Skip to main content
Log in

Understanding physics: ‘What?’, ‘Why?’, and ‘How?’

  • Paper in General Philosophy of Science
  • Published:
European Journal for Philosophy of Science Aims and scope Submit manuscript

Abstract

I want to combine two hitherto largely independent research projects, scientific understanding and mechanistic explanations. Understanding is not only achieved by answering why-questions, that is, by providing scientific explanations, but also by answering what-questions, that is, by providing what I call scientific descriptions. Based on this distinction, I develop three forms of understanding: understanding-what, understanding-why, and understanding-how. I argue that understanding-how is a particularly deep form of understanding, because it is based on mechanistic explanations, which answer why something happens in virtue of what it is made of. I apply the three forms of understanding to two case studies: first, to the historical development of thermodynamics and, second, to the differences between the Clausius and the Boltzmann entropy in explaining thermodynamic processes.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Notes

  1. In The Open Society and Its Enemies, Popper (1945/2020, p. 30) criticizes methodological essentialism and defends methodological nominalism, in which what-questions are not important: “The methodological nominalist will never think that a question like ‘What is energy?’ or ‘What is movement?’ or ‘What is an atom?’ is an important question for physics; but he will attach importance to a question like: ‘How can the energy of the sun be made useful?’ or ‘How does a planet move?’ or ’Under what condition does an atom radiate light?’ And to those philosophers who tell him that before having answered the ‘what is’ question he cannot hope to give exact answers to any of the ‘how’ questions, he will reply, if at all, by pointing out that he much prefers that modest degree of exactness which he can achieve by his methods to the pretentious muddle which they have achieved by theirs.” Obviously, I don’t agree with Popper, but he also uses what-questions and how-questions in a more general way than I do.

  2. I can envision a rough taxonomy for scientific descriptions into material, structural, and functional descriptions, but such a project is still undone and would be the subject of another paper.

    Fig. 1
    figure 1

    Relation between understanding, explanation, and description. Understanding can be gained in two different ways, by asking a what-question or a why-questions. Answers to what-questions are scientific descriptions, and answers to why-questions lead to a scientific explanations. Whereas scientific explanations occur in two different forms, namely, causal or non-causal, there is not yet a further fine-grained taxonomy of scientific descriptions. The dashed arrow indicates that a scientific description may be used (in different degrees) in scientific explanations

  3. The Greek word that Aristotle uses is αιτια (aitia), which is standardly translated as cause. Since Vlastos (1969) distinguishes between Aristotle’s causes and becauses, there is a debate whether to translate αιτια rather as explanation (see also Annas, 1982, section III for a defense of explanation). We will see in the following that explanation would be more suitable in my scheme.

  4. Probably the first extant discussion of understanding and explanation is Plato’s Phaedo, which Aristotle explicitly criticizes (Annas, 1982).

  5. As Hennig mentions in section VI by citing Avicenna and Aquinas, there have been many attempts to unify the four causes in the history of philosophy.

  6. Hennig (2009, section II) emphasizes that Aristotle has something more general in mind than what something consists of, namely, that which potentially is the result; therefore, there are other examples where the material cause is not only matter or not matter at all.

  7. Cartwright et al. (2020) distinguish two questions, which show the difference between my and Hennig’s approach:

    1. 1.

      An epistemological question: ”What kind of explanation is involved?”

    2. 2.

      An ontological question: ”What is going on in the world?”

  8. I don’t claim that this is what Aristotle indeed had in mind, but rather that this is a plausible way to interpret the passage.

  9. There are two other definitions of mechanisms often discussed, one presented by Machamer et al. (2000) and the other by Bechtel and Abrahamsen (2005), which are rather tailored to the biological sciences than to physics.

  10. Grimm does so in order to argue that, like in physics, understanding-as-grasping-structure is understanding enough for understanding people.

  11. The situation for electromagnetism looked a bit different: especially in Britain, physicists early on presupposed a mechanism for the electromagentic field, of which Maxwell’s vortex model was a prominent example (Siegel 1991, 2014), while in Germany Wilhelm Weber’s action-at-a-distance theory was popular until it was challenged by Helmholtz and Hertz at the end of the 19th century (Steinle, 2013; Buchwald, 2013).

  12. Max Abraham and Gustav Mie pursued a different strategy to build a unified theory of gravity by implementing gravity directly into special relativity (Norton, 2007; Smeenk & Martin, 2007). This project was finally abandoned for theoretical and empirical reasons. I thank Dennis Lehmkuhl for this example.

  13. Thanks to an anonymous reviewer for raising this concern.

  14. The concept of entropy found its way into thermodynamics only in 1854 in a paper by Rudolf Clausius (1822–1888), but it took another decade for Clausius to come up with the name “entropy” (Purrington, 1997, p. 95).

  15. This was also one of the main motivations for David Bohm to develop his own quantum theory, which he himself preferred to call the ontological interpretation, although it was empirically equivalent to standard quantum mechanics (see van Strien, 2020, for a detailed discussion).

  16. Max Planck later changed his mind when he used Boltzmann’s idea of quantizing energy for his work on blackbody radiation. Albert Einstein, although studying Boltzmann’s work, was still too young a student at the ETH Zürich at this time.

  17. Although often the same people worked also in electromagnetism, positivism didn’t seem to be so prevalent in this theory. A rival theory to the field theory of electromagnetism was Wilhelm Weber’s action-at-a-distance theory. It would be worthwhile to find out how this theory was influenced by positivism (see, for instance, Hon & Goldstein, 2016).

  18. I thank an anonymous reviewer for raising this idea.

  19. There are other definitions of entropy: the Gibbs entropy in statistical mechanics, the von Neumann entropy for quantum systems, and the Shannon entropy for information. Robertson (2021) shows how the Gibbs entropy relates to thermodynamics; Myrvold (2020) discusses how the von Neumann entropy relates to thermodynamics; and Carcassi et al. (2021) do so for the Shannon entropy.

  20. It suffices for our purposes to characterize reversible processes as minimally (that is, quasi-statically) changing the physical system so that they can be conducted in the reverse order (see Norton, 2016b, for a detailed discussion).

  21. A thermodynamic process is called cyclic, if it returns to its initial state (Norton, 2016b, p. 48).

  22. An adiabatic process is a process in which the physical system is closed, that is, it does not exchange heat or matter with an outside system. The word adiabatic comes from the Greek word adiábatos, which literally means impassable (from a “not” + diabatós “to be crossed”).

  23. Thermodynamic equilibrium is defined to be a state in which the macroscopic variables of a system, like pressure, volume, temperature, and entropy, do not change over (sufficiently long) time (Robertson, 2021, section 3.1).

  24. This definition of entropy is written on Boltzmann’s tomb, although Boltzmann actually defined entropy in a different mathematical way, even if similar in spirit. This formula for entropy was due to Max Planck (Darrigol & Renn, 2013, p. 783) and taken up as the foundation for the neo-Boltzmannian project of statistical mechanics (Callender 1999; Goldstein 2001; Lebowitz1993a, 1993b, 1994, 2008, Penrose 1989).

  25. There is a debate on whether one should rather interpret the measure as a typicality measure, but this should not concern us here (see, e.g., Volchan, 2007; Frigg, 2009; Werndl, 2013; Lazarovici & Reichert, 2015; Myrvold, 2019; Maudlin, 2020; Hubert, 2021).

  26. One may think that because of the imaginary ensembles Gibbsian statistical mechanics (which is the most popular formulation among physicists) does not specify a mechanism—I thank an anonymous reviewer for raising this concern. First, Robertson (2021) argues that ensembles are not essential for Gibbsian statistical mechanics; they are rather introduced to interpret the probabilities as hypothetical frequencies. Second, Gibbs introduces a probability measure over sets of microstates, which specify the probability in which particular microstate the system is. In this sense, Gibbs presupposes a mechanism or, at least, the existence of microstates. It seems, however, that a Newtonian mechanism does not play such a prominent role for Gibbs as it does for Boltzmann, whose entire project is to derive statistical mechanics from Newtonian physics. Since Gibbs follows a more pragmatic approach than Boltzmann, Frigg and Werndl (2019) call Gibbsian statistical mechanics an effective theory and Boltzmannian statistical mechanics a fundamental theory.

  27. I thank an anonymous reviewer for pushing me on this point.

References

  • Annas, J. (1982). Aristotle on inefficient causes. The Philosophical Quarterly, 32(129), 311–26.

    Article  Google Scholar 

  • Aristotle. (2008). Physics. Oxford University Press. Translated by Robin Waterfield, with an introduction and notes by David Bostock.

  • Bechtel, W., & Abrahamsen, A. (2005). Explanation: A mechanist alternative. Studies in History and Philosophy of Science Part C: Studies in History and Philosophy of Biological and Biomedical Sciences, 36(2), 421–41.

    Article  Google Scholar 

  • Becker, A. (2018). What is real? the unfinished quest for the meaning of quantum physics. Basic Books.

  • Bell, J. S. (2004). The theory of local beables. In Speakable and Unspeakable in Quantum Mechanics (pp. 52–62). Cambridge University Press.

  • Beller, M. (1999). Quantum dialogue: The making of a revolution. The University of Chicago Press.

  • Blackmore, J. T. (1972). Ernst mach: His life, work, and influence. University of California Press.

  • Brown, H. R., & Uffink, J. (2001). The origins of time-asymmetry in thermodynamics: The minus first law. Studies in History and Philosophy of Modern Physics, 32(4), 525–38.

    Article  Google Scholar 

  • Buchwald, J. Z. (2013). Electrodynamics from Thomson and Maxwell to Hertz. In The Oxford Handbook of History of Physics (pp. 571–83). Oxford University Press.

  • Callender, C. (1999). Reducing thermodynamics to statistical mechanics: The case of entropy. The Journal of Philosophy, 97(7), 348–373.

    Google Scholar 

  • Carcassi, G., Aidala, C. A., & Barbour, J. (2021). Variability as a better characterization of Shannon entropy. European Journal of Physics, Forthcoming.

  • Cardwell, D. S. L. (1971). From watt to clausius: The rise of thermodynamics in the early industrial age. Heinemann.

  • Carnot, S. (1897). Reflections on the motive power of heat. Wiley. Translated from the original French and edited by R. H. Thurston.

  • Cartwright, N., Pemberton, J., & Wieten, S. (2020). Mechanisms, laws and explanation. European Journal for Philosophy of Science, 10(3), 25. https://doi.org/10.1007/s13194-020-00284-y.

    Article  Google Scholar 

  • Cercignani, C. (1998). Ludwig boltzmann: The man who trusted atoms. Oxford University Press.

  • Chakravartty, A. (2017). Scientific ontology: Integrating naturalized metaphysics and voluntarist epistemology. Oxford University Press.

  • Chang, H. (2013). Thermal physics and thermodynamics. In J.Z. Buchwald R. Fox (Eds.) The Oxford Handbook of History of Physics (pp. 473–507). Oxford University Press.

  • Clausius, R. (1867). The mechanical theory of heat, with its applications to the steam engine, and to the physical properties of bodies. John van Voorst.

  • Cushing, J. T. (1994). Quantum mechanics: Historical contingency and the copenhagen hegemony. The University of Chicago Press.

  • Darrigol, O. (2018). Atoms, mechanics, and probability. Oxford University Press.

  • Darrigol, O., & Renn, J. (2013). The emergence of statistical mechanics. In J.Z. Buchwald R.s Fox (Eds.) The Oxford handbook of history of physics (pp. 765–88). Oxford University Press.

  • de Regt, H. W. (2014). Visualization as a tool for understanding. Perspectives on Science, 22(3), 377–96.

    Article  Google Scholar 

  • de Regt, H. W. (2017). Understanding scientific understanding. Oxford University Press.

  • de Regt, H. W. (2019). From explanation to understanding: Normativity lost?. Journal for General Philosophy of Science, 50(3), 327–43.

    Article  Google Scholar 

  • de Regt, H. W., & Dieks, D. (2005). A contextual approach to scientific understanding. Synthese, 144(1), 137–70.

    Article  Google Scholar 

  • de Regt, H. W., & Gijsbers, V. (2017). How false theories can yield genuine understanding. In S. R. Grimm, C. Baumberger, & S. Ammon (Eds.) Explaining Understanding: New Perspectives from Epistemology and Philosophy of Science (pp. 50–75). Routledge.

  • Deltete, R. J. (2012). Georg Helm’s Chemical Energetics. HYLE – International Journal for Philosophy of Chemistry, 18(1), 23–44.

    Google Scholar 

  • Einstein, A. (1905). Zur Elektrodynamik bewegter Körper. Annalen der Physik, 322(10), 891–921.

    Article  Google Scholar 

  • Einstein, A., Podolsky, B., & Rosen, N. (1935). Can quantum-mechanical description of physical reality be considered complete?. Physical Review, 47(10), 777–80.

    Article  Google Scholar 

  • Elliott, K. C., & McKaughan, D. J. (2014). Nonepistemic values and the multiple goals of science. Philosophy of Science, 81(1), 1–21.

    Article  Google Scholar 

  • Felline, L. (2019). Mechanistic explanation in physics. In Companion to the Philosophy of Physics. Routledge. Forthcoming.

  • Fine, A. (1993). Einstein’s interpretations of the quantum theory. Science in Context, 6, 257–73.

    Article  Google Scholar 

  • Fox, R. (1971). The caloric theory of gases from lavoisier to regnault. Clarendon Press.

  • Frigg, R. (2009). Typicality and the approach to equilibrium in Boltzmannian statistical mechanics. Philosophy of Science, 76(5), 997–1008.

    Article  Google Scholar 

  • Frigg, R., & Werndl, C. (2019). Statistical Mechanics: A Tale of Two Theories. The Monist, 102(4), 424–438. ISSN 0026-9662. https://doi.org/10.1093/monist/onz018.

    Article  Google Scholar 

  • Giovanelli, M. (2013). Talking at cross-purposes: How Einstein and the logical empiricists never agreed on what they were disagreeing about. Synthese, 190(17), 3819–63.

    Article  Google Scholar 

  • Glennan, S. (2017). The new mechanical philosophy. Oxford University Press.

  • Glennan, S, & Illari, P. (2018). The routledge handbook of mechanisms and mechanical philosophy. Routledge.

  • Glennan, S. S. (1996). Mechanisms and the nature of causation. Erkenntnis, 44(1), 49–71.

    Article  Google Scholar 

  • Goldstein, S. (2001). Boltzmann’s approach to statistical mechanics. In J. Bricmont, D. Dürr, M. C. Galavotti, G. Ghirardi, F. Petruccione, & N. Zanghì (Eds.) Chance in Physics: Foundations and Perspectives (pp. 39–54). Springer.

  • Greco, J. (2014). Episteme: Knowledge and understanding. In K. Timpe C. A. Boyd (Eds.) Virtues and Their Vices (pp. 285–302). Oxford University Press.

  • Grimm, S. R. (2008). Explanatory inquiry and the need for explanation. The British Journal for the Philosophy of Science, 59(3), 481–97.

    Article  Google Scholar 

  • Grimm, S. R. (2016). How understanding people differs from understanding the natural world. Philosophical Issues, 26(1), 209–25.

    Article  Google Scholar 

  • Harrigan, N., & Spekkens, R. W. (2010). Einstein, incompleteness, and the epistemic view of quantum states. Foundations of Physics, 40(2), 125–57.

    Article  Google Scholar 

  • Hempel, C. G. (1962). Explanation in science and history. In R. C. Colodny (Ed.) Frontiers of Science and Philosophy (pp. 9–19). University of Pittsburgh Press.

  • Hempel, C. G., & Oppenheim, P. (1948). Studies in the logic of explanation. Philosophy of Science, 15(2), 135–75.

    Article  Google Scholar 

  • Hennig, B. (2009). The four causes. The Journal of Philosophy, 106(3), 137–60.

    Article  Google Scholar 

  • Hoefer, C. (1994). Einstein’s struggle for a machian gravitation theory. Studies in History and Philosophy of Science Part A, 25 (3), 287–335. ISSN 0039-3681. https://doi.org/10.1016/0039-3681(94)90056-6, https://www.sciencedirect.com/science/article/pii/0039368194900566.

    Article  Google Scholar 

  • Holton, G. (1992). Ernst Mach and the fortunes of positivism in America. Isis, 83(1), 27–60.

    Article  Google Scholar 

  • Hon, G., & Goldstein, B. R. (2016). Maxwell’s methodological odyssey in electromagnetism. Manuscript at the Eighth Quadrennial Fellows Conference, Pittsburgh Center for Philosophy of Science, Lund University, Lund, Sweden. http://philsci-archive.pitt.edu/12241/.

  • Howard, D. (2005). Albert Einstein as a Philosopher of Science. Physics Today, 58(12), 34. https://doi.org/10.1063/1.2169442.

    Article  Google Scholar 

  • Hubert, M. (2021). Reviving frequentism. Synthese, online first. https://doi.org/10.1007/s11229-021-03024-8.

  • Illari, P. (2019). Mechanisms, models and laws in understanding supernovae. Journal for General Philosophy of Science, 50(1), 63–84.

    Article  Google Scholar 

  • Illari, P. M., & Williamson, J. (2011). Mechanisms are real and local. In P. M. Illari, F. Russo, & J. Williamson (Eds.) Causality in the Sciences (pp. 818–44). Oxford University Press,.

  • Janiak, A. (2008). Newton as philosopher. Cambridge University Press.

  • Kaiser, D. (2005). Drawing theories apart: The dispersion of feynman diagrams in postwar physics. The University of Chicago Press.

  • Kelp, C. (2015). Understanding phenomena. Synthese, 192(12), 3799–816.

    Article  Google Scholar 

  • Kuhlmann, M., & Glennan, S. (2014). On the relation between quantum mechanical and neo-mechanistic ontologies and explanatory strategies. European Journal for Philosophy of Science, 4(3), 337–59.

    Article  Google Scholar 

  • Kvanvig, J. (2003). The value of knowledge and the pursuit of understanding. Cambridge University Press.

  • Lazarovici, D., & Reichert, P. (2015). Typicality, irreversibility and the status of macroscopic laws. Erkenntnis, 80(4), 689–716.

    Article  Google Scholar 

  • Lebowitz, J. L. (1993). Boltzmann’s entropy and time’s arrow. Physics Today, 46, 32–38.

    Article  Google Scholar 

  • Lebowitz, J. L. (1993). Macroscopic laws, microscopic dynamics, time’s arrow and Boltzmann’s entropy. Physica A, 194, 1–27.

    Article  Google Scholar 

  • Lebowitz, J. L. (1994). Time’s arrow and Boltzmann’s entropy. In J. Halliwell W. H. Zurek (Eds.) Physical Origins of Time Asymmetry (pp .131–46). Cambridge University Press.

  • Lebowitz, J. L. (2008). Time’s arrow and Boltzmann’s entropy. Scholarpedia, 3(4), 3348. https://doi.org/10.4249/scholarpedia.3448.

    Article  Google Scholar 

  • Lemons, D. S. (2009). Mere thermodynamics. The Johns Hopikins Universtiy Press.

  • Lemons, D. S. (2019). Thermodynamic weirdness: From fahrenheit to clausius. MIT Press.

  • Lindley, D. (2001). Boltzmann’s atom: The great debate that launched a revolution in physics. The Free Press.

  • Machamer, P., Darden, L., & Craver, C. F. (2000). Thinking about mechanisms. Philosophy of Science, 67(1), 1–25.

    Article  Google Scholar 

  • Maudlin, T. (2019). Philosophy of physics: Quantum theory. Princeton University Press.

  • Maudlin, T. (2020). The grammar of typicality. In V. Allori (Ed.) Statistical Mechanics and Scientific Explanation: Determinism, Indeterminism and Laws of Nature (pp. 231–51). World Scientific.

  • Myrvold, W. C. (2019). Explaining thermodynamics: What remains to be done?. In V. Allori (Ed.) Statistical Mechanics and Scientific Explanation: Determinism, Indeterminism and Laws of Nature. World Scientific. Forthcoming.

  • Myrvold, W. C. (2020). The science of 𝜃δcs. Foundations of Physics, 50 (10), 1219–1251. https://doi.org/10.1007/s10701-020-00371-3.

    Article  Google Scholar 

  • Norsen, T. (2017). Foundations of quantum mechanics: An exploration of the physical meaning of quantum theory. Springer.

  • Norton, J. D. (1993). General covariance and the foundations of general relativity: Eight decades of dispute. Reports on Progress in Physics, 56(7), 791–858.

    Article  Google Scholar 

  • Norton, J. D. (2007). Einstein, Nordström, and the Early Demise of Scalar, Lorentz Covariant Theories of Gravitation. In M. Janssen, J. D. Norton, J. Renn, T. Sauer, & J. Stachel (Eds.) The Genesis of General Relativity, (Vol. 3 pp. 413–487). Springer Netherlands. https://doi.org/10.1007/978-1-4020-4000-9_27.

  • Norton, J. D. (2016). How Einstein Did Not Discover. Physics in Perspective, 18(3), 249–82.

    Article  Google Scholar 

  • Norton, J. D. (2016). The impossible process: Thermodynamic reversibility. Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics, 55, 43–61. https://doi.org/10.1016/j.shpsb.2016.08.001, https://www.sciencedirect.com/science/article/pii/S1355219815300563.

    Article  Google Scholar 

  • Nozick, R. (1981). Philosophical explanations. Harvard University Press.

  • Penrose, R. (1989). The emperor’s new mind: Concerning computers, minds and the laws of physics. Oxford University Press.

  • Popper, K. (1945/2020). The open society and its enemies. Princeton University Press.

  • Potochnik, A. (2015). The diverse aims of science. Studies in History and Philosophy of Science, 53, 71–80.

    Article  Google Scholar 

  • Potochnik, A. (2017). Idealization and the aims of science. The University of Chicago Press.

  • Purrington, R. D. (1997). Physics in the nineteenth century. Rutgers University Press.

  • Robertson, K. (2021). In search of the holy grail: How to reduce the second law of thermodynamics The British Journal for the Philosophy of Science, Forthcoming.

  • Salmon, W. C. (1998). The importance of scientific understanding. In Causality and Explanation (pp. 79–91). Oxford University Press.

  • Siegel, D. M. (1991). Innovation in maxwell’s electromagnetic theory: Molecular vortices, displacement current, and light. Cambridge University Press.

  • Siegel, D. M. (2014). Maxwell’s contributions to electricity and magnetism. In R. Flood, M. McCartney, & A. Whitaker (Eds.) James Clerk Maxwell: Perspectives on his Life and Work (pp. 187–203). Oxford University Press.

  • Smeenk, C., & Martin, C. (2007). Mie’s theories of matter and gravitation. In M. Janssen, J. D. Norton, J. Renn, T. Sauer, & J. Stachel (Eds.) The Genesis of General Relativity, (Vol. 4 pp. 623–632). Springer Netherlands.

  • Smith, C. (1976). Natural philosophy and thermodynamics: William thomson and ‘the dynamical theory of heat’. The British Journal for the History of Science, 9(3), 293–319. https://doi.org/10.1017/S0007087400014874.

    Article  Google Scholar 

  • Stein, N. (2011). Causation and explanation in aristotle. Philosophy Compass, 6(10), 699–707.

    Article  Google Scholar 

  • Steinle, F. (2013). Electromagnetism and field physics. In The Oxford Handbook of History of Physics (pp. 533–570). Oxford University Press.

  • Strevens, M. (2013). No understanding without explanation. Studies in History and Philosophy of Science, 44(3), 510–5.

    Article  Google Scholar 

  • Thomson, W. (1853). On the Dynamical Theory of Heat, with numerical results deduced from Mr Joule’s “Equivalent of a Thermal Unit”, and M. Regnault’s “Observations on Steam”. Transactions of the Royal Society of Edinburgh, 20(2), 261–288. https://doi.org/10.1017/S0080456800033172.

    Article  Google Scholar 

  • Trout, J. D. (2002). Scientific explanation and the sense of understanding. Philosophy of Science, 69(2), 212–33.

    Article  Google Scholar 

  • van Fraassen, B. (1980). The scientific image. Oxford University Press.

  • van Strien, M. (2020). Bohm’s theory of quantum mechanics and the notion of classicality. Studies in History and Philosophy of Modern Physics. https://doi.org/10.1016/j.shpsb.2020.04.005, http://www.sciencedirect.com/science/article/pii/S1355219819301716.

  • Vlastos, G. (1969). Reasons and causes in the Phaedo. The Philosophical Review, 78(3), 291–325.

    Article  Google Scholar 

  • Volchan, S. B. (2007). Probability as typicality. Studies in History and Philosophy of Modern Physics, 38(4), 801–14.

    Article  Google Scholar 

  • Werndl, C. (2013). Justifying typicality measures of boltzmannian statistical mechanics and dynamical systems. Studies in History and Philosophy of Modern Physics, 44(4), 470–9.

    Article  Google Scholar 

  • Woodward, J. (2003). Making things happen: A theory of causal explanation. Oxford University Press.

  • Zagzebski, L. (2001). Recovering understanding. In M. Steup (Ed.) Knowledge, Truth, and Duty: Essays on Epistemic Justification, Responsibility, and Virtue. Oxford University Press.

Download references

Acknowledgements

I wish to thank Frederick Eberhardt, Christopher Hitchcock, and Tim Maudlin for their support and many helpful discussions. I also thank the audience of the Caltech Philosophy of Physics Reading Group for their comments. Special thanks go to Joshua Eisenthal and Maaneli Derakhshani their help on the passages on Einstein and to Charles Sebens and Federica Malfatti for thoroughly reading and commenting on previous drafts of this paper. Lastly, I wish to thank two anonymous reviewers for their constructive criticism and many suggestions for improving the paper.

Funding

Funding for this research was partially provided by the Swiss National Science Foundation as part of the Early Postdoc.Mobility Fellowship, grant no. 174745.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mario Hubert.

Ethics declarations

Conflict of Interest

There is no conflict of interest.

Additional information

Ethical approval

Not applicable.

Informed consent

Not applicable.

Publisher’s note

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

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hubert, M. Understanding physics: ‘What?’, ‘Why?’, and ‘How?’. Euro Jnl Phil Sci 11, 85 (2021). https://doi.org/10.1007/s13194-021-00399-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s13194-021-00399-w

Keywords

Navigation