Skip to main content

Information-Based Physics and the Influence Network

  • Chapter
  • First Online:
It From Bit or Bit From It?

Part of the book series: The Frontiers Collection ((FRONTCOLL))

Abstract

I know about the universe because it influences me. Light excites the photoreceptors in my eyes, surfaces apply pressure to my touch receptors and my eardrums are buffeted by relentless waves of air molecules. My entire sensorium is excited by all that surrounds me. These experiences are all I have ever known, and for this reason, they comprise my reality. This essay considers a simple model of observers that are influenced by the world around them. Consistent quantification of information about such influences results in a great deal of familiar physics. The end result is a new perspective on relativistic quantum mechanics, which includes both a way of conceiving of spacetime as well as particle “properties” that may be amenable to a unification of quantum mechanics and gravity. Rather than thinking about the universe as a computer, perhaps it is more accurate to think about it as a network of influences where the laws of physics derive from both consistent descriptions and optimal information-based inferences made by embedded observers.

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

Access this chapter

eBook
USD 16.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 99.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Similar content being viewed by others

Notes

  1. 1.

    You may not like these assumptions—feel free to try others! For now, let’s see what physics these give rise to.

  2. 2.

    We are not going to worry whether an event on the observer chain constitutes a measurement or detection.

  3. 3.

    The signature of the metric, which determines where the minus sign goes, is in agreement with the particle physics tradition and opposite to that used in general relativity where one writes \(\Delta s^{2}=-\Delta t^{2}+\Delta x^{2}\). Here the signature is not arbitrary since the minus sign comes from the fact that the interval between chains is quantified by a pair that has opposite signs. Later, this gives rise to the mass-energy-momentum relation with the correct signature.

References

  1. B.A. Davey, H.A. Priestley, Introduction to Lattices and Order (Cambridge University Press, Cambridge, 2002)

    Book  MATH  Google Scholar 

  2. L. Bombelli, J.-H. Lee, D. Meyer, R. Sorkin, Space-time as a causal set. Phys. Rev. Lett. 59, 521–524 (1987)

    Article  ADS  MathSciNet  Google Scholar 

  3. K.H. Knuth, N. Bahreyni, A potential foundation for emergent space-time J. Math. Phys. 55, 112501 (2014). http://dx.doi.org/10.1063/1.4899081, arXiv:1209.0881 [math-ph]

  4. H. Bondi, Relativity and Common Sense (Dover, New York, 1980)

    Google Scholar 

  5. S. Sontag, At the Same Time: Essays and Speeches (Farrar Straus and Giroux, New York, 2007)

    Google Scholar 

  6. K. Huang, On the Zitterbewegung of the Dirac electron. Am. J. Phys. 20, 479 (1952)

    Google Scholar 

  7. D. Hestenes, The Zitterbewegung interpretation of quantum mechanics. Found. Phys. 20(10), 1213–1232 (1990)

    Google Scholar 

  8. D. Hestenes, Electron time, mass and Zitter. The Nature of Time Essay Contest, Foundational Questions Institute (2008)

    Google Scholar 

  9. E. Schrödinger, Über die kräftefreie Bewegung in der relativistischen Quantenmechanik (Akademie der wissenschaften in kommission bei W. de Gruyter u, Company, Berlin, 1930)

    Google Scholar 

  10. P. Catillon, N. Cue, M.J. Gaillard, R. Genre, M. Gouanère, R.G. Kirsch, M. Spighel, A search for the de Broglie particle internal clock by means of electron channeling. Found. Phys. 38(7), 659–664 (2008)

    Article  ADS  Google Scholar 

  11. R. Gerritsma, G. Kirchmair, F. Zähringer, E. Solano, R. Blatt, C.F. Roos, Quantum simulation of the Dirac equation. Nature 463(7277), 68–71 (2010)

    Article  ADS  Google Scholar 

  12. K.H. Knuth, J. Skilling, Foundations of inference. Axioms 1, 38–73 (2012)

    Article  MATH  Google Scholar 

  13. K.H. Knuth, Information-based physics: an observer-centric foundation. Contemp. Phys. 55(1), 12–32 (2014). doi:10.1080/00107514.2013.853426. arXiv:1310.1667 [quant-ph]

    Article  ADS  Google Scholar 

  14. P. Goyal, K.H. Knuth, J. Skilling, Why quantum theory is complex. Phys. Rev. A 81, 022109 (2010). arXiv:0907.0909v3 [quant-ph]

    Article  ADS  Google Scholar 

  15. P. Goyal, K.H. Knuth, Quantum theory and probability theory: their relationship and origin in symmetry. Symmetry 3(2), 171–206 (2011)

    Article  MathSciNet  Google Scholar 

  16. R.P. Feynman, A.R. Hibbs, Quantum Mechanics and Path Integrals (McGraw-Hill, New York, 1965)

    MATH  Google Scholar 

  17. J. Aczel, Lectures on Functional Equations and Their Applications (Academic Press, New York, 1966)

    MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kevin H. Knuth .

Editor information

Editors and Affiliations

Appendix

Appendix

The key idea behind employing coordinated chains is that they provide a means of delineating a specific 1 \(+\) 1-dimensional subspace in the non-dimensional poset. Events are defined to lie within the subspace defined by the two coordinated chains if the projection of the event onto one chain can be found by first projecting the event onto the other chain and then back to the first. This leads to a set of algebraic relations (for example\(Px=\overline{{P}}Qx\) and \(Qx=Q\overline{{P}}x)\) where we consider the projections of event x onto chains P and Q. This in turn leads to several different relationships between an event x and the pair of chains (for example, x can be on the P-side of PQ, the Q-side of PQ or between PQ). For example, we say that event x is between two coordinated chains P and Q if \(Px=P\overline{{Q}}x\), \(\overline{{P}}x=\overline{{P}}Qx\), \(\overline{{P}}x=\overline{{P}}Qx\), and \(\overline{{Q}}x=\overline{{Q}}Px\) [3].

Fig. 6.6
figure 6

Illustrates two consistently related chains. Chains Q and Q’ are omitted. By coordination we have \(\Delta \overline{{p}}=\Delta q\) and \(\Delta \overline{{p}}^{\prime }=\Delta q^{\prime }\)

The derivation of the scalar measure is based on a consistency requirement that any two chains that agree on the lengths of each others intervals (coordination) must agree on the lengths of every interval that both chains can quantify. We assume that the scalar measure is a non-trivial symmetric function of the pairwise measure. That is, \(s=\sigma (\Delta p,\Delta q)=\sigma (\Delta q,\Delta p)\), where \(\sigma (\cdot , \cdot )\)is a function to be determined. We can change our units of measure, so that we have \(\alpha s=\sigma (\alpha \Delta p,\alpha \Delta q)\). This is a special case of the homogeneity equation [17]

$$\begin{aligned} F(zx,zy)=z^{k}F(x,y) \end{aligned}$$
(6.11)

where in our problem the parameter \(k = 1\). The general symmetric solution is given by \(F(x,y)=\sqrt{xy}\;h(x/y)\), where h is an arbitrary function symmetric with respect to interchange of x and y. We can show that the function h is unity, and that lengths of intervals are given by \(\sqrt{\Delta p\Delta q}\), which leads to the interval scalar \(\Delta s^{2}=\Delta p\Delta q\) [3]. We can next consider chains that are consistently related where every interval of length \(\Delta p=k\) on chain P forward projects to an interval of length \(\Delta p^{\prime }=m\) on \(P^{\prime }\) and forward projects to an interval of length \(\Delta q^{\prime }=n\) on Q’. We now want to find a function L that takes the pair quantification of the interval I in the PQ frame to the P’Q’ frame:\(L_{PQ\rightarrow P^{\prime }Q^{\prime }} (\Delta p,\Delta q)_{PQ} =(\Delta p^{\prime },\Delta q^{\prime })_{P^{\prime }Q^{\prime }} \). (see Fig. 6.6). We note that we can write the projections of the interval I onto chain P in units of length k, so that the pairs can be written as \((\Delta p,\Delta q)=(\alpha k,\beta k)\)and \((\Delta p^{\prime },\Delta q^{\prime })=(\alpha m,\beta n)\). Preserving the scalar measure gives \(k^{2}=mn\) so that \(L_{PQ\rightarrow P^{\prime }Q^{\prime }} (\alpha k,\beta k)_{PQ}=(\alpha m,\beta n)_{P^{\prime }Q^{\prime }} \). It can then be shown that the general transform is given by \(L_{PQ\rightarrow P^{\prime }Q^{\prime }} (x,y)_{PQ}=(x\sqrt{m/n},y\sqrt{n/m})_{P^{\prime }Q^{\prime }} \) [3], which gives rise to the Lorentz transformation in (6.6) and (6.7).

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Knuth, K.H. (2015). Information-Based Physics and the Influence Network. In: Aguirre, A., Foster, B., Merali, Z. (eds) It From Bit or Bit From It?. The Frontiers Collection. Springer, Cham. https://doi.org/10.1007/978-3-319-12946-4_6

Download citation

Publish with us

Policies and ethics