What is a Scientific Theory?

What is a Scientific Theory?

The University of California, Berkley, defines a theory as “a broad, natural explanation for a wide range of phenomena. Theories are concise, coherent, systematic, predictive, and broadly applicable, often integrating and generalizing many hypotheses.” A scientific theory is an explanation of an aspect of the natural world that can be repeatedly tested and verified in accordance with the scientific method, using accepted protocols of observation, measurement, and evaluation of results. Where possible, theories are tested under controlled conditions in an experiment.[1][2] In circumstances not amenable to experimental testing, theories are evaluated through principles of abductive reasoning. Established scientific theories have withstood rigorous scrutiny and embody scientific knowledge.[3]

The meaning of the term scientific theory (often contracted to theory for brevity) as used in the disciplines of science is significantly different from the common vernacular usage of theory.[4] In everyday speech, theory can imply an explanation that represents an unsubstantiated and speculative guess,[4] whereas in science it describes an explanation that has been tested and widely accepted as valid. These different usages are comparable to the opposing usages of prediction in science versus common speech, where it denotes a mere hope.

The strength of a scientific theory is related to the diversity of phenomena it can explain and its simplicity. As additional scientific evidence is gathered, a scientific theory may be modified and ultimately rejected if it cannot be made to fit the new findings; in such circumstances, a more accurate theory is then required. That does not mean that all theories can be fundamentally changed (for example, well established foundational scientific theories such as evolution, heliocentric theory, cell theory, theory of plate tectonics, germ theory of disease, etc.). In certain cases, the less-accurate unmodified scientific theory can still be treated as a theory if it is useful (due to its sheer simplicity) as an approximation under specific conditions. A case in point is Newton’s laws of motion, which can serve as an approximation to special relativity at velocities that are small relative to the speed of light.

Scientific theories are testable and make falsifiable predictions.[5] They describe the causes of a particular natural phenomenon and are used to explain and predict aspects of the physical universe or specific areas of inquiry (for example, electricity, chemistry, and astronomy). Scientists use theories to further scientific knowledge, as well as to facilitate advances in technology or medicine.

As with other forms of scientific knowledge, scientific theories are both deductive and inductive,[6] aiming for predictive and explanatory power.

The paleontologist Stephen Jay Gould wrote that “…facts and theories are different things, not rungs in a hierarchy of increasing certainty. Facts are the world’s data. Theories are structures of ideas that explain and interpret facts.”[7]

About theories

Theories and laws

Both scientific laws and scientific theories are produced from the scientific method through the formation and testing of hypotheses, and can predict the behavior of the natural world. Both are typically well-supported by observations and/or experimental evidence.[29] However, scientific laws are descriptive accounts of how nature will behave under certain conditions.[30] Scientific theories are broader in scope, and give overarching explanations of how nature works and why it exhibits certain characteristics. Theories are supported by evidence from many different sources, and may contain one or several laws.[31]

A common misconception is that scientific theories are rudimentary ideas that will eventually graduate into scientific laws when enough data and evidence have been accumulated. A theory does not change into a scientific law with the accumulation of new or better evidence. A theory will always remain a theory; a law will always remain a law.[29][32][33] Both theories and laws could potentially be falsified by countervailing evidence.[34]

Theories and laws are also distinct from hypotheses. Unlike hypotheses, theories and laws may be simply referred to as scientific fact.[35][36] However, in science, theories are different from facts even when they are well supported.[37] For example, evolution is both a theory and a fact.[4]

Theories as axioms

The logical positivists thought of scientific theories as statements in a formal language. First-order logic is an example of a formal language. The logical positivists envisaged a similar scientific language. In addition to scientific theories, the language also included observation sentences (“the sun rises in the east”), definitions, and mathematical statements. The phenomena explained by the theories, if they could not be directly observed by the senses (for example, atoms and radio waves), were treated as theoretical concepts. In this view, theories function as axioms: predicted observations are derived from the theories much like theorems are derived in Euclidean geometry. However, the predictions are then tested against reality to verify the theories, and the “axioms” can be revised as a direct result.

The phrase “the received view of theories” is used to describe this approach. Terms commonly associated with it are “linguistic” (because theories are components of a language) and “syntactic” (because a language has rules about how symbols can be strung together). Problems in defining this kind of language precisely, e.g., are objects seen in microscopes observed or are they theoretical objects, led to the effective demise of logical positivism in the 1970s.

Theories as models

The semantic view of theories, which identifies scientific theories with models rather than propositions, has replaced the received view as the dominant position in theory formulation in the philosophy of science.[38][39][40] A model is a logical framework intended to represent reality (a “model of reality”), similar to the way that a map is a graphical model that represents the territory of a city or country.[41][42]

Precession of the perihelion of Mercury (exaggerated). The deviation in Mercury’s position from the Newtonian prediction is about 43 arc-seconds (about two-thirds of 1/60 of a degree) per century.[43][44]

In this approach, theories are a specific category of models that fulfill the necessary criteria (see above). One can use language to describe a model; however, the theory is the model (or a collection of similar models), and not the description of the model. A model of the solar system, for example, might consist of abstract objects that represent the sun and the planets. These objects have associated properties, e.g., positions, velocities, and masses. The model parameters, e.g., Newton’s Law of Gravitation, determine how the positions and velocities change with time. This model can then be tested to see whether it accurately predicts future observations; astronomers can verify that the positions of the model’s objects over time match the actual positions of the planets. For most planets, the Newtonian model’s predictions are accurate; for Mercury, it is slightly inaccurate and the model of general relativity must be used instead.

The word “semantic” refers to the way that a model represents the real world. The representation (literally, “re-presentation”) describes particular aspects of a phenomenon or the manner of interaction among a set of phenomena. For instance, a scale model of a house or of a solar system is clearly not an actual house or an actual solar system; the aspects of an actual house or an actual solar system represented in a scale model are, only in certain limited ways, representative of the actual entity. A scale model of a house is not a house; but to someone who wants to learn about houses, analogous to a scientist who wants to understand reality, a sufficiently detailed scale model may suffice.

Differences between theory and model

Several commentators[45] have stated that the distinguishing characteristic of theories is that they are explanatory as well as descriptive, while models are only descriptive (although still predictive in a more limited sense). Philosopher Stephen Pepper also distinguished between theories and models, and said in 1948 that general models and theories are predicated on a “root” metaphor that constrains how scientists theorize and model a phenomenon and thus arrive at testable hypotheses.

Engineering practice makes a distinction between “mathematical models” and “physical models”; the cost of fabricating a physical model can be minimized by first creating a mathematical model using a computer software package, such as a computer aided design tool. The component parts are each themselves modelled, and the fabrication tolerances are specified. An exploded view drawing is used to lay out the fabrication sequence. Simulation packages for displaying each of the subassemblies allow the parts to be rotated, magnified, in realistic detail. Software packages for creating the bill of materials for construction allows subcontractors to specialize in assembly processes, which spreads the cost of manufacturing machinery among multiple customers. See: Computer-aided engineering, Computer-aided manufacturing, and 3D printing.

From philosophers of science

Karl Popper described the characteristics of a scientific theory as follows:[5]

  1. It is easy to obtain confirmations, or verifications, for nearly every theory—if we look for confirmations.
  2. Confirmations should count only if they are the result of risky predictions; that is to say, if, unenlightened by the theory in question, we should have expected an event which was incompatible with the theory—an event which would have refuted the theory.
  3. Every “good” scientific theory is a prohibition: it forbids certain things to happen. The more a theory forbids, the better it is.
  4. A theory which is not refutable by any conceivable event is non-scientific. Irrefutability is not a virtue of a theory (as people often think) but a vice.
  5. Every genuine test of a theory is an attempt to falsify it, or to refute it. Testability is falsifiability; but there are degrees of testability: some theories are more testable, more exposed to refutation, than others; they take, as it were, greater risks.
  6. Confirming evidence should not count except when it is the result of a genuine test of the theory; and this means that it can be presented as a serious but unsuccessful attempt to falsify the theory. (I now speak in such cases of “corroborating evidence”.)
  7. Some genuinely testable theories, when found to be false, might still be upheld by their admirers—for example by introducing post hoc (after the fact) some auxiliary hypothesis or assumption, or by reinterpreting the theory post hoc in such a way that it escapes refutation. Such a procedure is always possible, but it rescues the theory from refutation only at the price of destroying, or at least lowering, its scientific status, by tampering with evidence. The temptation to tamper can be minimized by first taking the time to write down the testing protocol before embarking on the scientific work.

Popper summarized these statements by saying that the central criterion of the scientific status of a theory is its “falsifiability, or refutability, or testability“.[5] Echoing this, Stephen Hawking states, “A theory is a good theory if it satisfies two requirements: It must accurately describe a large class of observations on the basis of a model that contains only a few arbitrary elements, and it must make definite predictions about the results of future observations.” He also discusses the “unprovable but falsifiable” nature of theories, which is a necessary consequence of inductive logic, and that “you can disprove a theory by finding even a single observation that disagrees with the predictions of the theory”.[47]

Several philosophers and historians of science have, however, argued that Popper’s definition of theory as a set of falsifiable statements is wrong[48] because, as Philip Kitcher has pointed out, if one took a strictly Popperian view of “theory”, observations of Uranus when first discovered in 1781 would have “falsified” Newton’s celestial mechanics. Rather, people suggested that another planet influenced Uranus’ orbit—and this prediction was indeed eventually confirmed.

Kitcher agrees with Popper that “There is surely something right in the idea that a science can succeed only if it can fail.”[49] He also says that scientific theories include statements that cannot be falsified, and that good theories must also be creative. He insists we view scientific theories as an “elaborate collection of statements”, some of which are not falsifiable, while others—those he calls “auxiliary hypotheses”, are.

According to Kitcher, good scientific theories must have three features:[49]

  1. Unity: “A science should be unified…. Good theories consist of just one problem-solving strategy, or a small family of problem-solving strategies, that can be applied to a wide range of problems.”
  2. Fecundity: “A great scientific theory, like Newton’s, opens up new areas of research…. Because a theory presents a new way of looking at the world, it can lead us to ask new questions, and so to embark on new and fruitful lines of inquiry…. Typically, a flourishing science is incomplete. At any time, it raises more questions than it can currently answer. But incompleteness is not vice. On the contrary, incompleteness is the mother of fecundity…. A good theory should be productive; it should raise new questions and presume those questions can be answered without giving up its problem-solving strategies.”
  3. Auxiliary hypotheses that are independently testable: “An auxiliary hypothesis ought to be testable independently of the particular problem it is introduced to solve, independently of the theory it is designed to save.” (For example, the evidence for the existence of Neptune is independent of the anomalies in Uranus’s orbit.)

Like other definitions of theories, including Popper’s, Kitcher makes it clear that a theory must include statements that have observational consequences. But, like the observation of irregularities in the orbit of Uranus, falsification is only one possible consequence of observation. The production of new hypotheses is another possible and equally important result.

Analogies and metaphors

The concept of a scientific theory has also been described using analogies and metaphors. For instance, the logical empiricist Carl Gustav Hempel likened the structure of a scientific theory to a “complex spatial network:”

Its terms are represented by the knots, while the threads connecting the latter correspond, in part, to the definitions and, in part, to the fundamental and derivative hypotheses included in the theory. The whole system floats, as it were, above the plane of observation and is anchored to it by the rules of interpretation. These might be viewed as strings which are not part of the network but link certain points of the latter with specific places in the plane of observation. By virtue of these interpretive connections, the network can function as a scientific theory: From certain observational data, we may ascend, via an interpretive string, to some point in the theoretical network, thence proceed, via definitions and hypotheses, to other points, from which another interpretive string permits a descent to the plane of observation.[50]

Michael Polanyi made an analogy between a theory and a map:

A theory is something other than myself. It may be set out on paper as a system of rules, and it is the more truly a theory the more completely it can be put down in such terms. Mathematical theory reaches the highest perfection in this respect. But even a geographical map fully embodies in itself a set of strict rules for finding one’s way through a region of otherwise uncharted experience. Indeed, all theory may be regarded as a kind of map extended over space and time.[51]

A scientific theory can also be thought of as a book that captures the fundamental information about the world, a book that must be researched, written, and shared. In 1623, Galileo Galilei wrote:

Philosophy [i.e. physics] is written in this grand book—I mean the universe—which stands continually open to our gaze, but it cannot be understood unless one first learns to comprehend the language and interpret the characters in which it is written. It is written in the language of mathematics, and its characters are triangles, circles, and other geometrical figures, without which it is humanly impossible to understand a single word of it; without these, one is wandering around in a dark labyrinth.[52]

The book metaphor could also be applied in the following passage, by the contemporary philosopher of science Ian Hacking:

I myself prefer an Argentine fantasy. God did not write a Book of Nature of the sort that the old Europeans imagined. He wrote a Borgesian library, each book of which is as brief as possible, yet each book of which is inconsistent with every other. No book is redundant. For every book there is some humanly accessible bit of Nature such that that book, and no other, makes possible the comprehension, prediction and influencing of what is going on…Leibniz said that God chose a world which maximized the variety of phenomena while choosing the simplest laws. Exactly so: but the best way to maximize phenomena and have simplest laws is to have the laws inconsistent with each other, each applying to this or that but none applying to all.[53]

In physics

In physics, the term theory is generally used for a mathematical framework—derived from a small set of basic postulates (usually symmetries—like equality of locations in space or in time, or identity of electrons, etc.)—that is capable of producing experimental predictions for a given category of physical systems. A good example is classical electromagnetism, which encompasses results derived from gauge symmetry (sometimes called gauge invariance) in a form of a few equations called Maxwell’s equations. The specific mathematical aspects of classical electromagnetic theory are termed “laws of electromagnetism,” reflecting the level of consistent and reproducible evidence that supports them. Within electromagnetic theory generally, there are numerous hypotheses about how electromagnetism applies to specific situations. Many of these hypotheses are already considered to be adequately tested, with new ones always in the making and perhaps untested. An example of the latter might be the radiation reaction force. As of 2009, its effects on the periodic motion of charges are detectable in synchrotrons, but only as averaged effects over time. Some researchers are now considering experiments that could observe these effects at the instantaneous level (i.e. not averaged over time).[54][55]

Reference

  1. National Academy of Sciences (US) (1999). Science and Creationism: A View from the National Academy of Sciences(2nd ed.). National Academies Press. p. 2doi:10.17226/6024ISBN 978-0-309-06406-4PMID 25101403.
  2. The Structure of Scientific Theories. The Stanford Encyclopedia of Philosophy. Metaphysics Research Lab, Stanford University. 2016.
  3. Schafersman, Steven D. “An Introduction to Science”.
  4. “Is Evolution a Theory or a Fact?”National Academy of Sciences. 2008.
  5. Popper, Karl (1963), Conjectures and Refutations, Routledge and Kegan Paul, London, UK. Reprinted in Theodore Schick (ed., 2000), Readings in the Philosophy of Science, Mayfield Publishing Company, Mountain View, Calif.
  6. Andersen, Hanne; Hepburn, Brian (2015). Edward N. Zalta (ed.). Scientific Method. The Stanford Encyclopedia of Philosophy.
  7. The Devil in Dover,
  8. Howard, Don A. (23 June 2018). Zalta, Edward N. (ed.). The Stanford Encyclopedia of Philosophy. Metaphysics Research Lab, Stanford University – via Stanford Encyclopedia of Philosophy.
  9. Alan Baker (2010) [2004]. “Simplicity”. Stanford Encyclopedia of Philosophy. California: Stanford University. ISSN 1095-5054.
  10. Courtney A, Courtney M (2008). “Comments Regarding “On the Nature Of Science””. Physics in Canada. 64 (3): 7–8. arXiv:0812.4932.
  11. Elliott Sober, Let’s Razor Occam’s Razor, pp. 73–93, from Dudley Knowles (ed.) Explanation and Its Limits, Cambridge University Press (1994).
  12. National Academy of Sciences (2008), Science, Evolution, and Creationism.
  13. Hooke, Robert (1635–1703). Micrographia, Observation XVIII.
  14.  Misner, Charles W.; Thorne, Kip S.; Wheeler, John Archibald (1973). Gravitation, p. 1049. New York: W. H.Freeman and Company. ISBN 0-7167-0344-0.
  15. See Acid–base reaction.
  16. Bước lên tới:a b c “Chapter 1: The Nature of Science”. www.project2061.org.
  17. See, for example, Common descent and Evidence for common descent.
  18. For example, see the article on the discovery of Neptune; the discovery was based on an apparent violation of the orbit of Uranus as predicted by Newtonian mechanics. This explanation did not require any modification of the theory, but rather modification of the hypothesis that there were only seven planets in the Solar System.
  19. U. Le Verrier (1859), (in French), “Lettre de M. Le Verrier à M. Faye sur la théorie de Mercure et sur le mouvement du périhélie de cette planète”, Comptes rendus hebdomadaires des séances de l’Académie des sciences (Paris), vol. 49 (1859), pp. 379–83.
  20. For example, the modern theory of evolution (the modern evolutionary synthesis) incorporates significant contributions from R. A. FisherErnst MayrJ. B. S. Haldane, and many others.
  21. Weinberg S (1993). Dreams of a Final Theory: The Scientist’s Search for the Ultimate Laws of Nature.
  22. Maxwell, J. C., & Thompson, J. J. (1892). A treatise on electricity and magnetism. Clarendon Press series. Oxford: Clarendon.
  23. “How the Sun Shines”. www.nobelprize.org.
  24. The strong force, the electroweak force, and gravity. The electroweak force is the unification of electromagnetism and the weak force. All observed causal interactions are understood to take place through one or more of these three mechanisms, although most systems are far too complicated to account for these except through the successive approximations offered by other theories.
  25. Albert Einstein (1905) “Zur Elektrodynamik bewegter Körper Archived 2009-12-29 at the Wayback Machine“, Annalen der Physik 17: 891; English translation On the Electrodynamics of Moving Bodies by George Barker Jefferyand Wilfrid Perrett (1923); Another English translation On the Electrodynamics of Moving Bodies by Megh Nad Saha (1920).
  26. Schwarz, John H (Mar 1998). “Recent developments in superstring theory”. Proceedings of the National Academy of Sciences of the United States of America. 95 (6): 2750–57. Bibcode:1998PNAS…95.2750Sdoi:10.1073/pnas.95.6.2750PMC 19640PMID 9501161.
  27. See Tests of special relativity. Also, for example: Sidney Coleman, Sheldon L. Glashow, Cosmic Ray and Neutrino Tests of Special Relativity, Phys. Lett. B405 (1997) 249–52, found here [1]. An overview can be found here.
  28. Roberto Torretti, The Philosophy of Physics (Cambridge: Cambridge University Press, 1999), pp. 289–90.
  29. “Scientific Laws and Theories”.
  30. See the article on Physical law, for example.
  31. “Definitions of Fact, Theory, and Law in Scientific Work”. 16 March 2016.
  32. “Harding (1999)”.
  33. William F. McComas (30 December 2013). The Language of Science Education: An Expanded Glossary of Key Terms and Concepts in Science Teaching and Learning. Springer Science & Business Media. p. 107. ISBN 978-94-6209-497-0.
  34. “What’s the Difference Between a Scientific Hypothesis, Theory and Law?”.
  35.  Gould, Stephen Jay (1981-05-01). “Evolution as Fact and Theory”. Discover. 2 (5): 34–37.
  36. Further examples are here [2], and in the article on Evolution as fact and theory.
  37. “Essay”. ncse.com. Retrieved 25 March 2015.
  38. Suppe, Frederick (1998). “Understanding Scientific Theories: An Assessment of Developments, 1969–1998” (PDF). Philosophy of Science. 67: S102–S115. doi:10.1086/392812. Retrieved 14 February 2013.
  39. Halvorson, Hans (2012). “What Scientific Theories Could Not Be” (PDF). Philosophy of Science. 79 (2): 183–206. CiteSeerX 10.1.1.692.8455doi:10.1086/664745. Retrieved 14 February 2013.
  40. Frigg, Roman (2006). “Scientific Representation and the Semantic View of Theories” (PDF). Theoria. 55 (2): 183–206. Retrieved 14 February 2013.
  41. Hacking, Ian (1983). Representing and Intervening. Introductory Topics in the Philosophy of Natural Science. Cambridge University Press.
  42. Box, George E.P. & Draper, N.R. (1987). Empirical Model-Building and Response Surfaces. Wiley. p. 424
  43. Lorenzo Iorio (2005). “On the possibility of measuring the solar oblateness and some relativistic effects from planetary ranging”. Astronomy and Astrophysics. 433 (1): 385–93. arXiv:gr-qc/0406041Bibcode:2005A&A…433..385Idoi:10.1051/0004-6361:20047155.
  44. Myles Standish, Jet Propulsion Laboratory (1998)
  45. For example, Reese & Overto (1970); Lerner (1998); also Lerner & Teti (2005), in the context of modeling human behavior.
  46. Isaac Asimov, Understanding Physics (1966) pp. 4–5.
  47. Hawking, Stephen (1988). A Brief History of TimeBantam BooksISBN 978-0-553-38016-3.
  48. Hempel. C.G. 1951 “Problems and Changes in the Empiricist Criterion of Meaning” in Aspects of Scientific Explanation. Glencoe: the Free Press. Quine, W.V.O 1952 “Two Dogmas of Empiricism” reprinted in From a Logical Point of View. Cambridge: Harvard University Press
  49. Philip Kitcher 1982 Abusing Science: The Case Against Creationism, pp. 45–48. Cambridge: The MIT Press
  50. Hempel CG 1952. Fundamentals of Concept Formation in Empirical Science. (Volume 2, #7 of Foundations of the Unity of Science. Toward an International Encyclopedia of Unified Science). University of Chicago Press, p. 36.
  51. Polanyi M. 1958. Personal Knowledge. Towards a Post-Critical Philosophy. London: Routledge & Kegan Paul, p. 4.
  52. Galileo Galilei, The Assayer, as translated by Stillman Drake(1957), Discoveries and Opinions of Galileo pp. 237–38.
  53. Hacking I. 1983. Representing and Intervening. Cambridge University Press, p. 219.
  54. Koga J and Yamagiwa M (2006). Radiation reaction effects in ultrahigh irradiance laser pulse interactions with multiple electrons.
  55. Plass, G.N., 1956, The Carbon Dioxide Theory of Climatic Change, Tellus VIII, 2. (1956), pp. 140–54.

23 thoughts on “What is a Scientific Theory?

  1. bedava says:

    I am truly glad to glance at this wblog posts which includes lots of helpful information, thaznks for providing these data. Cosette Iain Klockau

  2. web-dl says:

    Hello. This post was extremely fascinating, particularly since I was searching for thoughts on this matter last Wednesday. Liane Natal Celtic

  3. bluray says:

    You made a few good points there. I did a search on the subject and found the majority of people will consent with your blog. Annalise Kermit Cleaves

  4. turkce says:

    Having read this I thought it was very informative. I appreciate you taking the time and energy to put this short article together. I once again find myself personally spending a significant amount of time both reading and commenting. But so what, it was still worth it! Ardine Mischa Lesya

Leave a Reply

Your email address will not be published. Required fields are marked *