Skip to main content

Quantum Field Theory

  • Chapter
  • First Online:
Book cover The Philosophy of Quantum Physics

Abstract

Quantum field theory (QFT) shares many of its philosophical problems with quantum mechanics. This applies in particular to the quantum measurement process and the connected interpretive problems, to which QFT contributes hardly any new aspects, let alone solutions. The question as to how the objects described by the theory are spatially embedded was already also discussed for quantum mechanics. However, the new mathematical structure of QFT promises new answers, which renders the spatiotemporal interpretation of QFT the pivotal question. In this chapter, we sketch the mathematical characteristics of QFT and show that a particle as well as a field interpretation breaks down.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 44.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 59.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 64.99
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

Notes

  1. 1.

    See, e.g. the 1925 Como lecture of Bohr (1961), in particular p. 54.

  2. 2.

    A note for more advanced readers: A manifestly relativistically invariant notation is also possible. Furthermore, we should also mention that the commutation relations (6.3) are valid for bosonic fields only, such as the electromagnetic field in particular. In the framework of QFT, however, one can in addition to these interaction fields also describe material “particles”, with half-integer spin, in terms of fields. For such fermionic fields, e.g. the Dirac field for electrons, one requires anti-commutation relations instead of Eq. 6.3. In the following, we also make use of the so-called Heisenberg picture, i.e. we will be working with time-dependent operators.

  3. 3.

    The Schrödinger equation violates the requirement of special relativity which states that the laws of nature must maintain their form when one goes via Lorentz transformations from the inertial frame of reference of one observer to that of another. The Maxwell equations, for example, fulfil this requirement: light, in particular, has the same velocity c in vacuum for all these observers.

  4. 4.

    As usual in QFT, we will call it the “free theory” in the following.

  5. 5.

    In particular, there are solutions with negative energies, which would lead to endless cascades towards energetically more favourable states at lower energies. In retrospect, one can argue that it cannot be expected that relativistic processes could be described by a single-particle theory, since the energy-mass equivalence \(E=mc^2\) in special relativity permits the formation of particle-antiparticle pairs (Peskin and Schroeder 1995, Chap. 2). An additional problem consists in the fact that the normalization of the states determined by the Klein–Gordon equation is no longer time-independent, which undermines their interpretation as probability densities. The Klein–Gordon equation indeed fulfils the requirements of special relativity, but not those of quantum mechanics (Srednicki 2007, Chap. 1).

  6. 6.

    Peskin and Schroeder (1995), Footnote on p. 19, and Redhead (1988).

  7. 7.

    In the Lagrange formalism, the relevant equation of motion (for fields the field equation) can be derived by subjecting the Lagrangian density to a variation procedure which is expressed by the Euler–Lagrange equation (or more precisely: which leads to that equation). Thus, for example, inserting the Lagrangian density of electrodynamics into the Euler–Lagrange equation permits one to derive the Maxwell equations.

  8. 8.

    The frequency \(\omega \) = 2\(\pi \)\(\nu \) is connected with the momentum via \(\hbar \omega _\mathbf {p} = \sqrt{|\mathbf {{p}}|^2c^2 + m^2c^4}\). In many textbooks on QFT, the solutions of the Klein–Gordon equation are not formulated in terms of the momenta \(\mathbf {{p}}\), but rather using the wavevectors \(\mathbf {{k}}\), which are related via \(\mathbf {{p}} = \hbar \mathbf {{k}}\) (and, if we use the above-mentioned unit convention \(\hbar =1\), are in fact identical).

  9. 9.

    The “0” on the right-hand side denotes the null vector, which should not be confused with the vacuum state \( | 0 \rangle \)!

  10. 10.

    In fact, at this point we are already working with the so-called Fock space representation of the commutation relations, which we will introduce systematically in the next Sect. 6.3.3.

  11. 11.

    For the relationship between commutation relations and the space of states, cf. Mandl and Shaw (2010), Sects. 1.2.2 and 3.1.

  12. 12.

    The possibility of describing variable numbers of particles does not entail that fundamental interactions themselves are described. We will continue to work with the so-called free theory from the previous section. Creation and annihilation operators do not describe dynamic processes in that theory. In fact, Haag’s theorem even says that the description of interactions is generally excluded within the framework of this theory. This limitation to a free theory has important consequences for its interpretation (see Sect. 6.4.2). For the actual treatment of scattering processes, this is not as important as it might seem, since there we are mainly dealing with asymptotically free states, i.e. far from the scattering process, and this “far from the scattering process” is attained almost immediately following the interaction.

  13. 13.

    This alternative means that the ontological significance of the states created is here not (yet) determined.

  14. 14.

    Since, according to Eq. 6.9, \(a^{\dagger } ({\mathbf {p}})\) and \(a^{\dagger } ({\mathbf {p'}})\) commute, the two-particle state \(a^{\dagger } ({\mathbf {p}}) a^{\dagger } (\mathbf{p'}) | 0 \rangle \) is identical to the state \(a^{\dagger } ({\mathbf {p'}}) a^{\dagger } ({\mathbf {p}}) | 0 \rangle \) with exchanged creation operators. Many-particle states are thus symmetric under permutations of the creation operators. Furthermore, arbitrarily many “Klein–Gordon particles” can be created with the same momentum or in the same field mode \({\mathbf {p}}\). As we have seen in Chap. 3, this means that “Klein–Gordon particles” must be bosons.

  15. 15.

    Of course there are also systems in quantum physics with distinguishable particles, namely many-particle systems with different types of particles. These are also described by the Hilbert-space formalism and therefore also permit the typical superpositions.

  16. 16.

    We use here \( {h}_1 \otimes _{s} {h}_2 \equiv {h}_1 \otimes {h}_2 + {h}_2 \otimes {h}_1\).

  17. 17.

    For a discussion of the significance of Feynman diagrams, see Kuhlmann (2010), Sects. 10.3–10.4, and Wüthrich (2012).

  18. 18.

    Since most operators are differential operators in Schrödinger’s version of quantum mechanics, we called this approach the “calculus approach”.

  19. 19.

    In the following, some still-unexplained concepts will occasionally appear in the text, and they will be placed in quotation marks or parentheses, either to ensure that the statements made are not incorrect, and/or to guarantee to readers who wish to pursue their interests further in the literature that they will be able to follow the concepts employed there. In a first reading of this chapter, these concepts can be ignored, however.

  20. 20.

    See Sect. 3.1.4.

  21. 21.

    One notorious example for the significance of inequivalent representations is the so-called Unruh effect, according to which what appears to be a vacuum for one observer takes the form of a thermal bath of particles for another, accelerated observer. The deeper reason for this apparent paradox is that different inequivalent representations are associated with the two observers, which means among other things that they experience different vacuum states. In fact, different inequivalent representations are even systematically related to different vacuum states. This is the basis of the so-called GNS construction, which plays an important role in the relation between AQFT and conventional QFT—because, beginning with observable algebras, different operator representations lead to different vacuum states.

  22. 22.

    Ruetsche (2003) discusses the philosophical consequences of this in detail.

  23. 23.

    Haag (1996) offers a complete treatment of AQFT. Buchholz (2000) emphasizes the basic features of AQFT. Halvorson and Müger (2007) give a medium-long introduction, which is directed especially to philosophers of physics.

  24. 24.

    This requirement does not contradict the possibility of non-local EPR correlations (see Chap. 4).

  25. 25.

    See e.g. Redhead (1995) , Halvorson and Clifton (2002), Earman und Fraser (2006) , Baker (2009), Kuhlmann (2010), and Ruetsche (2011).

  26. 26.

    A compact treatment of the various deficits of conventional QFT can be found in Kuhlmann (2012), Sect. 4.1.

  27. 27.

    An overview of the various interpretations of QFT can be found in Kuhlmann 2012, Sect. 5.1.2.

  28. 28.

    Wigner’s group-theoretical classification of the elementary particles (Wigner 1939) also gives no definition of the particle concept, in contrast to what is often assumed. What Wigner defines instead is “elementarity” (see Kuhlmann 2010, Sect. 8.1.2). This can be seen already by the fact that spatial localizability plays no role in Wigner’s definition.

  29. 29.

    Relativistic classical particles must obey the energy condition expressed by Eq. (6.4), owing to the equivalence of mass and energy.

  30. 30.

    Instead of “ordinally countable” sometimes one simply says “countable” and contrasts it with “aggregable” (here: “cardinally countable”). Teller (1995) in particular uses alternative terminology.

  31. 31.

    Fraser (2008) expresses this as follows: The countability and energy conditions are fulfilled. However, since countability is often understood in the ordinal sense—-for example by Teller —but in the present connection we are concerned with the cardinal sense, we speak instead of “discreteness”.

  32. 32.

    Baker (2013, p. 267) argues conversely that the situation could be similar to that of atoms in superposition states, which nevertheless does not cause us to doubt the existence of atoms.

  33. 33.

    Teller (1995) argues in opposition to this view that probabilistic statements are a generic feature of quantum physics, and that both problems can be resolved with a propensity interpretation of quantum-mechanical probabilities (see Sect.2.2.2). Teller (1995) discusses the first problem on pp. 31–33 and the second one on pp. 110–112. An accessible treatment and critical discussion of Teller’s arguments is given by Huggett and Weingard (1996) .

  34. 34.

    Note that Teller presumes that it is not reasonable to assume quantum objects to be primitively individuated. However, as we have seen from the discussion on the possibility of weak distinguishability in Sect. 3.2.3, this presumption has been frequently criticized, especially in recent years.

  35. 35.

    A simple example can be found in Huggett and Weingard (1996 p. 306) . Furthermore, in QFT on curved spacetimes, there are also infinitely many different inequivalent Fock space representations (Baker 2013).

  36. 36.

    This holds in spite of the fact that the Fock space is used in perturbation-theoretical calculations within conventional quantum mechanics.

  37. 37.

    Bain (2011) has formulated an alternative quanta interpretation for the asymptotically free theory.

  38. 38.

    According to Baker (2013) , the permutations refer to the order in which the charges are added (to “algebraic states”).

  39. 39.

    This was formulated somewhat more cautiously than in many other textbooks by Peskin and Schroeder (1995, p. 22): “It is quite natural to call these excitations particles, since they are discrete entities that have the proper relativistic energy-momentum relation. (By a particle, we do not mean something that must be localized in space; \(a_\mathbf{p}^{\dagger }\) creates particles in momentum eigenstates)”.

  40. 40.

    For this topic and other proofs, cf. Halvorson and Clifton (2002), Kuhlmann (2010), Chap. 8, and Kuhlmann (2012), Sect. 5.3.

  41. 41.

    In the mathematically correct algebraic formulation of QFT (see Sect. 6.3.5), the quantities are not associated with points, but rather with finite regions of spacetime; furthermore, they are not associated with individual operators, but with algebras of operators. For the following arguments, this makes no essential difference, however.

  42. 42.

    Baker (2009) has analysed some additional difficulties of a field interpretation.

  43. 43.

    Wayne (2008) , Morganti (2009) , and Kuhlmann (2010).

  44. 44.

    Maurin (2013) gives an up-to-date overview.

  45. 45.

    One concern is the boundary problem (Campbell 1990): Consider a blue sheet of paper that you tear apart. Are there suddenly two blue tropes now? Or were they “in” the original sheet already. If yes, then we have acquired a general problem: where does one colour trope end and another one begin on a macroscopic level?.

  46. 46.

    The term “particular” is less misleading here than “individual thing”, since single tropes are precisely not things.

  47. 47.

    Kuhlmann (2010), Chaps. 11–15.

  48. 48.

    Rossanese (2013) argues that the idea that the Unruh effect undermines the particle interpretation is not only transferable to the field interpretation (as asserted by Baker 2009 ), but may also be raised against a trope-ontological interpretation.

  49. 49.

    One proposed solution is the “Swiss army approach” by Ruetsche (2011) . However, it is not clear whether the problem is actually solved by such a multi-perspective approach, rather than merely explicitly formulated.

References

  • Audretsch, Jürgen. 1989. Vorläufige Physik und andere pragmatische Elemente physikalischer Naturerkenntnis. In Pragmatik. Handbuch des pragmatischen Denkens, vol. III, ed. H. Stachowiak, 373–392. Hamburg: Meiner.

    Google Scholar 

  • Bain, Jonathan. 2011. Quantum field theories in classical spacetimes and particles. Studies in History and Philosophy of Modern Physics 42: 98–106.

    Article  ADS  MathSciNet  Google Scholar 

  • Baker, David J. 2009. Against field interpretations of quantum field theory. British Journal for the Philosophy of Science 60: 585–609.

    Article  MathSciNet  Google Scholar 

  • Baker, David J. 2013. Identity, superselection theory, and the statistical properties of quantum fields. Philosophy of Science 80: 262–285.

    Article  MathSciNet  Google Scholar 

  • Bohr, Niels. 1961. The quantum postulate and the recent development of atomic theory. In Atomic Theory and the Description of Nature, ed. Niels Bohr, 52–91. Cambridge: Cambridge University Press.

    Google Scholar 

  • Brown, Harvey R., and Rom Harré (eds.). 1988. Philosophical Foundations of Quantum Field Theory. Oxford: Clarendon Press.

    MATH  Google Scholar 

  • Buchholz, Detlev. 2000. Current trends in axiomatic qantum field theory. In Quantum Field Theory. Proceedings of the Ringberg Workshop Held at Tegernsee, Germany, 21–24 June 1998 On the Occasion of Wolfhart Zimmermann’s 70th Birthday, vol. 558, ed. P. Breitenlohner, and D. Maison, 43–64, Lecture Notes in Physics Berlin, Heidelberg: Springer.

    Chapter  Google Scholar 

  • Butterfield, Jeremy, and Hans Halvorson (eds.). 2004. Quantum Entanglements - Selected Papers of Rob Clifton. Oxford: Oxford University Press.

    Google Scholar 

  • Campbell, Keith. 1990. Abstract Particulars. Oxford: Blackwell.

    Google Scholar 

  • Cao, Tian Y. 2010. From Current Algebra to Quantum Chromodynamics: A Case for Structural Realism. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Earman, John, and Doreen Fraser. 2006. Haag’s theorem and its implications for the foundations of quantum field theory. Erkenntnis 64: 305–344.

    Article  MathSciNet  Google Scholar 

  • Einstein, Albert. 1905. Über einen die Erzeugung und Verwandlung des Lichtes betreffenden heuristischen Gesichtspunkt. [On a heuristic point of view about the creation and conversion of light]. Annalen der Physik 17: 132–148.

    Article  ADS  Google Scholar 

  • Esfeld, Michael. 2011. Ontic structural realism as a metaphysics of objects. In Scientific Structuralism, ed. A. and P. Bukolich, Chap. 8. Dordrecht: Springer.

    Google Scholar 

  • Falkenburg, Brigitte. 2012. Was sind subatomare Teilchen? In Philosophie der Physik, ed. M. Esfeld, 158–184. Frankfurt: Suhrkamp.

    Google Scholar 

  • Fraser, Doreen. 2008. The fate of “particles” in quantum field theories with interactions. Studies in History and Philosophy of Modern Physics 39: 841–59.

    Article  ADS  MathSciNet  Google Scholar 

  • Haag, Rudolf. 1996. Local Quantum Physics: Fields, Particles, Algebras, 2nd ed. Berlin: Springer.

    Book  Google Scholar 

  • Halvorson, Hans and Rob Clifton (2002). No place for particles in relativistic quantum theories? Philosophy of Science 69, 1–28; See also in Butterfield and Halvorson (2004) and in Kuhlmann et al. (2002).

    Article  MathSciNet  Google Scholar 

  • Halvorson, Hans, and Michael Müger. 2007. Algebraic quantum field theory (with an appendix by Michael Müger). In Handbook of the Philosophy of Physics - Part A, ed. Jeremy Butterfield, and John Earman, 731–922. Amsterdam: Elsevier.

    Chapter  Google Scholar 

  • Huggett, Nick. 2000. Philosophical foundations of quantum field theory. The British Journal for the Philosophy Science 51: 617–637.

    Article  MathSciNet  Google Scholar 

  • Huggett, Nick, and Robert Weingard. 1996. Critical review: Paul Teller’s interpretive introduction to quantum field theory. Philosophy of Science 63: 302–314.

    Article  MathSciNet  Google Scholar 

  • Jammer, Max. 1966. The Conceptual Development of Quantum Mechanics. New York: McGraw Hill.

    Google Scholar 

  • Kuhlmann, Meinard. 2010. The Ultimate Constituents of the Material World - In Search of an Ontology for Fundamental Physics. Frankfurt: ontos Verlag.

    Book  Google Scholar 

  • Kuhlmann, Meinard (2012). Quantum Field Theory. In: E.N. Zalta (ed.). The Stanford Encyclopedia of Philosophy (Winter 2012).

    Google Scholar 

  • Kuhlmann, Meinard, Holger Lyre, and Andrew Wayne (eds.). 2002. Ontological Aspects of Quantum Field Theory. New Jersey: World Scientific.

    MATH  Google Scholar 

  • Ladyman, James. 1998. What is structural realism? Studies in History and Philosophy of Science 29: 409–424.

    Article  Google Scholar 

  • Lyre, Holger. 2012. Symmetrien, Strukturen, Realismus. In Philosophie der Physik, ed. M. Esfeld, 368–389. Frankfurt: Suhrkamp.

    Google Scholar 

  • Malament, David. 1996. In defense of dogma: Why there cannot be a relativistic quantum mechanics of (localizable) particles. In Perspectives on Quantum Reality: Non-Relativistic, Relativistic, and Field-Theoretic, ed. R. Clifton, 1–10. The University of Western Ontario Series in Philosophy of Science. Dordrecht: Kluwer Academic Publishers.

    Google Scholar 

  • Mandl, Franz, and Graham Shaw. 2010. Quantum Field Theory, 2nd ed. Chichester: Wiley.

    MATH  Google Scholar 

  • Maurin, Anna-Sofia (2013). Tropes. In: E. N. Zalta (ed.). The Stanford Encyclopedia of Philosophy (Fall 2013).

    Google Scholar 

  • Morganti, Matteo. 2009. Tropes and physics. Grazer Philosophische Studien 78: 185–205.

    Google Scholar 

  • Peskin, Michael E., and Daniel V. Schroeder. 1995. Introduction to Quantum Field Theory. Cambridge MA: Perseus Books.

    Google Scholar 

  • Redhead, Michael L.G. 1988. A philosopher looks at quantum field theory. In Brown and Harré 1988: 9–23.

    Google Scholar 

  • Redhead, Michael L.G. 1995. More ado about nothing. Foundations of Physics 25: 123–137.

    Article  ADS  MathSciNet  Google Scholar 

  • Reichenbach, Hans (1955). Die philosophische Bedeutung der Wellen-Korpuskel-Dualität. In Louis de Broglie und die Physiker, 79–94. Hamburg: Claassen Verlag. [Engl. translation: The philosophical significance of wave-particle dualism. In Selected Writings: 1909–1953, 279-289 (ed. by M. Reichenbach and R.S. Cohen). Dordrecht-Boston 1978: Reidel.]

    Google Scholar 

  • Rossanese, Emanuele. 2013. Trope ontology and algebraic quantum field theory: An evaluation of Kuhlmann’s proposal. Studies In History and Philosophy of Modern Physics 44: 417–423.

    Article  ADS  MathSciNet  Google Scholar 

  • Ruetsche, Laura. 2003. A matter of degree: Putting unitary inequivalence to work. Philosophy of Science 70: 1329–1342.

    Article  MathSciNet  Google Scholar 

  • Ruetsche, Laura. 2011. Interpreting Quantum theories: The Art of the Possible. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Ryder, Lewis H. 1996. Quantum Field Theory, 2nd ed. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Simons, Peter M. 1994. Particulars in particular clothing: Three trope theories of substance. Philosophy and Phenomenological Research LIV (3): 553–575.

    Article  Google Scholar 

  • Srednicki, Mark. 2007. Quantum Field Theory. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Stachel, John. 2002. ‘The relations between things’ versus ‘the things between relations’: The deeper meaning of the hole argument. In Reading Natural Philosophy: Essays in the History and Philosophy of Science and Mathematics, ed. D.B. Malament, 231–266. La Salle: Open Court.

    Google Scholar 

  • Teller, Paul. 1995. An Interpretive Introduction to Quantum Field Theory. Princeton: Princeton University Press.

    MATH  Google Scholar 

  • Wayne, Andrew. 2008. A trope-bundle ontology for field theory. In The Ontology of Spacetime II. Philosophy and Foundations of Physics 4, ed. D. Dieks, 1–15. Amsterdam: Elsevier.

    Google Scholar 

  • Wigner, Eugene P. 1939. On unitary representations of the inhomogeneous Lorentz group. Annals of Mathematics 40: 149–204.

    Article  ADS  MathSciNet  Google Scholar 

  • Wüthrich, Adrian. 2012. Zur Anwendung und Interpretation der Feynman-Diagramme. In Philosophie der Physik, ed. M. Esfeld, 227–246. Frankfurt: Suhrkamp.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Exercises

Exercises

  1. 1.

    Gather information about theories of light from the history of physics. Why was Newton’s theory of light not considered satisfactory? Compare the mathematical descriptions of particles and of waves. What is, in your opinion, the principal difference?

  2. 2.

    Is there anything in classical physics which is neither a particle nor a field?

  3. 3.

    State two arguments which in your opinion provide the strongest support for considering quantum field theory to be a theory of particles. What objections can be raised against these arguments?

  4. 4.

    Would it be helpful in your opinion to extend the concepts of “particles” and “fields”, i.e. by referring not only to those entities as “particles” or “fields” which exhibit all the characteristics of classical particles or classical fields?

Rights and permissions

Reprints and permissions

Copyright information

© 2018 Springer International Publishing AG, part of Springer Nature

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Kuhlmann, M., Stöckler, M. (2018). Quantum Field Theory. In: The Philosophy of Quantum Physics. Springer, Cham. https://doi.org/10.1007/978-3-319-78356-7_6

Download citation

Publish with us

Policies and ethics