Skip to main content
Log in

Is a bird in the hand worth two in the bush? Or, whether scientists should publish intermediate results

  • Published:
Synthese Aims and scope Submit manuscript

Abstract

A part of the scientific literature consists of intermediate results within a longer project. Scientists often publish a first result in the course of their work, while aware that they should soon achieve a more advanced result from this preliminary result. Should they follow the proverb “a bird in the hand is worth two in the bush”, and publish any intermediate result they get? This is the normative question addressed in this paper. My aim is to clarify, to refine, and to assess informal arguments about the choice whether to publish intermediate results. To this end, I adopt a rational decision framework, supposing some utility or preferences, and I propose a formal model. The best publishing strategy turns out to depend on the research situation. In some simple circumstances, even selfish and short-minded scientists should publish their intermediate results, and should thus behave like their altruistic peers, i. e. like society would like them to behave. In other research situations, with inhomogeneous reward or difficulty profiles, the best strategy is opposite. These results suggest qualified philosophical morals.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

Notes

  1. The historical material of this example is drawn from Bromberg (1991).

  2. The ethical dimension of this question is not tackled in this paper – only the epistemic and decision-theoretic dimensions are.

  3. (Kitcher (1993), p. 303).

  4. (Kitcher (1990), p. 6).

  5. (Strevens (2003), pp. 59–60), summarizing Dasgupta and David (1994). Note that “It was the sociologist Robert Merton who established the priority rule as a characteristic of the social organization of science worthy of study”, according to (Strevens (2003), p. 57). Strevens’s paper is a classical reference on understanding the role of the priority rule in science.

  6. It seems that the size of publications has changed through the decades, and a detailed study of its variations would be interesting. Before the creation of scientific journals in the XVIIth century, the unit of publication was the book. Nowadays, journals publish articles with various lengths. For example in physics, the Physical Review publishes regular articles, and also “Rapid Communications”, “Brief Reports”, and “Comments”.

  7. For a discussion on how citations indicate intellectual debts, cf. for instance (Collins (1974), p. 170).

  8. About historical contingency, see for example (Hacking (1999), chap. 3).

  9. Cf. (Pais (1982), p. 178).

  10. Aspect et al. (1981, 1982a, b).

  11. For instance: as scientists are pressured to have a great number of publications (the implicit “publish-or-perish” rule), several small papers are better than a big one. Conversely, top scientific journals like Science or Nature seem not to accept for publication a result of which an intermediate step has already been published. So, a scientist might not publish an intermediate step if she can expect to publish the whole result in such prestigious journals.

  12. My attempts to reuse and transpose some economic literature about patents to the question of intermediate publication have not been successful.

  13. Important contributions are for instance De Langhe and Greiff (2009), Goldman (2009), Kitcher (1990, 1993), Strevens (2003, 2006), Weisberg and Muldoon (2009) and Zollman (2007); Zollman (2009); Zollman (2010a, b).

  14. There are two ways to justify for it: either one of the scientists is considered to be actually the fastest to write to the journal and to get the whole reward, with probability \(1/2\); or simultaneity is interpreted in a strict way: both scientists are acknowledged as being independent discoverers, and both receive half of the reward—for example, as the amount of money in a Nobel Prize is divided into the laureates.

  15. I don’t want to argue for any particular conception of scientific progress here. My notion of progress is supposed to be compatible with almost any conception discussed in the literature.

  16. It must be noted that the hypothesis according to which scientists are interested in the sum of \(v\) is not particularly favorable to the strategy to publish. It doesn’t make a difference between a scientist who publishes 3 steps at one time, or the 3 steps one by one. For instance, counting the number of publications would favor the last option, and the strategy to publish in general.

  17. A risk-neutral agent has no preference if she is presented with the following choice: either receive \(v\) with certainty, or receive \(2v\) or 0 each with probability \(0.5\). A risk-averse agent prefers the first scenario, a risk-seeking agent prefers the second one.

  18. I thank Conor Mayo-Wilson for indicating me that the theorem is not valid for any strategy.

  19. The proofs can be found in the Appendix, Sects.  and , respectively. They rely on a backward induction proof, which shows some structural similarities between this model and the centipede game, for example.

  20. For Theorem 1, the hypothesis that the scientist is risk-neutral is not required in the proof and can be relaxed.

  21. Theorem 3 is limited to two steps, for computational reasons. With more steps, computations will be much more complicated, without in all likelihood changing the trend of the result.

  22. The proof can be found in the Appendix, Sect. . A technical point to note is that Theorem 3 is only proved on average, in contrast with Theorems 1 and 2, in Sect. 4.2. For some outcomes of \(A\) and \(B\)’s research, publishing the intermediate step might be better. In other words, the validity of Theorem 3 might depend on the hypothesis that scientists are risk-neutral (cf. Sect. 3.2).

  23. The proof can be found in the Appendix, Sect. . Note that Theorem 4 is only proved on average: its validity depends on the hypothesis that scientists are risk-neutral.

  24. Indeed, the hypothesis of the end of the game is nothing more than the insurance that everyone will publish, and this is all the role it plays in the above proof.

  25. Note that Theorem 1 can also be derived from Theorem 2 — with some subtleties. Thanks to Jan Sprenger for this hint.

  26. Note that Theorem 2 can also be derived from Theorem 1.

References

  • Aspect, A. (1976). Proposed experiment to test the nonseparability of quantum mechanics. Physical Review D, 14(8), 1944–1951.

    Article  Google Scholar 

  • Aspect, A., Dalibard, J., & Roger, G. (1982a). Experimental test of Bell’s inequalities using time-varying analysers. Physical Review Letters, 49(25), 1804–1807.

    Google Scholar 

  • Aspect, A., Grangier, P., & Roger, G. (1981). Experimental test of realistic local theories via Bell’s theorem. Physical Review Letters, 47(7), 460–463.

    Article  Google Scholar 

  • Aspect, A., Grangier, P., & Roger, G. (1982b). Experimental realization of Einstein–Podolsky–Rosen–Bohm Gedanken experiment: A new violation of Bell’s inequalities. Physical Review Letters, 49(2), 91–94.

    Google Scholar 

  • Bromberg, J. L. (1991). The laser in America 1950–1970. Cambridge, MA: MIT Press.

    Google Scholar 

  • Collins, H. M. (1974). The TEA set: Tacit knowledge and scientific networks. Science Studies, 4(2), 165–185.

    Article  Google Scholar 

  • Dasgupta, P., & David, P. (1994). Toward a new economics of science. Research Policy, XXIII, 487–521.

    Google Scholar 

  • De Langhe, R., & Greiff, M. (2009). Standards and the distribution of cognitive labour: A model of the dynamics of scientific activity. Logic Journal of the IGPL, 18(2), 278–294.

    Article  Google Scholar 

  • Goldman, A. (2009). Systems-oriented social epistemology. In T. Gendler & J. Hawthorne (Eds.), Oxford studies in epistemology (Vol. 3, pp. 189–214). Oxford: Oxford University Press.

    Google Scholar 

  • Kitcher, P. (1990). The division of cognitive labor. Journal of Philosophy, 87(1), 5–22.

    Article  Google Scholar 

  • Kitcher, P. (1993). The advancement of science: science without legend, objectivity without illusions. Oxford: Oxford University Press.

    Google Scholar 

  • Hacking, I. (1999). The social construction of what?. London: Harvard University Press.

    Google Scholar 

  • Pais, A. (1982). Subtle is the lord: The science and the life of Albert Einstein. Oxford: Oxford University Press.

    Google Scholar 

  • Strevens, M. (2003). The role of the priority rule in science. Journal of Philosophy, 100, 55–79.

    Google Scholar 

  • Strevens, M. (2006). The role of the Matthew effect in science. Studies in History and Philosophy of Science, 37, 159–173.

    Article  Google Scholar 

  • Weisberg, M., & Muldoon, R. (2009). Epistemic landscapes and the division of cognitive labor. Philosophy of Science, 76, 225–252.

    Article  Google Scholar 

  • Zollman, K. J. S. (2007). The communication structure of epistemic communities. Philosophy of Science, 74(5), 574–587.

    Article  Google Scholar 

  • Zollman, K. J. S. (2009). Optimal publishing strategies. Episteme: A Journal of Social Epistemology, 6(2), 185–199.

    Article  Google Scholar 

  • Zollman, K. J. S. (2010a). Social structure and the effects of conformity. Synthese, 172(3), 317–340.

    Article  Google Scholar 

  • Zollman, K. J. S. (2010b). The epistemic benefit of transient diversity. Erkenntnis, 72(1), 17–35.

    Article  Google Scholar 

Download references

Acknowledgments

Many thanks to Anouk Barberousse, Conor Mayo-Wilson, Jan Sprenger, Mikael Cozic, Philip Kitcher, Michael Strevens and an anonymous referee for their critical remarks and encouragements, which helped much improve this paper. Thanks also to participants in the EPSA conference in October 2011 in Athens (Greece), in the “Collective Dimension of Science” conference in December 2011 in Nancy (France), and in the “Progress of Science” conference in April 2012 in Tilburg (The Netherlands), for their remarks. The paper has been written while at the IHPST (UMR 8590, CNRS, Université de Paris 1 Panthéon-Sorbonne, ENS), at Columbia University (NY, USA), at the Archives Henri Poincaré (UMR 7117 CNRS, Université de Lorraine) and at Savoirs, Textes, Langage (UMR 8163 CNRS, Université Lille 1, Université Lille 3). A special financial support is acknowledged from the Université de Paris 1.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Thomas Boyer.

Appendix

Appendix

1.1 Proof of Theorem 1 in Sect. 4.2

Theorem 1 invites to consider two general strategies:

  • strategy I: always immediately publish a passed step,

  • strategy II: never publish anything before the end of the game.

The theorem is proved by induction on the number of remaining temporal intervals, for scientist \(A\). To start with, suppose there is just one temporal interval left. There are two cases: either \(A\) is ahead of \(B\), or the opposite (a similar position for \(A\) and \(B\) can be considered as a special case of the first or of the second case). Start with the first case, and define \(a\) and \(b\) such that: \(B\) is \(b\) steps ahead of the last published step (which might be step 0), and \(A\) is \(a\) steps ahead of \(B\).

Firstly, suppose \(B\) adopts strategy I. Let’s compare \(A\)’s reward according to the strategy she adopts. Suppose she adopts strategy I: she publishes \(b+ a\) steps, while \(B\) publishes \(b\) steps, so \(A\) immediately gets the reward \((a+b/2)v\). Then, \(A\) and \(B\) start on the same level for the last temporal interval. \(A\)’s final reward is given by Table 1, according to the research outcome of the final temporal interval.

Table 1 \(A\)’s reward during the last temporal interval, with strategy I

Suppose now \(A\) adopts strategy II: she doesn’t publish now (while \(B\) publishes her \(b\) steps) and waits for the last time. \(A\)’s final reward is given by Table 2. For each cell, Table 1 is larger than Table 2. This shows that, when \(A\) is in front of \(B\) and \(B\) adopts strategy I, then strategy I is better than strategy II for \(A\).

Table 2 \(A\)’s reward during the last temporal interval, with strategy II

Secondly, suppose \(B\) adopts strategy II. Suppose \(A\) adopts strategy I. She first publishes \(b+a\) steps, while \(B\) publishes nothing. So, her reward after the last temporal interval is given by \(bv/2\) plus the cells of Table 1. Suppose now \(A\) adopts strategy II. Then her reward after the last temporal interval is given by \(bv/2\) plus the cells of Table 2. So, here again, strategy I is better than strategy II for \(A\).

Consider now the second case: \(B\) is ahead of \(A\). Define \(a\) and \(b\) such that \(A\) is \(a\) steps ahead of the last published step, and \(B\) is \(b\) steps ahead of \(A\). First, suppose that \(B\) adopts strategy I. If \(A\) adopts strategy I too, her final reward is given by Table 3. If she adopts strategy II, it is given by Table 3 minus \(av/2\). So, strategy I is better for \(A\).

Table 3 Expected rewards for \(A\) and \(B\)

Suppose now that \(B\) adopts strategy II. If \(A\) adopts strategy I, her final reward is given by Table 4. If she adopts strategy II, it is given by Table 3 minus \(av/2\). So, strategy I is better for \(A\) in this case too. So in any case, strategy I is better than strategy II for \(A\), and this establishes the first case of the induction proof.

Table 4 \(A\)’s reward, according to the outcome of the first temporal interval

Suppose now it has been proved for some \(m<n\) that, when \(m\) temporal intervals remain, \(A\) should publish every step she has passed. Consider \(A\) when there remains \(m+1\) temporal intervals. By hypothesis, \(A\) should publish at the next temporal interval (when there remains only \(m\) temporal intervals). If \(B\) adopts strategy I, then she will publish at the next temporal interval, so \(A\) is in the case already provedFootnote 24, where \(m=1\). If \(B\) adopts strategy II, \(B\) doesn’t publish before the end, so nothing is changed whether \(A\) publishes right now or just at the next temporal interval. Hence, strategy I is still better than strategy II for \(A\), and Theorem 1 has been proved by induction.Footnote 25

Note that the conclusion that \(A\) should always publish has not only been shown on average, but also for every possible outcome of research — for every comparison of the corresponding cells in the various tables. So, the theorem doesn’t depend on the hypothesis that the scientist is risk-neutral.

1.2 Proof of Theorem 2 in Sect. 4.2

It is useful to distinguish between two kinds of steps: the steps both \(A\) and \(B\) have passed but not yet published (call them the “common steps”), and the steps \(A\) is the only one to have passed, if any (her “solitary steps”). I am going to show that solitary steps should be published. Then, taking into account the possibility of common steps cannot make publication a worse strategy than it is if there are only solitary steps (because nothing can be gained in not publishing them, in particular because the competitor’s strategy doesn’t depend on what is actually published, and a potential reward can be lost). For a result aiming at proving the superiority of publication, this is just fine. Strategies I and II are defined in the same way as in the proof of Theorem 1. Again, proving Theorem 2 amounts to showing that a scientist should prefer strategy I to strategy II.

As the chain is composed of \(l\) steps, a total reward \(lv\) will finally be distributed to the scientists. Because of their symmetric role, both scientists can expect a reward \(lv/2\) if they adopt the same strategy. Table 5 shows the expected rewards for scientists I will argue for. Going from this Table 5 to Theorem 2 — \(A\) should adopt strategy I — will not be difficult. Whatever strategy \(B\) adopts, \(A\) can expect a better reward if she adopts strategy I. And the same can be said for \(B\). So, both should choose it, and it is a Nash equilibrium.

Table 5 Expected rewards for \(A\) and \(B\) (respectively), according to the strategies they adopt

Table 5 could clearly be derived from the following proposition \(P(k)\), if it was true for any \(k\):

Proposition \(P(k)\): If \(B\) has adopted strategy I, if \(A\) has some solitary steps, and if there remain \(k\) steps before the end of the chain (i. e. \(A\) is at the position \(l-k\) ), then \(A\) ’s expected reward is greater if she publishes her solitary steps.

So, let’s prove \(P(k)\) by induction. In the initial case, \(k=1\), and \(A\) is at step \(l-1\). She has some (say \(a\)) solitary steps. Suppose \(A\) publishes her \(a\) steps now. She immediately gets an \(av\) reward, and both scientists are at step \(l-1\). Given the symmetry between \(A\) and \(B, A\)’s expected reward for the last step is \(v/2\). Summing, her expected reward in this case is

$$\begin{aligned} E = av + \frac{v}{2}. \end{aligned}$$
(1)

Suppose now \(A\) doesn’t publish her \(a\) steps now. I am going to build a table (Table 6) for \(A\)’s expected reward in the long run (called \(E_{6}\)), according to the results after this next temporal interval. Note that it is \(A\)’s expected reward from this specific situation until the end of the game, and not only for the next temporal interval. Here is how table 6 is filled in. If \(A\) succeeds, the game is over, and the rewards can be easily computed. If \(A\) doesn’t succeed, the game is not over, so computing a reward isn’t straightforward. If both \(A\) and \(B\) fail, then they still are at the same position. So, \(A\)’s expected reward in this case is \(E_{6}\) itself. In case \(A\) fails and \(B\) succeeds, \(A\) is in position \(l-1\), and \(B\) in \(l-a\), one step higher. Let’s call \(E_{6}^{\prime }\) the expected reward in this situation. In any case \(A\) cannot expect to publish more than \(a\) steps, and perhaps she will have to share some of these rewards, so \(E^{\prime }_{6}< av\) on average. This gives Table 6. To get the expected reward \(E_{6} \), this table is weighted with the probability of the cells outcome. Some easy computation shows that \(E_{6} < E\). In words: if \(A\) has some solitary steps and if she is at the position \(l-1\), her expected reward is larger if she publishes them now, than if she doesn’t. This proves \(P(1)\).

Table 6 Expected reward \(E_{6}\) for \(A\) after one temporal interval, with \(A\) starting \(a\) positions ahead of \(B\), who will publish after this temporal interval

Suppose now that \(P(k)\) has been proved for some \(k\). In the case \(k+1, A\) is at step \(l-k-1\). By \(P(k)\), she knows that she will publish her solitary steps, if she has some, when she is on \(l-k\), i. e. on the next step. Both \(A\) and \(B\) would publish if they reached step \(l-k\) (\(B\) would because she always publishes): we are tempted to say that this step plays the role of the end of a shorter chain, that we are in a situation where \(P(1)\) applies, in order to argue that \(A\) should publish now at \(l-k\). This reasoning implicitly assumes that \(A\) should maximize her reward when someone arrives at \(l-k\) (if it is to play the role of the end of chain). Actually, the real goal for \(A\) is to maximize her reward when either her or \(B\) arrives at step \(l\), the actual end of the chain. Hopefully, the only way to maximize the reward at \(l\) is to maximize it at \(l-k\) first: someone will publish at \(l-k\), the two scientists will be together at this step, so there is no way for \(A\) to keep any sort of better non-published lead for the real end of the chain. So, the argument of \(P(1)\) can be used, and \(P(k+1)\) is proved. As was argued before, this proves Table 5, and Theorem 2. Footnote 26

1.3 Proof of Theorem 3 in Sect. 4.3.1

As before, it is clear that \(A\) should publish a common step, and also that she cannot know in practice which steps are solitary or common. Strategies I and II are defined in the same way as in the proof of Theorem 1.

Theorem 3 is equivalent to stating that strategy II gives a larger average reward than strategy I. As the total reward distributed among scientists is constant at \(v_1 + v_2\), it is equivalent to stating Table 7.

Table 7 \(A\)’s expected reward, in case \(A\) is at step 1 and \(B\) at step 0

Let’s consider the case of the bottom left cell. Let’s call \(E_{4}\) the average reward that \(A\) can expect after the first temporal interval. Table 8 is filled in: if \(A\) and \(B\) succeed, only \(B\) publishes because \(A\) follows strategy II; then, they are in a symmetrical position for the last step: \(A\)’s expected reward is \(v_2/2\). If \(A\) fails and \(B\) succeeds, only \(B\) publishes too, and \(A\)’s expected reward is \(v_2/2\). If both scientists fail, they remain in the same position, and the reward is still \(E_{4}\). If \(A\) succeeds and \(B\) fails, \(A\)’s reward is called \(E_{6}\) and is computed below.

\(E_{4}\) is computed by weighting Table 8 with Table 9:

$$\begin{aligned} E_{4} = p^2 \frac{v_2}{2} + p(1-p)\frac{v_2}{2} +p(1-p) E_{6} + (1-p)^2 E_{4} \end{aligned}$$
(2)

The evaluation of \(E_{6}\) is made by filling in a similar Table 10.

Table 8 Expected rewards for \(A\) and \(B\), according to their strategies
Table 9 \(A\)’s expected reward, according to the outcome of the first temporal interval
Table 10 Expected reward \(E_{9}\) for \(A\), in case \(A\) is at step 1 and \(B\) at step 0

Weighting Table 10 with Table 9, one computes

$$\begin{aligned} E_{6} = p^2 \Big ( \frac{v_1}{2} + v_2 \Big ) + p(1-p) \Big (v_1 + v_2 + \frac{v_2}{2} \Big ) + (1-p)^2 E_{6}. \end{aligned}$$
(3)

\(E_{6}\) is extracted from it and injected in Eq. 2:

$$\begin{aligned} E_{4} = \frac{v_2}{2(2-p)} + \frac{1-p}{2(2-p)^2}\big [ (2-p) v_1 + (3-p) v_2 \big ] \end{aligned}$$
(4)

The condition expressed in the bottom left cell of Table 7 is that \(E_{4} > \frac{v_1+v_2}{2}\). With Eq. 4, it is equivalent to

$$\begin{aligned} v_2 > \frac{2-p}{1-p} v_1. \end{aligned}$$
(5)

1.4 Proof of Theorem 4 in Sect. 4.3.2

The proof of Theorem 4 is very similar to the proof in Sect. , and I only give the key differences here. The total distributed reward is \(2v\), so the expected rewards should be Tables 8 and 9.

\(E_{8}\) is computed by weighting Table 9 with Table 6:

$$\begin{aligned} E_{8} = p_1 \frac{v}{2} + p_1(1-p_1)E_{9} + (1-p_1)^2 E_{8} \end{aligned}$$
(6)

\(E_{9}\) is computed from Table 10, weighted with Table 11:

$$\begin{aligned} E_{9} = \frac{ p_1/2 + 2 p_2 - p_1 p_2}{p_1 + p_2 - p_1p_2} v. \end{aligned}$$
(7)
Table 11 Probability of \(A\) and \(B\)’s outcomes of research, when \(A\) is at step 1 and \(B\) at step \(0\)

From Eqs. 7 and 6, one gets

$$\begin{aligned} E_{8} = \frac{v}{2(2-p_1)} + \frac{1-p_1}{2-p_1} \times \frac{p_1/2+2p_2-p_1p_2}{p_1 + p_2 - p_1 p_2}v \end{aligned}$$
(8)

The condition expressed in the bottom left cell of Table 8 (\(E_{8}>v\)) leads to Theorem 4.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Boyer, T. Is a bird in the hand worth two in the bush? Or, whether scientists should publish intermediate results. Synthese 191, 17–35 (2014). https://doi.org/10.1007/s11229-012-0242-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11229-012-0242-4

Keywords

Navigation