Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-5nwft Total loading time: 0 Render date: 2024-05-03T20:48:37.196Z Has data issue: false hasContentIssue false

Part I - What Drives Humans to Seek Information?

Published online by Cambridge University Press:  19 May 2022

Irene Cogliati Dezza
Affiliation:
University College London
Eric Schulz
Affiliation:
Max-Planck-Institut für biologische Kybernetik, Tübingen
Charley M. Wu
Affiliation:
Eberhard-Karls-Universität Tübingen, Germany
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
The Drive for Knowledge
The Science of Human Information Seeking
, pp. 1 - 98
Publisher: Cambridge University Press
Print publication year: 2022

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

References

Anderson, C., Brion, S., Moore, D. A., & Kennedy, J. A. (2012). A status enhancement account of overconfidence. Journal of Personality and Social Psychology, 103(4), 718.Google Scholar
Aragones, E., Gilboa, I., Postlewaite, A., & Schmeidler, D. (2005). Fact-free learning. American Economic Review, 95(5), 13551368.Google Scholar
Ariely, D., Gneezy, U., Loewenstein, G., & Mazar, N. (2009). Large stakes and big mistakes. The Review of Economic Studies, 76(2), 451469.CrossRefGoogle Scholar
Bakshy, E., Messing, S., & Adamic, L. A. (2015). Exposure to ideologically diverse news and opinion on Facebook. Science, 348(6239), 11301132.Google Scholar
Bamman, D., O’Connor, B., & Smith, N. (2012). Censorship and deletion practices in Chinese social media. First Monday, 17(3). https://doi.org/10.5210/fm.v17i3.3943.Google Scholar
Batson, C. D. (1975). Rational processing or rationalization? The effect of disconfirming information on a stated religious belief. Journal of Personality and Social Psychology, 32(1), 176.Google Scholar
Beatty, M. J. (1988). Situational and predispositional correlates of public speaking anxiety. Communication Education, 37(1), 2839.CrossRefGoogle Scholar
Bechara, A., & Damasio, A. R. (2005). The somatic marker hypothesis: A neural theory of economic decision. Games and Economic Behavior, 52(2), 336372.Google Scholar
Bellman, R. E. (1957). Dynamic programming. Princeton University Press.Google ScholarPubMed
Blackwell, D. (1953). Equivalent comparisons of experiments. The annals of mathematical statistics, 24(2), 265272. www.jstor.org/stable/2236332.Google Scholar
Bramley, N. R., Dayan, P., Griffiths, T. L., & Lagnado, D. A. (2017). Formalizing Neurath’s ship: Approximate algorithms for online causal learning. Psychological Review, 124(3), 301.CrossRefGoogle ScholarPubMed
Bruder, M., Haffke, P., Neave, N., Nouripanah, N., & Imhoff, R. (2013). Measuring individual differences in generic beliefs in conspiracy theories across cultures: Conspiracy mentality questionnaire. Frontiers in Psychology, 4, 225.Google Scholar
Burda, Y., Edwards, H., Pathak, D., Storkey, A., Darrell, T., & Efros, A. A. (2018). Large-scale study of curiosity-driven learning. arXiv preprint arXiv:1808.04355.Google Scholar
Cabanac, M. (1971). Physiological role of pleasure. Science, 173(4002), 11031107.CrossRefGoogle ScholarPubMed
Cat, J. (2021). Otto Neurath. In Zalta, E. N. (Ed.), The Stanford encyclopedia of philosophy (Spring 2021 ed.). Metaphysics Research Lab, Stanford University. https://plato.stanford.edu/archives/spr2021/entries/neurath/.Google Scholar
Chater, N. (1996). Reconciling simplicity and likelihood principles in perceptual organization. Psychological Review, 103(3), 566.Google Scholar
Chater, N. (2019). The mind is flat. Yale University Press.Google Scholar
Chater, N., & Loewenstein, G. (2016). The under-appreciated drive for sensemaking. Journal of Economic Behavior & Organization, 126, 137154.CrossRefGoogle Scholar
Chater, N., & Vitányi, P. (2003). Simplicity: A unifying principle in cognitive science? Trends in Cognitive Sciences, 7(1), 1922.CrossRefGoogle ScholarPubMed
Chater, N., Zhu, J., Spicer, J., Sundh, J., León-Villagrá, P., & Sanborn, A. (2020). Probabilistic biases meet the Bayesian brain. Current Directions in Psychological Science, 29(5), 506512.CrossRefGoogle Scholar
Chen, S., & Heese, C. (2021). “Fishing for Good News: Motivated Information Acquisition.” CRC TR 224 Discussion Paper Series crctr224_2021_223v3, University of Bonn and University of Mannheim, Germany.Google Scholar
Cheng, X., Dale, C., & Liu, J. (2008). Statistics and social network of YouTube videos. In 2008 16th International Workshop on Quality of Service (pp. 229238). IEEE. https://ieeexplore.ieee.org/document/4539688.Google Scholar
Christensen, D. (1994). Conservatism in epistemology. Noûs, 28(1), 6989.Google Scholar
Colleoni, E., Rozza, A., & Arvidsson, A. (2014). Echo chamber or public sphere? Predicting political orientation and measuring political homophily in Twitter using big data. Journal of Communication, 64(2), 317332.CrossRefGoogle Scholar
Colman, A. M. (2003). Cooperation, psychological game theory, and limitations of rationality in social interaction. Behavioral and Brain Sciences, 26(2), 139153.CrossRefGoogle ScholarPubMed
Cook, J., & Lewandowsky, S. (2016). Rational irrationality: Modeling climate change belief polarization using Bayesian networks. Topics in Cognitive Science, 8(1), 160179.Google Scholar
Cosmides, L., & Tooby, J. (2000). Evolutionary psychology and the emotions. In Lewis, M. & Haviland-Jones, J. M. (eds.), Handbook of Emotions (pp. 91115). Guilford.Google Scholar
Coyle, K. (2006). Mass digitization of books. The Journal of Academic Librarianship, 32(6), 641645.Google Scholar
Cushman, F. (2020). Rationalization is rational. Behavioral and Brain Sciences, 43, E28.Google Scholar
Damasio, A. R. (2006). Descartes’ error. Random House.Google Scholar
Darwin, H., Neave, N., & Holmes, J. (2011). Belief in conspiracy theories. The role of paranormal belief, paranoid ideation and schizotypy. Personality and Individual Differences, 50(8), 12891293.Google Scholar
Dasgupta, I., Schulz, E., & Gershman, S. J. (2017). Where do hypotheses come from? Cognitive Psychology, 96, 125.Google Scholar
Dayan, P. (1998). A hierarchical model of binocular rivalry. Neural Computation, 10(5), 11191135.CrossRefGoogle ScholarPubMed
Deci, E. L., & Ryan, R. M. (1981). Curiosity and self-directed learning: The role of motivation in education. In Katz, L. (Ed.), Current topics in early childhood education (Vol. 4). Ablex Publishing Co.Google Scholar
DSM-5. (2013). Diagnostic and statistical manual of mental disorders. American Psychiatric Association.Google Scholar
Dubois, E., & Blank, G. (2018). The echo chamber is overstated: The moderating effect of political interest and diverse media. Information, Communication & Society, 21(5), 729745.Google Scholar
Ely, J., Frankel, A., & Kamenica, E. (2015). Suspense and surprise. Journal of Political Economy, 123(1), 215260.Google Scholar
Enke, B., & Zimmermann, F. (2019). Correlation neglect in belief formation. The Review of Economic Studies, 86(1), 313332.Google Scholar
Epley, N., & Gilovich, T. (2016). The mechanics of motivated reasoning. Journal of Economic perspectives, 30(3), 133140.Google Scholar
Eyster, E., & Rabin, M. (2014). Extensive imitation is irrational and harmful. The Quarterly Journal of Economics, 129(4), 18611898.Google Scholar
Festinger, L. (1957). A theory of cognitive dissonance (Vol. 2). Stanford University Press.Google Scholar
Flaxman, S., Goel, S., & Rao, J. M. (2016). Filter bubbles, echo chambers, and online news consumption. Public Opinion Quarterly, 80(S1), 298320.Google Scholar
Friston, K. (2010). The free-energy principle: A unified brain theory? Nature Reviews Neuroscience, 11(2), 127138.Google Scholar
Friston, K., & Kiebel, S. (2009). Predictive coding under the free-energy principle. Philosophical Transactions of the Royal Society B: Biological Sciences, 364(1521), 12111221.Google Scholar
Friston, K., Thornton, C., & Clark, A. (2012). Free-energy minimization and the dark-room problem. Frontiers in Psychology, 3, 130.Google Scholar
Gershman, S., Vul, E., & Tenenbaum, J. B. (2009). Perceptual multistability as Markov chain Monte Carlo inference. In Advances in neural information processing systems 22 (pp. 611619). https://proceedings.neurips.cc/paper/2009/hash/692f93be8c7a41525c0baf2076aecfb4-Abstract.html.Google Scholar
Gilovich, T. (2008). How we know what isn’t so. Simon and Schuster.Google Scholar
Goldfarb, A. (2014). What is different about online advertising? Review of Industrial Organization, 44(2), 115129.Google Scholar
Golman, R., & Loewenstein, G. (2018). Information gaps: A theory of preferences regarding the presence and absence of information. Decision, 5(3), 143.Google Scholar
Gopnik, A. (1998). Explanation as orgasm. Minds and Machines, 8(1), 101118.CrossRefGoogle Scholar
Gottlieb, J., Oudeyer, P.-Y., Lopes, M., & Baranes, A. (2013). Information seeking, curiosity, and attention: Computational and neural mechanisms. Trends in Cognitive Sciences, 17(11), 585593.Google Scholar
Griffiths, T. L., & Tenenbaum, J. B. (2003). Probability, algorithmic complexity, and subjective randomness. In Proceedings of the annual meeting of the cognitive science society (Vol. 25). https://escholarship.org/uc/item/6ts3j7bw.Google Scholar
Griffiths, T. L., Vul, E., & Sanborn, A. N. (2012). Bridging levels of analysis for probabilistic models of cognition. Current Directions in Psychological Science, 21(4), 263268.Google Scholar
Hanoch, Y. (2002). “Neither an angel nor an ant”: Emotion as an aid to bounded rationality. Journal of Economic Psychology, 23(1), 125.Google Scholar
Harman, G. (2003). Skepticism and foundations. In Luper, S. (ed.), The skeptics: Contemporary essays (pp. 111). Routledge.Google Scholar
Heath, C., & Heath, D. (2007). Made to stick: Why some ideas survive and others die. Random House.Google Scholar
Jacob, F. (1977). Evolution and tinkering. Science, 196(4295), 11611166.Google Scholar
Jeffrey, R. C. (2004). Subjective probability: The real thing. Cambridge University Press.Google Scholar
Jepma, M., Verdonschot, R. G., Van Steenbergen, H., Rombouts, S. A., & Nieuwenhuis, S. (2012). Neural mechanisms underlying the induction and relief of perceptual curiosity. Frontiers in Behavioral Neuroscience, 6, 5. www.frontiersin.org/article/10.3389/fnbeh.2012.00005.Google Scholar
Jern, A., Chang, K.-M. K., & Kemp, C. (2014). Belief polarization is not always irrational. Psychological Review, 121(2), 206.Google Scholar
John, L. K., Loewenstein, G., & Prelec, D. (2012). Measuring the prevalence of questionable research practices with incentives for truth telling. Psychological Science, 23(5), 524532.Google Scholar
Kahneman, D., & Lovallo, D. (1993). Timid choices and bold forecasts: A cognitive perspective on risk taking. Management Science, 39(1), 1731.Google Scholar
Kang, M. J., Hsu, M., Krajbich, I. M., Loewenstein, G., McClure, S. M., Wang, J. T.-y., & Camerer, C. F. (2009). The wick in the candle of learning: Epistemic curiosity activates reward circuitry and enhances memory. Psychological Science, 20(8), 963973.CrossRefGoogle ScholarPubMed
Kehlmann, D. (2009). Measuring the world: A novel. Vintage.Google Scholar
Kidd, C., & Hayden, B. Y. (2015). The psychology and neuroscience of curiosity. Neuron, 88(3), 449460.Google Scholar
Kleine, D. (1990). Anxiety and sport performance: A meta-analysis. Anxiety Research, 2(2), 113131.Google Scholar
Kramer, A. D., Guillory, J. E., & Hancock, J. T. (2014). Experimental evidence of massive-scale emotional contagion through social networks. Proceedings of the National Academy of Sciences, 111(24), 87888790.Google Scholar
Kunda, Z. (1990). The case for motivated reasoning. Psychological Bulletin, 108(3), 480.Google Scholar
Kurzban, R., Duckworth, A., Kable, J. W., & Myers, J. (2013). An opportunity cost model of subjective effort and task performance. Behavioral and Brain Sciences, 36(6), 661679.Google Scholar
Kwisthout, J. (2011). Most probable explanations in Bayesian networks: Complexity and tractability. International Journal of Approximate Reasoning, 52(9), 14521469.Google Scholar
Levy, R. P., Reali, F., & Griffiths, T. L. (2009). Modeling the effects of memory on human online sentence processing with particle filters. In Advances in neural information processing systems (pp. 937944). https://cocosci.princeton.edu/tom/papers/sentencepf1.pdf.Google Scholar
Liberman, A., & Chaiken, S. (1992). Defensive processing of personally relevant health messages. Personality and Social Psychology Bulletin, 18(6), 669679.Google Scholar
Litman, J. (2005). Curiosity and the pleasures of learning: Wanting and liking new information. Cognition & Emotion, 19(6), 793814.Google Scholar
Loewenstein, G. (1994). The psychology of curiosity: A review and reinterpretation. Psychological Bulletin, 116(1), 75.Google Scholar
Loewenstein, G., & Molnar, A. (2018). The renaissance of belief-based utility in economics. Nature Human Behaviour, 2(3), 166167.Google Scholar
Lopes, M., Lang, T., Toussaint, M., & Oudeyer, P.-Y. (2012). Exploration in model-based reinforcement learning by empirically estimating learning progress. In Pereira, F., Burges, C. J. C., Bottou, L. & Weinberger, K. Q. (eds.), Advances in neural information processing systems (pp. 206214). Curran Associates, Inc.Google Scholar
Lord, C. G., Ross, L., & Lepper, M. R. (1979). Biased assimilation and attitude polarization: The effects of prior theories on subsequently considered evidence. Journal of Personality and Social Psychology, 37(11), 2098.Google Scholar
MacLeod, W. B. (1996). Decision, contract, and emotion: Some economics for a complex and confusing world. Canadian Journal of Economics, 29(4), 788810.Google Scholar
Markey, A., & Loewenstein, G. (2014). Curiosity. In Linnenbrink-Garcia, L. (ed.) International handbook of emotions in education (pp. 228245). Routledge.Google Scholar
Marshall, G. (2015). Don’t even think about it: Why our brains are wired to ignore climate change. Bloomsbury Publishing USA.Google Scholar
Matz, S. C., Kosinski, M., Nave, G., & Stillwell, D. J. (2017). Psychological targeting as an effective approach to digital mass persuasion. Proceedings of the National Academy of Sciences, 114(48), 1271412719.Google Scholar
McCabe, M. P. (2005). The role of performance anxiety in the development and maintenance of sexual dysfunction in men and women. International Journal of Stress Management, 12(4), 379.Google Scholar
McLean, B. F., Mattiske, J. K., & Balzan, R. P. (2017). Association of the jumping to conclusions and evidence integration biases with delusions in psychosis: A detailed meta-analysis. Schizophrenia Bulletin, 43(2), 344354.Google Scholar
McPherson, M., Smith-Lovin, L., & Cook, J. M. (2001). Birds of a feather: Homophily in social networks. Annual Review of Sociology, 27(1), 415444.Google Scholar
Meder, B., Nelson, J. D., Jones, M., & Ruggeri, A. (2019). Stepwise versus globally optimal search in children and adults. Cognition, 191, 103965.Google Scholar
Mercier, H., & Sperber, D. (2011). Why do humans reason? Arguments for an argumentative theory. Behavioral and Brain Sciences, 34(2), 5774.Google Scholar
Moreno-Bote, R., Knill, D. C., & Pouget, A. (2011). Bayesian sampling in visual perception. Proceedings of the National Academy of Sciences, 108(30), 1249112496.Google Scholar
Muramatsu, R., & Hanoch, Y. (2005). Emotions as a mechanism for boundedly rational agents: The fast and frugal way. Journal of Economic Psychology, 26(2), 201221.Google Scholar
Pashler, H. (1999). The psychology of attention. MIT Press.Google Scholar
Pathak, D., Agrawal, P., Efros, A. A., & Darrell, T. (2017). Curiosity-driven exploration by self-supervised prediction. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition Workshops (pp. 16– 17). http://proceedings.mlr.press/v70/pathak17a/pathak17a.pdf.Google Scholar
Pluck, G., & Johnson, H. (2011). Stimulating curiosity to enhance learning. GESJ: Education Sciences and Psychology, 2(19). ISSN 1512-1801.Google Scholar
Pronin, E., Lin, D. Y., & Ross, L. (2002). The bias blind spot: Perceptions of bias in self versus others. Personality and Social Psychology Bulletin, 28(3), 369381.Google Scholar
Rabin, M., & Schrag, J. L. (1999). First impressions matter: A model of confirmatory bias. The Quarterly Journal of Economics, 114(1), 3782.Google Scholar
Rapoport, A., & Chammah, A. M. (1966). The game of chicken. American Behavioral Scientist, 10(3), 1028.Google Scholar
Ruan, B., Hsee, C. K., & Lu, Z. Y. (2018). The teasing effect: An underappreciated benefit of creating and resolving an uncertainty. Journal of Marketing Research, 55(4), 556570.Google Scholar
Samuelson, L., & Swinkels, J. M. (2006). Information, evolution and utility. Theoretical Economics, 1(1), 119142.Google Scholar
Sanborn, A. N., & Chater, N. (2016). Bayesian brains without probabilities. Trends in Cognitive Sciences, 20(12), 883893.Google Scholar
Savage, L. J. (1972). The foundations of statistics. Courier Corporation.Google Scholar
Schelling, T. C. (1980). The strategy of conflict: With a new preface by the author. Harvard University Press.Google Scholar
Schmidhuber, J. (1991). A possibility for implementing curiosity and boredom in model-building neural controllers. In Meyer, J. A. and Wilson, S. W. (eds.), Proc. of the international conference on simulation of adaptive behavior: From animals to animats (pp. 222227). MIT Press/Bradford Books.Google Scholar
Shenhav, A., Musslick, S., Lieder, F., Kool, W., Griffiths, T. L., Cohen, J. D., & Botvinick, M. M. (2017). Toward a rational and mechanistic account of mental effort. Annual Review of Neuroscience, 40, 99124.Google Scholar
Simmons, J. P., Nelson, L. D., & Simonsohn, U. (2011). False-positive psychology: Undisclosed flexibility in data collection and analysis allows presenting anything as significant. Psychological Science, 22(11), 13591366.Google Scholar
Simon, H. A. (1972). Complexity and the representation of patterned sequences of symbols. Psychological Review, 79(5), 369.Google Scholar
Sorg, J., Singh, S. P., & Lewis, R. L. (2010). Internal rewards mitigate agent boundedness. In Proceedings of the 27th International Conference on Machine Learning (ICML-10) (pp. 10071014). https://icml.cc/Conferences/2010/papers/442.pdf.Google Scholar
Sun, Z., & Firestone, C. (2020). The dark room problem. Trends in Cognitive Sciences, 24. https://doi.org/10.1016/j.tics.2020.02.006.Google Scholar
Sunstein, C. R. (2002). The law of group polarization. The Journal of Political Philosophy, 10(2), 175195.Google Scholar
Sutton, R. S., & Barto, A. G. (2018). Reinforcement learning: An introduction. MIT press.Google Scholar
Tenenbaum, J. B., Griffiths, T. L., & Niyogi, S. (2007). Intuitive theories as grammars for causal inference. In Causal learning: Psychology, philosophy, and computation, 301322. https://oxford.universitypressscholarship.com/view/10.1093/acprof:oso/9780195176803.001.0001/acprof-9780195176803-chapter-20.Google Scholar
Vallone, R. P., Ross, L., & Lepper, M. R. (1985). The hostile media phenomenon: Biased perception and perceptions of media bias in coverage of the Beirut massacre. Journal of Personality and Social Psychology, 49(3), 577585.Google Scholar
Van Rooij, I. (2008). The tractable cognition thesis. Cognitive Science, 32(6), 939984.Google Scholar
Voß, J. (2005). Measuring Wikipedia. In International Conference of the International Society for Scientometrics and Informetrics: 10th, Stockholm (Sweden), 24–28 July 2005 (pp. 221231). http://eprints.rclis.org/6207/.Google Scholar
Vul, E., Goodman, N., Griffiths, T. L., & Tenenbaum, J. B. (2014). One and done? Optimal decisions from very few samples. Cognitive Science, 38(4), 599637.Google Scholar
Vul, E., & Pashler, H. (2008). Measuring the crowd within: Probabilistic representations within individuals. Psychological Science, 19(7), 645647.Google Scholar
Wade, S., & Kidd, C. (2019). The role of prior knowledge and curiosity in learning. Psychonomic Bulletin & Review, 26(4), 13771387.Google Scholar
Wojtowicz, Z., Chater, N., & Loewenstein, G. (2020). Boredom and flow: An opportunity cost theory of attention-directing motivational states. Available at SSRN 3339123. https://papers.ssrn.com/sol3/papers.cfm?abstract_id=3339123.Google Scholar
Wojtowicz, Z., & DeDeo, S. (2020). From probability to consilience: How explanatory values implement Bayesian reasoning. Trends in Cognitive Sciences, 24(12), 981993.CrossRefGoogle ScholarPubMed
Wojtowicz, Z., & Loewenstein, G. (2020). Curiosity and the economics of attention. Current Opinion in Behavioral Sciences, 35, 135140.Google Scholar
Yuille, A., & Kersten, D. (2006). Vision as Bayesian inference: Analysis by synthesis? Trends in Cognitive Sciences, 10(7), 301308.Google Scholar
Zarouali, B., Dobber, T., De Pauw, G., & de Vreese, C. (2020). Using a personality-profiling algorithm to investigate political microtargeting: Assessing the persuasion effects of personality-tailored ads on social media. Communication Research, 0093650220961965.Google Scholar
Zeidner, M. (2010). Test anxiety. In Weiner, I. B., & Craighead, W. E. (eds.), The Corsini encyclopedia of psychology, pp. 13. John Wiley.Google Scholar

References

Alexander, P. A., Jetton, T. L., & Kulikowich, J. M. (1995). Interrelationship of knowledge, interest, and recall: Assessing a model of domain learning. Journal of Educational Psychology, 87(4), 559575. https://doi.org/10.1037/0022-0663.87.4.559.Google Scholar
Alexander, P. A., Kulikowich, J. M., & Schulze, S. K. (1994). How subject-matter knowledge affects recall and interest. American Educational Research Journal, 31(2), 313337.Google Scholar
Berlyne, D. E. (1949). “Interest” as a psychological concept. British Journal of Psychology: General Section, 39(4), 184195. https://doi.org/10.1111/j.2044-8295.1949.tb00219.x.Google Scholar
Berlyne, D. E. (1950). Novelty and curiosity as determinants of exploratory behaviour. British Journal of Psychology, 41(1), 6880.Google Scholar
Berlyne, D. E. (1960). Conflict, Arousal, and Curiosity. McGraw-Hill Book Company. https://doi.org/10.1037/11164-000.CrossRefGoogle Scholar
Berridge, K. C. (2000). Reward learning: Reinforcement, incentives, and expectations. Psychology of Learning and Motivation – Advances in Research and Theory, 40, 223278. https://doi.org/10.1016/s0079-7421(00)80022-5.Google Scholar
Charpentier, C. J., Bromberg-Martin, E. S., & Sharot, T. (2018). Valuation of knowledge and ignorance in mesolimbic reward circuitry. Proceedings of the National Academy of Sciences of the United States of America, 115(31), E7255E7264. https://doi.org/10.1073/pnas.1800547115.Google Scholar
Chu, L., Tsai, J. L., & Fung, H. H. (2020). Association between age and intellectual curiosity: The mediating roles of future time perspective and importance of curiosity. European Journal of Ageing. https://doi.org/10.1007/s10433-020-00567-6.Google Scholar
Conway, M. A., & Pleydell-Pearce, C. W. (2000). The construction of autobiographical memories in the self-memory system. Psychological Review, 107(2), 261288. https://doi.org/10.1037/0033-295X.107.2.261.Google Scholar
Day, H. I. (1982). Curiosity and the interested explorer. Performance & Instruction, 21(4), 1922. https://doi.org/10.1002/pfi.4170210410.Google Scholar
Dodson, C. S., Bawa, S., & Krueger, L. E. (2007). Aging, metamemory, and high-confidence errors: A misrecollection account. Psychology and Aging, 22(1), 122133. https://doi.org/10.1037/0882-7974.22.1.122.Google Scholar
Donnellan, E., Aslan, S., Fastrich, G. M., & Murayama, K., (2021). How are curiosity and interest different? Naïve Bayes classification of people’s beliefs. Educational Psychology Review. https://doi.org/10.1007/s10648-021-09622-9.CrossRefGoogle Scholar
Dubey, R., Mehta, H., & Lombrozo, T. (2021). Curiosity is contagious: A social influence intervention to induce curiosity. Cognitive Science, 45(2). https://doi.org/10.1111/cogs.12937.Google Scholar
Fastrich, G. M., & Murayama, K. (2020). Development of interest and role of choice during sequential knowledge acquisition. AERA Open, 6(2), 233285842092998. https://doi.org/10.1177/2332858420929981.CrossRefGoogle Scholar
FitzGibbon, L., Komiya, A., & Murayama, K. (2021). The lure of counterfactual curiosity: People incur a cost to experience regret. Psychological Science, 32(2), 241255. https://doi.org/10.1177/0956797620963615.Google Scholar
Frenzel, A. C., Pekrun, R., Dicke, A. L., & Goetz, T. (2012). Beyond quantitative decline: Conceptual shifts in adolescents’ development of interest in mathematics. Developmental Psychology, 48(4), 10691082. https://doi.org/10.1037/a0026895.Google Scholar
Gershman, S. J., & Niv, Y. (2015). Novelty and inductive generalization in human reinforcement learning. Topics in Cognitive Science, 7(3), 391415. https://doi.org/10.1111/tops.12138.Google Scholar
Gottlieb, J., Oudeyer, P.-Y., Lopes, M., & Baranes, A. (2013). Information-seeking, curiosity, and attention: computational and neural mechanisms. Trends in Cognitive Sciences, 17(11), 585593. https://doi.org/10.1016/j.tics.2013.09.001.Google Scholar
Grossnickle, E. M. (2016). Disentangling curiosity: Dimensionality, definitions, and distinctions from interest in educational contexts. Educational Psychology Review, 28(1), 2360. https://doi.org/10.1007/s10648-014-9294-y.Google Scholar
Gruber, M. J., Gelman, B. D., & Ranganath, C. (2014). States of curiosity modulate hippocampus-dependent learning via the dopaminergic circuit. Neuron, 84(2), 486496. https://doi.org/10.1016/j.neuron.2014.08.060.Google Scholar
Gruber, M. J., & Ranganath, C. (2019). How curiosity enhances hippocampus-dependent memory: The prediction, appraisal, curiosity, and exploration (PACE) framework. Trends in Cognitive Sciences, 23(12), 10141025. https://doi.org/10.1016/j.tics.2019.10.003.Google Scholar
Gruber, M. J., Ritchey, M., Wang, S. F., Doss, M. K., & Ranganath, C. (2016). Post-learning hippocampal dynamics promote preferential retention of rewarding events. Neuron, 89(5), 11101120. https://doi.org/10.1016/j.neuron.2016.01.017.Google Scholar
Harackiewicz, J. M., Durik, A. M., Barron, K. E., Linnenbrink-Garcia, L., & Tauer, J. M. (2008). The role of achievement goals in the development of interest: Reciprocal relations between achievement goals, interest, and performance. Journal of Educational Psychology, 100(1), 105122. https://doi.org/10.1037/0022-0663.100.1.105.CrossRefGoogle Scholar
Harrison, S. H., Sluss, D. M., & Ashforth, B. E. (2011). Curiosity adapted the cat: The role of trait curiosity in newcomer adaptation. Journal of Applied Psychology, 96(1), 211220. https://doi.org/10.1037/a0021647.Google Scholar
Hidi, S. E., & Renninger, K. A. (2006). The four-phase model of interest development. Educational Psychologist, 41(2), 111127. https://doi.org/10.1207/s15326985ep4102_4.Google Scholar
Hsee, C. K., & Ruan, B. (2016). The Pandora Effect: The power and peril of curiosity. Psychological Science, 27(5), 659666. https://doi.org/10.1177/0956797616631733.Google Scholar
Iigaya, K., Hauser, T. U., Kurth-Nelson, Z., O’Doherty, J. P., Dayan, P., & Dolan, R. J. (2020). The value of what’s to come: Neural mechanisms coupling prediction error and the utility of anticipation. Science Advances, 6(25), eaba3828. https://doi.org/10.1126/sciadv.aba3828.Google Scholar
Jepma, M., Verdonschot, R. G., van Steenbergen, H., Rombouts, S. A. R. B., & Nieuwenhuis, S. (2012). Neural mechanisms underlying the induction and relief of perceptual curiosity. Frontiers in Behavioral Neuroscience, 6, 19. https://doi.org/10.3389/fnbeh.2012.00005.Google Scholar
John, O. P., & Srivastava, S. (1999). The Big Five Trait taxonomy: History, measurement, and theoretical perspectives. In Pervin, L. A. & John, O. P. (Eds.), Handbook of personality: Theory and research (2nd edition) (pp. 102138). Guildford.Google Scholar
Kang, M. J., Hsu, M., Krajbich, I. M., Loewenstein, G., McClure, S. M., Wang, J. T., & Camerer, C. F. (2009). The wick in the candle of learning. Psychological Science, 20(8), 963973. https://doi.org/10.1111/j.1467-9280.2009.02402.x.Google Scholar
Kashdan, T. B., Rose, P., & Fincham, F. D. (2004). Curiosity and exploration: Facilitating positive subjective experiences and personal growth opportunities. Journal of Personality Assessment, 82(3), 291305. https://doi.org/10.1207/s15327752jpa8203_05.Google Scholar
Kavé, G., & Halamish, V. (2015). Doubly blessed: Older adults know more vocabulary and know better what they know. Psychology and Aging, 30(1), 6873. https://doi.org/10.1037/a0038669.Google Scholar
Kidd, C., & Hayden, B. Y. (2015). The psychology and neuroscience of curiosity. Neuron, 88(3), 449460. https://doi.org/10.1016/j.neuron.2015.09.010.Google Scholar
Kobayashi, K., & Hsu, M. (2019). Common neural code for reward and information value. Proceedings of the National Academy of Sciences of the United States of America, 116(26), 1306113066. https://doi.org/10.1073/pnas.1820145116.Google Scholar
Kobayashi, K., Ravaioli, S., Baranès, A., Woodford, M., & Gottlieb, J. (2019). Diverse motives for human curiosity. Nature Human Behaviour, 3(6), 587595. https://doi.org/10.1038/s41562-019-0589-3.Google Scholar
Krapp, A. (2000). Interest and human development during adolescence: An educational-psychological approach. In Heckhausen, J. (Ed.), Motivational psychology of human development: Developing motivation and motivating development (pp. 109128). Elsevier Science. https://doi.org/10.1016/S0166-4115(00)80008-4.Google Scholar
Lanzetta, J. T., & Driscoll, J. M. (1966). Preference for information about an uncertain but unavoidable outcome. Journal of Personality and Social Psychology. https://doi.org/10.1037/h0022674.Google Scholar
Lau, J. K. L., Ozono, H., Kuratomi, K., Komiya, A., & Murayama, K. (2020). Shared striatal activity in decisions to satisfy curiosity and hunger at the risk of electric shocks. Nature Human Behaviour. https://doi.org/10.1038/s41562-020-0848-3.Google Scholar
Litman, J. A., Collins, R. P., & Spielberger, C. D. (2005). The nature and measurement of sensory curiosity. Personality and Individual Differences, 39(6), 11231133. https://doi.org/10.1016/j.paid.2005.05.001.Google Scholar
Litman, J. A., Hutchins, T., & Russon, R. (2005). Epistemic curiosity, feeling-of-knowing, and exploratory behaviour. Cognition & Emotion, 19(4), 559582. https://doi.org/10.1080/02699930441000427.Google Scholar
Litman, J. A., & Spielberger, C. D. (2003). Measuring epistemic curiosity and its diversive and specific components. Journal of Personality Assessment, 80(1), 7586. http://dx.doi.org/10.1207/S15327752JPA8001_16.Google Scholar
Loewenstein, G. (1994). The psychology of curiosity: A review and reinterpretation. Psychological Bulletin, 116(1), 7598. https://doi.org/10.1037/0033-2909.116.1.75.Google Scholar
Mata, R., Wilke, A., & Czienskowski, U. (2013). Foraging across the life span: is there a reduction in exploration with aging? Frontiers in Neuroscience, 7, 17. https://doi.org/10.3389/fnins.2013.00053.Google Scholar
McCrae, R. R., & Costa, P. T. (1987). Validation of the five-factor model of personality across instruments and observers. Journal of Personality and Social Psychology, 52(1), 8190. https://doi.org/10.1037/0022-3514.52.1.81.Google Scholar
Metcalfe, J., Schwartz, B. L., & Eich, T. S. (2020). Epistemic curiosity and the region of proximal learning. Current Opinion in Behavioral Sciences, 35, 4047. https://doi.org/10.1016/j.cobeha.2020.06.007.Google Scholar
Mirolli, M., & Baldassarre, G. (2013). Functions and mechanisms of intrinsic motivations. In Baldassarre, G. & Mirolli, M. (Eds.), Intrinsically motivating learning in natural and artificial systems (pp. 4972). Springer.Google Scholar
Murayama, K. (2022). A reward-learning framework of knowledge acquisition: An integrated account of curiosity, interest, and intrinsic-extrinsic rewards. Psychological Review, https://doi.org/10.1037/rev0000349.Google Scholar
Murayama, K., FitzGibbon, L., & Sakaki, M. (2019). Process account of curiosity and interest: A reward-learning perspective. Educational Psychology Review, 31(4), 875895. https://doi.org/10.1007/s10648-019-09499-9.Google Scholar
O’Doherty, J. P. (2004). Reward representations and reward-related learning in the human brain: Insights from neuroimaging. Current Opinion in Neurobiology, 14(6), 769776. https://doi.org/10.1016/j.conb.2004.10.016.Google Scholar
Oosterwijk, S. (2017). Choosing the negative: A behavioral demonstration of morbid curiosity. PloS One, 12(7), 120. https://doi.org/10.1371/journal.pone.0178399.Google Scholar
Oudeyer, P.-Y., & Kaplan, F. (2009). What is intrinsic motivation? A typology of computational approaches. Frontiers in Neurorobotics, 1, 114. https://doi.org/10.3389/neuro.12.006.2007.Google Scholar
Peterson, E. G., & Cohen, J. (2019). A case for domain-specific curiosity in mathematics. Educational Psychology Review, 31(4), 807832. https://doi.org/10.1007/s10648-019-09501-4.Google Scholar
Peterson, E. G., & Hidi, S. E. (2019). Curiosity and interest: Current perspectives. Educational Psychology Review, 31(4), 781788. https://doi.org/10.1007/s10648-019-09513-0.Google Scholar
Prenzel, M. (1992). Selective persistence of interest. In Renninger, K. A., Hidi, S. E., & Krapp, A. (Eds.), The role of interest in learning and development (pp. 7198). Erlbaum.Google Scholar
Renninger, K. A., & Hidi, S. E. (2016). The power of interest for motivation and engagement. Routledge. www.taylorfrancis.com/books/9781317674214.Google Scholar
Robinson, O. C., Demetre, J. D., & Litman, J. A. (2017). Adult life stage and crisis as predictors of curiosity and authenticity. International Journal of Behavioral Development, 41(3), 426431. https://doi.org/10.1177/0165025416645201.Google Scholar
Rodriguez Cabrero, J. A. M., Zhu, J.-Q., & Ludvig, E. A. (2019). Costly curiosity: People pay a price to resolve an uncertain gamble early. Behavioural Processes, 160, 2025. https://doi.org/10.1016/j.beproc.2018.12.015.Google Scholar
Rotgans, J. I., & Schmidt, H. G. (2017). Interest development: Arousing situational interest affects the growth trajectory of individual interest. Contemporary Educational Psychology, 49, 175184. https://doi.org/10.1016/j.cedpsych.2017.02.003.Google Scholar
Rushworth, M. F. S., Mars, R. B., & Summerfield, C. (2009). General mechanisms for making decisions? Current Opinion in Neurobiology, 19(1), 7583. https://doi.org/10.1016/j.conb.2009.02.005.Google Scholar
Sakaki, M., Yagi, A., & Murayama, K. (2018). Curiosity in old age: A possible key to achieving adaptive aging. Neuroscience and Biobehavioral Reviews, 88, 106116. https://doi.org/10.1016/j.neubiorev.2018.03.007.CrossRefGoogle ScholarPubMed
Sansone, C., & Thoman, D. B. (2005). Interest as the missing motivator in self-regulation. European Psychologist, 10(3), 175186. https://doi.org/10.1027/1016-9040.10.3.175.Google Scholar
Schiefele, U. (2009). Situational and individual interest. In Wenzel, K. R. & Wigfield, A. (Eds.), Educational psychology handbook series: Handbook of motivation at school (pp. 197222). Routledge/Taylor & Francis Group.Google Scholar
Schulz, E., Bhui, R., Love, B. C., Brier, B., Todd, M. T., & Gershman, S. J. (2019). Structured, uncertainty-driven exploration in real-world consumer choice. Proceedings of the National Academy of Sciences, 116(28), 1390313908. https://doi.org/10.1073/pnas.1821028116.Google Scholar
Schulz, E., Wu, C. M., Ruggeri, A., & Meder, B. (2019). Searching for rewards like a child means less generalization and more directed exploration. Psychological Science, 30(11), 15611572. https://doi.org/10.1177/0956797619863663.Google Scholar
Sedikides, C., & Strube, M. J. (1997). Self-evaluation: To thine own self be good, to thine own self be sure, to thine own self be true, and to thine own self be better. Advances in Experimental Social Psychology, 29, 209269. https://doi.org/10.1016/S0065-2601(08)60018-0.Google Scholar
Sharot, T., & Sunstein, C. R. (2020). How people decide what they want to know. Nature Human Behaviour, 4(1), 1419. https://doi.org/10.1038/s41562-019-0793-1.Google Scholar
Somerville, L. H., Sasse, S. F., Garrad, M. C., Drysdale, A. T., Abi Akar, N., Insel, C., & Wilson, R. C. (2017). Charting the expansion of strategic exploratory behavior during adolescence. Journal of Experimental Psychology: General, 146(2), 155164. https://doi.org/10.1037/xge0000250.Google Scholar
Van Lieshout, L. L. F., Vandenbroucke, A. R. E., Müller, N. C. J., Cools, R., & de Lange, F. P. (2018). Induction and relief of curiosity elicit parietal and frontal activity. Journal of Neuroscience, 38(10), 25792588. https://doi.org/10.1523/JNEUROSCI.2816-17.2018.Google Scholar
Witherby, A. E., & Carpenter, S. K. (2021). The rich-get-richer effect: Prior knowledge predicts new learning of domain-relevant information. Journal of Experimental Psychology: Learning, Memory, and Cognition. https://doi.org/10.1037/xlm0000996.Google Scholar
Wu, C. M., Schulz, E., Speekenbrink, M., Nelson, J. D., & Meder, B. (2018). Generalization guides human exploration in vast decision spaces. Nature Human Behaviour, 2(12), 915924. https://doi.org/10.1038/s41562-018-0467-4.Google Scholar

References

Andreae, P. M., & Andreae, J. H. (1978). A teachable machine in the real world. International Journal of Man-Machine Studies, 10(3), 301312.Google Scholar
Andrychowicz, M., Wolski, F., Ray, A., Schneider, J., Fong, R., Welinder, P., … & Zaremba, W. (2018). Hindsight experience replay. arXiv preprint arXiv:1707.01495.Google Scholar
Aubret, A., Matignon, L., & Hassas, S. (2019). A survey on intrinsic motivation in reinforcement learning. arXiv preprint arXiv:1908.06976.Google Scholar
Baker, B., Kanitscheider, I., Markov, T., Wu, Y., Powell, G., McGrew, B., & Mordatch, I. (2020). Emergent tool use from multi-agent autocurricula. arXiv preprint arXiv:1909.07528.Google Scholar
Baranes, A., & Oudeyer, P. Y. (2009). R-iac: Robust intrinsically motivated exploration and active learning. IEEE Transactions on Autonomous Mental Development, 1(3), 155169.Google Scholar
Baranes, A., & Oudeyer, P. Y. (2013). Active learning of inverse models with intrinsically motivated goal exploration in robots. Robotics and Autonomous Systems, 61(1), 4973.Google Scholar
Barron, A. B., Hebets, E. A., Cleland, T. A., Fitzpatrick, C. L., Hauber, M. E., & Stevens, J. R. (2015). Embracing multiple definitions of learning. Trends in Neurosciences, 38(7), 405407.Google Scholar
Bazhydai, M., Twomey, K., & Westermann, G. (2021). Curiosity and Exploration. In Benson, J. B. (Ed.), Encyclopedia of Infant and Early Childhood Development (2nd ed.). Elsevier, pp. 370378.Google Scholar
Bellemare, M., Srinivasan, S., Ostrovski, G., Schaul, T., Saxton, D., & Munos, R. (2016). Unifying count-based exploration and intrinsic motivation. Advances in Neural Information Processing Systems, 29, 14711479.Google Scholar
Benureau, F. C., & Oudeyer, P. Y. (2016). Behavioral diversity generation in autonomous exploration through reuse of past experience. Frontiers in Robotics and AI, 3, 8.Google Scholar
Berseth, G., Geng, D., Devin, C., Rhinehart, N., Finn, C., Jayaraman, D., & Levine, S. (2021). SMiRL: Surprise Minimizing Reinforcement Learning in Unstable Environments. arXiv preprint arXiv:1912.05510.Google Scholar
Bougie, N., & Ichise, R. (2020). Skill-based curiosity for intrinsically motivated reinforcement learning. Machine Learning, 109(3), 493512.Google Scholar
Bougie, N., & Ichise, R. (2021). Fast and slow curiosity for high-level exploration in reinforcement learning. Applied Intelligence, 51(2), 10861107.Google Scholar
Burda, Y., Edwards, H., Pathak, D., Storkey, A., Darrell, T., & Efros, A. A. (2018). Large-scale study of curiosity-driven learning. arXiv preprint arXiv:1808.04355.Google Scholar
Caligiore, D., Ferrauto, T., Parisi, D., Accornero, N., Capozza, M., & Baldassarre, G. (2008). Using motor babbling and hebb rules for modeling the development of reaching with obstacles and grasping. In International Conference on Cognitive Systems (Vol. 13, pp. 2223). www.researchgate.net/publication/227945187_Using_Motor_Babbling_and_Hebb_Rules_for_Modeling_the_Development_of_Reaching_with_Obstacles_and_Grasping.Google Scholar
Chu, J., & Schulz, L. E. (2020). Play, curiosity, and cognition. Annual Review of Developmental Psychology, 2, 317343.Google Scholar
Clement, B., Oudeyer, P. Y., & Lopes, M. (2016). A Comparison of Automatic Teaching Strategies for Heterogeneous Student Populations. Proceedings of the 9th International Conference on Educational Data Mining, Raleigh, USA.Google Scholar
Clement, B., Roy, D., Oudeyer, P. Y., & Lopes, M. (2015). Multi-armed bandits for intelligent tutoring systems. arXiv preprint arXiv:1310.3174.Google Scholar
Cohn, D. A., Ghahramani, Z., & Jordan, M. I. (1996). Active learning with statistical models. Journal of Artificial Intelligence Research, 4, 129145.Google Scholar
Colas, C., Fournier, P., Chetouani, M., Sigaud, O., & Oudeyer, P. Y. (2019, May). CURIOUS: intrinsically motivated modular multi-goal reinforcement learning. In International conference on machine learning (pp. 13311340). PMLR. http://proceedings.mlr.press/v97/colas19a.html.Google Scholar
Colas, C., Karch, T., Lair, N., Dussoux, J. M., Moulin-Frier, C., Dominey, P. F., & Oudeyer, P. Y. (2020). Language as a cognitive tool to imagine goals in curiosity-driven exploration. arXiv preprint arXiv:2002.09253.Google Scholar
Colas, C., Karch, T., Sigaud, O., & Oudeyer, P. Y. (2021). Intrinsically motivated goal-conditioned reinforcement learning: a short survey. arXiv preprint arXiv:2012.09830.Google Scholar
Colas, C., Sigaud, O., & Oudeyer, P. Y. (2018). Gep-pg: Decoupling exploration and exploitation in deep reinforcement learning algorithms. In International conference on machine learning (pp. 10391048). PMLR. http://proceedings.mlr.press/v80/colas18a.html.Google Scholar
Delmas, A., Clement, B., Oudeyer, P. Y., & Sauzéon, H. (2018). Fostering health education with a serious game in children with asthma: pilot studies for assessing learning efficacy and automatized learning personalization. In Frontiers in Education (Vol. 3, p. 99). Frontiers. https://doi.org/10.3389/feduc.2018.00099.Google Scholar
Dobzhansky, T. (1973). Nothing in biology makes sense except in the light of evolution. The American Biology Teacher, 75(2), 8791.Google Scholar
Etcheverry, M., Moulin-Frier, C., & Oudeyer, P. Y. (2021). Hierarchically organized latent modules for exploratory search in morphogenetic systems. arXiv preprint arXiv:2007.01195.Google Scholar
Florensa, C., Held, D., Geng, X., & Abbeel, P. (2018, July). Automatic goal generation for reinforcement learning agents. In International conference on machine learning (pp. 15151528). PMLR. http://proceedings.mlr.press/v80/florensa18a.html.Google Scholar
Fogarty, L., & Creanza, N. (2017). The niche construction of cultural complexity: interactions between innovations, population size and the environment. Philosophical Transactions of the Royal Society B: Biological Sciences, 372(1735), 20160428.Google Scholar
Forestier, S., Portelas, R., Mollard, Y., & Oudeyer, P. Y. (2017). Intrinsically motivated goal exploration processes with automatic curriculum learning. arXiv preprint arXiv:1708.02190.Google Scholar
Gershman, S. J. (2019). Uncertainty and exploration. Decision, 6(3), 277.Google Scholar
Gopnik, A. (2020). Childhood as a solution to explore–exploit tensions. Philosophical Transactions of the Royal Society B, 375(1803), 20190502.Google Scholar
Gottlieb, J., & Oudeyer, P. Y. (2018). Towards a neuroscience of active sampling and curiosity. Nature Reviews Neuroscience, 19(12), 758770.Google Scholar
Gottlieb, J., Oudeyer, P. Y., Lopes, M., & Baranes, A. (2013). Information-seeking, curiosity, and attention: computational and neural mechanisms. Trends in Cognitive Sciences, 17(11), 585593.Google Scholar
Grizou, J., Points, L. J., Sharma, A., & Cronin, L. (2020). A curious formulation robot enables the discovery of a novel protocell behavior. Science Advances, 6(5), eaay4237.Google Scholar
Gross, M. E., Zedelius, C. M., & Schooler, J. W. (2020). Cultivating an understanding of curiosity as a seed for creativity. Current Opinion in Behavioral Sciences, 35, 7782.Google Scholar
Haber, N., Mrowca, D., Fei-Fei, L., & Yamins, D. L. (2018). Learning to play with intrinsically-motivated self-aware agents. arXiv preprint arXiv:1802.07442.Google Scholar
Harlow, H. F., Harlow, M. K., & Meyer, D. R. (1950). Learning motivated by a manipulation drive. Journal of Experimental Psychology, 40(2), 228.Google Scholar
Hidi, S., & Renninger, K. A. (2006). The four-phase model of interest development. Educational Psychologist, 41(2), 111127.Google Scholar
Holm, L., Wadenholt, G., & Schrater, P. (2019). Episodic curiosity for avoiding asteroids: Per-trial information gain for choice outcomes drive information seeking. Scientific Reports, 9(1), 116.Google Scholar
Hull, C. L. (1943). Principles of behavior: An introduction to behavior theory. Appleton-Century.Google Scholar
Jaderberg, M., Mnih, V., Czarnecki, W. M., Schaul, T., Leibo, J. Z., Silver, D., & Kavukcuoglu, K. (2016). Reinforcement learning with unsupervised auxiliary tasks. arXiv preprint arXiv:1611.05397.Google Scholar
Jaques, N., Lazaridou, A., Hughes, E., Gulcehre, C., Ortega, P., Strouse, D. J., … & De Freitas, N. (2019, May). Social influence as intrinsic motivation for multi-agent deep reinforcement learning. In International Conference on Machine Learning (pp. 30403049). PMLR. http://proceedings.mlr.press/v97/jaques19a.html.Google Scholar
Jordan, M. I., & Mitchell, T. M. (2015). Machine learning: Trends, perspectives, and prospects. Science, 349(6245), 255260.Google Scholar
Jordan, M. I., & Rumelhart, D. E. (1992). Forward models: Supervised learning with a distal teacher. Cognitive Science, 16(3), 307354.Google Scholar
Kaplan, F., & Oudeyer, P. Y. (2007). In search of the neural circuits of intrinsic motivation. Frontiers in Neuroscience, 1, 17.Google Scholar
Kim, K., Sano, M., De Freitas, J., Haber, N., & Yamins, D. (2020). Active world model learning with progress curiosity. In International conference on machine learning (pp. 53065315). PMLR. https://proceedings.mlr.press/v119/kim20e.html.Google Scholar
Laversanne-Finot, A., Péré, A., & Oudeyer, P. Y. (2018). Curiosity driven exploration of learned disentangled goal spaces. In Conference on Robot Learning (pp. 487504). PMLR. https://proceedings.mlr.press/v87/laversanne-finot18a.html.Google Scholar
Laversanne-Finot, A., Péré, A., & Oudeyer, P. Y. (2021). Intrinsically motivated exploration of learned goal spaces. Frontiers in Neurorobotics, 14, 109.Google Scholar
Lefort, M., & Gepperth, A. (2015). Active learning of local predictable representations with artificial curiosity. In 2015 Joint IEEE International Conference on Development and Learning and Epigenetic Robotics (ICDL-EpiRob) (pp. 228233). IEEE. https://ieeexplore.ieee.org/abstract/document/7346145.Google Scholar
Leibo, J. Z., Hughes, E., Lanctot, M., & Graepel, T. (2019). Autocurricula and the emergence of innovation from social interaction: A manifesto for multi-agent intelligence research. arXiv preprint arXiv:1903.00742.Google Scholar
Lin, B., Cecchi, G., Bouneffouf, D., Reinen, J., & Rish, I. (2019). A story of two streams: Reinforcement learning models from human behavior and neuropsychiatry. arXiv preprint arXiv:1906.11286.Google Scholar
Linke, C., Ady, N. M., White, M., Degris, T., & White, A. (2020). Adapting behavior via intrinsic reward: A survey and empirical study. Journal of Artificial Intelligence Research, 69, 12871332.Google Scholar
Loewenstein, G. (1994). The psychology of curiosity: A review and reinterpretation. Psychological Bulletin, 116(1), 75.CrossRefGoogle Scholar
Lucas, R. E. (2004). The industrial revolution: Past and future. Economic Education Bulletin, 44(8), 18. www.aier.org/wp-content/uploads/2013/11/EEB-8.04-IndustRev.pdf.Google Scholar
Maturana, H. R., & Varela, F. J. (1980). Autopoiesis and cognition: The realization of the living (Vol. 42). Springer Science & Business Media.Google Scholar
McClelland, J. L. (2009). The place of modeling in cognitive science. Topics in Cognitive Science, 1(1), 1138.Google Scholar
Mirolli, M., & Baldassarre, G. (2013). Functions and mechanisms of intrinsic motivations. In Baldassarre, G & Mirolli, M (Eds.), Intrinsically Motivated Learning in Natural and Artificial Systems (pp. 4972). Springer.Google Scholar
Mnih, V., Kavukcuoglu, K., Silver, D., Rusu, A. A., Veness, J., Bellemare, M. G., … & Hassabis, D. (2015). Human-level control through deep reinforcement learning. Nature, 518(7540), 529533.Google Scholar
Moulin-Frier, C., Nguyen, S. M., & Oudeyer, P. Y. (2014). Self-organization of early vocal development in infants and machines: the role of intrinsic motivation. Frontiers in Psychology, 4, 1006.Google Scholar
Moulin-Frier, C., & Oudeyer, P. Y. (2013, August). Exploration strategies in developmental robotics: A unified probabilistic framework. In 2013 IEEE Third Joint International Conference on Development and Learning and Epigenetic Robotics (ICDL) (pp. 16). IEEE. https://doi.org/10.1109/DevLrn.2013.6652535.Google Scholar
Nair, A., Pong, V., Dalal, M., Bahl, S., Lin, S., & Levine, S. (2018). Visual reinforcement learning with imagined goals. arXiv preprint arXiv:1807.04742.Google Scholar
Nake, F. (1976). Ästhetik als Informationsverarbeitung: Grundlagen und Anwendungen der Informatik im Bereich ästhetischer Produktion und Kritik. Journal of Aesthetics and Art Criticism, 34(3).Google Scholar
Nguyen, S. M., & Oudeyer, P. Y. (2012). Active choice of teachers, learning strategies and goals for a socially guided intrinsic motivation learner. Paladyn, 3(3), 136146.Google Scholar
Eleni Nisioti, Katia Jodogne-del Litto, Clément Moulin-Frier. Grounding an Ecological Theory of Artificial Intelligence in Human Evolution. NeurIPS 2021 - Conference on Neural Information Processing Systems / Workshop: Ecological Theory of Reinforcement Learning, Dec 2021, virtual event, France. (hal-03446961v2)Google Scholar
Oller, D. K. (2000). The emergence of the speech capacity. Psychology Press.Google Scholar
Oudeyer, P. Y. (2018). Computational theories of curiosity-driven learning. arXiv preprint arXiv:1802.10546.Google Scholar
Oudeyer, P. Y., & Kaplan, F. (2006). Discovering communication. Connection Science, 18(2), 189206.Google Scholar
Oudeyer, P. Y., & Kaplan, F. (2009). What is intrinsic motivation? A typology of computational approaches. Frontiers in Neurorobotics, 1, 6.Google Scholar
Oudeyer, P. Y., Kaplan, F., & Hafner, V. V. (2007). Intrinsic motivation systems for autonomous mental development. IEEE Transactions on Evolutionary Computation, 11(2), 265286.Google Scholar
Oudeyer, P. Y., & Smith, L. B. (2016). How evolution may work through curiosity‐driven developmental process. Topics in Cognitive Science, 8(2), 492502.Google Scholar
Pan, M., Huang, A., Wang, G., Zhang, T., & Li, X. (2020). Reinforcement learning based curiosity-driven testing of android applications. In Proceedings of the 29th ACM SIGSOFT International Symposium on Software Testing and Analysis (pp. 153164). https://doi.org/10.1145/3395363.3397354.Google Scholar
Pathak, D., Agrawal, P., Efros, A. A., & Darrell, T. (2017). Curiosity-driven exploration by self-supervised prediction. In International conference on machine learning (pp. 27782787). PMLR. http://proceedings.mlr.press/v70/pathak17a.html.Google Scholar
Poli, F., Serino, G., Mars, R. B., & Hunnius, S. (2020). Infants tailor their attention to maximize learning. Science Advances, 6(39), eabb5053.Google Scholar
Pong, V. H., Dalal, M., Lin, S., Nair, A., Bahl, S., & Levine, S. (2020). Skew-fit: State-covering self-supervised reinforcement learning. arXiv preprint arXiv:1903.03698.Google Scholar
Potts, R. (2013). Hominin evolution in settings of strong environmental variability. Quaternary Science Reviews, 73, 113.Google Scholar
Reinke, C., Etcheverry, M., & Oudeyer, P. Y. (2020). Intrinsically motivated discovery of diverse patterns in self-organizing systems. arXiv preprint arXiv:1908.06663.Google Scholar
Rolf, M., Steil, J. J., & Gienger, M. (2010). Goal babbling permits direct learning of inverse kinematics. IEEE Transactions on Autonomous Mental Development, 2(3), 216229.Google Scholar
Ryan, R. M., & Deci, E. L. (2000). Self-determination theory and the facilitation of intrinsic motivation, social development, and well-being. American Psychologist, 55(1), 68.Google Scholar
Saegusa, R., Metta, G., Sandini, G., & Sakka, S. (2009). Active motor babbling for sensorimotor learning. In 2008 IEEE International Conference on Robotics and Biomimetics (pp. 794799). IEEE. https://doi.org/10.1109/ROBIO.2009.4913101.Google Scholar
Santucci, V. G., Baldassarre, G., & Mirolli, M. (2013). Which is the best intrinsic motivation signal for learning multiple skills? Frontiers in Neurorobotics, 7, 22.Google Scholar
Schaul, T., Quan, J., Antonoglou, I., & Silver, D. (2016). Prioritized experience replay. arXiv preprint arXiv:1511.05952.Google Scholar
Schmidhuber, J. (1991a). A possibility for implementing curiosity and boredom in model-building neural controllers. In Proc. of the International Conference on Simulation of Adaptive Behavior: From Animals to Animats (pp. 222227). https://doi.org/10.7551/mitpress/3115.003.0030.Google Scholar
Schmidhuber, J. (1991b). Curious model-building control systems. In Proc. International Joint Conference on Neural Networks, Singapore City, November 18–21, 1991, (pp. 14581463). www.scirp.org/(S(lz5mqp453ed%20snp55rrgjct55))/reference/referencespapers.aspx?referenceid=1385254.Google Scholar
Schueller, W., Loreto, V., & Oudeyer, P. Y. (2018). Complexity reduction in the negotiation of new lexical conventions. arXiv preprint arXiv:1805.05631.Google Scholar
Singh, S., Lewis, R. L., Barto, A. G., & Sorg, J. (2010). Intrinsically motivated reinforcement learning: An evolutionary perspective. IEEE Transactions on Autonomous Mental Development, 2(2), 7082.Google Scholar
Stout, A., & Barto, A. G. (2010, August). Competence progress intrinsic motivation. In 2010 IEEE 9th International Conference on Development and Learning (pp. 257262). IEEE. http://citeseerx.ist.psu.edu/viewdoc/summary;jsessionid=837FA22F4803E257348D38A7397B2774?doi=10.1.1.224.71.Google Scholar
Sutton, R. S., & Barto, A. G. (2018). Reinforcement learning: An introduction. MIT press.Google Scholar
Takahashi, K., Ogata, T., Nakanishi, J., Cheng, G., & Sugano, S. (2017). Dynamic motion learning for multi-DOF flexible-joint robots using active–passive motor babbling through deep learning. Advanced Robotics, 31(18), 10021015.Google Scholar
Tang, H., Houthooft, R., Foote, D., Stooke, A., Chen, X., Duan, Y., … & Abbeel, P. (2017). #Exploration: A study of count-based exploration for deep reinforcement learning. Advances in Neural Information Processing Systems. Presented at the 31st Conference on Neural Information Processing Systems (NIPS), 2017. In 31st Conference on Neural Information Processing Systems (NIPS) (Vol. 30, pp. 118).Google Scholar
Ten, A., Kaushik, P., Oudeyer, P. Y., & Gottlieb, J. (2021). Humans monitor learning progress in curiosity-driven exploration. Nature Communications, 12(1), 110.Google Scholar
Thrun, S. (1995). A lifelong learning perspective for mobile robot control. In Intelligent robots and systems (pp. 201214). Elsevier Science BV.Google Scholar
Twomey, K. E., & Westermann, G. (2018). Curiosity‐based learning in infants: a neurocomputational approach. Developmental Science, 21(4), e12629.Google Scholar

References

Ackermann, L., Hepach, R., & Mani, N. (2020). Children learn words easier when they are interested in the category to which the word belongs. Developmental Science, 23(3), Article e12915. https://doi.org/10.1111/desc.12915.Google Scholar
Ainsworth, M. D. S. (1992). A consideration of social referencing in the context of attachment theory and research. In Feinman, S. (Ed.), Social referencing and the social construction of reality in infancy (pp. 349367). Plenum.Google Scholar
Alvarez, A. L., & Booth, A. E. (2014). Motivated by meaning: Testing the effect of knowledge-infused rewards on preschoolers’ persistence. Child Development, 85, 783791. https://doi.org/10.1111/cdev.12151.Google Scholar
Bazhydai, M., Westermann, G., & Parise, E. (2020). “I don’t know but I know who to ask”: 12‐month‐olds actively seek information from knowledgeable adults. Developmental Science, 23(5), Article e12938. https://doi.org/10.1111/desc.12938.Google Scholar
Begus, K., & Bonawitz, E. (2020). The rhythm of learning: Theta oscillations as an index of active learning in infancy. Developmental Cognitive Neuroscience, 45, Article 100810. https://doi.org/10.1016/j.dcn.2020.100810.Google Scholar
Begus, K., & Southgate, V. (2012). Infant pointing serves an interrogative function. Developmental Science, 15, 611617. https://doi.org/10.1111/j.1467-7687.2012.01160.x.Google Scholar
Begus, K., & Southgate, V. (2018). Curious learners: How infants’ motivation to learn shapes and is shaped by infants’ interactions with the social world. In Saylor, M. & Ganea, P. (Eds.), Active learning from infancy to childhood (pp. 1337). Springer. https://doi.org/10.1007/978-3-319-77182-3_2.Google Scholar
Betsch, T., Lang, A., Lehmann, A., & Axmann, J. M. (2014). Utilizing probabilities as decision weights in closed and open information boards: A comparison of children and adults. Acta Psychologica, 153, 7486. https://doi.org/10.1016/j.actpsy.2014.09.008.Google Scholar
Betsch, T., Lehmann, A., Lindow, S., Lang, A., & Schoemann, M. (2016). Lost in search: (Mal-)adaptation to probabilistic decision environments in children and adults. Developmental Psychology, 52(2), 311325. https://doi.org/10.1037/dev0000077.Google Scholar
Callanan, M. A., & Oakes, L. M. (1992). Preschoolers’ questions and parents’ explanations: Causal thinking in everyday activity. Cognitive Development, 7(2), 213233. https://doi.org/10.1016/0885-2014(92)90012-G.Google Scholar
Chouinard, M. M., Harris, P. L., & Maratsos, M. P. (2007). Children’s questions: A mechanism for cognitive development. Monographs of the Society for Research in Child Development, 72(1), i179. https://doi.org/10.1111/j.1540-5834.2007.00412.x.Google Scholar
Cook, C., Goodman, N. D., & Schulz, L. E. (2011). Where science starts: Spontaneous experiments in preschoolers’ exploratory play. Cognition, 120(3), 341349. https://doi.org/10.1016/j.cognition.2011.03.003.Google Scholar
Cooper, R. P., & Aslin, R. N. (1990). Preferences for infant-directed speech in the first month after birth. Child Development, 61(5), 15841595. https://doi.org/10.2307/1130766.Google Scholar
Csibra, G., & Gergely, G. (2009). Natural pedagogy. Trends in Cognitive Sciences, 13(4), 148153. https://doi.org/10.1016/j.tics.2009.01.005.Google Scholar
Delgado, B., Gómez, J. C., & Sarriá, E. (2011). Pointing gestures as a cognitive tool in young children: Experimental evidence. Journal of Experimental Child Psychology, 110(3), 299312. https://doi.org/10.1016/j.jecp.2011.04.010.Google Scholar
De Simone, C., Battisti, A., & Ruggeri, A. (2021). Differential impact of web habits and active navigation on adolescents’ online learning. PsyArXiv. https://doi.org/10.31234/osf.io/hsvc4.Google Scholar
De Simone, C., & Ruggeri, A. (2021). What is a good question asker better at? From unsystematic generalization to adult-like selectivity across childhood. Cognitive Development, 59, 101082. https://doi.org/10.1016/j.cogdev.2021.101082.Google Scholar
Dewey, J. (1986). Experience and education. The Educational Forum, 50(3), 241252. https://doi.org/10.1080/00131728609335764.Google Scholar
Dickstein, S., Thompson, R. A., Estes, D., Malkin, C., & Lamb, M. E. (1984). Social referencing and the security of attachment. Infant Behavior and Development, 7(4), 507516. https://doi.org/10.1016/S0163-6383(84)80009-0.Google Scholar
Domberg, A., Koskuba, K., Rothe, A., & Ruggeri, A. (2020). Goal-adaptiveness in children’s cue-based information search. In Denison., S., Mack, M., Xu, Y., & Armstrong, B.C. (Eds.), Proceedings of the 42nd Annual Conference of the Cognitive Science Society (pp. 14371443). Cognitive Science Society. https://cogsci.mindmodeling.org/2020/papers/0298/0298.pdf.Google Scholar
Donnellan, E., Bannard, C., McGillion, M. L., Slocombe, K. E., & Matthews, D. (2020). Infants’ intentionally communicative vocalizations elicit responses from caregivers and are the best predictors of the transition to language: A longitudinal investigation of infants’ vocalizations, gestures and word production. Developmental Science, 23, Article e12843. https://doi.org/10.1111/desc.12843.Google Scholar
Duarte Torres, S., & Weber, I. (2011). What and how children search on the web. In Proceedings of the 20th ACM International Conference on Information and Knowledge Management (pp. 393402). Association for Computing Machinery. https://doi.org/10.1145/2063576.2063638.Google Scholar
Dunn, K., & Bremner, J. G. (2017). Investigating looking and social looking measures as an index of infant violation of expectation. Developmental Science, 20, Article e12452. https://doi.org/10.1111/desc.12452.Google Scholar
Einav, S., & Robinson, E. J. (2010). Children’s sensitivity to error magnitude when evaluating informants. Cognitive Development, 25(3), 218232. https://doi.org/10.1016/j.cogdev.2010.04.002.Google Scholar
Elsner, C., & Wertz, A. E. (2019). The seeds of social learning: Infants exhibit more social looking for plants than other object types. Cognition, 183, 244255. https://doi.org/10.1016/j.cognition.2018.09.016.Google Scholar
Fitneva, S. A., Lam, N. H. L., & Dunfield, K. A. (2013). The development of children’s information gathering: To look or to ask? Developmental Psychology, 49(3), 533542. https://doi.org/10.1037/a0031326.Google Scholar
Frazier, B. N., Gelman, S. A., & Wellman, H. M. (2009). Preschoolers’ search for explanatory information within adult–child conversation. Child Development, 80(6), 15921611. https://doi.org/10.1111/j.1467-8624.2009.01356.x.Google Scholar
Freeman, J. L., Caldwell, P. H., Bennett, P. A., & Scott, K. M. (2018). How adolescents search for and appraise online health information: A systematic review. The Journal of Pediatrics, 195, 244255. https://doi.org/10.1016/j.jpeds.2017.11.031.Google Scholar
Goldin‐Meadow, S. (2007). Pointing sets the stage for learning language – and creating language. Child Development, 78, 741745. https://doi.org/10.1111/j.1467-8624.2007.01029.x.Google Scholar
Gopnik, A., & Sobel, D. M. (2000). Detecting blickets: How young children use information about novel causal powers in categorization and induction. Child Development, 71(5), 12051222. https://doi.org/10.1111/1467-8624.00224.Google Scholar
Gopnik, A., Frankenhuis, W. E., & Tomasello, M. (2020). Introduction to special issue: “Life history and learning: how childhood, caregiving and old age shape cognition and culture in humans and other animals.” Philosophical Transactions of the Royal Society B. https://doi.org/10.1098/rstb.2019.0489.Google Scholar
Gossen, T., Low, T., & Nürnberger, A. (2011). What are the real differences of children’s and adults’ web search. In Proceedings of the 34th International ACM SIGIR Conference on Research and Development in Information Retrieval (pp. 11151116). Association for Computing Machinery. https://doi.org/10.1145/2009916.2010076.Google Scholar
Gottlieb, J., Oudeyer, P.-Y., Lopes, M., & Baranes, A. (2013). Information-seeking, curiosity, and attention: Computational and neural mechanisms. Trends in Cognitive Sciences, 17(11), 585593. https://doi.org/10.1016/j.tics.2013.09.001.Google Scholar
Goupil, L., Romand-Monnier, M., & Kouider, S. (2016). Infants ask for help when they know they don’t know. Proceedings of the National Academy of Sciences of the United States of America, 113(13), 34923496. https://doi.org/10.1073/pnas.1515129113.Google Scholar
Gregan-Paxton, J., & John, D. R. (1995). Are young children adaptive decision makers? A study of age differences in information search behavior. Journal of Consumer Research, 21(4), 567580. https://doi.org/10.1086/209419.Google Scholar
Greif, M. L., Kemler Nelson, D. G., Keil, F. C., & Gutierrez, F. (2006). What do children want to know about animals and artifacts? Domain-specific requests for information. Psychological Science, 17(6), 455459. https://doi.org/10.1111/j.1467-9280.2006.01727.x.Google Scholar
Harris, P. L., Koenig, M. A., Corriveau, K. H., & Jaswal, V. K. (2018). Cognitive foundations of learning from testimony. Annual Review of Psychology, 69, 251273. https://doi.org/10.1146/annurev-psych-122216-011710.Google Scholar
Hembacher, E., deMayo, B. & Frank, M. C. (2017). Children’s social referencing reflects sensitivity to graded uncertainty. In Gunzelmann, G, Howes, A., Tenbrink, T., & Davelaar, E (Eds.), CogSci 2017: Proceedings of the 39th Annual Meeting of the Cognitive Science Society (pp. 496500). Cognitive Science Society. https://cogsci.mindmodeling.org/2017/papers/0101/paper0101.pdf.Google Scholar
Henderson, A. M., & Woodward, A. L. (2011). “Let’s work together”: What do infants understand about collaborative goals? Cognition, 121(1), 1221. https://doi.org/10.1016/j.cognition.2011.05.008.Google Scholar
Herwig, J. E. (1982). Effects of age, stimuli, and category recognition factors in children’s inquiry behavior. Journal of Experimental Child Psychology, 33(2), 196206. https://doi.org/10.1016/0022-0965(82)90015-7.Google Scholar
Hickling, A. K., & Wellman, H. M. (2001). The emergence of children’s causal explanations and theories: Evidence from everyday conversation. Developmental Psychology, 37(5), 668683. https://doi.org/10.1037/0012-1649.37.5.668.Google Scholar
Howse, R. B., Best, D. L., & Stone, E. R. (2003). Children’s decision making: The effects of training, reinforcement, and memory aids. Cognitive Development, 18(2), 247268. https://doi.org/10.1016/S0885-2014(03)00023-6.Google Scholar
Jones, A., Swaboda, N., & Ruggeri, A. (2020). Developmental changes in question-asking. In Butler, L., Ronfard, S., & Corriveau, K. (Eds.), The questioning child: Insights from psychology and education (pp. 118143). Cambridge University Press. https://doi.org/10.1017/9781108553803.007.Google Scholar
Kaldy, Z., & Blaser, E. (2013). Red to green or fast to slow? Infants’ visual working memory for “just salient differences.Child Development, 84, 18551862. https://doi.org/10.1111/cdev.12086.Google Scholar
Kampe, K. K., Frith, C. D., & Frith, U. (2003). “Hey John”: Signals conveying communicative intention toward the self activate brain regions associated with “mentalizing,” regardless of modality. Journal of Neuroscience, 23(12), 52585263. https://doi.org/10.1523/JNEUROSCI.23-12-05258.2003.Google Scholar
Kelemen, D., Callanan, M. A., Casler, K., & Pérez-Granados, D. R. (2005). Why things happen: Teleological explanation in parent-child conversations. Developmental Psychology, 41(1), 251264. https://doi.org/10.1037/0012-1649.41.1.251.Google Scholar
Kemler Nelson, D. G., Egan, L. C., & Holt, M. B. (2004). When children ask, “What is it?” what do they want to know about artifacts? Psychological Science, 15(6), 384389. https://doi.org/10.1111/j.0956-7976.2004.00689.x.Google Scholar
Kidd, C., & Hayden, B. Y. (2015). The psychology and neuroscience of curiosity. Neuron, 88(3), 449460. https://doi.org/10.1016/j.neuron.2015.09.010.Google Scholar
Kidd, C., Piantadosi, S. T., & Aslin, R. N. (2012) The Goldilocks effect: Human infants allocate attention to visual sequences that are neither too simple nor too complex. PLoS One, 7(5), Article e36399. https://doi.org/10.1371/journal.pone.0036399.Google Scholar
Kidd, C., Piantadosi, S. T., & Aslin, R. N. (2014). The Goldilocks effect in infant auditory attention. Child Development, 85, 17951804. https://doi.org/10.1111/cdev.12263.Google Scholar
Koenig, M. A., & Harris, P. L. (2005). Preschoolers mistrust ignorant and inaccurate speakers. Child Development, 76(6), 12611277. https://doi.org/10.1111/j.1467-8624.2005.00849.x.Google Scholar
Köksal-Tuncer, Ö., & Sodian, B. (2018). The development of scientific reasoning: Hypothesis testing and argumentation from evidence in young children. Cognitive Development, 48, 135145. https://doi.org/10.1016/j.cogdev.2018.06.011.Google Scholar
Kurkul, K. E., & Corriveau, K. H. (2018). Question, explanation, follow‐up: A mechanism for learning from others? Child Development, 89(1), 280294. https://doi.org/10.1111/cdev.12726.Google Scholar
Kutsuki, A., Egami, S., Ogura, T., Nakagawa, K., Kuroki, M., & Itakura, S. (2007). Developmental changes of referential looks in 7- and 9-month-olds: A transition from dyadic to proto-referential looks. Psychologia, 50(4), 319329. https://doi.org/10.2117/psysoc.2007.319.Google Scholar
Legare, C. H., Mills, C. M., Souza, A. L., Plummer, L. E., & Yasskin, R. (2013). The use of questions as problem-solving strategies during early childhood. Journal of Experimental Child Psychology, 114(1), 6376. https://doi.org/10.1016/j.jecp.2012.07.002.Google Scholar
Loewenstein, G. (1994). The psychology of curiosity: A review and reinterpretation. Psychological Bulletin, 116(1), 7598. https://doi.org/10.1037/0033-2909.116.1.75.Google Scholar
Lucca, K., & Wilbourn, M. P. (2019). The what and the how: Information-seeking pointing gestures facilitate learning labels and functions. Journal of Experimental Child Psychology, 178, 417436. https://doi.org/10.1016/j.jecp.2018.08.003.Google Scholar
Lutz, D. J., & Keil, F. C. (2002). Early understanding of the division of cognitive labor. Child Development, 73(4), 10731084. https://doi.org/10.1111/1467-8624.00458.Google Scholar
Macedo‐Rouet, M., Potocki, A., Scharrer, L., Ros, C., Stadtler, M., Salmerón, L., & Rouet, J.‐F. (2019). How good is this page? Benefits and limits of prompting on adolescents’ evaluation of web information quality. Reading Research Quarterly, 54(3), 299321. https://doi.org/10.1002/rrq.241.Google Scholar
McCormack, T., Frosch, C., Patrick, F., & Lagnado, D. (2015). Temporal and statistical information in causal structure learning. Journal of Experimental Psychology: Learning, Memory, and Cognition, 41(2), 395416. https://doi.org/10.1037/a0038385.Google Scholar
Meder, B., Wu, C. M., Schulz, E., & Ruggeri, A. (2021). Development of directed and random exploration in children. Developmental Science. Advance online publication. https://doi.org/10.1111/desc.13095.Google Scholar
Mills, C. M. (2013). Knowing when to doubt: Developing a critical stance when learning from others. Developmental Psychology, 49(3), 404418. https://doi.org/10.1037/a0029500.Google Scholar
Mosher, F. A., Hornsby, J. R., Bruner, J., & Oliver, R. (1966). On asking questions. In Bruner, J. (Ed.), Studies in cognitive growth (pp. 86102). Wiley.Google Scholar
Muentener, P., & Schulz, L. (2014). Toddlers infer unobserved causes for spontaneous events. Frontiers in Psychology, 5, Article 1496. https://doi.org/10.3389/fpsyg.2014.01496.Google Scholar
Nelson, J. D., Divjak, B., Gudmundsdottir, G., Martignon, L. F., & Meder, B. (2014). Children’s sequential information search is sensitive to environmental probabilities. Cognition, 130(1), 7480. https://doi.org/10.1016/j.cognition.2013.09.007.Google Scholar
Nyhout, A., & Ganea, P. A. (2019). Mature counterfactual reasoning in 4- and 5-year-olds. Cognition, 183, 5766. https://doi.org/10.1016/j.cognition.2018.10.027.Google Scholar
Oakes, L. M. (2010). Using habituation of looking time to assess mental processes in infancy. Journal of Cognition and Development, 11(3), 255268. https://doi.org/10.1080/15248371003699977.Google Scholar
Oakes, L. M. (2017). Sample size, statistical power, and false conclusions in infant looking‐time research. Infancy, 22, 436469. https://doi.org/10.1111/infa.12186.Google Scholar
Palmquist, C. M., & Jaswal, V. K. (2015). Preschoolers’ inferences about pointers and labelers: The modality matters. Cognitive Development, 35, 178185. https://doi.org/10.1016/j.cogdev.2015.06.003.Google Scholar
Pelz, M., & Kidd, C. (2020). The elaboration of exploratory play. Philosophical Transactions of the Royal Society B, 375(1803), 20190503. https://doi.org/10.1098/rstb.2019.0503.Google Scholar
Perez, J., & Feigenson, L. (2020). Violations of expectation trigger infants to search for explanations. PsyArXiv. https://doi.org/10.31234/osf.io/eahjd.Google Scholar
Piaget, J. (1952). The origins of intelligence in children. International University Press.Google Scholar
Ronfard, S., Zambrana, I. M., Hermansen, T. K., & Kelemen, D. (2018). Question-asking in childhood: A review of the literature and a framework for understanding its development. Developmental Review, 49, 101120. https://doi.org/10.1016/j.dr.2018.05.002.Google Scholar
Ruggeri, A., & Feufel, M. (2015). How basic-level objects facilitate question-asking in a categorization task. Frontiers in Psychology, 6, Article 918. https://doi.org/10.3389/fpsyg.2015.00918.Google Scholar
Ruggeri, A., & Lombrozo, T. (2015). Children adapt their questions to achieve efficient search. Cognition, 143, 203216. https://doi.org/10.1016/j.cognition.2015.07.004.Google Scholar
Ruggeri, A., Lombrozo, T., Griffiths, T. L., & Xu, F. (2016). Sources of developmental change in the efficiency of information search. Developmental Psychology, 52(12), 21592173. https://doi.org/10.1037/dev0000240.Google Scholar
Ruggeri, A., Sim, Z. L., & Xu, F. (2017). “Why is Toma late to school again?” Preschoolers identify the most informative questions. Developmental Psychology, 53(9), 16201632. https://doi.org/10.1037/dev0000340.Google Scholar
Ruggeri, A., Swaboda, N., Sim, Z. L., & Gopnik, A. (2019). Shake it baby, but only when needed: Preschoolers adapt their exploratory strategies to the information structure of the task. Cognition, 193, Article 104013. https://doi.org/10.1016/j.cognition.2019.104013.Google Scholar
Ruggeri, A., Walker, C. M., Lombrozo, T., & Gopnik, A. (2021). How to help young children ask better questions? Frontiers in Psychology, 11, Article 2908. https://doi.org/10.3389/fpsyg.2020.586819.Google Scholar
Sabbagh, M. A., & Baldwin, D. A. (2001). Learning words from knowledgeable versus ignorant speakers: Links between preschoolers’ theory of mind and semantic development. Child Development, 72(4), 10541070. https://doi.org/10.1111/1467-8624.00334.Google Scholar
Schulz, E., Pelz, M., Gopnik, A., & Ruggeri, A. (2019). Preschoolers search longer when there is more information to be gained. PsyArXiv. https://doi.org/10.31234/osf.io/5wegk.Google Scholar
Schulz, E., Wu, C. M., Ruggeri, A., & Meder, B. (2019). Searching for rewards like a child means less generalization and more directed exploration. Psychological Science, 30(11), 15611572. https://doi.org/10.1177/0956797619863663.Google Scholar
Schulz, L. E., & Bonawitz, E. B. (2007). Serious fun: Preschoolers engage in more exploratory play when evidence is confounded. Developmental Psychology, 43(4), 10451050. https://doi.org/10.1037/0012-1649.43.4.1045.Google Scholar
Senju, A., & Csibra, G. (2008). Gaze following in human infants depends on communicative signals. Current Biology, 18(9), 668671. https://doi.org/10.1016/j.cub.2008.03.059.Google Scholar
Siegel, M., Pelz, M., Magid, M., Tenenbaum, J. B., & Schulz, L. E. (2021). Children’s exploratory play tracks the discriminability of hypotheses. Nature Communications, 12(3598). https://doi.org/10.1038/s41467-021-23431-2.Google Scholar
Sim, Z. L., & Xu, F. (2017). Learning higher-order generalizations through free play: Evidence from 2- and 3-year-old children. Developmental Psychology, 53(4), 642651. http://dx.doi.org/10.1037/dev0000278.Google Scholar
Stahl, A. E., & Feigenson, L. (2015). Observing the unexpected enhances infants’ learning and exploration. Science, 348(6230), 9194. https://doi.org/10.1126/science.aaa3799.Google Scholar
Stahl, A. E., & Feigenson, L. (2019). Violations of core knowledge shape early learning. Topics in Cognitive Science, 11, 136153. https://doi.org/10.1111/tops.12389.Google Scholar
Stenberg, G. (2009). Selectivity in infant social referencing. Infancy, 14(4), 457473. https://doi.org/10.1080/15250000902994115.Google Scholar
Swaboda, N., Meder, B., & Ruggeri, A. (2020). Finding the (most efficient) way out of a maze is easier than asking (good) questions. PsyArXiv. https://doi.org/10.31234/osf.io/tdaqg.Google Scholar
Vaish, A., Demir, Ö. E., & Baldwin, D. (2011). Thirteen‐ and 18‐month‐old infants recognize when they need referential information. Social Development, 20, 431449. https://doi.org/10.1111/j.1467-9507.2010.00601.x.Google Scholar
van den Bos, W., & Hertwig, R. (2017). Adolescents display distinctive tolerance to ambiguity and to uncertainty during risky decision making. Scientific Reports, 7, Article 40962. https://doi.org/10.1038/srep40962.Google Scholar
VanderBorght, M., & Jaswal, V. K. (2009). Who knows best? Preschoolers sometimes prefer child informants over adult informants. Infant and Child Development, 18, 6171. https://doi.org/10.1002/icd.591.Google Scholar
van Schijndel, T. J., Visser, I., van Bers, B. M., & Raijmakers, M. E. (2015). Preschoolers perform more informative experiments after observing theory-violating evidence. Journal of Experimental Child Psychology, 131, 104119. http://doi.org/10.1016/j.jecp.2014.11.008.Google Scholar
Vygotsky, L. S. (1987). Problems of general psychology. In Rieber, R. & Carton, A. (Eds.), The collected works of L. S. Vygotsky. Plenum Press. https://doi.org/10.1007/978-1-4613-1655-8.Google Scholar
Walden, T., Kim, G., McCoy, C., & Karrass, J. (2007). Do you believe in magic? Infants’ social looking during violations of expectations. Developmental Science, 10, 654663. https://doi.org/10.1111/j.1467-7687.2007.00607.x.Google Scholar
Wu, Y., & Gweon, H. (2019). Preschool-aged children jointly consider others’ emotional expressions and prior knowledge to decide when to explore. PsyArXiv. https://doi.org/10.1111/cdev.13585.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×