If you're seeing this message, it means we're having trouble loading external resources on our website.

If you're behind a web filter, please make sure that the domains *.kastatic.org and *.kasandbox.org are unblocked.

To log in and use all the features of Khan Academy, please enable JavaScript in your browser.

Biology library

Course: biology library   >   unit 1, the scientific method.

  • Controlled experiments
  • The scientific method and experimental design

Introduction

  • Make an observation.
  • Ask a question.
  • Form a hypothesis , or testable explanation.
  • Make a prediction based on the hypothesis.
  • Test the prediction.
  • Iterate: use the results to make new hypotheses or predictions.

Scientific method example: Failure to toast

1. make an observation..

  • Observation: the toaster won't toast.

2. Ask a question.

  • Question: Why won't my toaster toast?

3. Propose a hypothesis.

  • Hypothesis: Maybe the outlet is broken.

4. Make predictions.

  • Prediction: If I plug the toaster into a different outlet, then it will toast the bread.

5. Test the predictions.

  • Test of prediction: Plug the toaster into a different outlet and try again.
  • If the toaster does toast, then the hypothesis is supported—likely correct.
  • If the toaster doesn't toast, then the hypothesis is not supported—likely wrong.

Logical possibility

Practical possibility, building a body of evidence, 6. iterate..

  • Iteration time!
  • If the hypothesis was supported, we might do additional tests to confirm it, or revise it to be more specific. For instance, we might investigate why the outlet is broken.
  • If the hypothesis was not supported, we would come up with a new hypothesis. For instance, the next hypothesis might be that there's a broken wire in the toaster.

Want to join the conversation?

  • Upvote Button navigates to signup page
  • Downvote Button navigates to signup page
  • Flag Button navigates to signup page

Incredible Answer

University of Lethbridge

Science Toolkit

What is a Hypothesis?

A hypothesis (plural: hypotheses) is in its simplest form nothing more than an idea about how the world works. For example, “the moon is made of green cheese” is a valid hypothesis. But there are several characteristics which separate useful scientific hypotheses from those which are impractical. First and foremost, a hypothesis must be testable. We must (at least in principle) be able to design an experiment which will allow us to determine whether the hypothesis is false. Keep in mind that we can never prove any hypothesis is completely true because we can always

English chemist Robert Boyle (1627-1691) was one of the first scientists to explicitly adopt a program of hypothesis testing. imagine new circumstances in which it has not been tested, or other possible explanations for the results we have obtained. It is much easier to show, with a high degree of confidence, that a hypothesis is false. If it is not consistent with the results of a well designed and executed experiment, we are forced to accept that the hypothesis is false. If a hypothesis is not falsifiable, it is outside the realm of science. Note that the “green-cheese hypothesis” meets this test. A hypothesis should also be plausible. That is, the hypothesis, should be consistent with what we already know about the subject being investigated, and its parts should be logically and mathematically sound. We often celebrate the creative spark by which new hypotheses come to light. But typically that moment of inspiration follows a great deal of perspiration racked up in a thorough review of previous research in the subject. A hypothesis may in the end be a guess, but it should be the best guess possible. Given what we know about astronomy (and cheese production) the”green-cheese hypothesis” is not plausible, and not worth investing much of our time and resources in testing. What are predictions?

The predictions of a hypothesis set out what we expect to see if the hypothesis is true. (This is where we use deductive “If-Then” logic.) Experiments are designed to test specific predictions of the hypothesis. The “green-cheese hypothesis” predicts that material collected from the moon would contain milk proteins and fungi. These predictions could be tested by bringing material back from the moon, and testing its chemical structure. The hypothesis also makes predictions about the wavelengths of light reflected from the moon, a field called spectroscopy. (These predictions have actually been tested, believe it or not. Needless to say the hypothesis was not supported!) Three main factors make a prediction useful in testing a hypothesis: The prediction should be specific to the hypothesis (i.e. no other hypotheses make the same prediction). If several hypotheses predict the same outcome of an experiment, we will need to do further experiments to distinguish between them. The prediction should provide results which are unambiguous. It should be practical and economically feasible to run the experiment. A prediction is really nothing more than a simpler hypothesis — practical to test — derived from a larger hypothesis. Note that a prediction does not have be about the future, but it does have to apply to a situation we have not looked at yet. We are free to use the results of previous experiments to develop a new hypothesis, but we can’t then test our predictions against the results of those old experiments — to do so would be arguing in circles. Are theories different from hypotheses?

A “theory” has no formal definition in science (Style Manual Committee, CBE 1994). Hypotheses which have considerable support from experiments, and which are useful in explaining a fairly wide range of phenomena, are “upgraded” to theories, for example Darwin’s Theory of Evolution by Natural Selection, or the Theory of Plate Tectonics (which explains the movement of continents). So a theory is simply a well-tested and widely useful hypothesis, and there’s no strict rules defining when a hypothesis becomes a theory. Theories which are extremely well supported by experiments, particularly those which can be expressed as simple mathematical equations, are often called laws, e.g. Newton’s Law of Gravitation, Kepler’s Laws of Motion, or Mendel’s Laws of Genetics. Again, no one has yet laid out a strict set of rules for defining a natural law. A final term that scientists use to describe their ideas is a model. This dates back to the time when physical models were one of the few tools researchers had in investigating phenomena which were too big or too small to manipulate directly. Physical models are still used in science. Francis Crick and James Watson used a scale model of a DNA molecule to help them deduce its structure (Giere 1997). But scientists also use mathematical models to help them understand how different factors will interact. The development of computers has vastly increased the scope of mathematical models, and made them accessible even to non-mathematicians. Why are hypotheses important?

Philosopher of science Karl Popper likened a hypothesis to a searchlight, which the researcher shines on the relevant portion of nature (Davies 1973). It tells us which experiments are the important ones to perform, and which observations the important ones to make, out of an infinite number of possibilities. Without hypotheses scientists would be reduced to bean counters, and science to a collection of facts without organization or purpose. A hypothesis is the cornerstone used in building an elegant, structured body of knowledge from the apparent chaos of nature. (So they’re pretty important, eh!)

Library homepage

  • school Campus Bookshelves
  • menu_book Bookshelves
  • perm_media Learning Objects
  • login Login
  • how_to_reg Request Instructor Account
  • hub Instructor Commons
  • Download Page (PDF)
  • Download Full Book (PDF)
  • Periodic Table
  • Physics Constants
  • Scientific Calculator
  • Reference & Cite
  • Tools expand_more
  • Readability

selected template will load here

This action is not available.

Social Sci LibreTexts

2.4: Developing a Hypothesis

  • Last updated
  • Save as PDF
  • Page ID 16104

Learning Objectives

  • Distinguish between a theory and a hypothesis.
  • Discover how theories are used to generate hypotheses and how the results of studies can be used to further inform theories.
  • Understand the characteristics of a good hypothesis.

Theories and Hypotheses

Before describing how to develop a hypothesis it is important to distinguish between a theory and a hypothesis. A theory is a coherent explanation or interpretation of one or more phenomena. Although theories can take a variety of forms, one thing they have in common is that they go beyond the phenomena they explain by including variables, structures, processes, functions, or organizing principles that have not been observed directly. Consider, for example, Zajonc’s theory of social facilitation and social inhibition. He proposed that being watched by others while performing a task creates a general state of physiological arousal, which increases the likelihood of the dominant (most likely) response. So for highly practiced tasks, being watched increases the tendency to make correct responses, but for relatively unpracticed tasks, being watched increases the tendency to make incorrect responses. Notice that this theory—which has come to be called drive theory—provides an explanation of both social facilitation and social inhibition that goes beyond the phenomena themselves by including concepts such as “arousal” and “dominant response,” along with processes such as the effect of arousal on the dominant response.

Outside of science, referring to an idea as a theory often implies that it is untested—perhaps no more than a wild guess. In science, however, the term theory has no such implication. A theory is simply an explanation or interpretation of a set of phenomena. It can be untested, but it can also be extensively tested, well supported, and accepted as an accurate description of the world by the scientific community. The theory of evolution by natural selection, for example, is a theory because it is an explanation of the diversity of life on earth—not because it is untested or unsupported by scientific research. On the contrary, the evidence for this theory is overwhelmingly positive and nearly all scientists accept its basic assumptions as accurate. Similarly, the “germ theory” of disease is a theory because it is an explanation of the origin of various diseases, not because there is any doubt that many diseases are caused by microorganisms that infect the body.

A hypothesis , on the other hand, is a specific prediction about a new phenomenon that should be observed if a particular theory is accurate. It is an explanation that relies on just a few key concepts. Hypotheses are often specific predictions about what will happen in a particular study. They are developed by considering existing evidence and using reasoning to infer what will happen in the specific context of interest. Hypotheses are often but not always derived from theories. So a hypothesis is often a prediction based on a theory but some hypotheses are a-theoretical and only after a set of observations have been made, is a theory developed. This is because theories are broad in nature and they explain larger bodies of data. So if our research question is really original then we may need to collect some data and make some observation before we can develop a broader theory.

Theories and hypotheses always have this if-then relationship. “ If drive theory is correct, then cockroaches should run through a straight runway faster, and a branching runway more slowly, when other cockroaches are present.” Although hypotheses are usually expressed as statements, they can always be rephrased as questions. “Do cockroaches run through a straight runway faster when other cockroaches are present?” Thus deriving hypotheses from theories is an excellent way of generating interesting research questions.

But how do researchers derive hypotheses from theories? One way is to generate a research question using the techniques discussed in this chapter and then ask whether any theory implies an answer to that question. For example, you might wonder whether expressive writing about positive experiences improves health as much as expressive writing about traumatic experiences. Although this question is an interesting one on its own, you might then ask whether the habituation theory—the idea that expressive writing causes people to habituate to negative thoughts and feelings—implies an answer. In this case, it seems clear that if the habituation theory is correct, then expressive writing about positive experiences should not be effective because it would not cause people to habituate to negative thoughts and feelings. A second way to derive hypotheses from theories is to focus on some component of the theory that has not yet been directly observed. For example, a researcher could focus on the process of habituation—perhaps hypothesizing that people should show fewer signs of emotional distress with each new writing session.

Among the very best hypotheses are those that distinguish between competing theories. For example, Norbert Schwarz and his colleagues considered two theories of how people make judgments about themselves, such as how assertive they are (Schwarz et al., 1991) [1] . Both theories held that such judgments are based on relevant examples that people bring to mind. However, one theory was that people base their judgments on the number of examples they bring to mind and the other was that people base their judgments on how easily they bring those examples to mind. To test these theories, the researchers asked people to recall either six times when they were assertive (which is easy for most people) or 12 times (which is difficult for most people). Then they asked them to judge their own assertiveness. Note that the number-of-examples theory implies that people who recalled 12 examples should judge themselves to be more assertive because they recalled more examples, but the ease-of-examples theory implies that participants who recalled six examples should judge themselves as more assertive because recalling the examples was easier. Thus the two theories made opposite predictions so that only one of the predictions could be confirmed. The surprising result was that participants who recalled fewer examples judged themselves to be more assertive—providing particularly convincing evidence in favor of the ease-of-retrieval theory over the number-of-examples theory.

Theory Testing

The primary way that scientific researchers use theories is sometimes called the hypothetico-deductive method (although this term is much more likely to be used by philosophers of science than by scientists themselves). A researcher begins with a set of phenomena and either constructs a theory to explain or interpret them or chooses an existing theory to work with. He or she then makes a prediction about some new phenomenon that should be observed if the theory is correct. Again, this prediction is called a hypothesis. The researcher then conducts an empirical study to test the hypothesis. Finally, he or she reevaluates the theory in light of the new results and revises it if necessary. This process is usually conceptualized as a cycle because the researcher can then derive a new hypothesis from the revised theory, conduct a new empirical study to test the hypothesis, and so on. As Figure 2.2 shows, this approach meshes nicely with the model of scientific research in psychology presented earlier in the textbook—creating a more detailed model of “theoretically motivated” or “theory-driven” research.

4.4.png

As an example, let us consider Zajonc’s research on social facilitation and inhibition. He started with a somewhat contradictory pattern of results from the research literature. He then constructed his drive theory, according to which being watched by others while performing a task causes physiological arousal, which increases an organism’s tendency to make the dominant response. This theory predicts social facilitation for well-learned tasks and social inhibition for poorly learned tasks. He now had a theory that organized previous results in a meaningful way—but he still needed to test it. He hypothesized that if his theory was correct, he should observe that the presence of others improves performance in a simple laboratory task but inhibits performance in a difficult version of the very same laboratory task. To test this hypothesis, one of the studies he conducted used cockroaches as subjects (Zajonc, Heingartner, & Herman, 1969) [2] . The cockroaches ran either down a straight runway (an easy task for a cockroach) or through a cross-shaped maze (a difficult task for a cockroach) to escape into a dark chamber when a light was shined on them. They did this either while alone or in the presence of other cockroaches in clear plastic “audience boxes.” Zajonc found that cockroaches in the straight runway reached their goal more quickly in the presence of other cockroaches, but cockroaches in the cross-shaped maze reached their goal more slowly when they were in the presence of other cockroaches. Thus he confirmed his hypothesis and provided support for his drive theory. (Zajonc also showed that drive theory existed in humans (Zajonc & Sales, 1966) [3] in many other studies afterward).

Incorporating Theory into Your Research

When you write your research report or plan your presentation, be aware that there are two basic ways that researchers usually include theory. The first is to raise a research question, answer that question by conducting a new study, and then offer one or more theories (usually more) to explain or interpret the results. This format works well for applied research questions and for research questions that existing theories do not address. The second way is to describe one or more existing theories, derive a hypothesis from one of those theories, test the hypothesis in a new study, and finally reevaluate the theory. This format works well when there is an existing theory that addresses the research question—especially if the resulting hypothesis is surprising or conflicts with a hypothesis derived from a different theory.

To use theories in your research will not only give you guidance in coming up with experiment ideas and possible projects, but it lends legitimacy to your work. Psychologists have been interested in a variety of human behaviors and have developed many theories along the way. Using established theories will help you break new ground as a researcher, not limit you from developing your own ideas.

Characteristics of a Good Hypothesis

There are three general characteristics of a good hypothesis. First, a good hypothesis must be testable and falsifiable . We must be able to test the hypothesis using the methods of science and if you’ll recall Popper’s falsifiability criterion, it must be possible to gather evidence that will disconfirm the hypothesis if it is indeed false. Second, a good hypothesis must be logical. As described above, hypotheses are more than just a random guess. Hypotheses should be informed by previous theories or observations and logical reasoning. Typically, we begin with a broad and general theory and use deductive reasoning to generate a more specific hypothesis to test based on that theory. Occasionally, however, when there is no theory to inform our hypothesis, we use inductive reasoning which involves using specific observations or research findings to form a more general hypothesis. Finally, the hypothesis should be positive. That is, the hypothesis should make a positive statement about the existence of a relationship or effect, rather than a statement that a relationship or effect does not exist. As scientists, we don’t set out to show that relationships do not exist or that effects do not occur so our hypotheses should not be worded in a way to suggest that an effect or relationship does not exist. The nature of science is to assume that something does not exist and then seek to find evidence to prove this wrong, to show that really it does exist. That may seem backward to you but that is the nature of the scientific method. The underlying reason for this is beyond the scope of this chapter but it has to do with statistical theory.

Key Takeaways

  • A theory is broad in nature and explains larger bodies of data. A hypothesis is more specific and makes a prediction about the outcome of a particular study.
  • Working with theories is not “icing on the cake.” It is a basic ingredient of psychological research.
  • Like other scientists, psychologists use the hypothetico-deductive method. They construct theories to explain or interpret phenomena (or work with existing theories), derive hypotheses from their theories, test the hypotheses, and then reevaluate the theories in light of the new results.
  • Practice: Find a recent empirical research report in a professional journal. Read the introduction and highlight in different colors descriptions of theories and hypotheses.
  • Schwarz, N., Bless, H., Strack, F., Klumpp, G., Rittenauer-Schatka, H., & Simons, A. (1991). Ease of retrieval as information: Another look at the availability heuristic. Journal of Personality and Social Psychology, 61 , 195–202.
  • Zajonc, R. B., Heingartner, A., & Herman, E. M. (1969). Social enhancement and impairment of performance in the cockroach. Journal of Personality and Social Psychology, 13 , 83–92.
  • Zajonc, R.B. & Sales, S.M. (1966). Social facilitation of dominant and subordinate responses. Journal of Experimental Social Psychology, 2 , 160-168.

Language selection

  • Français fr

WxT Search form

Science shorts 2: the scientific method, what is science.

Science is a way of seeking to explain the world around us, either by simply observing the world or by experimentally manipulating it in some way. Science proceeds by posing familiar questions: What? Where? When? How? and Why? Observations, estimates, and patterns are three different kinds of factual claims that describe what occurred, where it occurred or when it occurred. Answers to how and why questions are factual claims of the fourth kind, namely causal hypotheses .

Scientific hypotheses

What makes a causal hypothesis “scientific”? For both scientists and the courts, Footnote 1 scientific hypotheses are those that can be empirically tested.

Empirically testable hypotheses satisfy two conditions. First, the hypothesis must be refutable. Footnote 2 A refutable hypothesis is one for which there exists the logical possibility of observations that would be considered inconsistent with the hypothesis and hence, would lead us to conclude the hypothesis is false .

Second, empirical testability requires that contradictory observations be not only logically possible but capable of being collected in practice . Footnote 3

The question of whether an hypothesis is scientific is a different question than whether it is true . For example, the claim that God made life on earth could be true. But whatever its truth, this factual claim (of the fourth kind) is not scientific because there are no observations that are, even in principle, inconsistent with it: it fails to satisfy the criterion of refutability.

Testing scientific hypotheses: The scientific method

Testing a scientific hypothesis proceeds by three basic steps. The first involves formulation of a clear, unambiguous scientific hypothesis. The second step involves the design and prosecution of a study in which the hypothesis under investigation generates at least one prediction .

Predictions are simply the study results one expects if the hypothesis under investigation is true.

Third, scientists compare the study results to those predicted. If study results match the predictions sufficiently well, then the hypothesis is supported — the results are consistent with the hypothesis being true . On the other hand, if they do not match, the hypothesis is not supported — the results are consistent with the hypothesis being false .

We can think of scientific hypotheses and predictions as a type of IF-THEN statement: IF the hypothesis is true THEN the study results should match those predicted. (Fig. 1).

Fig. 1. A simple example of the scientific method.

Fig. 1. A simple example of the scientific method. We want to figure out why a light doesn’t work. One hypothesis is that the bulb is burnt out. To test this hypothesis, there are several different experiments that could be conducted, each of which yields a specific prediction. For example, in experiment 2, if the cause of the non-functioning light is a burnt-out bulb (the causal hypothesis), then the light should work when the old bulb is replaced (prediction). On the other hand, if it still doesn’t work, then the correct explanation is unlikely to be a burnt-out bulb — perhaps the light is unplugged!

We tend to think of science as being concerned only with the physical or natural world. But much of the social sciences is concerned with understanding the reasons why people behave (or misbehave) as they do. Explanations for individual or group behavior can be formulated as scientific hypotheses and tested accordingly.

Hypothesis testing as refutation

Suppose observed results match predictions. Can we conclude that the hypothesis is true? No, for a simple reason: there are always alternative explanations (hypotheses) for observed results. For example, although the light working after the bulb is replaced is consistent with the hypothesis of a burnt-out bulb, it is also consistent with the hypothesis that the power to the house was off and was restored between the time the old bulb was removed and the new one installed. So it is entirely possible that observed results match predictions yet the hypothesis is nonetheless false.

Scientific hypotheses cannot be proven because for any set of results, there are always alternate hypotheses that generate the same predictions, and scientists cannot test all possible hypotheses. This means that scientific hypotheses that scientists accept as “facts” are simply those that have been subjected to the most rigorous and exhausting testing and have failed to be refuted. In the words of the late paleontologist Stephen J. Gould:

“In science, ‘fact’ can only mean ‘confirmed to such a degree that it would be perverse to withhold provisional assent.’” Footnote 4

Note the “provisional assent” here: even scientific hypotheses for which there is compelling supporting evidence may turn out to be false, or at least incomplete. In science, there are no absolute truths.

Science for decision-making

All decisions are based on (usually implicit) causal hypotheses that connect the decision with desired or undesired outcomes. It is these underlying hypotheses that give rise to the predicted effects of alternative decisions.

For decision-makers, explicit enumeration of all underlying causal hypotheses is important because it:

  • clearly identifies the relevant science, that is, science that provides evidence concerning the truth, or otherwise, of at least one underlying hypothesis. This dramatically increases the efficiency of evidence gathering as well as clearly proscribing irrelevant science.
  • reduces the risk of “unanticipated” consequences of policy decisions. Theories of change explicitly identify the underlying causal pathways — simply collections of hypotheses — that link candidate decisions with desired and undesired outcomes. Explicit consideration of these causal pathways can bring to light potential effects of candidate decisions that might not otherwise have been considered, reducing the risk of unanticipated consequences. Footnote 5

Library homepage

  • school Campus Bookshelves
  • menu_book Bookshelves
  • perm_media Learning Objects
  • login Login
  • how_to_reg Request Instructor Account
  • hub Instructor Commons
  • Download Page (PDF)
  • Download Full Book (PDF)
  • Periodic Table
  • Physics Constants
  • Scientific Calculator
  • Reference & Cite
  • Tools expand_more
  • Readability

selected template will load here

This action is not available.

Biology LibreTexts

1.5: Scientific Investigations

  • Last updated
  • Save as PDF
  • Page ID 16715

  • Suzanne Wakim & Mandeep Grewal
  • Butte College

What Turned the Water Orange?

If you were walking in the woods and saw this stream, you probably would wonder what made the water turn orange. Is the water orange because of something growing in it? Is it polluted with some kind of chemicals? To answer these questions, you might do a little research. For example, you might ask local people if they know why the water is orange, or you might try to learn more about it online. If you still haven't found answers, you could undertake a scientific investigation. In short, you could "do" science.

Yellow water flowing in the Rio Tinto, Spain

"Doing" Science

Science is more about doing than knowing. Scientists are always trying to learn more and gain a better understanding of the natural world. There are basic methods of gaining knowledge that is common to all of science. At the heart of science is the scientific investigation. A scientific investigation is a plan for asking questions and testing possible answers in order to advance scientific knowledge.

Figure \(\PageIndex{2}\) outlines the steps of the scientific method. Science textbooks often present this simple, linear "recipe" for a scientific investigation. This is an oversimplification of how science is actually done, but it does highlight the basic plan and purpose of any scientific investigation: testing ideas with evidence. We will use this flowchart to help explain the overall format for scientific inquiry.

Science is actually a complex endeavor that cannot be reduced to a single, linear sequence of steps, like the instructions on a package of cake mix. Real science is nonlinear, iterative (repetitive), creative, unpredictable, and exciting. Scientists often undertake the steps of an investigation in a different sequence, or they repeat the same steps many times as they gain more information and develop new ideas. Scientific investigations often raise new questions as old ones are answered. Successive investigations may address the same questions but at ever-deeper levels. Alternatively, an investigation might lead to an unexpected observation that sparks a new question and takes the research in a completely different direction.

Knowing how scientists "do" science can help you in your everyday life, even if you aren't a scientist. Some steps of the scientific process — such as asking questions and evaluating evidence — can be applied to answering real-life questions and solving practical problems.

Scientific method flow chart. described in text of page

Making Observations

A scientific investigation typically begins with observations. An observation is anything that is detected through human senses or with instruments and measuring devices that enhance human senses. We usually think of observations as things we see with our eyes, but we can also make observations with our sense of touch, smell, taste, or hearing. In addition, we can extend and improve our own senses with instruments such as thermometers and microscopes. Other instruments can be used to sense things that human senses cannot detect at all, such as ultraviolet light or radio waves.

Sometimes chance observations lead to important scientific discoveries. One such observation was made by the Scottish biologist Alexander Fleming (Figure \(\PageIndex{3}\)) in the 1920s. Fleming's name may sound familiar to you because he is famous for the discovery in question. Fleming had been growing a certain type of bacteria on glass plates in his lab when he noticed that one of the plates had been contaminated with mold. On closer examination, Fleming observed that the area around the mold was free of bacteria.

Alexander Fleming looking at a Petri Dish with growth on it

Asking Questions

Observations often lead to interesting questions. This is especially true if the observer is thinking like a scientist. Having scientific training and knowledge is also useful. Relevant background knowledge and logical thinking help make sense of observations so the observer can form particularly salient questions. Fleming, for example, wondered whether the mold — or some substance it produced — had killed bacteria on the plate. Fortunately for us, Fleming didn't just throw out the mold-contaminated plate. Instead, he investigated his question and in so doing, discovered the antibiotic penicillin.

Hypothesis Formation

To find the answer to a question, the next step in a scientific investigation typically is to form a hypothesis. A hypothesis is a possible answer to a scientific question. But it isn’t just any answer. A hypothesis must be based on scientific knowledge. In other words, it shouldn't be at odds with what is already known about the natural world. A hypothesis also must be logical, and it is beneficial if the hypothesis is relatively simple. In addition, to be useful in science, a hypothesis must be testable and falsifiable. In other words, it must be possible to subject the hypothesis to a test that generates evidence for or against it, and it must be possible to make observations that would disprove the hypothesis if it really is false.

A hypothesis is often expressed in the form of prediction: If the hypothesis is true, then B will happen to the dependent variable . Fleming's hypothesis might have been: "If a certain type of mold is introduced to a particular kind of bacteria growing on a plate, the bacteria will die." Is this a good and useful hypothesis? The hypothesis is logical and based directly on observations. The hypothesis is also simple, involving just one type each of mold and bacteria growing on a glass plate. This makes it easy to test. In addition, the hypothesis is falsifiable. If bacteria were to grow in the presence of the mold, it would disprove the hypothesis if it really is false.

Hypothesis Testing

Hypothesis testing is at the heart of a scientific investigation. How would Fleming test his hypothesis? He would gather relevant data as evidence. Evidence is any type of data that may be used to test a hypothesis. Data (singular, datum) are essentially just observations. The observations may be measurements in an experiment or just something the researcher notices. Testing a hypothesis then involves using the data to answer two basic questions:

  • If my hypothesis is true, what would I expect to observe?
  • Does what I actually observe match what predicted?

A hypothesis is supported if the actual observations (data) match the expected observations. A hypothesis is refuted if the actual observations differ from the expected observations.

Testing Fleming's Hypothesis

To test his hypothesis that the mold kills bacteria, Fleming grew colonies of bacteria on several glass plates and introduced mold to just some of the plates. He subjected all of the plates to the same conditions except for the introduction of mold. Any differences in the growth of bacteria on the two groups of plates could then be reasonably attributed to the presence/absence of mold. Fleming's data might have included actual measurements of bacterial colony size, like the data shown in the data table below, or they might have been just an indication of the presence or absence of bacteria growing near the mold. Data like the former, which can be expressed numerically, are called quantitative data. Data like the latter, which can only be expressed in words, such as present or absent, are called qualitative data.

Analyzing and Interpreting Data

The data scientists gather in their investigations are raw data. These are the actual measurements or other observations that are made in an investigation, like the measurements of bacterial growth shown in the data table above. Raw data usually must be analyzed and interpreted before they become evidence to test a hypothesis. To make sense of raw data and decide whether they support a hypothesis, scientists generally use statistics.

There are two basic types of statistics: descriptive statistics and inferential statistics. Both types are important in scientific investigations.

  • Descriptive statistics describe and summarize the data. They include values such as the mean, or average, value in the data. Another basic descriptive statistic is the standard deviation, which gives an idea of the spread of data values around the mean value. Descriptive statistics make it easier to use and discuss the data and also to spot trends or patterns in the data.
  • Inferential statistics help interpret data to test hypotheses. They determine how likely it is that the actual results obtained in an investigation occurred just by chance rather than for the reason posited by the hypothesis. For example, if inferential statistics show that the results of an investigation would happen by chance only 5 percent of the time, then the hypothesis has a 95 percent chance of being correctly supported by the results. An example of a statistical hypothesis test is a t-test. It can be used to compare the mean value of the actual data with the expected value predicted by the hypothesis. Alternatively, a t-test can be used to compare the mean value of one group of data with the mean value of another group to determine whether the mean values are significantly different or just different by chance.

Assume that Fleming obtained the raw data shown in the data table above. We could use a descriptive statistic such as the mean area of bacterial growth to describe the raw data. Based on these data, the mean area of bacterial growth for plates with mold is 56 mm 2 , and the mean area for plates without mold is 69 mm 2 . Is this difference in bacterial growth significant? In other words, does it provide convincing evidence that bacteria are killed by the mold or something produced by the mold? Or could the difference in mean values between the two groups of plates be due to chance alone? What is the likelihood that this outcome could have occurred even if mold or one of its products does not kill bacteria? A t-test could be done to answer this question. The p-value for the t-test analysis of the data above is less than 0.05. This means that one can say with 95% confidence that the means of the above data are statistically different.

Drawing Conclusions

A statistical analysis of Fleming's evidence showed that it did indeed support his hypothesis. Does this mean that the hypothesis is true? No, not necessarily. That's because a hypothesis can never be proven conclusively to be true. Scientists can never examine all of the possible evidence, and someday evidence might be found that disproves the hypothesis. In addition, other hypotheses, as yet unformed, may be supported by the same evidence. For example, in Fleming's investigation, something else introduced onto the plates with the mold might have been responsible for the death of the bacteria. Although a hypothesis cannot be proven true without a shadow of a doubt, the more evidence that supports a hypothesis, the more likely the hypothesis is to be correct. Similarly, the better the match between actual observations and expected observations, the more likely a hypothesis is to be true.

Many times, competing hypotheses are supported by evidence. When that occurs, how do scientists conclude which hypothesis is better? There are several criteria that may be used to judge competing hypotheses. For example, scientists are more likely to accept a hypothesis that:

  • explains a wider variety of observations.
  • explains observations that were previously unexplained.
  • generates more expectations and is thus more testable.
  • is more consistent with well-established theories.
  • is more parsimonious, that is, is a simpler and less convoluted explanation.

Correlation-Causation Fallacy

Many statistical tests used in scientific research calculate correlations between variables. Correlation refers to how closely related two data sets are, which may be a useful starting point for further investigation. However, correlation is also one of the most misused types of evidence, primarily because of the logical fallacy that correlation implies causation. In reality, just because two variables are correlated does not necessarily mean that either variable causes the other.

A simple example can be used to demonstrate the correlation-causation fallacy. Assume a study found that both ice cream sales and burglaries are correlated; that is, rates of both events increase together. If correlation really did imply causation, then you could conclude that ice cream sales cause burglaries or vice versa. It is more likely, however, that a third variable, such as the weather, influences rates of both ice cream sales and burglaries. Both might increase when the weather is sunny.

An actual example of the correlation-causation fallacy occurred during the latter half of the 20th century. Numerous studies showed that women taking hormone replacement therapy (HRT) to treat menopausal symptoms also had a lower-than-average incidence of coronary heart disease (CHD). This correlation was misinterpreted as evidence that HRT protects women against CHD. Subsequent studies that controlled other factors related to CHD disproved this presumed causal connection. The studies found that women taking HRT were more likely to come from higher socio-economic groups, with better-than-average diets and exercise regimens. Rather than HRT causing lower CHD incidence, these studies concluded that HRT and lower CHD were both effects of higher socioeconomic status and related lifestyle factors.

Communicating Results

The last step in a scientific investigation is communicating the results to other scientists. This is a very important step because it allows other scientists to try to repeat the investigation and see if they can produce the same results. If other researchers get the same results, it adds support to the hypothesis. If they get different results, it may disprove the hypothesis. When scientists communicate their results, they should describe their methods and point out any possible problems with the investigation. This allows other researchers to identify any flaws in the method or think of ways to avoid possible problems in future studies.

Repeating a scientific investigation and reproducing the same results is called replication . It is a cornerstone of scientific research. Replication is not required for every investigation in science, but it is highly recommended for those that produce surprising or particularly consequential results. In some scientific fields, scientists routinely try to replicate their own investigations to ensure the reproducibility of the results before they communicate them.

Scientists may communicate their results in a variety of ways. The most rigorous way is to write up the investigation and results in the form of an article and submit it to a peer-reviewed scientific journal for publication. The editor of the journal provides copies of the article to several other scientists who work in the same field. These are the peers in the peer-review process. The reviewers study the article and tell the editor whether they think it should be published, based on the validity of the methods and significance of the study. The article may be rejected outright, or it may be accepted, either as is or with revisions. Only articles that meet high scientific standards are ultimately published.

  • Outline the steps of a typical scientific investigation.
  • What is a scientific hypothesis? What characteristics must a hypothesis have to be useful in science?
  • Explain how you could do a scientific investigation to answer this question: Which of the following surfaces in my home has the most bacteria: the house phone, TV remote, bathroom sink faucet, or outside door handle? Form a hypothesis and state what results would support it and what results would refute it.
  • Look at the areas of bacterial growth for the plates in just one group – either with mold (plates 1-5) or without mold (plates 6-10). Is there a variation within the group? What do you think could be possible sources of variation within the group?
  • Compare the area of bacterial growth for plate 1 vs. plate 7. Does this appear to be more of a difference between the mold group vs. the no mold group than if you compared plate 5 vs. plate 6? Using these differences among the individual data points, explain why it is important to find the mean of each group when analyzing the data.
  • Why do you think it would be important for other researchers to try to replicate the findings in this study?
  • Is the energy level of the mice treated with the drug a qualitative or quantitative observation?
  • At the end of the study, the scientist measures the size of the tumors. Is this qualitative or quantitative data?
  • Would the size of each tumor be considered raw data or descriptive statistics?
  • The scientist determines the average decrease in tumor size for the drug-treated group. Is this raw data, descriptive statistics, or inferential statistics?
  • The average decrease in tumor size in the drug-treated group is larger than the average decrease in the untreated group. Can the scientist assume that the drug shrinks tumors? If not, what do they need to do next?
  • Do you think results published in a peer-reviewed scientific journal are more or less likely to be scientifically valid than those in a self-published article or book? Why or why not
  • Explain why real science is usually “nonlinear”?

Explore More

Watch this TED talk for a lively discussion of why the standard scientific method is an inadequate model of how science is really done.

Attributions

  • Rio Tinto River by Carol Stoker, NASA, public domain via Wikimedia Commons
  • Scientific Method by OpenStax, licensed CC BY 4.0
  • Alexander Flemming by Ministry of Information Photo Division Photographer, public domain via Wikimedia Commons
  • Text adapted from Human Biology by CK-12 licensed CC BY-NC 3.0

1.2 The Scientific Methods

Section learning objectives.

By the end of this section, you will be able to do the following:

  • Explain how the methods of science are used to make scientific discoveries
  • Define a scientific model and describe examples of physical and mathematical models used in physics
  • Compare and contrast hypothesis, theory, and law

Teacher Support

The learning objectives in this section will help your students master the following standards:

  • (A) know the definition of science and understand that it has limitations, as specified in subsection (b)(2) of this section;
  • (B) know that scientific hypotheses are tentative and testable statements that must be capable of being supported or not supported by observational evidence. Hypotheses of durable explanatory power which have been tested over a wide variety of conditions are incorporated into theories;
  • (C) know that scientific theories are based on natural and physical phenomena and are capable of being tested by multiple independent researchers. Unlike hypotheses, scientific theories are well-established and highly-reliable explanations, but may be subject to change as new areas of science and new technologies are developed;
  • (D) distinguish between scientific hypotheses and scientific theories.

Section Key Terms

[OL] Pre-assessment for this section could involve students sharing or writing down an anecdote about when they used the methods of science. Then, students could label their thought processes in their anecdote with the appropriate scientific methods. The class could also discuss their definitions of theory and law, both outside and within the context of science.

[OL] It should be noted and possibly mentioned that a scientist , as mentioned in this section, does not necessarily mean a trained scientist. It could be anyone using methods of science.

Scientific Methods

Scientists often plan and carry out investigations to answer questions about the universe around us. These investigations may lead to natural laws. Such laws are intrinsic to the universe, meaning that humans did not create them and cannot change them. We can only discover and understand them. Their discovery is a very human endeavor, with all the elements of mystery, imagination, struggle, triumph, and disappointment inherent in any creative effort. The cornerstone of discovering natural laws is observation. Science must describe the universe as it is, not as we imagine or wish it to be.

We all are curious to some extent. We look around, make generalizations, and try to understand what we see. For example, we look up and wonder whether one type of cloud signals an oncoming storm. As we become serious about exploring nature, we become more organized and formal in collecting and analyzing data. We attempt greater precision, perform controlled experiments (if we can), and write down ideas about how data may be organized. We then formulate models, theories, and laws based on the data we have collected, and communicate those results with others. This, in a nutshell, describes the scientific method that scientists employ to decide scientific issues on the basis of evidence from observation and experiment.

An investigation often begins with a scientist making an observation . The scientist observes a pattern or trend within the natural world. Observation may generate questions that the scientist wishes to answer. Next, the scientist may perform some research about the topic and devise a hypothesis . A hypothesis is a testable statement that describes how something in the natural world works. In essence, a hypothesis is an educated guess that explains something about an observation.

[OL] An educated guess is used throughout this section in describing a hypothesis to combat the tendency to think of a theory as an educated guess.

Scientists may test the hypothesis by performing an experiment . During an experiment, the scientist collects data that will help them learn about the phenomenon they are studying. Then the scientists analyze the results of the experiment (that is, the data), often using statistical, mathematical, and/or graphical methods. From the data analysis, they draw conclusions. They may conclude that their experiment either supports or rejects their hypothesis. If the hypothesis is supported, the scientist usually goes on to test another hypothesis related to the first. If their hypothesis is rejected, they will often then test a new and different hypothesis in their effort to learn more about whatever they are studying.

Scientific processes can be applied to many situations. Let’s say that you try to turn on your car, but it will not start. You have just made an observation! You ask yourself, "Why won’t my car start?" You can now use scientific processes to answer this question. First, you generate a hypothesis such as, "The car won’t start because it has no gasoline in the gas tank." To test this hypothesis, you put gasoline in the car and try to start it again. If the car starts, then your hypothesis is supported by the experiment. If the car does not start, then your hypothesis is rejected. You will then need to think up a new hypothesis to test such as, "My car won’t start because the fuel pump is broken." Hopefully, your investigations lead you to discover why the car won’t start and enable you to fix it.

A model is a representation of something that is often too difficult (or impossible) to study directly. Models can take the form of physical models, equations, computer programs, or simulations—computer graphics/animations. Models are tools that are especially useful in modern physics because they let us visualize phenomena that we normally cannot observe with our senses, such as very small objects or objects that move at high speeds. For example, we can understand the structure of an atom using models, without seeing an atom with our own eyes. Although images of single atoms are now possible, these images are extremely difficult to achieve and are only possible due to the success of our models. The existence of these images is a consequence rather than a source of our understanding of atoms. Models are always approximate, so they are simpler to consider than the real situation; the more complete a model is, the more complicated it must be. Models put the intangible or the extremely complex into human terms that we can visualize, discuss, and hypothesize about.

Scientific models are constructed based on the results of previous experiments. Even still, models often only describe a phenomenon partially or in a few limited situations. Some phenomena are so complex that they may be impossible to model them in their entirety, even using computers. An example is the electron cloud model of the atom in which electrons are moving around the atom’s center in distinct clouds ( Figure 1.12 ), that represent the likelihood of finding an electron in different places. This model helps us to visualize the structure of an atom. However, it does not show us exactly where an electron will be within its cloud at any one particular time.

As mentioned previously, physicists use a variety of models including equations, physical models, computer simulations, etc. For example, three-dimensional models are often commonly used in chemistry and physics to model molecules. Properties other than appearance or location are usually modelled using mathematics, where functions are used to show how these properties relate to one another. Processes such as the formation of a star or the planets, can also be modelled using computer simulations. Once a simulation is correctly programmed based on actual experimental data, the simulation can allow us to view processes that happened in the past or happen too quickly or slowly for us to observe directly. In addition, scientists can also run virtual experiments using computer-based models. In a model of planet formation, for example, the scientist could alter the amount or type of rocks present in space and see how it affects planet formation.

Scientists use models and experimental results to construct explanations of observations or design solutions to problems. For example, one way to make a car more fuel efficient is to reduce the friction or drag caused by air flowing around the moving car. This can be done by designing the body shape of the car to be more aerodynamic, such as by using rounded corners instead of sharp ones. Engineers can then construct physical models of the car body, place them in a wind tunnel, and examine the flow of air around the model. This can also be done mathematically in a computer simulation. The air flow pattern can be analyzed for regions smooth air flow and for eddies that indicate drag. The model of the car body may have to be altered slightly to produce the smoothest pattern of air flow (i.e., the least drag). The pattern with the least drag may be the solution to increasing fuel efficiency of the car. This solution might then be incorporated into the car design.

Using Models and the Scientific Processes

Be sure to secure loose items before opening the window or door.

In this activity, you will learn about scientific models by making a model of how air flows through your classroom or a room in your house.

  • One room with at least one window or door that can be opened
  • Work with a group of four, as directed by your teacher. Close all of the windows and doors in the room you are working in. Your teacher may assign you a specific window or door to study.
  • Before opening any windows or doors, draw a to-scale diagram of your room. First, measure the length and width of your room using the tape measure. Then, transform the measurement using a scale that could fit on your paper, such as 5 centimeters = 1 meter.
  • Your teacher will assign you a specific window or door to study air flow. On your diagram, add arrows showing your hypothesis (before opening any windows or doors) of how air will flow through the room when your assigned window or door is opened. Use pencil so that you can easily make changes to your diagram.
  • On your diagram, mark four locations where you would like to test air flow in your room. To test for airflow, hold a strip of single ply tissue paper between the thumb and index finger. Note the direction that the paper moves when exposed to the airflow. Then, for each location, predict which way the paper will move if your air flow diagram is correct.
  • Now, each member of your group will stand in one of the four selected areas. Each member will test the airflow Agree upon an approximate height at which everyone will hold their papers.
  • When you teacher tells you to, open your assigned window and/or door. Each person should note the direction that their paper points immediately after the window or door was opened. Record your results on your diagram.
  • Did the airflow test data support or refute the hypothetical model of air flow shown in your diagram? Why or why not? Correct your model based on your experimental evidence.
  • With your group, discuss how accurate your model is. What limitations did it have? Write down the limitations that your group agreed upon.
  • Yes, you could use your model to predict air flow through a new window. The earlier experiment of air flow would help you model the system more accurately.
  • Yes, you could use your model to predict air flow through a new window. The earlier experiment of air flow is not useful for modeling the new system.
  • No, you cannot model a system to predict the air flow through a new window. The earlier experiment of air flow would help you model the system more accurately.
  • No, you cannot model a system to predict the air flow through a new window. The earlier experiment of air flow is not useful for modeling the new system.

This Snap Lab! has students construct a model of how air flows in their classroom. Each group of four students will create a model of air flow in their classroom using a scale drawing of the room. Then, the groups will test the validity of their model by placing weathervanes that they have constructed around the room and opening a window or door. By observing the weather vanes, students will see how air actually flows through the room from a specific window or door. Students will then correct their model based on their experimental evidence. The following material list is given per group:

  • One room with at least one window or door that can be opened (An optimal configuration would be one window or door per group.)
  • Several pieces of construction paper (at least four per group)
  • Strips of single ply tissue paper
  • One tape measure (long enough to measure the dimensions of the room)
  • Group size can vary depending on the number of windows/doors available and the number of students in the class.
  • The room dimensions could be provided by the teacher. Also, students may need a brief introduction in how to make a drawing to scale.
  • This is another opportunity to discuss controlled experiments in terms of why the students should hold the strips of tissue paper at the same height and in the same way. One student could also serve as a control and stand far away from the window/door or in another area that will not receive air flow from the window/door.
  • You will probably need to coordinate this when multiple windows or doors are used. Only one window or door should be opened at a time for best results. Between openings, allow a short period (5 minutes) when all windows and doors are closed, if possible.

Answers to the Grasp Check will vary, but the air flow in the new window or door should be based on what the students observed in their experiment.

Scientific Laws and Theories

A scientific law is a description of a pattern in nature that is true in all circumstances that have been studied. That is, physical laws are meant to be universal , meaning that they apply throughout the known universe. Laws are often also concise, whereas theories are more complicated. A law can be expressed in the form of a single sentence or mathematical equation. For example, Newton’s second law of motion , which relates the motion of an object to the force applied ( F ), the mass of the object ( m ), and the object’s acceleration ( a ), is simply stated using the equation

Scientific ideas and explanations that are true in many, but not all situations in the universe are usually called principles . An example is Pascal’s principle , which explains properties of liquids, but not solids or gases. However, the distinction between laws and principles is sometimes not carefully made in science.

A theory is an explanation for patterns in nature that is supported by much scientific evidence and verified multiple times by multiple researchers. While many people confuse theories with educated guesses or hypotheses, theories have withstood more rigorous testing and verification than hypotheses.

[OL] Explain to students that in informal, everyday English the word theory can be used to describe an idea that is possibly true but that has not been proven to be true. This use of the word theory often leads people to think that scientific theories are nothing more than educated guesses. This is not just a misconception among students, but among the general public as well.

As a closing idea about scientific processes, we want to point out that scientific laws and theories, even those that have been supported by experiments for centuries, can still be changed by new discoveries. This is especially true when new technologies emerge that allow us to observe things that were formerly unobservable. Imagine how viewing previously invisible objects with a microscope or viewing Earth for the first time from space may have instantly changed our scientific theories and laws! What discoveries still await us in the future? The constant retesting and perfecting of our scientific laws and theories allows our knowledge of nature to progress. For this reason, many scientists are reluctant to say that their studies prove anything. By saying support instead of prove , it keeps the door open for future discoveries, even if they won’t occur for centuries or even millennia.

[OL] With regard to scientists avoiding using the word prove , the general public knows that science has proven certain things such as that the heart pumps blood and the Earth is round. However, scientists should shy away from using prove because it is impossible to test every single instance and every set of conditions in a system to absolutely prove anything. Using support or similar terminology leaves the door open for further discovery.

Check Your Understanding

  • Models are simpler to analyze.
  • Models give more accurate results.
  • Models provide more reliable predictions.
  • Models do not require any computer calculations.
  • They are the same.
  • A hypothesis has been thoroughly tested and found to be true.
  • A hypothesis is a tentative assumption based on what is already known.
  • A hypothesis is a broad explanation firmly supported by evidence.
  • A scientific model is a representation of something that can be easily studied directly. It is useful for studying things that can be easily analyzed by humans.
  • A scientific model is a representation of something that is often too difficult to study directly. It is useful for studying a complex system or systems that humans cannot observe directly.
  • A scientific model is a representation of scientific equipment. It is useful for studying working principles of scientific equipment.
  • A scientific model is a representation of a laboratory where experiments are performed. It is useful for studying requirements needed inside the laboratory.
  • The hypothesis must be validated by scientific experiments.
  • The hypothesis must not include any physical quantity.
  • The hypothesis must be a short and concise statement.
  • The hypothesis must apply to all the situations in the universe.
  • A scientific theory is an explanation of natural phenomena that is supported by evidence.
  • A scientific theory is an explanation of natural phenomena without the support of evidence.
  • A scientific theory is an educated guess about the natural phenomena occurring in nature.
  • A scientific theory is an uneducated guess about natural phenomena occurring in nature.
  • A hypothesis is an explanation of the natural world with experimental support, while a scientific theory is an educated guess about a natural phenomenon.
  • A hypothesis is an educated guess about natural phenomenon, while a scientific theory is an explanation of natural world with experimental support.
  • A hypothesis is experimental evidence of a natural phenomenon, while a scientific theory is an explanation of the natural world with experimental support.
  • A hypothesis is an explanation of the natural world with experimental support, while a scientific theory is experimental evidence of a natural phenomenon.

Use the Check Your Understanding questions to assess students’ achievement of the section’s learning objectives. If students are struggling with a specific objective, the Check Your Understanding will help identify which objective and direct students to the relevant content.

As an Amazon Associate we earn from qualifying purchases.

This book may not be used in the training of large language models or otherwise be ingested into large language models or generative AI offerings without OpenStax's permission.

Want to cite, share, or modify this book? This book uses the Creative Commons Attribution License and you must attribute Texas Education Agency (TEA). The original material is available at: https://www.texasgateway.org/book/tea-physics . Changes were made to the original material, including updates to art, structure, and other content updates.

Access for free at https://openstax.org/books/physics/pages/1-introduction
  • Authors: Paul Peter Urone, Roger Hinrichs
  • Publisher/website: OpenStax
  • Book title: Physics
  • Publication date: Mar 26, 2020
  • Location: Houston, Texas
  • Book URL: https://openstax.org/books/physics/pages/1-introduction
  • Section URL: https://openstax.org/books/physics/pages/1-2-the-scientific-methods

© Jan 19, 2024 Texas Education Agency (TEA). The OpenStax name, OpenStax logo, OpenStax book covers, OpenStax CNX name, and OpenStax CNX logo are not subject to the Creative Commons license and may not be reproduced without the prior and express written consent of Rice University.

Science and the scientific method: Definitions and examples

Here's a look at the foundation of doing science — the scientific method.

Kids follow the scientific method to carry out an experiment.

The scientific method

Hypothesis, theory and law, a brief history of science, additional resources, bibliography.

Science is a systematic and logical approach to discovering how things in the universe work. It is also the body of knowledge accumulated through the discoveries about all the things in the universe. 

The word "science" is derived from the Latin word "scientia," which means knowledge based on demonstrable and reproducible data, according to the Merriam-Webster dictionary . True to this definition, science aims for measurable results through testing and analysis, a process known as the scientific method. Science is based on fact, not opinion or preferences. The process of science is designed to challenge ideas through research. One important aspect of the scientific process is that it focuses only on the natural world, according to the University of California, Berkeley . Anything that is considered supernatural, or beyond physical reality, does not fit into the definition of science.

When conducting research, scientists use the scientific method to collect measurable, empirical evidence in an experiment related to a hypothesis (often in the form of an if/then statement) that is designed to support or contradict a scientific theory .

"As a field biologist, my favorite part of the scientific method is being in the field collecting the data," Jaime Tanner, a professor of biology at Marlboro College, told Live Science. "But what really makes that fun is knowing that you are trying to answer an interesting question. So the first step in identifying questions and generating possible answers (hypotheses) is also very important and is a creative process. Then once you collect the data you analyze it to see if your hypothesis is supported or not."

Here's an illustration showing the steps in the scientific method.

The steps of the scientific method go something like this, according to Highline College :

  • Make an observation or observations.
  • Form a hypothesis — a tentative description of what's been observed, and make predictions based on that hypothesis.
  • Test the hypothesis and predictions in an experiment that can be reproduced.
  • Analyze the data and draw conclusions; accept or reject the hypothesis or modify the hypothesis if necessary.
  • Reproduce the experiment until there are no discrepancies between observations and theory. "Replication of methods and results is my favorite step in the scientific method," Moshe Pritsker, a former post-doctoral researcher at Harvard Medical School and CEO of JoVE, told Live Science. "The reproducibility of published experiments is the foundation of science. No reproducibility — no science."

Some key underpinnings to the scientific method:

  • The hypothesis must be testable and falsifiable, according to North Carolina State University . Falsifiable means that there must be a possible negative answer to the hypothesis.
  • Research must involve deductive reasoning and inductive reasoning . Deductive reasoning is the process of using true premises to reach a logical true conclusion while inductive reasoning uses observations to infer an explanation for those observations.
  • An experiment should include a dependent variable (which does not change) and an independent variable (which does change), according to the University of California, Santa Barbara .
  • An experiment should include an experimental group and a control group. The control group is what the experimental group is compared against, according to Britannica .

The process of generating and testing a hypothesis forms the backbone of the scientific method. When an idea has been confirmed over many experiments, it can be called a scientific theory. While a theory provides an explanation for a phenomenon, a scientific law provides a description of a phenomenon, according to The University of Waikato . One example would be the law of conservation of energy, which is the first law of thermodynamics that says that energy can neither be created nor destroyed. 

A law describes an observed phenomenon, but it doesn't explain why the phenomenon exists or what causes it. "In science, laws are a starting place," said Peter Coppinger, an associate professor of biology and biomedical engineering at the Rose-Hulman Institute of Technology. "From there, scientists can then ask the questions, 'Why and how?'"

Laws are generally considered to be without exception, though some laws have been modified over time after further testing found discrepancies. For instance, Newton's laws of motion describe everything we've observed in the macroscopic world, but they break down at the subatomic level.

This does not mean theories are not meaningful. For a hypothesis to become a theory, scientists must conduct rigorous testing, typically across multiple disciplines by separate groups of scientists. Saying something is "just a theory" confuses the scientific definition of "theory" with the layperson's definition. To most people a theory is a hunch. In science, a theory is the framework for observations and facts, Tanner told Live Science.

This Copernican heliocentric solar system, from 1708, shows the orbit of the moon around the Earth, and the orbits of the Earth and planets round the sun, including Jupiter and its moons, all surrounded by the 12 signs of the zodiac.

The earliest evidence of science can be found as far back as records exist. Early tablets contain numerals and information about the solar system , which were derived by using careful observation, prediction and testing of those predictions. Science became decidedly more "scientific" over time, however.

1200s: Robert Grosseteste developed the framework for the proper methods of modern scientific experimentation, according to the Stanford Encyclopedia of Philosophy. His works included the principle that an inquiry must be based on measurable evidence that is confirmed through testing.

1400s: Leonardo da Vinci began his notebooks in pursuit of evidence that the human body is microcosmic. The artist, scientist and mathematician also gathered information about optics and hydrodynamics.

1500s: Nicolaus Copernicus advanced the understanding of the solar system with his discovery of heliocentrism. This is a model in which Earth and the other planets revolve around the sun, which is the center of the solar system.

1600s: Johannes Kepler built upon those observations with his laws of planetary motion. Galileo Galilei improved on a new invention, the telescope, and used it to study the sun and planets. The 1600s also saw advancements in the study of physics as Isaac Newton developed his laws of motion.

1700s: Benjamin Franklin discovered that lightning is electrical. He also contributed to the study of oceanography and meteorology. The understanding of chemistry also evolved during this century as Antoine Lavoisier, dubbed the father of modern chemistry , developed the law of conservation of mass.

1800s: Milestones included Alessandro Volta's discoveries regarding electrochemical series, which led to the invention of the battery. John Dalton also introduced atomic theory, which stated that all matter is composed of atoms that combine to form molecules. The basis of modern study of genetics advanced as Gregor Mendel unveiled his laws of inheritance. Later in the century, Wilhelm Conrad Röntgen discovered X-rays , while George Ohm's law provided the basis for understanding how to harness electrical charges.

1900s: The discoveries of Albert Einstein , who is best known for his theory of relativity, dominated the beginning of the 20th century. Einstein's theory of relativity is actually two separate theories. His special theory of relativity, which he outlined in a 1905 paper, " The Electrodynamics of Moving Bodies ," concluded that time must change according to the speed of a moving object relative to the frame of reference of an observer. His second theory of general relativity, which he published as " The Foundation of the General Theory of Relativity ," advanced the idea that matter causes space to curve.

In 1952, Jonas Salk developed the polio vaccine , which reduced the incidence of polio in the United States by nearly 90%, according to Britannica . The following year, James D. Watson and Francis Crick discovered the structure of DNA , which is a double helix formed by base pairs attached to a sugar-phosphate backbone, according to the National Human Genome Research Institute .

2000s: The 21st century saw the first draft of the human genome completed, leading to a greater understanding of DNA. This advanced the study of genetics, its role in human biology and its use as a predictor of diseases and other disorders, according to the National Human Genome Research Institute .

  • This video from City University of New York delves into the basics of what defines science.
  • Learn about what makes science science in this book excerpt from Washington State University .
  • This resource from the University of Michigan — Flint explains how to design your own scientific study.

Merriam-Webster Dictionary, Scientia. 2022. https://www.merriam-webster.com/dictionary/scientia

University of California, Berkeley, "Understanding Science: An Overview." 2022. ​​ https://undsci.berkeley.edu/article/0_0_0/intro_01  

Highline College, "Scientific method." July 12, 2015. https://people.highline.edu/iglozman/classes/astronotes/scimeth.htm  

North Carolina State University, "Science Scripts." https://projects.ncsu.edu/project/bio183de/Black/science/science_scripts.html  

University of California, Santa Barbara. "What is an Independent variable?" October 31,2017. http://scienceline.ucsb.edu/getkey.php?key=6045  

Encyclopedia Britannica, "Control group." May 14, 2020. https://www.britannica.com/science/control-group  

The University of Waikato, "Scientific Hypothesis, Theories and Laws." https://sci.waikato.ac.nz/evolution/Theories.shtml  

Stanford Encyclopedia of Philosophy, Robert Grosseteste. May 3, 2019. https://plato.stanford.edu/entries/grosseteste/  

Encyclopedia Britannica, "Jonas Salk." October 21, 2021. https://www.britannica.com/ biography /Jonas-Salk

National Human Genome Research Institute, "​Phosphate Backbone." https://www.genome.gov/genetics-glossary/Phosphate-Backbone  

National Human Genome Research Institute, "What is the Human Genome Project?" https://www.genome.gov/human-genome-project/What  

‌ Live Science contributor Ashley Hamer updated this article on Jan. 16, 2022.

Sign up for the Live Science daily newsletter now

Get the world’s most fascinating discoveries delivered straight to your inbox.

Alina Bradford

How much does Earth weigh?

Largest gold nugget ever found in England unearthed with faulty metal detector

Space photo of the week: James Webb telescope reveals surprising starburst in ancient galaxy

Most Popular

By Anna Gora December 27, 2023

By Anna Gora December 26, 2023

By Anna Gora December 25, 2023

By Emily Cooke December 23, 2023

By Victoria Atkinson December 22, 2023

By Anna Gora December 16, 2023

By Anna Gora December 15, 2023

By Anna Gora November 09, 2023

By Donavyn Coffey November 06, 2023

By Anna Gora October 31, 2023

By Anna Gora October 26, 2023

  • 2 James Webb telescope confirms there is something seriously wrong with our understanding of the universe
  • 3 'You could almost see and smell their world': Remnants of 'Britain's Pompeii' reveal details of life in Bronze Age village
  • 4 Where does the solar system end?
  • 5 Fiber-optic data transfer speeds hit a rapid 301 Tbps — 1.2 million times faster than your home broadband connection
  • 2 Polar vortex is 'spinning backwards' above Arctic after major reversal event
  • 3 The 7 most powerful supercomputers in the world right now
  • 4 Fiber-optic data transfer speeds hit a rapid 301 Tbps — 1.2 million times faster than your home broadband connection
  • 5 Powerful X-class solar flare slams Earth, triggering radio blackout over the Pacific Ocean

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • Microb Biotechnol
  • v.15(11); 2022 Nov

On the role of hypotheses in science

Harald brüssow.

1 Laboratory of Gene Technology, Department of Biosystems, KU Leuven, Leuven Belgium

Associated Data

Scientific research progresses by the dialectic dialogue between hypothesis building and the experimental testing of these hypotheses. Microbiologists as biologists in general can rely on an increasing set of sophisticated experimental methods for hypothesis testing such that many scientists maintain that progress in biology essentially comes with new experimental tools. While this is certainly true, the importance of hypothesis building in science should not be neglected. Some scientists rely on intuition for hypothesis building. However, there is also a large body of philosophical thinking on hypothesis building whose knowledge may be of use to young scientists. The present essay presents a primer into philosophical thoughts on hypothesis building and illustrates it with two hypotheses that played a major role in the history of science (the parallel axiom and the fifth element hypothesis). It continues with philosophical concepts on hypotheses as a calculus that fits observations (Copernicus), the need for plausibility (Descartes and Gilbert) and for explicatory power imposing a strong selection on theories (Darwin, James and Dewey). Galilei introduced and James and Poincaré later justified the reductionist principle in hypothesis building. Waddington stressed the feed‐forward aspect of fruitful hypothesis building, while Poincaré called for a dialogue between experiment and hypothesis and distinguished false, true, fruitful and dangerous hypotheses. Theoretical biology plays a much lesser role than theoretical physics because physical thinking strives for unification principle across the universe while biology is confronted with a breathtaking diversity of life forms and its historical development on a single planet. Knowledge of the philosophical foundations on hypothesis building in science might stimulate more hypothesis‐driven experimentation that simple observation‐oriented “fishing expeditions” in biological research.

Short abstract

Scientific research progresses by the dialectic dialogue between hypothesis building and the experimental testing of these hypotheses. Microbiologists can rely on an increasing set of sophisticated experimental methods for hypothesis testing but the importance of hypothesis building in science should not be neglected. This Lilliput offers a primer on philosophical concepts on hypotheses in science.

INTRODUCTION

Philosophy of science and the theory of knowledge (epistemology) are important branches of philosophy. However, philosophy has over the centuries lost its dominant role it enjoyed in antiquity and became in Medieval Ages the maid of theology (ancilla theologiae) and after the rise of natural sciences and its technological applications many practising scientists and the general public doubt whether they need philosophical concepts in their professional and private life. This is in the opinion of the writer of this article, an applied microbiologist, shortsighted for several reasons. Philosophers of the 20th century have made important contributions to the theory of knowledge, and many eminent scientists grew interested in philosophical problems. Mathematics which plays such a prominent role in physics and increasingly also in other branches of science is a hybrid: to some extent, it is the paradigm of an exact science while its abstract aspects are deeply rooted in philosophical thinking. In the present essay, the focus is on hypothesis and hypothesis building in science, essentially it is a compilation what philosophers and scientists thought about this subject in past and present. The controversy between the mathematical mind and that of the practical mind is an old one. The philosopher, physicist and mathematician Pascal ( 1623 –1662a) wrote in his Pensées : “Mathematicians who are only mathematicians have exact minds, provided all things are explained to them by means of definitions and axioms; otherwise they are inaccurate. They are only right when the principles are quite clear. And men of intuition cannot have the patience to reach to first principles of things speculative and conceptional, which they have never seen in the world and which are altogether out of the common. The intellect can be strong and narrow, and can be comprehensive and weak.” Hypothesis building is an act both of intuition and exact thinking and I hope that theoretical knowledge about hypothesis building will also profit young microbiologists.

HYPOTHESES AND AXIOMS IN MATHEMATICS

In the following, I will illustrate the importance of hypothesis building for the history of science and the development of knowledge and illustrate it with two famous concepts, the parallel axiom in mathematics and the five elements hypothesis in physics.

Euclidean geometry

The prominent role of hypotheses in the development of science becomes already clear in the first science book of the Western civilization: Euclid's The Elements written about 300 BC starts with a set of statements called Definitions, Postulates and Common Notions that lay out the foundation of geometry (Euclid,  c.323‐c.283 ). This axiomatic approach is very modern as exemplified by the fact that Euclid's book remained for long time after the Bible the most read book in the Western hemisphere and a backbone of school teaching in mathematics. Euclid's twenty‐three definitions start with sentences such as “1. A point is that which has no part; 2. A line is breadthless length; 3. The extremities of a line are points”; and continues with the definition of angles (“8. A plane angle is the inclination to one another of two lines in a plane which meet one another and do not lie in a straight line”) and that of circles, triangles and quadrilateral figures. For the history of science, the 23rd definition of parallels is particularly interesting: “Parallel straight lines are straight lines which, being in the same plane and being produced indefinitely in both directions, do not meet one another in either direction”. This is the famous parallel axiom. It is clear that the parallel axiom cannot be the result of experimental observations, but must be a concept created in the mind. Euclid ends with five Common Notions (“1. Things which are equal to the same thing are also equal to one another, to 5. The whole is greater than the part”). The establishment of a contradiction‐free system for a branch of mathematics based on a set of axioms from which theorems were deduced was revolutionary modern. Hilbert ( 1899 ) formulated a sound modern formulation for Euclidian geometry. Hilbert's axiom system contains the notions “point, line and plane” and the concepts of “betweenness, containment and congruence” leading to five axioms, namely the axioms of Incidence (“Verknüpfung”), of Order (“Anordnung”), of Congruence, of Continuity (“Stetigkeit”) and of Parallels.

Origin of axioms

Philosophers gave various explanations for the origin of the Euclidean hypotheses or axioms. Plato considered geometrical figures as related to ideas (the true things behind the world of appearances). Aristoteles considered geometric figures as abstractions of physical bodies. Descartes perceived geometric figures as inborn ideas from extended bodies ( res extensa ), while Pascal thought that the axioms of Euclidian geometry were derived from intuition. Kant reasoned that Euclidian geometry represented a priori perceptions of space. Newton considered geometry as part of general mechanics linked to theories of measurement. Hilbert argued that the axioms of mathematical geometry are neither the result of contemplation (“Anschauung”) nor of psychological source. For him, axioms were formal propositions (“formale Aussageformen”) characterized by consistency (“Widerspruchsfreiheit”, i.e. absence of contradiction) (Mittelstrass,  1980a ).

Definitions

Axioms were also differently defined by philosophers. In Topics , Aristoteles calls axioms the assumptions taken up by one partner of a dialogue to initiate a dialectic discussion. Plato states that an axiom needs to be an acceptable or credible proposition, which cannot be justified by reference to other statements. Yet, a justification is not necessary because an axiom is an evident statement. In modern definition, axioms are methodical first sentences in the foundation of a deductive science (Mittelstrass,  1980a ). In Posterior Analytics , Aristotle defines postulates as positions which are at least initially not accepted by the dialogue partners while hypotheses are accepted for the sake of reasoning. In Euclid's book, postulates are construction methods that assure the existence of the geometric objects. Today postulates and axioms are used as synonyms while the 18th‐century philosophy made differences: Lambert defined axioms as descriptive sentences and postulates as prescriptive sentences. According to Kant, mathematical postulates create (synthesize) concepts (Mittelstrass,  1980b ). Definitions then fix the use of signs; they can be semantic definitions that explain the proper meaning of a sign in common language use (in a dictionary style) or they can be syntactic definitions that regulate the use of these signs in formal operations. Nominal definitions explain the words, while real definitions explain the meaning or the nature of the defined object. Definitions are thus essential for the development of a language of science, assuring communication and mutual understanding (Mittelstrass,  1980c ). Finally, hypotheses are also frequently defined as consistent conjectures that are compatible with the available knowledge. The truth of the hypothesis is only supposed in order to explain true observations and facts. Consequences of this hypothetical assumptions should explain the observed facts. Normally, descriptive hypotheses precede explanatory hypotheses in the development of scientific thought. Sometimes only tentative concepts are introduced as working hypotheses to test whether they have an explanatory capacity for the observations (Mittelstrass,  1980d ).

The Euclidian geometry is constructed along a logical “if→then” concept. The “if‐clause” formulates at the beginning the supposition, the “then clause” formulates the consequences from these axioms which provides a system of geometric theorems or insights. The conclusions do not follow directly from the hypothesis; this would otherwise represent self‐evident immediate conclusions. The “if‐then” concept in geometry is not used as in other branches of science where the consequences deduced from the axioms are checked against reality whether they are true, in order to confirm the validity of the hypothesis. The task in mathematics is: what can be logically deduced from a given set of axioms to build a contradiction‐free system of geometry. Whether this applies to the real world is in contrast to the situation in natural sciences another question and absolutely secondary to mathematics (Syntopicon,  1992 ).

Pascal's rules for hypotheses

In his Scientific Treatises on Geometric Demonstrations , Pascal ( 1623‐1662b ) formulates “Five rules are absolutely necessary and we cannot dispense with them without an essential defect and frequently even error. Do not leave undefined any terms at all obscure or ambiguous. Use in definitions of terms only words perfectly well known or already explained. Do not fail to ask that each of the necessary principles be granted, however clear and evident it may be. Ask only that perfectly self‐evident things be granted as axioms. Prove all propositions, using for their proof only axioms that are perfectly self‐evident or propositions already demonstrated or granted. Never get caught in the ambiguity of terms by failing to substitute in thought the definitions which restrict or define them. One should accept as true only those things whose contradiction appears to be false. We may then boldly affirm the original statement, however incomprehensible it is.”

Kant's rules on hypotheses

Kant ( 1724–1804 ) wrote that the analysis described in his book The Critique of Pure Reason “has now taught us that all its efforts to extend the bounds of knowledge by means of pure speculation, are utterly fruitless. So much the wider field lies open to hypothesis; as where we cannot know with certainty, we are at liberty to make guesses and to form suppositions. Imagination may be allowed, under the strict surveillance of reason, to invent suppositions; but these must be based on something that is perfectly certain‐ and that is the possibility of the object. Such a supposition is termed a hypothesis. We cannot imagine or invent any object or any property of an object not given in experience and employ it in a hypothesis; otherwise we should be basing our chain of reasoning upon mere chimerical fancies and not upon conception of things. Thus, we have no right to assume of new powers, not existing in nature and consequently we cannot assume that there is any other kind of community among substances than that observable in experience, any kind of presence than that in space and any kind of duration than that in time. The conditions of possible experience are for reason the only conditions of the possibility of things. Otherwise, such conceptions, although not self‐contradictory, are without object and without application. Transcendental hypotheses are therefore inadmissible, and we cannot use the liberty of employing in the absence of physical, hyperphysical grounds of explanation because such hypotheses do not advance reason, but rather stop it in its progress. When the explanation of natural phenomena happens to be difficult, we have constantly at hand a transcendental ground of explanation, which lifts us above the necessity of investigating nature. The next requisite for the admissibility of a hypothesis is its sufficiency. That is it must determine a priori the consequences which are given in experience and which are supposed to follow from the hypothesis itself.” Kant stresses another aspect when dealing with hypotheses: “It is our duty to try to discover new objections, to put weapons in the hands of our opponent, and to grant him the most favorable position. We have nothing to fear from these concessions; on the contrary, we may rather hope that we shall thus make ourselves master of a possession which no one will ever venture to dispute.”

For Kant's analytical and synthetical judgements and Difference between philosophy and mathematics (Kant, Whitehead) , see Appendices  S1 and S2 , respectively.

Poincaré on hypotheses

The mathematician‐philosopher Poincaré ( 1854 –1912a) explored the foundation of mathematics and physics in his book Science and Hypothesis . In the preface to the book, he summarizes common thinking of scientists at the end of the 19th century. “To the superficial observer scientific truth is unassailable, the logic of science is infallible, and if scientific men sometimes make mistakes, it is because they have not understood the rules of the game. Mathematical truths are derived from a few self‐evident propositions, by a chain of flawless reasoning, they are imposed not only by us, but on Nature itself. This is for the minds of most people the origin of certainty in science.” Poincaré then continues “but upon more mature reflection the position held by hypothesis was seen; it was recognized that it is as necessary to the experimenter as it is to the mathematician. And then the doubt arose if all these constructions are built on solid foundations.” However, “to doubt everything or to believe everything are two equally convenient solutions: both dispense with the necessity of reflection. Instead, we should examine with the utmost care the role of hypothesis; we shall then recognize not only that it is necessary, but that in most cases it is legitimate. We shall also see that there are several kinds of hypotheses; that some are verifiable and when once confirmed by experiment become truths of great fertility; that others may be useful to us in fixing our ideas; and finally that others are hypotheses only in appearance, and reduce to definitions or to conventions in disguise.” Poincaré argues that “we must seek mathematical thought where it has remained pure‐i.e. in arithmetic, in the proofs of the most elementary theorems. The process is proof by recurrence. We first show that a theorem is true for n  = 1; we then show that if it is true for n –1 it is true for n; and we conclude that it is true for all integers. The essential characteristic of reasoning by recurrence is that it contains, condensed in a single formula, an infinite number of syllogisms.” Syllogism is logical argument that applies deductive reasoning to arrive at a conclusion. Poincaré notes “that here is a striking analogy with the usual process of induction. But an essential difference exists. Induction applied to the physical sciences is always uncertain because it is based on the belief in a general order of the universe, an order which is external to us. Mathematical induction‐ i.e. proof by recurrence – is on the contrary, necessarily imposed on us, because it is only the affirmation of a property of the mind itself. No doubt mathematical recurrent reasoning and physical inductive reasoning are based on different foundations, but they move in parallel lines and in the same direction‐namely, from the particular to the general.”

Non‐Euclidian geometry: from Gauss to Lobatschewsky

Mathematics is an abstract science that intrinsically does not request that the structures described reflect a physical reality. Paradoxically, mathematics is the language of physics since the founder of experimental physics Galilei used Euclidian geometry when exploring the laws of the free fall. In his 1623 treatise The Assayer , Galilei ( 1564 –1642a) famously formulated that the book of Nature is written in the language of mathematics, thus establishing a link between formal concepts in mathematics and the structure of the physical world. Euclid's parallel axiom played historically a prominent role for the connection between mathematical concepts and physical realities. Mathematicians had doubted that the parallel axiom was needed and tried to prove it. In Euclidian geometry, there is a connection between the parallel axiom and the sum of the angles in a triangle being two right angles. It is therefore revealing that the famous mathematician C.F. Gauss investigated in the early 19th century experimentally whether this Euclidian theorem applies in nature. He approached this problem by measuring the sum of angles in a real triangle by using geodetic angle measurements of three geographical elevations in the vicinity of Göttingen where he was teaching mathematics. He reportedly measured a sum of the angles in this triangle that differed from 180°. Gauss had at the same time also developed statistical methods to evaluate the accuracy of measurements. Apparently, the difference of his measured angles was still within the interval of Gaussian error propagation. He did not publish the reasoning and the results for this experiment because he feared the outcry of colleagues about this unorthodox, even heretical approach to mathematical reasoning (Carnap,  1891 ‐1970a). However, soon afterwards non‐Euclidian geometries were developed. In the words of Poincaré, “Lobatschewsky assumes at the outset that several parallels may be drawn through a point to a given straight line, and he retains all the other axioms of Euclid. From these hypotheses he deduces a series of theorems between which it is impossible to find any contradiction, and he constructs a geometry as impeccable in its logic as Euclidian geometry. The theorems are very different, however, from those to which we are accustomed, and at first will be found a little disconcerting. For instance, the sum of the angles of a triangle is always less than two right angles, and the difference between that sum and two right angles is proportional to the area of the triangle. Lobatschewsky's propositions have no relation to those of Euclid, but are none the less logically interconnected.” Poincaré continues “most mathematicians regard Lobatschewsky's geometry as a mere logical curiosity. Some of them have, however, gone further. If several geometries are possible, they say, is it certain that our geometry is true? Experiments no doubt teaches us that the sum of the angles of a triangle is equal to two right angles, but this is because the triangles we deal with are too small” (Poincaré,  1854 ‐1912a)—hence the importance of Gauss' geodetic triangulation experiment. Gauss was aware that his three hills experiment was too small and thought on measurements on triangles formed with stars.

Poincaré vs. Einstein

Lobatschewsky's hyperbolic geometry did not remain the only non‐Euclidian geometry. Riemann developed a geometry without the parallel axiom, while the other Euclidian axioms were maintained with the exception of that of Order (Anordnung). Poincaré notes “so there is a kind of opposition between the geometries. For instance the sum of the angles in a triangle is equal to two right angles in Euclid's geometry, less than two right angles in that of Lobatschewsky, and greater than two right angles in that of Riemann. The number of parallel lines that can be drawn through a given point to a given line is one in Euclid's geometry, none in Riemann's, and an infinite number in the geometry of Lobatschewsky. Let us add that Riemann's space is finite, although unbounded.” As further distinction, the ratio of the circumference to the diameter of a circle is equal to π in Euclid's, greater than π in Lobatschewsky's and smaller than π in Riemann's geometry. A further difference between these geometries concerns the degree of curvature (Krümmungsmass k) which is 0 for a Euclidian surface, smaller than 0 for a Lobatschewsky and greater than 0 for a Riemann surface. The difference in curvature can be roughly compared with plane, concave and convex surfaces. The inner geometric structure of a Riemann plane resembles the surface structure of a Euclidean sphere and a Lobatschewsky plane resembles that of a Euclidean pseudosphere (a negatively curved geometry of a saddle). What geometry is true? Poincaré asked “Ought we then, to conclude that the axioms of geometry are experimental truths?” and continues “If geometry were an experimental science, it would not be an exact science. The geometric axioms are therefore neither synthetic a priori intuitions as affirmed by Kant nor experimental facts. They are conventions. Our choice among all possible conventions is guided by experimental facts; but it remains free and is only limited by the necessity of avoiding contradictions. In other words, the axioms of geometry are only definitions in disguise. What then are we to think of the question: Is Euclidean geometry true? It has no meaning. One geometry cannot be more true than another, it can only be more convenient. Now, Euclidean geometry is, and will remain, the most convenient, 1 st because it is the simplest and 2 nd because it sufficiently agrees with the properties of natural bodies” (Poincaré,  1854 ‐1912a).

Poincaré's book was published in 1903 and only a few years later Einstein published his general theory of relativity ( 1916 ) where he used a non‐Euclidean, Riemann geometry and where he demonstrated a structure of space that deviated from Euclidean geometry in the vicinity of strong gravitational fields. And in 1919, astronomical observations during a solar eclipse showed that light rays from a distant star were indeed “bent” when passing next to the sun. These physical observations challenged the view of Poincaré, and we should now address some aspects of hypotheses in physics (Carnap,  1891 ‐1970b).

HYPOTHESES IN PHYSICS

The long life of the five elements hypothesis.

Physical sciences—not to speak of biological sciences — were less developed in antiquity than mathematics which is already demonstrated by the primitive ideas on the elements constituting physical bodies. Plato and Aristotle spoke of the four elements which they took over from Thales (water), Anaximenes (air) and Parmenides (fire and earth) and add a fifth element (quinta essentia, our quintessence), namely ether. Ether is imagined a heavenly element belonging to the supralunar world. In Plato's dialogue Timaios (Plato,  c.424‐c.348 BC a ), the five elements were associated with regular polyhedra in geometry and became known as Platonic bodies: tetrahedron (fire), octahedron (air), cube (earth), icosahedron (water) and dodecahedron (ether). In regular polyhedra, faces are congruent (identical in shape and size), all angles and all edges are congruent, and the same number of faces meet at each vertex. The number of elements is limited to five because in Euclidian space there are exactly five regular polyhedral. There is in Plato's writing even a kind of geometrical chemistry. Since two octahedra (air) plus one tetrahedron (fire) can be combined into one icosahedron (water), these “liquid” elements can combine while this is not the case for combinations with the cube (earth). The 12 faces of the dodecahedron were compared with the 12 zodiac signs (Mittelstrass,  1980e ). This geometry‐based hypothesis of physics had a long life. As late as 1612, Kepler in his Mysterium cosmographicum tried to fit the Platonic bodies into the planetary shells of his solar system model. The ether theory even survived into the scientific discussion of the 19th‐century physics and the idea of a mathematical structure of the universe dominated by symmetry operations even fertilized 20th‐century ideas about symmetry concepts in the physics of elementary particles.

Huygens on sound waves in air

The ether hypothesis figures prominently in the 1690 Treatise on Light from Huygens ( 1617‐1670 ). He first reports on the transmission of sound by air when writing “this may be proved by shutting up a sounding body in a glass vessel from which the air is withdrawn and care was taken to place the sounding body on cotton that it cannot communicate its tremor to the glass vessel which encloses it. After having exhausted all the air, one hears no sound from the metal though it is struck.” Huygens comes up with some foresight when suspecting “the air is of such a nature that it can be compressed and reduced to a much smaller space than that it normally occupies. Air is made up of small bodies which float about and which are agitated very rapidly. So that the spreading of sound is the effort which these little bodies make in collisions with one another, to regain freedom when they are a little more squeezed together in the circuit of these waves than elsewhere.”

Huygens on light waves in ether

“That is not the same air but another kind of matter in which light spreads; since if the air is removed from the vessel the light does not cease to traverse it as before. The extreme velocity of light cannot admit such a propagation of motion” as sound waves. To achieve the propagation of light, Huygens invokes ether “as a substance approaching to perfect hardness and possessing springiness as prompt as we choose. One may conceive light to spread successively by spherical waves. The propagation consists nowise in the transport of those particles but merely in a small agitation which they cannot help communicate to those surrounding.” The hypothesis of an ether in outer space fills libraries of physical discussions, but all experimental approaches led to contradictions with respect to postulated properties of this hypothetical material for example when optical experiments showed that light waves display transversal and not longitudinal oscillations.

The demise of ether

Mechanical models for the transmission of light or gravitation waves requiring ether were finally put to rest by the theory of relativity from Einstein (Mittelstrass,  1980f ). This theory posits that the speed of light in an empty space is constant and does not depend on movements of the source of light or that of an observer as requested by the ether hypothesis. The theory of relativity also provides an answer how the force of gravitation is transmitted from one mass to another across an essentially empty space. In the non‐Euclidian formulation of the theory of relativity (Einstein used the Riemann geometry), there is no gravitation force in the sense of mechanical or electromagnetic forces. The gravitation force is in this formulation simply replaced by a geometric structure (space curvature near high and dense masses) of a four‐dimensional space–time system (Carnap,  1891 ‐1970c; Einstein & Imfeld,  1956 ) Gravitation waves and gravitation lens effects have indeed been experimental demonstrated by astrophysicists (Dorfmüller et al.,  1998 ).

For Aristotle's on physical hypotheses , see Appendix  S3 .

PHILOSOPHICAL THOUGHTS ON HYPOTHESES

In the following, the opinions of a number of famous scientists and philosophers on hypotheses are quoted to provide a historical overview on the subject.

Copernicus' hypothesis: a calculus which fits observations

In his book Revolutions of Heavenly Spheres Copernicus ( 1473–1543 ) reasoned in the preface about hypotheses in physics. “Since the newness of the hypotheses of this work ‐which sets the earth in motion and puts an immovable sun at the center of the universe‐ has already received a great deal of publicity, I have no doubt that certain of the savants have taken great offense.” He defended his heliocentric thesis by stating “For it is the job of the astronomer to use painstaking and skilled observations in gathering together the history of the celestial movements‐ and then – since he cannot by any line of reasoning reach the true causes of these movements‐ to think up or construct whatever causes or hypotheses he pleases such that, by the assumption of these causes, those same movements can be calculated from the principles of geometry for the past and the future too. This artist is markedly outstanding in both of these respects: for it is not necessary that these hypotheses should be true, or even probable; but it is enough if they provide a calculus which fits the observations.” This preface written in 1543 sounds in its arguments very modern physics. However, historians of science have discovered that it was probably written by a theologian friend of Copernicus to defend the book against the criticism by the church.

Bacon's intermediate hypotheses

In his book Novum Organum , Francis Bacon ( 1561–1626 ) claims for hypotheses and scientific reasoning “that they augur well for the sciences, when the ascent shall proceed by a true scale and successive steps, without interruption or breach, from particulars to the lesser axioms, thence to the intermediates and lastly to the most general.” He then notes “that the lowest axioms differ but little from bare experiments, the highest and most general are notional, abstract, and of no real weight. The intermediate are true, solid, full of life, and up to them depend the business and fortune of mankind.” He warns that “we must not then add wings, but rather lead and ballast to the understanding, to prevent its jumping and flying, which has not yet been done; but whenever this takes place we may entertain greater hopes of the sciences.” With respect to methodology, Bacon claims that “we must invent a different form of induction. The induction which proceeds by simple enumeration is puerile, leads to uncertain conclusions, …deciding generally from too small a number of facts. Sciences should separate nature by proper rejections and exclusions and then conclude for the affirmative, after collecting a sufficient number of negatives.”

Gilbert and Descartes for plausible hypotheses

William Gilbert introduced in his book On the Loadstone (Gilbert,  1544‐1603 ) the argument of plausibility into physical hypothesis building. “From these arguments, therefore, we infer not with mere probability, but with certainty, the diurnal rotation of the earth; for nature ever acts with fewer than with many means; and because it is more accordant to reason that the one small body, the earth, should make a daily revolution than the whole universe should be whirled around it.”

Descartes ( 1596‐1650 ) reflected on the sources of understanding in his book Rules for Direction and distinguished what “comes about by impulse, by conjecture, or by deduction. Impulse can assign no reason for their belief and when determined by fanciful disposition, it is almost always a source of error.” When speaking about the working of conjectures he quotes thoughts of Aristotle: “water which is at a greater distance from the center of the globe than earth is likewise less dense substance, and likewise the air which is above the water, is still rarer. Hence, we hazard the guess that above the air nothing exists but a very pure ether which is much rarer than air itself. Moreover nothing that we construct in this way really deceives, if we merely judge it to be probable and never affirm it to be true; in fact it makes us better instructed. Deduction is thus left to us as the only means of putting things together so as to be sure of their truth. Yet in it, too, there may be many defects.”

Care in formulating hypotheses

Locke ( 1632‐1704 ) in his treatise Concerning Human Understanding admits that “we may make use of any probable hypotheses whatsoever. Hypotheses if they are well made are at least great helps to the memory and often direct us to new discoveries. However, we should not take up any one too hastily.” Also, practising scientists argued against careless use of hypotheses and proposed remedies. Lavoisier ( 1743‐1794 ) in the preface to his Element of Chemistry warned about beaten‐track hypotheses. “Instead of applying observation to the things we wished to know, we have chosen rather to imagine them. Advancing from one ill‐founded supposition to another, we have at last bewildered ourselves amidst a multitude of errors. These errors becoming prejudices, are adopted as principles and we thus bewilder ourselves more and more. We abuse words which we do not understand. There is but one remedy: this is to forget all that we have learned, to trace back our ideas to their sources and as Bacon says to frame the human understanding anew.”

Faraday ( 1791–1867 ) in a Speculation Touching Electric Conduction and the Nature of Matter highlighted the fundamental difference between hypotheses and facts when noting “that he has most power of penetrating the secrets of nature, and guessing by hypothesis at her mode of working, will also be most careful for his own safe progress and that of others, to distinguish that knowledge which consists of assumption, by which I mean theory and hypothesis, from that which is the knowledge of facts and laws; never raising the former to the dignity or authority of the latter.”

Explicatory power justifies hypotheses

Darwin ( 1809 –1882a) defended the conclusions and hypothesis of his book The Origin of Species “that species have been modified in a long course of descent. This has been affected chiefly through the natural selection of numerous, slight, favorable variations.” He uses a post hoc argument for this hypothesis: “It can hardly be supposed that a false theory would explain, to so satisfactory a manner as does the theory of natural selection, the several large classes of facts” described in his book.

The natural selection of hypotheses

In the concluding chapter of The Descent of Man Darwin ( 1809 –1882b) admits “that many of the views which have been advanced in this book are highly speculative and some no doubt will prove erroneous.” However, he distinguished that “false facts are highly injurious to the progress of science for they often endure long; but false views do little harm for everyone takes a salutory pleasure in proving their falseness; and when this is done, one path to error is closed and the road to truth is often at the same time opened.”

The American philosopher William James ( 1842–1907 ) concurred with Darwin's view when he wrote in his Principles of Psychology “every scientific conception is in the first instance a spontaneous variation in someone'’s brain. For one that proves useful and applicable there are a thousand that perish through their worthlessness. The scientific conceptions must prove their worth by being verified. This test, however, is the cause of their preservation, not of their production.”

The American philosopher J. Dewey ( 1859‐1952 ) in his treatise Experience and Education notes that “the experimental method of science attaches more importance not less to ideas than do other methods. There is no such thing as experiment in the scientific sense unless action is directed by some leading idea. The fact that the ideas employed are hypotheses, not final truths, is the reason why ideas are more jealously guarded and tested in science than anywhere else. As fixed truths they must be accepted and that is the end of the matter. But as hypotheses, they must be continuously tested and revised, a requirement that demands they be accurately formulated. Ideas or hypotheses are tested by the consequences which they produce when they are acted upon. The method of intelligence manifested in the experimental method demands keeping track of ideas, activities, and observed consequences. Keeping track is a matter of reflective review.”

The reductionist principle

James ( 1842‐1907 ) pushed this idea further when saying “Scientific thought goes by selection. We break the solid plenitude of fact into separate essences, conceive generally what only exists particularly, and by our classifications leave nothing in its natural neighborhood. The reality exists as a plenum. All its part are contemporaneous, but we can neither experience nor think this plenum. What we experience is a chaos of fragmentary impressions, what we think is an abstract system of hypothetical data and laws. We must decompose each chaos into single facts. We must learn to see in the chaotic antecedent a multitude of distinct antecedents, in the chaotic consequent a multitude of distinct consequents.” From these considerations James concluded “even those experiences which are used to prove a scientific truth are for the most part artificial experiences of the laboratory gained after the truth itself has been conjectured. Instead of experiences engendering the inner relations, the inner relations are what engender the experience here.“

Following curiosity

Freud ( 1856–1939 ) considered curiosity and imagination as driving forces of hypothesis building which need to be confronted as quickly as possible with observations. In Beyond the Pleasure Principle , Freud wrote “One may surely give oneself up to a line of thought and follow it up as far as it leads, simply out of scientific curiosity. These innovations were direct translations of observation into theory, subject to no greater sources of error than is inevitable in anything of the kind. At all events there is no way of working out this idea except by combining facts with pure imagination and thereby departing far from observation.” This can quickly go astray when trusting intuition. Freud recommends “that one may inexorably reject theories that are contradicted by the very first steps in the analysis of observation and be aware that those one holds have only a tentative validity.”

Feed‐forward aspects of hypotheses

The geneticist Waddington ( 1905–1975 ) in his essay The Nature of Life states that “a scientific theory cannot remain a mere structure within the world of logic, but must have implications for action and that in two rather different ways. It must involve the consequence that if you do so and so, such and such result will follow. That is to say it must give, or at least offer, the possibility of controlling the process. Secondly, its value is quite largely dependent on its power of suggesting the next step in scientific advance. Any complete piece of scientific work starts with an activity essentially the same as that of an artist. It starts by asking a relevant question. The first step may be a new awareness of some facet of the world that no one else had previously thought worth attending to. Or some new imaginative idea which depends on a sensitive receptiveness to the oddity of nature essentially similar to that of the artist. In his logical analysis and manipulative experimentation, the scientist is behaving arrogantly towards nature, trying to force her into his categories of thought or to trick her into doing what he wants. But finally he has to be humble. He has to take his intuition, his logical theory and his manipulative skill to the bar of Nature and see whether she answers yes or no; and he has to abide by the result. Science is often quite ready to tolerate some logical inadequacy in a theory‐or even a flat logical contradiction like that between the particle and wave theories of matter‐so long as it finds itself in the possession of a hypothesis which offers both the possibility of control and a guide to worthwhile avenues of exploration.”

Poincaré: the dialogue between experiment and hypothesis

Poincaré ( 1854 –1912b) also dealt with physics in Science and Hypothesis . “Experiment is the sole source of truth. It alone can teach us certainty. Cannot we be content with experiment alone? What place is left for mathematical physics? The man of science must work with method. Science is built up of facts, as a house is built of stones, but an accumulation of facts is no more a science than a heap of stones is a house. It is often said that experiments should be made without preconceived concepts. That is impossible. Without the hypothesis, no conclusion could have been drawn; nothing extraordinary would have been seen; and only one fact the more would have been catalogued, without deducing from it the remotest consequence.” Poincaré compares science to a library. Experimental physics alone can enrich the library with new books, but mathematical theoretical physics draw up the catalogue to find the books and to reveal gaps which have to be closed by the purchase of new books.

Poincaré: false, true, fruitful and dangerous hypotheses

Poincaré continues “we all know that there are good and bad experiments. The latter accumulate in vain. Whether there are hundred or thousand, one single piece of work will be sufficient to sweep them into oblivion. Bacon invented the term of an experimentum crucis for such experiments. What then is a good experiment? It is that which teaches us something more than an isolated fact. It is that which enables us to predict and to generalize. Experiments only gives us a certain number of isolated points. They must be connected by a continuous line and that is true generalization. Every generalization is a hypothesis. It should be as soon as possible submitted to verification. If it cannot stand the test, it must be abandoned without any hesitation. The physicist who has just given up one of his hypotheses should rejoice, for he found an unexpected opportunity of discovery. The hypothesis took into account all the known factors which seem capable of intervention in the phenomenon. If it is not verified, it is because there is something unexpected. Has the hypothesis thus rejected been sterile? Far from it. It has rendered more service than a true hypothesis.” Poincaré notes that “with a true hypothesis only one fact the more would have been catalogued, without deducing from it the remotest consequence. It may be said that the wrong hypothesis has rendered more service than a true hypothesis.” However, Poincaré warns that “some hypotheses are dangerous – first and foremost those which are tacit and unconscious. And since we make them without knowing them, we cannot get rid of them.” Poincaré notes that here mathematical physics is of help because by its precision one is compelled to formulate all the hypotheses, revealing also the tacit ones.

Arguments for the reductionist principle

Poincaré also warned against multiplying hypotheses indefinitely: “If we construct a theory upon multiple hypotheses, and if experiment condemns it, which of the premisses must be changed?” Poincaré also recommended to “resolve the complex phenomenon given directly by experiment into a very large number of elementary phenomena. First, with respect to time. Instead of embracing in its entirety the progressive development of a phenomenon, we simply try to connect each moment with the one immediately preceding. Next, we try to decompose the phenomenon in space. We must try to deduce the elementary phenomenon localized in a very small region of space.” Poincaré suggested that the physicist should “be guided by the instinct of simplicity, and that is why in physical science generalization so readily takes the mathematical form to state the problem in the form of an equation.” This argument goes back to Galilei ( 1564 –1642b) who wrote in The Two Sciences “when I observe a stone initially at rest falling from an elevated position and continually acquiring new increments of speed, why should I not believe that such increases take place in a manner which is exceedingly simple and rather obvious to everybody? If now we examine the matter carefully we find no addition or increment more simple than that which repeats itself always in the same manner. It seems we shall not be far wrong if we put the increment of speed as proportional to the increment of time.” With a bit of geometrical reasoning, Galilei deduced that the distance travelled by a freely falling body varies as the square of the time. However, Galilei was not naïve and continued “I grant that these conclusions proved in the abstract will be different when applied in the concrete” and considers disturbances cause by friction and air resistance that complicate the initially conceived simplicity.

Four sequential steps of discovery…

Some philosophers of science attributed a fundamental importance to observations for the acquisition of experience in science. The process starts with accidental observations (Aristotle), going to systematic observations (Bacon), leading to quantitative rules obtained with exact measurements (Newton and Kant) and culminating in observations under artificially created conditions in experiments (Galilei) (Mittelstrass,  1980g ).

…rejected by Popper and Kant

In fact, Newton wrote that he had developed his theory of gravitation from experience followed by induction. K. Popper ( 1902‐1994 ) in his book Conjectures and Refutations did not agree with this logical flow “experience leading to theory” and that for several reasons. This scheme is according to Popper intuitively false because observations are always inexact, while theory makes absolute exact assertions. It is also historically false because Copernicus and Kepler were not led to their theories by experimental observations but by geometry and number theories of Plato and Pythagoras for which they searched verifications in observational data. Kepler, for example, tried to prove the concept of circular planetary movement influenced by Greek theory of the circle being a perfect geometric figure and only when he could not demonstrate this with observational data, he tried elliptical movements. Popper noted that it was Kant who realized that even physical experiments are not prior to theories when quoting Kant's preface to the Critique of Pure Reason : “When Galilei let his globes run down an inclined plane with a gravity which he has chosen himself, then a light dawned on all natural philosophers. They learnt that our reason can only understand what it creates according to its own design; that we must compel Nature to answer our questions, rather than cling to Nature's apron strings and allow her to guide us. For purely accidental observations, made without any plan having been thought out in advance, cannot be connected by a law‐ which is what reason is searching for.” From that reasoning Popper concluded that “we ourselves must confront nature with hypotheses and demand a reply to our questions; and that lacking such hypotheses, we can only make haphazard observations which follow no plan and which can therefore never lead to a natural law. Everyday experience, too, goes far beyond all observations. Everyday experience must interpret observations for without theoretical interpretation, observations remain blind and uninformative. Everyday experience constantly operates with abstract ideas, such as that of cause and effect, and so it cannot be derived from observation.” Popper agreed with Kant who said “Our intellect does not draw its laws from nature…but imposes them on nature”. Popper modifies this statement to “Our intellect does not draw its laws from nature, but tries‐ with varying degrees of success – to impose upon nature laws which it freely invents. Theories are seen to be free creations of our mind, the result of almost poetic intuition. While theories cannot be logically derived from observations, they can, however, clash with observations. This fact makes it possible to infer from observations that a theory is false. The possibility of refuting theories by observations is the basis of all empirical tests. All empirical tests are therefore attempted refutations.”

OUTLOOK: HYPOTHESES IN BIOLOGY

Is biology special.

Waddington notes that “living organisms are much more complicated than the non‐living things. Biology has therefore developed more slowly than sciences such as physics and chemistry and has tended to rely on them for many of its basic ideas. These older physical sciences have provided biology with many firm foundations which have been of the greatest value to it, but throughout most of its history biology has found itself faced with the dilemma as to how far its reliance on physics and chemistry should be pushed” both with respect to its experimental methods and its theoretical foundations. Vitalism is indeed such a theory maintaining that organisms cannot be explained solely by physicochemical laws claiming specific biological forces active in organisms. However, efforts to prove the existence of such vital forces have failed and today most biologists consider vitalism a superseded theory.

Biology as a branch of science is as old as physics. If one takes Aristotle as a reference, he has written more on biology than on physics. Sophisticated animal experiments were already conducted in the antiquity by Galen (Brüssow, 2022 ). Alertus Magnus displayed biological research interest during the medieval time. Knowledge on plants provided the basis of medical drugs in early modern times. What explains biology's decreasing influence compared with the rapid development of physics by Galilei and Newton? One reason is the possibility to use mathematical equations to describe physical phenomena which was not possible for biological phenomena. Physics has from the beginning displayed a trend to few fundamental underlying principles. This is not the case for biology. With the discovery of new continents, biologists were fascinated by the diversity of life. Diversity was the conducting line of biological thinking. This changed only when taxonomists and comparative anatomists revealed recurring pattern in this stunning biological variety and when Darwin provided a theoretical concept to understand variation as a driving force in biology. Even when genetics and molecular biology allowed to understand biology from a few universally shared properties, such as a universal genetic code, biology differed in fundamental aspects from physics and chemistry. First, biology is so far restricted to the planet earth while the laws of physic and chemistry apply in principle to the entire universe. Second, biology is to a great extent a historical discipline; many biological processes cannot be understood from present‐day observations because they are the result of historical developments in evolution. Hence, the importance of Dobzhansky's dictum that nothing makes sense in biology except in the light of evolution. The great diversity of life forms, the complexity of processes occurring in cells and their integration in higher organisms and the importance of a historical past for the understanding of extant organisms, all that has delayed the successful application of mathematical methods in biology or the construction of theoretical frameworks in biology. Theoretical biology by far did not achieve a comparable role as theoretical physics which is on equal foot with experimental physics. Many biologists are even rather sceptical towards a theoretical biology and see progress in the development of ever more sophisticated experimental methods instead in theoretical concepts expressed by new hypotheses.

Knowledge from data without hypothesis?

Philosophers distinguish rational knowledge ( cognitio ex principiis ) from knowledge from data ( cognitio ex data ). Kant associates these two branches with natural sciences and natural history, respectively. The latter with descriptions of natural objects as prominently done with systematic classification of animals and plants or, where it is really history, when describing events in the evolution of life forms on earth. Cognitio ex data thus played a much more prominent role in biology than in physics and explains why the compilation of data and in extremis the collection of museum specimen characterizes biological research. To account for this difference, philosophers of the logical empiricism developed a two‐level concept of science languages consisting of a language of observations (Beobachtungssprache) and a language of theories (Theoriesprache) which are linked by certain rules of correspondence (Korrespondenzregeln) (Carnap,  1891 –1970d). If one looks into leading biological research journals, it becomes clear that biology has a sophisticated language of observation and a much less developed language of theories.

Do we need more philosophical thinking in biology or at least a more vigorous theoretical biology? The breathtaking speed of progress in experimental biology seems to indicate that biology can well develop without much theoretical or philosophical thinking. At the same time, one could argue that some fields in biology might need more theoretical rigour. Microbiologists might think on microbiome research—one of the breakthrough developments of microbiology research in recent years. The field teems with fascinating, but ill‐defined terms (our second genome; holobionts; gut–brain axis; dysbiosis, symbionts; probiotics; health benefits) that call for stricter definitions. One might also argue that biologists should at least consider the criticism of Goethe ( 1749–1832 ), a poet who was also an active scientist. In Faust , the devil ironically teaches biology to a young student.

“Wer will was Lebendigs erkennen und beschreiben, Sucht erst den Geist herauszutreiben, Dann hat er die Teile in seiner Hand, Fehlt, leider! nur das geistige Band.” (To docket living things past any doubt. You cancel first the living spirit out: The parts lie in the hollow of your hand, You only lack the living thing you banned).

We probably need both in biology: more data and more theory and hypotheses.

CONFLICT OF INTEREST

The author reports no conflict of interest.

FUNDING INFORMATION

No funding information provided.

Supporting information

Appendix S1

Brüssow, H. (2022) On the role of hypotheses in science . Microbial Biotechnology , 15 , 2687–2698. Available from: 10.1111/1751-7915.14141 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]

  • Bacon, F. (1561. –1626) Novum Organum. In: Adler, M.J. (Ed.) (editor‐in‐chief) Great books of the western world . Chicago, IL: Encyclopaedia Britannica, Inc. 2nd edition 1992 vol 1–60 (abbreviated below as GBWW) here: GBWW vol. 28: 128. [ Google Scholar ]
  • Brüssow, H. (2022) What is Truth – in science and beyond . Environmental Microbiology , 24 , 2895–2906. [ PubMed ] [ Google Scholar ]
  • Carnap, R. (1891. ‐1970a) Philosophical foundations of physics. Ch. 14 . Basic Books, Inc., New York, 1969. [ Google Scholar ]
  • Carnap, R. (1891. ‐1970b) Philosophical foundations of physics. Ch. 15 . Basic Books, Inc., New York, 1969. [ Google Scholar ]
  • Carnap, R. (1891. ‐1970c) Philosophical foundations of physics. Ch. 16 . Basic Books, Inc., New York, 1969. [ Google Scholar ]
  • Carnap, R. (1891. ‐1970d) Philosophical foundations of physics. Ch. 27–28 . Basic Books, Inc., New York, 1969. [ Google Scholar ]
  • Copernicus . (1473. ‐1543) Revolutions of heavenly spheres . GBWW , vol. 15 , 505–506. [ Google Scholar ]
  • Darwin, C. (1809. ‐1882a) The origin of species . GBWW , vol. 49 : 239. [ Google Scholar ]
  • Darwin, C. (1809. ‐1882b) The descent of man . GBWW , vol. 49 : 590. [ Google Scholar ]
  • Descartes, R. (1596. ‐1650) Rules for direction . GBWW , vol. 28 , 245. [ Google Scholar ]
  • Dewey, J. (1859. –1952) Experience and education . GBWW , vol. 55 , 124. [ Google Scholar ]
  • Dorfmüller, T. , Hering, W.T. & Stierstadt, K. (1998) Bergmann Schäfer Lehrbuch der Experimentalphysik: Band 1 Mechanik, Relativität, Wärme. In: Was ist Schwerkraft: Von Newton zu Einstein . Berlin, New York: Walter de Gruyter, pp. 197–203. [ Google Scholar ]
  • Einstein, A. (1916) Relativity . GBWW , vol. 56 , 191–243. [ Google Scholar ]
  • Einstein, A. & Imfeld, L. (1956) Die Evolution der Physik . Hamburg: Rowohlts deutsche Enzyklopädie, Rowohlt Verlag. [ Google Scholar ]
  • Euclid . (c.323‐c.283) The elements . GBWW , vol. 10 , 1–2. [ Google Scholar ]
  • Faraday, M. (1791. –1867) Speculation touching electric conduction and the nature of matter . GBWW , 42 , 758–763. [ Google Scholar ]
  • Freud, S. (1856. –1939) Beyond the pleasure principle . GBWW , vol. 54 , 661–662. [ Google Scholar ]
  • Galilei, G. (1564. ‐1642a) The Assayer, as translated by S. Drake (1957) Discoveries and Opinions of Galileo pp. 237–8 abridged pdf at Stanford University .
  • Galilei, G. (1564. ‐1642b) The two sciences . GBWW vol. 26 : 200. [ Google Scholar ]
  • Gilbert, W. (1544. ‐1603) On the Loadstone . GBWW , vol. 26 , 108–110. [ Google Scholar ]
  • Goethe, J.W. (1749. –1832) Faust . GBWW , vol. 45 , 20. [ Google Scholar ]
  • Hilbert, D. (1899) Grundlagen der Geometrie . Leipzig, Germany: Verlag Teubner. [ Google Scholar ]
  • Huygens, C. (1617. ‐1670) Treatise on light . GBWW , vol. 32 , 557–560. [ Google Scholar ]
  • James, W. (1842. –1907) Principles of psychology . GBWW , vol. 53 , 862–866. [ Google Scholar ]
  • Kant, I. (1724. –1804) Critique of pure reason . GBWW , vol. 39 , 227–230. [ Google Scholar ]
  • Lavoisier, A.L. (1743. ‐1794) Element of chemistry . GBWW , vol. 42 , p. 2, 6‐7, 9‐10. [ Google Scholar ]
  • Locke, J. (1632. ‐1704) Concerning Human Understanding . GBWW , vol. 33 , 317–362. [ Google Scholar ]
  • Mittelstrass, J. (1980a) Enzyklopädie Philosophie und Wissenschaftstheorie Bibliographisches Institut Mannheim, Wien, Zürich B.I. Wissenschaftsverlag Vol. 1: 239–241 .
  • Mittelstrass, J. (1980b) Enzyklopädie Philosophie und Wissenschaftstheorie Bibliographisches Institut Mannheim, Wien, Zürich B.I. Wissenschaftsverlag Vol. 3: 307 .
  • Mittelstrass, J. (1980c) Enzyklopädie Philosophie und Wissenschaftstheorie Bibliographisches Institut Mannheim, Wien, Zürich B.I. Wissenschaftsverlag Vol. 1: 439–442 .
  • Mittelstrass, J. (1980d) Enzyklopädie Philosophie und Wissenschaftstheorie Bibliographisches Institut Mannheim, Wien, Zürich B.I. Wissenschaftsverlag Vol. 2: 157–158 .
  • Mittelstrass, J. (1980e) Enzyklopädie Philosophie und Wissenschaftstheorie Bibliographisches Institut Mannheim, Wien, Zürich B.I. Wissenschaftsverlag Vol. 3: 264‐267, 449.450 .
  • Mittelstrass, J. (1980f) Enzyklopädie Philosophie und Wissenschaftstheorie Bibliographisches Institut Mannheim, Wien, Zürich B.I. Wissenschaftsverlag Vol. 1: 209–210 .
  • Mittelstrass, J. (1980g) Enzyklopädie Philosophie und Wissenschaftstheorie Bibliographisches Institut Mannheim, Wien, Zürich B.I. Wissenschaftsverlag Vol. 1: 281–282 .
  • Pascal, B. (1623. ‐1662a) Pensées GBWW vol. 30 : 171–173. [ Google Scholar ]
  • Pascal, B. (1623. ‐1662b) Scientific treatises on geometric demonstrations . GBWW vol. 30 : 442–443. [ Google Scholar ]
  • Plato . (c.424‐c.348 BC a) Timaeus . GBWW , vol. 6 , 442–477. [ Google Scholar ]
  • Poincaré, H. (1854. ‐1912a) Science and hypothesis GBWW , vol. 56 : XV‐XVI, 1–5, 10–15 [ Google Scholar ]
  • Poincaré, H. (1854. ‐1912b) Science and hypothesis GBWW , vol. 56 : 40–52. [ Google Scholar ]
  • Popper, K. (1902. ‐1994) Conjectures and refutations . London and New York, 2002: The Growth of Scientific Knowledge Routledge Classics, pp. 249–261. [ Google Scholar ]
  • Syntopicon . (1992) Hypothesis . GBWW , vol. 1 , 576–587. [ Google Scholar ]
  • Waddington, C.H. (1905. –1975) The nature of life . GBWW , vol. 56 , 697–699. [ Google Scholar ]

1 The ideas in the previous two paragraphs are from the fascinating course "Philosophy of Science" by Jeffery L. Kasser, published by The Teaching Company

[ Chapter 1 Objectives ]

Talk to our experts

1800-120-456-456

  • Scientific Hypothesis

ffImage

Define Hypothesis in Science

A scientific hypothesis is a tentative statement that may not have a surety. A hypothesis in science is a paper-made explanation about a scientific phenomenon or a narrow set of phenomena observed in nature.

Scientists work on the previous knowledge and they believe that to be useful in a science a hypothesis must be able to satisfy the previous observations.

To be scientifically valid a hypothesis must be testable and the result must be accurate. In this article, we will discuss hypothesis definition science, hypothesis definition scientific method in detail.

Hypothesis in Science

A scientific hypothesis must be supportive to experiments and observations, otherwise, it remains a fairy tale just like you used to listen to from your grandmother. 

Now, let’s understand the hypothesis definition of science:

A scientific hypothesis is the first brick of the building in the scientific method. Many scientists describe the hypothesis definition scientific method as an "educated guess," based on previous knowledge and observation. While this is true, the scientific hypothesis definition can be expanded; a hypothesis explains why the guess may be correct. 

Evolution of Scientific Hypothesis

Most formal hypotheses comprise some concepts that can be connected to something and their relationships can be tested. A group of hypotheses comes together to form a conceptual argument. As sufficient data and evidence are collected to support a hypothesis, it becomes a working or a valid hypothesis, which is an important task on the way to becoming a theory. 

Though hypotheses and theories are often confusing; however, theories are the result of a tested hypothesis, while hypotheses are ideas or theories that explain the findings of the testing of those ideas or theories. 

According to Tanner: Scientific Hypotheses are theories or the ways that make sense of what we observe in the natural world. Theories are structures of ideas that explain and descriptively interpret facts.

Basics of Hypothesis

A hypothesis is a defined set of solutions for an unexplainable occurrence of something in the past that does not fit into the present accepted scientific theory/method. The basic idea of a hypothesis is that there is no predetermined result/outcome. To be scientifically valid a hypothesis must be supported or refuted through carefully crafted experimentation or observation.

Key Function of Scientific Method

A key function in the scientific method is generating predictions or premises from the hypotheses about the results of future experiments, and then performing experiments on these premises to see whether they support the predictions/premises and arguments.

We write a hypothesis in the form of an if or then statement, according to the University of California. This statement gives an ‘if’ possibility and explains what may/may not happen because of the possibility (then). The statement includes "may."

[Image will be Uploaded Soon]

Examples Supporting Scientific Hypothesis

Below is the list of statements supporting the scientific hypothesis:

If garlic leaf repels fleas, then a cat that is given garlic every day will not get fleas.

Bacterial growth may get affected by higher/lower moisture levels in the air.

If sugar can cause cavities, then people who eat a lot of candy or chocolate may be more prone to cavities.

If UV rays can damage your eyes, then maybe UV light may cause blindness.

Testing a Scientific Hypothesis

A hypothesis is examined by multiple scientists to ensure the integrity and accuracy of the experiment. This process can take a long period of years, and in many cases, the hypothesis scientific method does not undergo any further in the scientific method as it is difficult to collect sufficient supporting evidence.

Also, notice that all of the statements mentioned above are testable. The fundamental trait of a hypothesis is that something can be tested and that these tests can be reproduced, according to Midwestern State University.

An illustration of the untestable statement is, "Everyone love to visit Australia at least once in their life." The definition of love is subjective. Also, it would be an impossible task to poll every human about their travel history. An untestable statement can still be reworded to make it testable, though. 

For instance, the previous statement can be changed to, "If love is an important emotion and “visit” is one of the interests, some may believe that everyone should visit Australia once.

With the above statement, the researcher can work on a group of people to see how many love to visit Australia at least once in their lifetime.

Hypothesis Definition Science

To be useful in science a hypothesis must be supported with arguments made; however, scientists encounter the following issues:

During testing, a scientist may encounter two types of errors, i.e., Type I and Type II errors.

A Type I error occurs when the null hypothesis is rejected though it is true. 

A Type II error encounters when the null hypothesis is not rejected even though it is false.

arrow-right

FAQs on Scientific Hypothesis

Q1: State the Steps Involved in the Scientific Method.

Ans: In the scientific method, whether it is researching in the field of psychology, biology, or some other area, a hypothesis represents what the researchers think may happen truly in an experiment.

The scientific method involves the following eight steps:

 Forming a question

Performing background research

Creating a new hypothesis

Designing an experiment on the hypothesis

Collecting data after the experiment

Analyzing those results

Drawing conclusions

Communicating the results with the teammates

So, while studying and exploring the effects of a particular drug, the researchers might expect that hypothesis of the drug to have some type of effect on the symptoms of a specific illness. In psychology, the hypothesis may focus on how a certain aspect of the surroundings might influence a person’s behaviour.

Q2: State the Elements of a Good Hypothesis.

Ans: Scientists while trying to come up with a good hypothesis for their research or experiments, ask themselves the following questions:

Is their hypothesis based on their research on a topic?

Can the hypothesis be tested?

Does their hypothesis include either independent or dependent variables or both?

To form a hypothesis, scientists follow the steps like, they gather as many observations about a topic or problem as they can, evaluate observations and look for possible reasons for the problem, create a list of possible experimentations that can be performed on these.

After they develop a few possible hypotheses, they think of ways to confirm or disprove each hypothesis through experimentation. This process is known as falsifiability.

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • CAREER Q&A
  • 26 March 2024

The beauty of what science can do when urgently needed

  • Katherine Bourzac 0

Katherine Bourzac is a freelance journalist in San Francisco, California.

You can also search for this author in PubMed   Google Scholar

A woman sits in an office room with a blue wall. A chart is shown on the glowing screen behind her.

Cultivarium chief scientific officer Nili Ostrov works to make model organisms more useful and accessible for scientific research Credit: Donis Perkins

Nili Ostrov has always been passionate about finding ways to use biology for practical purposes. So perhaps it wasn’t surprising that, when the COVID-19 pandemic hit during her postdoctoral studies, she went in the opposite direction from most people, moving to New York City to work as the director of molecular diagnostics in the Pandemic Response Lab, providing COVID-19 tests and surveilling viral variants. She was inspired by seeing what scientists could accomplish and how much they could help when under pressure.

Now the chief scientific officer at Cultivarium in Watertown, Massachusetts, Ostrov is bringing that sense of urgency to fundamental problems in synthetic biology. Cultivarium is a non-profit focused research organization, a structure that comes with a finite amount of time and funding to pursue ‘moonshot’ scientific goals, which would usually be difficult for academic laboratories or start-up companies to achieve. Cultivarium has five years of funding, which started in 2022, to develop tools to make it possible for scientists to genetically engineer unconventional model organisms — a group that includes most microbes.

Typically, scientists are limited to working with yeast, the bacterium Escherichia coli and other common lab organisms, because the necessary conditions to grow and manipulate them are well understood. Ostrov wants to make it easier to engineer other microbes, such as soil bacteria or microorganisms that live in extreme conditions, for scientific purposes. This could open up new possibilities for biomanufacturing drugs or transportation fuels and solving environmental problems.

What is synthetic biology and what drew you to it?

Synthetic biology melds biology and engineering — it is the level at which you say, “I know how this part works. What can I do with it?” Synthetic biologists ask questions such as, what is this part useful for? How can it benefit people or the environment in some way?

During my PhD programme at Columbia University in New York City, my team worked with the yeast that is used for brewing beer — but we asked, can you use these yeast cells as sensors? Because yeast cells can sense their environment, we could engineer them to detect a pathogen in a water sample. In my postdoctoral work at Harvard University in Cambridge, Massachusetts, we investigated a marine bacterium, Vibrio natriegens . A lot of time during research is spent waiting for cells to grow. V. natriegens doubles in number about every ten minutes — the fastest growth rate of any organism.

Could we use it to speed up research? But using V. natriegens and other uncommon research organisms is hard work. You have to develop the right genetic-engineering tools.

How did the COVID-19 pandemic alter your career trajectory?

It pushed me to do something that I otherwise would not have done. During my postdoctoral programme, I met Jef Boeke, a synthetic biologist at New York University. In 2020, he asked me whether I wanted to help with the city’s Pandemic Response Lab, because of my expertise in DNA technology. I’m probably one of the only people with a newborn baby who moved into Manhattan when COVID-19 hit.

That was an amazing experience: I took my science and skills and used them for something essential and urgent. In a couple of months, we set up a lab that supported the city’s health system. We monitored for new variants of the virus using genomic sequencing and ran diagnostic tests.

Seeing what science can do when needed — it was beautiful. It showed me how effective science can be, and how fast science can move with the right set-up.

How did that influence what you’re doing now with Cultivarium?

COVID-19 showed me how urgently needed science can be done. It’s about bringing together the right people from different disciplines. Cultivarium is addressing fundamental problems in science, which is usually done in academic settings, with the fast pace and the dynamic of a start-up company.

We need to make progress on finding ways to use unconventional microbes to advance science. A lot of bioproduction of industrial and therapeutic molecules is done in a few model organisms, such as E. coli and yeast. Imagine what you could achieve if you had 100 different organisms. If you’re looking to produce a protein that needs to be made in high temperatures or at an extreme pH, you can’t use E. coli , because it won’t grow.

How is Cultivarium making unconventional microbes research-friendly?

It took my postdoctoral lab team six years to get to the point where we could take V. natriegens , which we initially didn’t know how to grow well or engineer, and knock out every gene in its genome.

At Cultivarium, we’re taking a more systematic approach to provide those culturing and engineering tools for researchers to use in their organism of choice. This kind of topic gets less funding, because it’s foundational science.

So, we develop and distribute the tools to reproducibly culture microorganisms, introduce DNA into them and genetically engineer them. Only then can the organism be used in research and engineering.

Developing these tools takes many years and a lot of money and skills. It takes a lot of people in the room: a biologist, a microbiologist, an automation person, a computational biologist, an engineer. As a non-profit company, we try to make our tools available to all scientists to help them to use their organism of choice for a given application.

We have funding for five years from Schmidt Futures, a non-profit organization in New York City. We’re already releasing and distributing tools and information online. We’re building a portal where all data for non-standard model organisms will be available.

Which appeals to you more — academic research or the private sector?

I like the fast pace of start-up companies. I like the accessibility of expertise: you can bring the engineer into the room with the biologists. I like that you can build a team of people who all work for the same goal with the same motivation and urgency.

Academia is wonderful, and I think it’s very important for people to get rigorous training. But I think we should also showcase other career options for early-career researchers. Before the pandemic, I didn’t know what it was like to work in a non-academic set-up. And once I got a taste of it, I found that it worked well for me.

doi: https://doi.org/10.1038/d41586-024-00928-6

This interview has been edited for length and clarity.

Related Articles

to be useful a scientific hypothesis must

  • Synthetic biology
  • Research management

How scientists are making the most of Reddit

How scientists are making the most of Reddit

Career Feature 01 APR 24

Overcoming low vision to prove my abilities under pressure

Overcoming low vision to prove my abilities under pressure

Career Q&A 28 MAR 24

How a spreadsheet helped me to land my dream job

How a spreadsheet helped me to land my dream job

Career Column 28 MAR 24

More than 4,000 plastic chemicals are hazardous, report finds

More than 4,000 plastic chemicals are hazardous, report finds

News 14 MAR 24

‘This is my calling’: building point-of-care diagnostic tools to fight tuberculosis

‘This is my calling’: building point-of-care diagnostic tools to fight tuberculosis

Career Q&A 01 MAR 24

How co-working labs reduce costs and accelerate progress for biotech start-ups

How co-working labs reduce costs and accelerate progress for biotech start-ups

Career Feature 29 JAN 24

Proteome-scale discovery of protein degradation and stabilization effectors

Proteome-scale discovery of protein degradation and stabilization effectors

Article 20 MAR 24

Mammalian cells repress random DNA that yeast transcribes

Mammalian cells repress random DNA that yeast transcribes

News & Views 06 MAR 24

Synthetic reversed sequences reveal default genomic states

Synthetic reversed sequences reveal default genomic states

Article 06 MAR 24

Supervisory Bioinformatics Specialist, CTG Program Head

National Institutes of Health (NIH) National Library of Medicine (NLM) National Center for Biotechnology Information (NCBI) Information Engineering...

Washington D.C. (US)

National Library of Medicine, National Center for Biotechnology Information

Postdoc Research Associates in Single Cell Multi-Omics Analysis and Molecular Biology

The Cao Lab at UT Dallas is seeking for two highly motivated postdocs in Single Cell Multi-Omics Analysis and Molecular Biology to join us.

Dallas, Texas (US)

the Department of Bioengineering, UT Dallas

to be useful a scientific hypothesis must

Expression of Interest – Marie Skłodowska-Curie Actions – Postdoctoral Fellowships 2024 (MSCA-PF)

Academic institutions in Brittany are looking for excellent postdoctoral researchers willing to apply for a Marie S. Curie Postdoctoral Fellowship.

France (FR)

Plateforme projets européens (2PE) -Bretagne

to be useful a scientific hypothesis must

Tenure-track Assistant Professor in Ecological and Evolutionary Modeling

Tenure-track Assistant Professor in Ecosystem Ecology linked to IceLab’s Center for modeling adaptive mechanisms in living systems under stress

Umeå, Sweden

Umeå University

to be useful a scientific hypothesis must

Faculty Positions in Westlake University

Founded in 2018, Westlake University is a new type of non-profit research-oriented university in Hangzhou, China, supported by public a...

Hangzhou, Zhejiang, China

Westlake University

to be useful a scientific hypothesis must

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

NASA Logo

Scientific Consensus

to be useful a scientific hypothesis must

It’s important to remember that scientists always focus on the evidence, not on opinions. Scientific evidence continues to show that human activities ( primarily the human burning of fossil fuels ) have warmed Earth’s surface and its ocean basins, which in turn have continued to impact Earth’s climate . This is based on over a century of scientific evidence forming the structural backbone of today's civilization.

NASA Global Climate Change presents the state of scientific knowledge about climate change while highlighting the role NASA plays in better understanding our home planet. This effort includes citing multiple peer-reviewed studies from research groups across the world, 1 illustrating the accuracy and consensus of research results (in this case, the scientific consensus on climate change) consistent with NASA’s scientific research portfolio.

With that said, multiple studies published in peer-reviewed scientific journals 1 show that climate-warming trends over the past century are extremely likely due to human activities. In addition, most of the leading scientific organizations worldwide have issued public statements endorsing this position. The following is a partial list of these organizations, along with links to their published statements and a selection of related resources.

American Scientific Societies

Statement on climate change from 18 scientific associations.

"Observations throughout the world make it clear that climate change is occurring, and rigorous scientific research demonstrates that the greenhouse gases emitted by human activities are the primary driver." (2009) 2

American Association for the Advancement of Science

"Based on well-established evidence, about 97% of climate scientists have concluded that human-caused climate change is happening." (2014) 3

AAAS emblem

American Chemical Society

"The Earth’s climate is changing in response to increasing concentrations of greenhouse gases (GHGs) and particulate matter in the atmosphere, largely as the result of human activities." (2016-2019) 4

ACS emblem

American Geophysical Union

"Based on extensive scientific evidence, it is extremely likely that human activities, especially emissions of greenhouse gases, are the dominant cause of the observed warming since the mid-20th century. There is no alterative explanation supported by convincing evidence." (2019) 5

AGU emblem

American Medical Association

"Our AMA ... supports the findings of the Intergovernmental Panel on Climate Change’s fourth assessment report and concurs with the scientific consensus that the Earth is undergoing adverse global climate change and that anthropogenic contributions are significant." (2019) 6

AMA emblem

American Meteorological Society

"Research has found a human influence on the climate of the past several decades ... The IPCC (2013), USGCRP (2017), and USGCRP (2018) indicate that it is extremely likely that human influence has been the dominant cause of the observed warming since the mid-twentieth century." (2019) 7

AMS emblem

American Physical Society

"Earth's changing climate is a critical issue and poses the risk of significant environmental, social and economic disruptions around the globe. While natural sources of climate variability are significant, multiple lines of evidence indicate that human influences have had an increasingly dominant effect on global climate warming observed since the mid-twentieth century." (2015) 8

APS emblem

The Geological Society of America

"The Geological Society of America (GSA) concurs with assessments by the National Academies of Science (2005), the National Research Council (2011), the Intergovernmental Panel on Climate Change (IPCC, 2013) and the U.S. Global Change Research Program (Melillo et al., 2014) that global climate has warmed in response to increasing concentrations of carbon dioxide (CO2) and other greenhouse gases ... Human activities (mainly greenhouse-gas emissions) are the dominant cause of the rapid warming since the middle 1900s (IPCC, 2013)." (2015) 9

GSA emblem

Science Academies

International academies: joint statement.

"Climate change is real. There will always be uncertainty in understanding a system as complex as the world’s climate. However there is now strong evidence that significant global warming is occurring. The evidence comes from direct measurements of rising surface air temperatures and subsurface ocean temperatures and from phenomena such as increases in average global sea levels, retreating glaciers, and changes to many physical and biological systems. It is likely that most of the warming in recent decades can be attributed to human activities (IPCC 2001)." (2005, 11 international science academies) 1 0

U.S. National Academy of Sciences

"Scientists have known for some time, from multiple lines of evidence, that humans are changing Earth’s climate, primarily through greenhouse gas emissions." 1 1

UNSAS emblem

U.S. Government Agencies

U.s. global change research program.

"Earth’s climate is now changing faster than at any point in the history of modern civilization, primarily as a result of human activities." (2018, 13 U.S. government departments and agencies) 12

USGCRP emblem

Intergovernmental Bodies

Intergovernmental panel on climate change.

“It is unequivocal that the increase of CO 2 , methane, and nitrous oxide in the atmosphere over the industrial era is the result of human activities and that human influence is the principal driver of many changes observed across the atmosphere, ocean, cryosphere, and biosphere. “Since systematic scientific assessments began in the 1970s, the influence of human activity on the warming of the climate system has evolved from theory to established fact.” 1 3-17

IPCC emblem

Other Resources

List of worldwide scientific organizations.

The following page lists the nearly 200 worldwide scientific organizations that hold the position that climate change has been caused by human action. http://www.opr.ca.gov/facts/list-of-scientific-organizations.html

U.S. Agencies

The following page contains information on what federal agencies are doing to adapt to climate change. https://www.c2es.org/site/assets/uploads/2012/02/climate-change-adaptation-what-federal-agencies-are-doing.pdf

Technically, a “consensus” is a general agreement of opinion, but the scientific method steers us away from this to an objective framework. In science, facts or observations are explained by a hypothesis (a statement of a possible explanation for some natural phenomenon), which can then be tested and retested until it is refuted (or disproved).

As scientists gather more observations, they will build off one explanation and add details to complete the picture. Eventually, a group of hypotheses might be integrated and generalized into a scientific theory, a scientifically acceptable general principle or body of principles offered to explain phenomena.

1. K. Myers, et al, "Consensus revisited: quantifying scientific agreement on climate change and climate expertise among Earth scientists 10 years later", Environmental Research Letters Vol.16 No. 10, 104030 (20 October 2021); DOI:10.1088/1748-9326/ac2774 M. Lynas, et al, "Greater than 99% consensus on human caused climate change in the peer-reviewed scientific literature", Environmental Research Letters Vol.16 No. 11, 114005 (19 October 2021); DOI:10.1088/1748-9326/ac2966 J. Cook et al., "Consensus on consensus: a synthesis of consensus estimates on human-caused global warming", Environmental Research Letters Vol. 11 No. 4, (13 April 2016); DOI:10.1088/1748-9326/11/4/048002 J. Cook et al., "Quantifying the consensus on anthropogenic global warming in the scientific literature", Environmental Research Letters Vol. 8 No. 2, (15 May 2013); DOI:10.1088/1748-9326/8/2/024024 W. R. L. Anderegg, “Expert Credibility in Climate Change”, Proceedings of the National Academy of Sciences Vol. 107 No. 27, 12107-12109 (21 June 2010); DOI: 10.1073/pnas.1003187107 P. T. Doran & M. K. Zimmerman, "Examining the Scientific Consensus on Climate Change", Eos Transactions American Geophysical Union Vol. 90 Issue 3 (2009), 22; DOI: 10.1029/2009EO030002 N. Oreskes, “Beyond the Ivory Tower: The Scientific Consensus on Climate Change”, Science Vol. 306 no. 5702, p. 1686 (3 December 2004); DOI: 10.1126/science.1103618

2. Statement on climate change from 18 scientific associations (2009)

3. AAAS Board Statement on Climate Change (2014)

4. ACS Public Policy Statement: Climate Change (2016-2019)

5. Society Must Address the Growing Climate Crisis Now (2019)

6. Global Climate Change and Human Health (2019)

7. Climate Change: An Information Statement of the American Meteorological Society (2019)

8. American Physical Society (2021)

9. GSA Position Statement on Climate Change (2015)

10. Joint science academies' statement: Global response to climate change (2005)

11. Climate at the National Academies

12. Fourth National Climate Assessment: Volume II (2018)

13. IPCC Fifth Assessment Report, Summary for Policymakers, SPM 1.1 (2014)

14. IPCC Fifth Assessment Report, Summary for Policymakers, SPM 1 (2014)

15. IPCC Sixth Assessment Report, Working Group 1 (2021)

16. IPCC Sixth Assessment Report, Working Group 2 (2022)

17. IPCC Sixth Assessment Report, Working Group 3 (2022)

Discover More Topics From NASA

Explore Earth Science

to be useful a scientific hypothesis must

Earth Science in Action

Earth Action

Earth Science Data

The sum of Earth's plants, on land and in the ocean, changes slightly from year to year as weather patterns shift.

Facts About Earth

to be useful a scientific hypothesis must

March 19, 2024

Evidence Does Not Support the Use of the Death Penalty

Capital punishment must come to an end. It does not deter crime, is not humane and has no moral or medical basis

By The Editors

A woman protesting, holding a sign showing the Ruth Bader Ginsburg.

A death penalty vigil, held in 2021 outside an Indiana penitentiary.

Bryan Woolston/Reuters/Redux

It is long past time to abolish the death penalty in the U.S.

Capital punishment was halted in the U.S. in 1972 but reinstated in 1976, and since then, nearly 1,600 people have been executed. To whose gain? Study after study shows that the death penalty does not deter crime, puts innocent people to death , is racially biased , and is cruel and inhumane. It is state-sanctioned homicide, wholly ineffective, often botched, and a much more expensive punishment than life imprisonment. There is no ethical, scientifically supported, medically acceptable or morally justifiable way to carry it out.

The recent execution of Kenneth Eugene Smith demonstrates this barbarity. After a failed attempt at lethal injection by prison officials seemingly inexperienced in the placement of an IV, the state of Alabama killed Smith in January using nitrogen gas . The Alabama attorney general claimed that this method of execution was fast and humane , despite no supporting evidence. Eyewitnesses recounted that Smith thrashed during the nitrogen administration and took more than 20 minutes to die.

On supporting science journalism

If you're enjoying this article, consider supporting our award-winning journalism by subscribing . By purchasing a subscription you are helping to ensure the future of impactful stories about the discoveries and ideas shaping our world today.

Opposition to the death penalty is growing among the American public , and the Biden administration must follow through on its promise to end this horror. The Department of Justice must heed its own admission that the death penalty doesn’t stop crime, and our legislators must continue to take up the issue on the congressional floor. The few states that still condemn people to death must follow the lead of states that have considered the evidence and rejected capital punishment.

Programs such as the Innocence Project have shown, over and over, that innocent people have been sentenced to death. Since 1973 nearly 200 people on death row have been exonerated, based on appeals, the reopening of cases, and the entrance of new and sometimes previously suppressed evidence. People have recanted testimony, and supposedly airtight cases have been poked full of evidentiary holes.

Through the death penalty, the criminal justice system has killed at least 20 people now believed to have been innocent and uncounted others whose cases have not been reexamined . Too many of these victims have been Black or Hispanic. This is not justice. These are state-sanctioned hate crimes.

Using rigorous statistical and experimental control methods, both economics and criminal justice studies have consistently found that there is no evidence for deterrence of violent crimes in states that allow capital punishment. One such study, a 2009 paper by criminology researchers at the University of Dallas, outlines experimental and statistical flaws in econometrics-based death penalty studies that claim to find a correlated reduction in violent crime. The death penalty does not stop people from killing. Executions don’t make us safer.

The methods used to kill prisoners are inhumane. Electrocution fails , causing significant pain and suffering. Joel Zivot, an anesthesiologist who criticizes the use of medicines in carrying out the death penalty, has found (at the request of lawyers of death row inmates) that the lungs of prisoners who were killed by lethal injection were often heavy with fluid and froth that suggested they were struggling to breathe and felt like they were drowning. Nitrogen gas is used in some veterinary euthanasia, but based in part on the behavior of rats in its presence, it is “unacceptable” for mammals , according to the American Veterinary Medical Association. This means that Smith, as his lawyers claimed in efforts to stop his execution, became a human subject in an immoral experiment.

Courts have often decided, against the abundant evidence, that these killings are constitutional and do not fall under the “cruel and unusual punishment” clause of the 8th Amendment or, in Smith’s appeal , both the 8th Amendment and the due process protection clause of the 14th amendment.

A small number of prosecutors and judges in a few states, mostly in the South, are responsible for most of the death sentences being handed down in the U.S. today. It’s a power they should not be able to wield. Smith was sentenced to life in prison by a jury before the judge in his case overruled the jury and gave him the death sentence.

A furious urge for vengeance against those who have done wrong—or those we think have done wrong—is the biggest motivation for the death penalty. But this desire for violent retribution is the very impulse that our criminal justice system is made to check, not abet. Elected officials need to reform this aspect of our justice system at both the state and federal levels. Capital punishment does not stop crime and mocks both justice and humanity. The death penalty in the U.S. must come to an end.

This is an opinion and analysis article, and the views expressed by the author or authors are not necessarily those of Scientific American .

IMAGES

  1. How to Write a Strong Hypothesis in 6 Simple Steps

    to be useful a scientific hypothesis must

  2. How to Write a Hypothesis

    to be useful a scientific hypothesis must

  3. Scientific hypothesis

    to be useful a scientific hypothesis must

  4. Research Hypothesis: Definition, Types, Examples and Quick Tips

    to be useful a scientific hypothesis must

  5. 🏷️ Formulation of hypothesis in research. How to Write a Strong

    to be useful a scientific hypothesis must

  6. Research Hypothesis: Definition, Types, Examples and Quick Tips

    to be useful a scientific hypothesis must

VIDEO

  1. Bunnies Useful Scientific Facts

  2. Are we living in a Simulation #science #sciencefacts

  3. In the scientific method, a hypothesis is an a observation b measurement c test d propos

  4. Breakthrough Science Hypothesis

  5. Hypothesis

  6. Research Hypothesis and its Types with examples /urdu/hindi

COMMENTS

  1. Scientific hypothesis

    hypothesis. science. scientific hypothesis, an idea that proposes a tentative explanation about a phenomenon or a narrow set of phenomena observed in the natural world. The two primary features of a scientific hypothesis are falsifiability and testability, which are reflected in an "If…then" statement summarizing the idea and in the ...

  2. Scientific Hypotheses: Writing, Promoting, and Predicting Implications

    A snapshot analysis of citation activity of hypothesis articles may reveal interest of the global scientific community towards their implications across various disciplines and countries. As a prime example, Strachan's hygiene hypothesis, published in 1989,10 is still attracting numerous citations on Scopus, the largest bibliographic database ...

  3. On the scope of scientific hypotheses

    2. The scientific hypothesis. In this section, we will describe a functional and descriptive role regarding how scientists use hypotheses. Jeong & Kwon [] investigated and summarized the different uses the concept of 'hypothesis' had in philosophical and scientific texts.They identified five meanings: assumption, tentative explanation, tentative cause, tentative law, and prediction.

  4. Formulating Hypotheses for Different Study Designs

    A scientific hypothesis should have a sound basis on previously published literature as well as the scientist's observations. Randomly generated (a priori) hypotheses are unlikely to be proven. ... It must be considered that subsequent experiments to prove or disprove a hypothesis have an equal chance of failing or succeeding, akin to tossing a ...

  5. The scientific method (article)

    The scientific method. At the core of biology and other sciences lies a problem-solving approach called the scientific method. The scientific method has five basic steps, plus one feedback step: Make an observation. Ask a question. Form a hypothesis, or testable explanation. Make a prediction based on the hypothesis.

  6. What is a scientific hypothesis?

    A useful hypothesis should be testable and falsifiable. That means that it should be possible to prove it wrong. A theory that can't be proved wrong is nonscientific, according to Karl Popper's ...

  7. How to Write a Strong Hypothesis

    Developing a hypothesis (with example) Step 1. Ask a question. Writing a hypothesis begins with a research question that you want to answer. The question should be focused, specific, and researchable within the constraints of your project. Example: Research question.

  8. What is a Hypothesis?

    A hypothesis (plural: hypotheses) is in its simplest form nothing more than an idea about how the world works. For example, "the moon is made of green cheese" is a valid hypothesis. But there are several characteristics which separate useful scientific hypotheses from those which are impractical. First and foremost, a hypothesis must be ...

  9. Hypothesis

    The hypothesis of Andreas Cellarius, showing the planetary motions in eccentric and epicyclical orbits.. A hypothesis (pl.: hypotheses) is a proposed explanation for a phenomenon.For a hypothesis to be a scientific hypothesis, the scientific method requires that one can test it. Scientists generally base scientific hypotheses on previous observations that cannot satisfactorily be explained ...

  10. 2.4: Developing a Hypothesis

    The primary way that scientific researchers use theories is sometimes called the hypothetico-deductive method ... There are three general characteristics of a good hypothesis. First, a good hypothesis must be testable and falsifiable. We must be able to test the hypothesis using the methods of science and if you'll recall Popper's ...

  11. 1.1: Scientific Investigation

    Forming a Hypothesis. The next step in a scientific investigation is forming a hypothesis.A hypothesis is a possible answer to a scientific question, but it isn't just any answer. A hypothesis must be based on scientific knowledge, and it must be logical. A hypothesis also must be falsifiable. In other words, it must be possible to make observations that would disprove the hypothesis if it ...

  12. Science Shorts 2: The Scientific Method

    Testing a scientific hypothesis proceeds by three basic steps. The first involves formulation of a clear, unambiguous scientific hypothesis. The second step involves the design and prosecution of a study in which the hypothesis under investigation generates at least one prediction. Predictions are simply the study results one expects if the ...

  13. 1.5: Scientific Investigations

    A hypothesis must be based on scientific knowledge. In other words, it shouldn't be at odds with what is already known about the natural world. A hypothesis also must be logical, and it is beneficial if the hypothesis is relatively simple. In addition, to be useful in science, a hypothesis must be testable and falsifiable.

  14. PDF Topic #6: Hypothesis

    The scientific method requires that one can test a scientific hypothesis. Scientists generally base such ... the hypothesis is not yet useful, and must wait for others who might come afterward to make possible the needed observations. For example, a new technology or theory might make the necessary experiments feasible.

  15. 1.2 The Scientific Methods

    The hypothesis must be validated by scientific experiments. The hypothesis must not include any physical quantity. The hypothesis must be a short and concise statement. The hypothesis must apply to all the situations in the universe. 10. What is a scientific theory?

  16. Science and the scientific method: Definitions and examples

    Some key underpinnings to the scientific method: The hypothesis must be testable and falsifiable, according to North Carolina State University. Falsifiable means that there must be a possible ...

  17. On the role of hypotheses in science

    Scientific research progresses by the dialectic dialogue between hypothesis building and the experimental testing of these hypotheses. Microbiologists as biologists in general can rely on an increasing set of sophisticated experimental methods for hypothesis testing such that many scientists maintain that progress in biology essentially comes with new experimental tools.

  18. Scientific method

    A scientific hypothesis must be falsifiable, ... the use of scientific modelling and reliance on abstract typologies and theories is normally accepted. The scientific method counters claims that revelation, political or religious dogma, ... Any useful hypothesis will enable predictions, ...

  19. Hypotheses

    A Scientific Hypothesis Must Be "Falsifiable". A scientific hypothesis must be testable, but there is a much stronger requirement that a testable hypothesis must meet before it can really be considered scientific. This criterion comes primarily from the work of the philosopher of science Karl Popper, and is called "falsifiability".

  20. Physical Science-Chapter 1 Flashcards

    In the 17th century Galileo and the English philosopher Francis Bacon formalized the scientific method which includes the following steps: For a hypothesis to be considered scientific it must be_________--it must, in principle, be capable of being proven_________. Science is restricted to the observable natural world.

  21. BIO 111 Study Part 1 Flashcards

    a) The hypothesis must be proven correct. b) The hypothesis must be reproducible. c) The hypothesis represents established facts. d) The hypothesis must be popularly accepted. e) The hypothesis is testable and falsifiable. e) The hypothesis is testable and falsifiable. A controlled experiment is one in which. a) the experiment is repeated many ...

  22. Scientific Hypothesis

    A hypothesis is a defined set of solutions for an unexplainable occurrence of something in the past that does not fit into the present accepted scientific theory/method. The basic idea of a hypothesis is that there is no predetermined result/outcome. To be scientifically valid a hypothesis must be supported or refuted through carefully crafted ...

  23. Bio 111 P1 Flashcards

    Study with Quizlet and memorize flashcards containing terms like Science is, Superstitions, To be useful in the scientific method, an observation must be and more. Try Magic Notes and save time. Try it free

  24. The beauty of what science can do when urgently needed

    Cultivarium chief scientific officer Nili Ostrov works to make model organisms more useful and accessible for scientific research Credit: Donis Perkins. Nili Ostrov has always been passionate ...

  25. Scientific Consensus

    Scientific Consensus. Temperature data showing rapid warming in the past few decades, the latest data going up through 2023. According to NASA, Earth's average surface temperature in 2023 was the warmest on record since recordkeeping began in 1880, continuing a long-term trend of rising global temperatures. On top of that, the 10 most recent ...

  26. Evidence Does Not Support the Use of the Death Penalty

    This article was originally published with the title " The Death Penalty Must Be Abolished " in Scientific American Magazine Vol. 330 No. 4 (April 2024), p. 66 doi:10.1038 ...