Jump to content

Scientific theory: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
m Reverted edits by 97.100.33.36 (talk) to last revision by Lethe (HG)
Description on talk page
Line 3: Line 3:
A '''scientific theory''' is a set of [[Scientific law|principles]] that explain and predict [[Phenomenon|phenomena]].<ref>{{cite book|last=Myers|first=David G.|title=Psychology|year=2007|publisher=Worth Publishers|location=New York, NY|isbn=978-0-7167-6428-1|pages=24–25}}</ref> Scientists create scientific theories with the [[scientific method]], when they are originally proposed as [[hypothesis|hypotheses]] and tested for accuracy through [[Observation|observations]] and [[Experiment|experiments]].<ref>{{cite web|last=Schafersman|first=Steven D.|title=An Introduction to Science|url=http://www.geo.sunysb.edu/esp/files/scientific-method.html}}</ref>
A '''scientific theory''' is a set of [[Scientific law|principles]] that explain and predict [[Phenomenon|phenomena]].<ref>{{cite book|last=Myers|first=David G.|title=Psychology|year=2007|publisher=Worth Publishers|location=New York, NY|isbn=978-0-7167-6428-1|pages=24–25}}</ref> Scientists create scientific theories with the [[scientific method]], when they are originally proposed as [[hypothesis|hypotheses]] and tested for accuracy through [[Observation|observations]] and [[Experiment|experiments]].<ref>{{cite web|last=Schafersman|first=Steven D.|title=An Introduction to Science|url=http://www.geo.sunysb.edu/esp/files/scientific-method.html}}</ref>


==Characteristics of theories==

===Essential criteria===
The defining characteristic of a scientific theory is that it makes [[falsifiable]] or testable [[Predictive power|predictions]]. The relevance and specificity of those predictions determine how potentially useful the theory is. A would-be theory that makes no predictions is not a useful theory. Predictions not sufficiently specific to be tested are similarly not useful. In both cases, the term "theory" is hardly applicable.

In practice a body of descriptions of [[knowledge]] is usually only called a theory once it has a minimum empirical basis, according to certain criteria:
* It is consistent with pre-existing theory, to the extent the pre-existing theory was experimentally verified, though it will often show pre-existing theory to be wrong in an exact sense.
* It is supported by many strands of evidence, rather than a single foundation, ensuring it is probably a good approximation, if not totally correct.

===Non-essential criteria===
Additionally, a theory is generally only taken seriously if:
* It is tentative, correctable, and dynamic in allowing for changes as new facts are discovered, rather than asserting certainty.
* It is among the most [[Parsimony#Science|parsimonious]] explanations, sparing in proposed entities or explanations—commonly referred to as passing the [[Occam's razor]] test. (Since there is no generally accepted objective definition of parsimony, this is not a strict , but some theories are much less economical than others.)

This is true of such established theories as [[special relativity|special]] and [[general relativity]], [[quantum mechanics]], [[plate tectonics]], [[evolution]], etc. Theories considered scientific meet at least most, but ideally all, of these extra criteria.

In [[chemistry]], there are many [[acid-base theories]] which, highly divergent explanations of what "really" makes acids acids and bases bases are very useful for describing the phenomenology of certain chemical reactions which fall under the concept of "acid-base reaction". In a sense, the notion of ''generalized acid-base reaction'' is not precisely defined, and therefore theories about what gives rise to acid-base chemistry are "inexact"; nonetheless, they are useful scientific theories.

===Definitions from scientific organizations===
In pedagogical contexts or in official pronouncements by scientific organizations a definition such as the following may be promulgated

<blockquote>The formal scientific definition of theory is quite different from the everyday meaning of the word. It refers to a comprehensive explanation of some aspect of nature that is supported by a vast body of evidence. Many scientific theories are so well established that no new evidence is likely to alter them substantially. For example, no new evidence will demonstrate that the Earth does not orbit around the sun (heliocentric theory), or that living things are not made of cells (cell theory), that matter is not composed of atoms, or that the surface of the Earth is not divided into solid plates that have moved over geological timescales (the theory of plate tectonics). One of the most useful properties of scientific theories is that they can be used to make predictions about natural events or phenomena that have not yet been observed.<ref>[[United States National Academy of Sciences|National Academy of Sciences]] (2005), ''Science, Evolution, and Creationism'', a [http://www.nap.edu/catalog.php?record_id=11876#toc brochure on the book of the same title].</ref></blockquote>

According to this definition, a theory must be well supported by evidence. Furthermore, the term theory would not be appropriate for describing untested but intricate hypotheses or even scientific models. Consumers of science may find the above definition useful when evaluating the validity and/or efficacy of a theory.


==About theories==
==About theories==
Line 28: Line 52:


===Assumptions to formulate a theory===
===Assumptions to formulate a theory===
This is a view shared by [[Isaac Asimov]]. In ''Understanding Physics'', Asimov spoke of theories as "arguments" where one deduces a "scheme" or model. Arguments or theories always begin with some premises—"arbitrary elements" as Hawking calls them (see above)—which are here described as "assumptions". An assumption according to Asimov is...
This is a view shared by [[Isaac Asimov]]. In ''Understanding Physics'', Asimov spoke of theories as "arguments" where one deduces a "scheme" or model. Arguments or theories always begin with some premises—"arbitrary elements" as Hawking calls them (see above)—which are here described as "assumptions". An assumption according to Asimov is


<blockquote>...something accepted without proof, and it is incorrect to speak of an assumption as either true or false, since there is no way of proving it to be either (If there were, it would no longer be an assumption). It is better to consider assumptions as either useful or useless, depending on whether deductions made from them corresponded to reality. ... On the other hand, it seems obvious that assumptions are the weak points in any argument, as they have to be accepted on faith in a [[philosophy]] of science that prides itself on its rationalism. Since we must start somewhere, we must have assumptions, but at least let us have as few assumptions as possible.</blockquote>
<blockquote>...something accepted without proof, and it is incorrect to speak of an assumption as either true or false, since there is no way of proving it to be either (If there were, it would no longer be an assumption). It is better to consider assumptions as either useful or useless, depending on whether deductions made from them corresponded to reality. ... On the other hand, it seems obvious that assumptions are the weak points in any argument, as they have to be accepted on faith in a [[philosophy]] of science that prides itself on its rationalism. Since we must start somewhere, we must have assumptions, but at least let us have as few assumptions as possible.</blockquote>
Line 52: Line 76:


Engineering practice makes a distinction between "mathematical models" and "physical models."{{Citation needed|date=January 2012}}
Engineering practice makes a distinction between "mathematical models" and "physical models."{{Citation needed|date=January 2012}}

===Essential criteria===
The defining characteristic of a scientific theory is that it makes [[falsifiable]] or testable [[Predictive power|predictions]]. The relevance and specificity of those predictions determine how potentially useful the theory is. A would-be theory that makes no predictions that can be observed is not a useful theory. Predictions not sufficiently specific to be tested are similarly not useful. In both cases, the term "theory" is hardly applicable.

In practice a body of descriptions of [[knowledge]] is usually only called a theory once it has a minimum empirical basis, according to certain criteria:
* It is consistent with pre-existing theory, to the extent the pre-existing theory was experimentally verified, though it will often show pre-existing theory to be wrong in an exact sense.
* It is supported by many strands of evidence, rather than a single foundation, ensuring it is probably a good approximation, if not totally correct.

===Non-essential criteria===
Additionally, a theory is generally only taken seriously if:
* It is tentative, correctable, and dynamic in allowing for changes as new facts are discovered, rather than asserting certainty.
* It is among the most [[Parsimony#Science|parsimonious]] explanations, sparing in proposed entities or explanations—commonly referred to as passing the [[Occam's razor]] test. (Since there is no generally accepted objective definition of parsimony, this is not a strict criteria, but some theories are much less economical than others.)

This is true of such established theories as [[special relativity|special]] and [[general relativity]], [[quantum mechanics]], [[plate tectonics]], [[evolution]], etc. Theories considered scientific meet at least most, but ideally all, of these extra criteria.

Theories do not have to be perfectly accurate to be scientifically useful.
* The predictions made by [[classical mechanics]] are known to be inaccurate in the relatistivic realm, but they suffice at the low velocities of common experience.
* In [[chemistry]], there are many [[acid-base theories]] which, while providing highly divergent explanations of what "really" makes acids acids and bases bases, they are very useful for describing the phenomenology of certain chemical reactions which fall under the concept of "acid-base reaction". In a sense, the notion of ''generalized acid-base reaction'' is not precisely defined, and therefore theories about what gives rise to acid-base chemistry are "inexact"; nonetheless, they are useful scientific theories.


===Images, analogies, and metaphors of theory===
===Images, analogies, and metaphors of theory===
Line 110: Line 116:
==In physics==
==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.)—which 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]]. Note that the specific theoretical aspects of classical electromagnetic theory, which have been consistently and successfully replicated for well over a century, are termed "laws of electromagnetism", reflecting that they are today taken for granted. 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, it has never been observed directly, but its effects on periodic motion of charges ''in a time-averaged sense'' is detectable in [[synchrotron]]s. Some researchers are now considering the possibility of experiments that could observe the effects of this force at the instantaneous (i.e. not averaged over periods of cyclical motion) level. <ref>http://epsppd.epfl.ch/Roma/pdf/P1_031.pdf</ref><ref>http://scitation.aip.org/getabs/servlet/GetabsServlet?prog=normal&id=PHPAEN000013000011113106000001&idtype=cvips&gifs=yes&ref=no</ref>
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.)—which 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]]. Note that the specific theoretical aspects of classical electromagnetic theory, which have been consistently and successfully replicated for well over a century, are termed "laws of electromagnetism", reflecting that they are today taken for granted. 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, it has never been observed directly, but its effects on periodic motion of charges ''in a time-averaged sense'' is detectable in [[synchrotron]]s. Some researchers are now considering the possibility of experiments that could observe the effects of this force at the instantaneous (i.e. not averaged over periods of cyclical motion) level. <ref>http://epsppd.epfl.ch/Roma/pdf/P1_031.pdf</ref><ref>http://scitation.aip.org/getabs/servlet/GetabsServlet?prog=normal&id=PHPAEN000013000011113106000001&idtype=cvips&gifs=yes&ref=no</ref>

==Pedagogical definition==
In pedagogical contexts or in official pronouncements by scientific organizations a definition such as the following may be promulgated.

According to the [[United States National Academy of Sciences]],
<blockquote>The formal scientific definition of theory is quite different from the everyday meaning of the word. It refers to a comprehensive explanation of some aspect of nature that is supported by a vast body of evidence. Many scientific theories are so well established that no new evidence is likely to alter them substantially. For example, no new evidence will demonstrate that the Earth does not orbit around the sun (heliocentric theory), or that living things are not made of cells (cell theory), that matter is not composed of atoms, or that the surface of the Earth is not divided into solid plates that have moved over geological timescales (the theory of plate tectonics). One of the most useful properties of scientific theories is that they can be used to make predictions about natural events or phenomena that have not yet been observed.<ref>[[United States National Academy of Sciences|National Academy of Sciences]] (2005), ''Science, Evolution, and Creationism'', a [http://www.nap.edu/catalog.php?record_id=11876#toc brochure on the book of the same title].</ref></blockquote>

According to this definition, a theory must be well supported by evidence. Furthermore, the term theory would not be appropriate for describing untested but intricate hypotheses or even scientific models. Consumers of science may find the above definition useful when evaluating the validity and/or efficacy of a theory.


==The term ''theoretical''==
==The term ''theoretical''==

Revision as of 07:55, 8 May 2012

A scientific theory is a set of principles that explain and predict phenomena.[1] Scientists create scientific theories with the scientific method, when they are originally proposed as hypotheses and tested for accuracy through observations and experiments.[2]

Characteristics of theories

Essential criteria

The defining characteristic of a scientific theory is that it makes falsifiable or testable predictions. The relevance and specificity of those predictions determine how potentially useful the theory is. A would-be theory that makes no observable predictions is not a useful theory. Predictions not sufficiently specific to be tested are similarly not useful. In both cases, the term "theory" is hardly applicable.

In practice, a body of descriptions of knowledge is usually only called a theory once it has a minimum empirical basis, according to certain criteria:

  • It is consistent with pre-existing theory, to the extent the pre-existing theory was experimentally verified, though it will often show pre-existing theory to be wrong in an exact sense.
  • It is supported by many strands of evidence, rather than a single foundation, ensuring it is probably a good approximation, if not totally correct.

Non-essential criteria

Additionally, a theory is generally only taken seriously if:

  • It is tentative, correctable, and dynamic in allowing for changes as new facts are discovered, rather than asserting certainty.
  • It is among the most parsimonious explanations, sparing in proposed entities or explanations—commonly referred to as passing the Occam's razor test. (Since there is no generally accepted objective definition of parsimony, this is not a strict criterion, but some theories are much less economical than others.)

This is true of such established theories as special and general relativity, quantum mechanics, plate tectonics, evolution, etc. Theories considered scientific meet at least most, but ideally all, of these extra criteria.

Theories do not have to be perfectly accurate to be scientifically useful. For example, the predictions made by classical mechanics are known to be inaccurate in the relatistivic realm, but they suffice at the low velocities of common experience. In chemistry, there are many acid-base theories which, provide highly divergent explanations of what "really" makes acids acids and bases bases but are very useful for describing the phenomenology of certain chemical reactions which fall under the concept of "acid-base reaction". In a sense, the notion of generalized acid-base reaction is not precisely defined, and therefore theories about what gives rise to acid-base chemistry are "inexact"; nonetheless, they are useful scientific theories.

Definitions from scientific organizations

In pedagogical contexts or in official pronouncements by scientific organizations, a definition such as the following (from the United States National Academy of Sciences) may be promulgated:

The formal scientific definition of theory is quite different from the everyday meaning of the word. It refers to a comprehensive explanation of some aspect of nature that is supported by a vast body of evidence. Many scientific theories are so well established that no new evidence is likely to alter them substantially. For example, no new evidence will demonstrate that the Earth does not orbit around the sun (heliocentric theory), or that living things are not made of cells (cell theory), that matter is not composed of atoms, or that the surface of the Earth is not divided into solid plates that have moved over geological timescales (the theory of plate tectonics). One of the most useful properties of scientific theories is that they can be used to make predictions about natural events or phenomena that have not yet been observed.[3]

According to this definition, a theory must be well supported by evidence. Furthermore, the term theory would not be appropriate for describing untested but intricate hypotheses or even scientific models. Consumers of science may find the above definition useful when evaluating the validity and/or efficacy of a theory.

About theories

Theories as axioms

The logical positivists thought of scientific theories as statements in a kind of formal language. Mathematics 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. Theoretical concepts such as atoms and radio waves that cannot be directly observed by humans are incorporated in the theories. These theories function as axioms; predicted observation sentences are derived from the theories much like theorems are derived in Euclidean geometry. Observation sentences are then tested to verify the theories.

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 1970's. Also, scientists may often find it easier to think in terms of models than in terms of equations.

Theories as models

The previously dominant position in philosophy of science -- the received view of theories -- which was prevalent in logical empiricism has in the course of the second half of the 20th century been replaced by the semantic view of theories which identifies scientific theories with models rather than propositions.[4]

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. Functions defined on this set of objects, e.g., Newton's Law of Gravitation, determine how the positions and velocities change with time. This model can be tested; astronomers can verify that the positions of the model's objects over time match the actual positions of the planets. For most planets, they do match; for Mercury, if the model is based on Newton's Law of Gravitation, they don't.

In this approach, the theory is the model. More precisely, a theory is a collection of similar models. One can use language to describe a model; however, the theory is the model, not the description of the model.

The word "semantic" refers to the way that a model represents the real world.

Description and prediction

Echoing the scientific philosopher Karl Popper, Stephen Hawking in A Brief History of Time 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 goes on to state, "Any physical theory is always provisional, in the sense that it is only a hypothesis; you can never prove it. No matter how many times the results of experiments agree with some theory, you can never be sure that the next time the result will not contradict the theory. On the other hand, you can disprove a theory by finding even a single observation that disagrees with the predictions of the theory." The "unprovable but falsifiable" nature of theories is a necessary consequence of using inductive logic.

Assumptions to formulate a theory

This is a view shared by Isaac Asimov. In Understanding Physics, Asimov spoke of theories as "arguments" where one deduces a "scheme" or model. Arguments or theories always begin with some premises—"arbitrary elements" as Hawking calls them (see above)—which are here described as "assumptions". An assumption according to Asimov is:

...something accepted without proof, and it is incorrect to speak of an assumption as either true or false, since there is no way of proving it to be either (If there were, it would no longer be an assumption). It is better to consider assumptions as either useful or useless, depending on whether deductions made from them corresponded to reality. ... On the other hand, it seems obvious that assumptions are the weak points in any argument, as they have to be accepted on faith in a philosophy of science that prides itself on its rationalism. Since we must start somewhere, we must have assumptions, but at least let us have as few assumptions as possible.

Example: Special Theory of Relativity

As an example of the use of assumptions to formulate a theory, consider how Albert Einstein put forth his Special Theory of Relativity. He took two phenomena that had been observed — that the "addition of velocities" is valid (Galilean transformation), and that light did not appear to have an "addition of velocities" (Michelson-Morley experiment). He assumed both observations to be correct, and formulated his theory, based on these assumptions, by simply altering the Galilean transformation to accommodate the lack of addition of velocities with regard to the speed of light. The model created in his theory is, therefore, based on the assumption that light maintains a constant velocity (or more commonly: the speed of light is a constant).

Example: Ptolemy

An example of how theories are models can be seen from theories on the planetary system. The Greeks formulated theories, which the astronomer Ptolemy recorded. In Ptolemy's planetary model, the earth was at the center, the planets and the sun made circular orbits around the earth, and the stars were on a sphere outside of the orbits of the planet and the earth. Retrograde motion of the planets was explained by smaller circular orbits of individual planets. This could be illustrated as a model, and could even be built into a literal model. Mathematical calculations could be made that predicted, to a great degree of accuracy, where the planets would be. His model of the planetary system survived for over 1500 years until the time of Copernicus. So one can see that a theory is a "model of reality" that explains certain scientific facts; yet, the theory may not be a satisfactory picture of reality. Another, more acceptable, theory can later replace the previous model. For example, compare the early Ptolemaic theory, with its circles within circles, fine adjustments, and numerous ad hoc assumptions, to the Copernican theory; the former is overly complex, while the latter is simple and parsimonious. Or a new theory can be used to modify an older theory as when Einstein modified Newtonian mechanics (which is still used for computing planetary orbits or modeling spacecraft trajectories) with his theories of relativity.

Differences between theory and model

Central to the nature of models, from general models to scale models, is the employment of representation (literally, "re-presentation") to describe 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. In most ways that matter, the scale model of a house is not a house. Several commentators (e.g., Reese & Overton 1970; Lerner, 1998; Lerner & Teti, 2005, in the context of modeling human behavior) have stated that the important difference between theories and models is that the first is explanatory as well as descriptive, while the second is only descriptive (although still predictive in a more limited sense). General models and theories, according to philosopher Stephen Pepper (1948)—who also distinguishes between theories and models—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."[citation needed]

Images, analogies, and metaphors of theory

Sometimes images, analogies, and metaphors are more powerful for motivating our understanding of a highly abstract topic than rigid and detached philosophical analysis. Philosophers realize this, and capitalize on it. For instance, regarding the structure of scientific theory, the influential logical empiricist Carl Gustav Hempel likened a theory to a network:

A scientific theory might therefore be likened 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. [5]

Michael Polanyi made a powerful 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. [6]

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. Galileo Galilei's 1623 polemic The Assayer asserts that the universe is an open book; however, the language of this book is mathematics, and those ignorant of this are doomed to stumble in a darkened labyrinth. And it is this metaphor that the contemporary philosopher of science Ian Hacking seems to allude to in the following multi-faceted passage:

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. Far from being untidy, this is New World Leibnizianism. 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.[7]

Criteria for scientific status

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

  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, are still upheld by their admirers—for example by introducing ad hoc some auxiliary assumption, or by reinterpreting the theory ad 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. (I later describe such a rescuing operation as a "conventionalist twist" or a "conventionalist stratagem".)

One can sum up all this by saying that according to Popper, the criterion of the scientific status of a theory is its falsifiability, or refutability, or testability.[citation needed]

Several philosophers and historians of science have, however, argued that Popper's definition of theory as a set of falsifiable statements is wrong [8] 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." [9] He also takes into account Hempel and Quine's critiques of Popper, to the effect that scientific theories include statements that cannot be falsified (presumably what Hawking alluded to as arbitrary elements), and the point 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:

  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" (1982: 47).
  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" (1982: 47–48).
  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" (1982: 46) (e.g. 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 good theory includes statements that have (in his terms) "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—observational consequence.

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.)—which 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. Note that the specific theoretical aspects of classical electromagnetic theory, which have been consistently and successfully replicated for well over a century, are termed "laws of electromagnetism", reflecting that they are today taken for granted. 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, it has never been observed directly, but its effects on periodic motion of charges in a time-averaged sense is detectable in synchrotrons. Some researchers are now considering the possibility of experiments that could observe the effects of this force at the instantaneous (i.e. not averaged over periods of cyclical motion) level. [10][11]

The term theoretical

The term theoretical is sometimes informally used in lieu of hypothetical to describe a result that is predicted by theory but has not yet been adequately tested by observation or experiment. It is not uncommon for a theory to produce predictions that are later confirmed or proven incorrect by experiment. By inference, a prediction proved incorrect by experiment demonstrates the hypothesis is invalid. This either means the theory is incorrect, or the experimental conjecture was wrong and the theory did not predict the hypothesis.

Scientific laws

Scientific laws are similar to scientific theories in that they are principles that can be used to predict the behavior of the natural world. Both scientific laws and scientific theories are typically well-supported by observations and/or experimental evidence. Usually scientific laws refer to rules for how nature will behave under certain conditions.[12] Scientific theories are more overarching explanations of how nature works and why it exhibits certain characteristics.

A common misconception is that scientific theories are rudimentary ideas that will eventually graduate into scientific laws when enough data and evidence has 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.[13]

See also

References

  1. ^ Myers, David G. (2007). Psychology. New York, NY: Worth Publishers. pp. 24–25. ISBN 978-0-7167-6428-1.
  2. ^ Schafersman, Steven D. "An Introduction to Science".
  3. ^ National Academy of Sciences (2005), Science, Evolution, and Creationism, a brochure on the book of the same title.
  4. ^ Suppe, Frederick 1977 The Structure of Scientific Theories p. 221f
  5. ^ 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.
  6. ^ Polanyi M. 1958. Personal Knowledge. Towards a Post-Critical Philosophy. London: Routledge & Kegan Paul, p. 4.
  7. ^ Hacking I. 1983. Representing and Intervening. Cambridge University Press, p. 219.
  8. ^ 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
  9. ^ Philip Kitcher 1982 Abusing Science: The Case Against Creationism. Page 45 Cambridge: The MIT Press
  10. ^ http://epsppd.epfl.ch/Roma/pdf/P1_031.pdf
  11. ^ http://scitation.aip.org/getabs/servlet/GetabsServlet?prog=normal&id=PHPAEN000013000011113106000001&idtype=cvips&gifs=yes&ref=no
  12. ^ See the article on Physical law, for example.
  13. ^ theory

References

  • Popper, Karl (1963), Conjectures and Refutations, Routledge and Kegan Paul, London, UK, pp. 33–39. Reprinted in Theodore Schick (ed., 2000), Readings in the Philosophy of Science, Mayfield Publishing Company, Mountain View, Calif., pp. 9–13.
  • Chairman of Biology and Kennesaw State Ronald Matson's webpage comparing scientific laws and theories
  • Hawking, Stephen (1996). "The Illustrated A Brief History of Time" (Updated and expanded ed.). New York: Bantam Books, p. 15.
  • Mohr, Johnathon (2008). "Revelations and Implications of the Failure of Pragmatism: The Hijacking of Knowledge Creation by the Ivory Tower". New York: Ballantine Books. pp. 87–192.