22)
$$\begin{aligned} {\dot{y}}^i= & {} F^i(y) + J^i(\xi , y), \end{aligned}$$
(23)

where we denoted the points of \(\Gamma _{\Psi }\) by \(y^i\); i is a discrete index that labels the mental degrees of freedom. The vector \(F^i\) on \(\Gamma _{\Psi }\) describes self-dynamics on \(\Gamma _{\Psi }\). Interaction is described by the vector fields \(G^a\) on \(\Gamma _{\Phi }\) and \(J^i\) on \(\Gamma _{\Psi }\).

Eq. (22) implies that

$$\begin{aligned} {\dot{H}} = G^a \partial _a H. \end{aligned}$$
(24)

Hence, energy is conserved if \(G^a \partial _a H = 0\), i.e., if \(G^a\) is tangent to the surfaces of constant energy. To motivate our subsequent analysis, let us first assume that physical degrees of freedom satisfy some form of Hamilton equations, even in the presence of MBI. This implies that \(G^a\) is a Hamiltonian vector field for each y, i.e., that

$$\begin{aligned} G^a(\xi , y) = \mathcal{P}^{ab} \partial _b S(\xi , y) \end{aligned}$$
(25)

for some scalar function S. Hence, energy is conserved if \(\mathcal{P}^{ab}\partial _aS \partial _bH = 0\).

A Hamiltonian system with a large number of degrees of freedom may have constants of the motion other than the Hamiltonian. However, it is highly implausible that we can ever relate such constants, when they are defined for very different systems, for example, a room with one person inside and a concert hall with one thousand persons.

We expect that energy-conserving dynamics of sufficient generality are possible only if S depends on \(\xi\) solely through the Hamiltonian, i.,e., if \(S(\xi , y) = \phi (H(\xi ), y)\) for some function \(\phi : {\pmb R} \times \Gamma _{\Psi }\rightarrow {\pmb R}\). Then, the evolution equation on \(\Gamma _{\Phi }\) becomes

$$\begin{aligned} {\dot{\xi }}^a = \tilde{\mathcal{P}}^{ab}( \xi , y) \partial _bH(\xi ) \end{aligned}$$
(26)

where \(\tilde{\mathcal{P}}^{ab}(\xi , y) = [1 + \phi '[H(\xi ), y]]\mathcal{P}^{ab}(\xi )\), and the prime denotes the derivative of \(\phi\) with respect to its first argument. Hence, energy is conserved if the mental degrees of freedom are coupled to the physical degrees of freedom through an antisymmetric tensor \(\tilde{\mathcal{P}}^{ab}\).

The tensor \(\tilde{\mathcal{P}}^{ab}\) is not an actual Poisson tensor, because it does not satisfy the Jacobi identity, Eq. (20). However, a small modification of our previous analysis can justify a Poisson tensor \(\tilde{\mathcal{P}}^{ab}\) in Eq. (26), leading to a \(\Psi \Phi\)I theory with a much more interesting mathematical structure.

To this end, let us assume that there is a submanifold \(\Omega \subset \Gamma _{\Psi }\) that corresponds to the mental vacuum of Sect. 5.2, i.e., if \(y \in \Omega\) then no mental processes are present. Consider now a Poisson tensor \(\tilde{\mathcal{P}}^{ab}\) that satisfies \(\tilde{\mathcal{P}}^{ab} = \mathcal{P}^{ab}\), for any \(y \in \Omega\). Then, Eqs. (26) and (23) define a dynamical system for MBI that reduces to the classical equations of motions in absence of mental processes. Hence, they define a deterministic \(\Psi \Phi\)I theory with energy conservation.

The difference from our previous analysis is that \(G^a\) is not a Hamiltonian vector field, i.e., Eq. (25) does not hold. Instead, \(G^a = \Delta \mathcal{P}^{ab} \partial _b H\), where \(\Delta \mathcal{P}^{ab} = \tilde{\mathcal{P}}^{ab} - \mathcal{P}^{ab}\).

The implementation of energy conservation constraints only one of the two equations that describe the dynamical system, namely, Eq. (26) Thus, it persists even if the term \(F^i(y)\) of Eq. (23) is a ‘random force’, i.e., if the values of \(F^i\) at each moment of time are distributed probabilistically. Thus, energy conservation also applies to stochastic \(\Psi \Phi\)I theories—in fact it is completely insensitive to the dynamics of the mental degrees of freedom—, as long as Eq. (26) is satisfied.

The same structure can be employed in order to define energy-conserving dynamics for MBI theories, where matter is treated quantum mechanically. Schrödinger’s equation on a Hilbert space \(\mathcal{H}\) is equivalent to Hamilton’s equation on the projective Hilbert space \(\mathcal{PH}\) [66]. Hence, the above analysis can straightforwardly be transferred into a quantum context.

1.2 Generalised Energy

In mechanical systems, energy conservation arises as a consequence of the time-translation symmetry, i.e., the fact that dynamics are unchanged under a transformation that moves the time of the various events. It is plausible that the time translation symmetry of the extended dynamics (2223) corresponds to a conserved quantity \(K(\xi , y)\) that reduces to the Hamiltonian H in a limit where the mind–body interaction term can be ignored. For example, \(K(\xi , y)\) may be of the form \(L(y) + H(\xi )\), where L(y) is invariant under the self dynamics of the mental degrees of freedom, \(F^i \partial L = 0\). Then,

$$\begin{aligned} {\dot{K}} = G^a \partial _a H + J^i\partial _i L, \end{aligned}$$
(27)

and the conservation of K leads to a relation between the coupling terms \(G^a\) and \(J^i\): \(G^a \partial _a H + J^i\partial _i L = 0\). If this condition holds, the Hamiltonian is part of a more general conserved quantity K. Since K is assumed to follow from the time-translation symmetry of the dynamics, its values should be identified with energy. Thus, L is a new form of energy associated to mental processes, and the notion of energy has to be extended in order to account for mind–body interactions.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Anastopoulos, C. Mind–Body Interaction and Modern Physics. Found Phys 51, 65 (2021). https://doi.org/10.1007/s10701-021-00472-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10701-021-00472-7

Navigation