Top Banner
How learning biases and cultural transmission structure language: Iterated learning in Bayesian agents and human subjects. Vanessa Ferdinand Research Master Cognitive Science Institute for Interdisciplinary Studies University of Amsterdam August 2008
63

How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

Oct 17, 2020

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

How learning biases and cultural transmissionstructure language:

Iterated learning in Bayesian agents and humansubjects.

Vanessa Ferdinand

Research Master Cognitive ScienceInstitute for Interdisciplinary Studies

University of Amsterdam

August 2008

Page 2: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

2

Abstract

What is the mechanism that translates the individual properties of learners into theproperties of the language they speak? This thesis will investigate culturaltransmission as this mechanism and will take up the Iterated Learning Model as aformal framework in which to address this claim. This model describes language as aspecial learning problem, where the output of one generation is the input for the next.Previous research has shown that universal properties of human language emergefrom the process of cultural transmission. However, particular biases are alsonecessary to obtain these properties, and the exact interplay between individual biasesand cultural transmission is still an open question. In the present research, acomputational, Bayesian iterated learning model is constructed to analyze therelationship between learning biases and what additional structure culturaltransmission adds to language. An iterated learning experiment with human subjectsis also conducted, to obtain a better understanding of the model’s results for humanlearners. Many new insights are gained, which attest to the merits of combiningcomputational and experimental iterated learning models to explain the properties oflanguage.

Page 3: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

3

Table of Contents

Abstract 2

Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.1 The Iterated Learning Model . . . . . . . . . . . . . . . . . . . . . . 71.2 Bayesian Iterated Learning Models . . . . . . . . . . . . . . . . . . . 10

Chapter 2 A Bayesian Iterated Learning Model . . . . . . . . . . . . . . . . . . . . . 13

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132.2 Model Description . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.2.1 The Hypotheses . . . . . . . . . . . . . . . . . . . . . 132.2.2 The Data . . . . . . . . . . . . . . . . . . . . . . . . 142.2.3 The Prior . . . . . . . . . . . . . . . . . . . . . . . . 142.2.4 The Social Structure. . . . . . . . . . . . . . . . . . . 152.2.5 Bayesian Inference . . . . . . . . . . . . . . . . . . . 152.2.6 Data Production . . . . . . . . . . . . . . . . . . . . . 162.2.7 Iteration . . . . . . . . . . . . . . . . . . . . . . . . . 172.2.8 Model Parameters . . . . . . . . . . . . . . . . . . . . 17

2.3 Model Analyses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182.3.1 Overview of Assessment Methods . . . . . . . . . . . 182.3.2 Q Matrix Calculations . . . . . . . . . . . . . . . . . . 182.3.3 Stationary Distribution Calculations . . . . . . . . . . 232.3.4 Summary. . . . . . . . . . . . . . . . . . . . . . . . . 24

2.4 Model Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242.4.1 Basic Sampler Behavior . . . . . . . . . . . . . . . . . 242.4.2 Basic MAP Behavior . . . . . . . . . . . . . . . . . . 242.4.3 The Bottleneck Effect . . . . . . . . . . . . . . . . . . 282.4.4 Population Size . . . . . . . . . . . . . . . . . . . . . 312.4.5 Heterogeneity . . . . . . . . . . . . . . . . . . . . . . 35

2.5 Model Discussion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

Chapter 3 An Experiment in Iterated Function Learning . . . . . . . . . . . . . . . . 43

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 433.2 Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

3.1.1 Participants . . . . . . . . . . . . . . . . . . . . . . . . 473.1.2 Apparatus and Stimuli . . . . . . . . . . . . . . . . . . 473.1.3 Procedure . . . . . . . . . . . . . . . . . . . . . . . . . 483.1.4 Data Collection and Analyses . . . . . . . . . . . . . . 49

3.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 493.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

Page 4: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

4

Chapter 4 General Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

Appendix A Model Code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

Appendix B Experiment Instructions . . . . . . . . . . . . . . . . . . . . . . . . 58

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

Page 5: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

5

Chapter 1

Introduction

The language that you speak is not a product of your mind alone. As language istransmitted from person to person and generation to generation, it adapts to the mindsit propagates through, and they adapt to it. This makes the evolution of language aspecial problem, because the output of one learner is the input for the next.

The field of linguistics has made great advances in describing human language.Through the description of language universals and from animal comparative studies,we have a good picture of what human language is, and what it is not. Influentialparadigms in 20th century linguistics, such as the generative program, concentrate onin-depth studies of particular languages or the variations and constraints on variationfound in the world’s languages, to infer what the innate biases of human learners mustbe. Notwithstanding fundamental differences, most of these research programs havetended to ignore the socio-cultural and historical dimensions of language.Additionally they fail to provide an account of how the innate biases of individualstranslate to the universals witnessed in the world’s languages. This problem oflinkage can be overcome by identifying the mechanism which translates the propertiesof individual learners into the properties of human language (Kirby, 1999). Thistheses will argue in favor of claims that cultural transmission itself, may indeed bethis mechanism.

The crucial next step, where linguistics has arguably made much less progress, is toprovide a mechanistic explanation of why language is the way that it is, and not someother way. This explanation necessitates a description at a level below languageitself: What are the constraints that shape language? Where do they come from andhow do they interact?

The constraints of language arise from two systems: the embodied cognitive agent andthe socio-cultural system in which these agents communicate with one another. Thefirst is the domain of cognitive science and psycholinguistics. Here, the mainconstraints lie in perception, processing and representation, production. How is thedata constrained as it enters the cognitive system, how is it cognitively processed, andhow is it constrained as utterances are produced? Some examples of perceptual biasesare purely physical and constrained by the human senses, such as the range of soundsone can perceive. Others form with cognitive development, such as the phenomena ofcategorical perception (Lieberman et al., 1967; Kuhl, 2004). Likewise, productionbiases are constrained by the physical limitations of human anatomy, such as thefrequency range of vocalizations and the degree of motor control we have over ourvocal tracts. The processing which mediates what is perceived and what is ultimatelyproduced includes high-level cognitive processes such as reasoning, induction, and

Page 6: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

6

learning, each of which come with their own biases. These processes may also besubject to constraints on how linguistic knowledge is represented in the brain; whatkind of representations are possible in a network of neurons, and what kinds are not?In short, the constraints which shape a cognitive agent’s production of language areboth shaped by both biological evolution and individual development, which includeslearning from one’s environment.

The second system, how cognitive agents interact, has been pioneered bycomputational modeling and mathematics. The social structure characterizes how theproduction and perception components of the cognitive agent link up. Particular typesof social structures involve different constraints on what kind of access the agentshave to the external data that constitutes language. For example, a population with nogenerational turnover (i.e. no agents are born or die) would conceivably have a verydifferent language than language as we know it. Or for a more intuitive example, ifthe future of the English language becomes confined to nothing but emailcommunication, its developmental trajectory would be very different than if itremained a spoken language. Therefore, if we want to explain why this new “emailEnglish” is the way that it is, and not some other way (like the old spoken English),we would have to describe the constraints of email English – in terms of theconstraints of its social system (the network of computers and how this shapes humaninteraction) and in terms of the cognitive constraints which the new system engages(such as production biases associated with typing).

In this light, language is a complex, dynamical system in its own right. This meansthat the behavior of the system is a product of both its components (the embodiedcognitive agents) and how they interact (their social system). However, as stated insystems theory, no systems have true boundaries, and the borders we impose when westudy them are purely artificial constructs (Weisbuch, 1991). There are multiple waysto carve up the systems and their constraints in order to guide our search. The mostcommon delineation, among those who computationally model language evolution, isthat language sits at the crux of 3 complex, dynamical systems: biological evolution,cultural transmission, and individual learning. (Christiansen & Kirby, 2003). Thistripartite division of these separate, but interacting, systems is misleading because itimplies that evolution acts directly on learning as an adaptive system. This viewessentially deletes cognition from the picture, because it is the embodied cognitiveagent that ultimately roots its high-level process of language induction within thebiologically evolved wet-ware that is the true processor of language.

By viewing language as a product of cognitive agents and the cultural transmissionsystem which propagates it, we would expect the constraints of language to be rootedin these two aspects. However, organizing the problem in this way has the side effectof losing any direct linguistic consequences of biological evolution within theembodied cognitive agent, and rightly so. Undoubtedly, the biological endowmentwhich makes us human places hard constraints on the possibilities of ontogeneticdevelopment. But the structure of cultural transmission is in the position to placeadditional constraints on this biological potential, father defining language into itsultimate form, as we witness it in the world. The fact that human language isculturally transmitted is just as universal to our language system as is the sharedgenetics which makes all humans, human. Though logically, this places both as

Page 7: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

7

candidates for the explanatory burden of language universals, the most informativeexplanation will be the one that cuts language the closest.

So how do the properties of cognitive agents determine the properties of the languagethey speak, and what does cultural transmission add to this explanation? A good wayto proceed with this question is to create a formal framework for testing hypothesesabout how cultural transmission mechanistically translates the properties of cognitiveagents into the properties of human language, and whether or not the dynamics of thiscultural transmission place additional constraints on the ultimate form of language.This thesis will take up one such framework, the iterated learning model, in order toformally address the socio-cultural constraints on human language.

1.1 The Iterated Learning Model

The iterated learning model (ILM) was first formalized for the study of languageevolution by Kirby (1998) and provides a framework for the empirical study ofcultural transmission and how it effects the information being transmitted. ILMs canbe implemented in a variety of ways, but they all contain these fundamentalcomponents:

1) A learning algorithm2) Some form of information which is the input/output of the algorithm3) Structured transmission of the information, where the output of one learner

serves as the input for the next.Some learning algorithms commonly used in ILMs are symbolic grammar inductionalgorithms (Brighton & Kirby, 2001), neural networks (Smith, 2002), Bayesian agents(Kalish et al., 2007), and even human subjects (Cornish, 2006; Griffiths et al., 2006).The data can be linguistic input or numerical values and the transmission format couldbe any conceivable social structure, but is commonly kept to a parent-child chain foranalytical ease.

Possibly, the first study of this kind was Bartlett’s (1932) psychological experiment in“serial reproduction”. A subject would be shown a picture, for example a nice sketchof an elk, and then be asked to re-draw it from memory. Then, this copy would begiven to another subject to re-draw, and so on. Over the course of this serialreproduction, the information present in the elk would change. The shading woulddisappear, the complexity of the antlers would diminish, until all that was left was theoutline of a cat. Although this is a nice illustration that information can be shaped bythe very process of its transmission, Bartlett’s stimuli, pictures and stories, were notcontrolled and therefore do not lend themselves well to empirical study.

The first computational ILMs were developed by Hare & Elman (1995), Batali(1998), and Kirby (1998) as computer programs of agents in a simulated population.Here, agents were simple language-learning algorithms that paired meanings withstrings of letters, and one agent would learn its language from another. Their result isthat the signal-meaning system became increasingly regular as it passed through moreand more agents. The regular structure which emerged was also compositional, wherespecific letters or letter combinations designated specific parts of the overall meaning,as words do in human language. However, these effects only occurred when therewas a transmission bottleneck. This means that the agents cannot pass their language

Page 8: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

8

on in totality to the next generation. Humans have an infinite capacity for linguisticexpression, however, we can only express a finite amount of linguistic utterances.The transmission bottleneck mirrors this by limiting the number of signal productionsto below the number of possible meanings in the meaning space. Only under aspecific range of bottleneck size do regularity and compositionality emerge. ManyILM studies which followed these, each using a different learning algorithm anddifferent assumptions regarding the signal-meaning spaces, consistently reported thesame result; the emergence of regularity and compositionality due to the learningbottleneck (Brighton, 2002; Hurford, 2000; Kirby, 2000; Smith, 2003; Vogt, 2003).

For a concrete example of regularity emerging from a bottleneck, we can look at theEnglish past tense, which has both regular (verb+ed) and irregular (go – went) pasttense rules. A regular rule is also a general rule, which is applied every time alanguage learner uses the past tense of a regular rule. Irregular rules, on the otherhand, have to be learned one by one, when the learner comes into contact with theirregular verb. Looking at regular and irregular rules separately, regular rules have amuch higher chance of being transmitted to the next generation when the bottleneck issmall, because they apply to more verbs and therefore have a higher chance of beingproduced. Irregular verbs, on the other hand, can only survive over the generationswhen the verbs they apply to are high frequency verbs (Kirby, 2001). In fact, this isexactly the case with the English past tense; the top 10? most frequent verbs are allirregular. Additionally, it is well-documented historically that low-frequencyirregular verbs in English are gradually adopting the regular rule (Lieberman, et al.,2007).

The ILM research demonstrates that, through cultural transmission and the constraintimposed by the bottleneck, the information in language compresses in a self-organizing way (Brighton et al., 2005). Additionally, the language itself adapts tobecome learnable by the agents which transmit it, and not the other way around(Zuidema, 2003). Agents can only produce what they were able to learn, and when allagents in the population are similar, this makes the task easier for the next agent in thetransmission line. Some of the hard claims of ILM proponents are that culturaltransmission inevitably leads to regularization, an increase in learnability, andcompositionality. In most ILM implementations, no biological criterion of fitness isimposed which selects agents according to the goodness of their language use. Thus,regularization, learnability, and compositionality are all claimed as properties oflinguistic evolution, and not biological evolution.

The fact that diverse learning algorithms all produce similar results when iterated,shows that these results are most likely due to the properties of the iteration, and thebottleneck effect, rather than to something inherent in the learning algorithms.However, every learning algorithm has its bias and it is still possible that all of thelearning algorithms that were used do share some bias which allows for theemergence of regularity and compositionality. It is possible that some learningalgorithms are structured in such a way that they cannot support compositionalbehavior. In this sense, the bias of a learning algorithm defines what behaviors it canand cannot yield, as well as what behaviors its structure encourages. Smith (2003)carried out a comparative study of the ILM algorithms and determined that they doshare two basic biases: a bias toward one-to-one signal-meaning mappings and a biastoward exploiting regularities in the input data. Therefore, these two biases can be

Page 9: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

9

seen as two components of the learning algorithm’s structure which are necessary forthe algorithm to display compositionality. And indeed, these are two biases whichhuman learners likely bring to the task of language induction (Pinker, 1984).

This raises the question of how much the of outcome of iterated learning isdetermined by cultural transmission and how much is determined by the biases. Onthe one hand, if the process of cultural transmission completely determines theoutcome of iterated learning, we could expect to see the same results for learningalgorithms which have nothing in common. Additionally, we might even expect theseproperties to hold for data compression algorithms which have no plausibilitywhatsoever as a cognitive model of language learning. Take, for example, this toymodel of an interpolation algorithm which transmits data over a bottleneck of 5 datapoints (Figure 2). Even here, the function which describes the data increases inregularity and stability with more and more iterations. However, compositionalitywas not obtained this model. It is even hard to say what compositionality would looklike in terms of this model’s capabilities. Clearly, all ILMs do not universally yieldcompositionality. Therefore, we still must need a certain type of bias to obtaincompositionally through iterated learning.

Figure 2A toy ILM I constructed with a linear interpolation learning algorithm. An initial function is randomlygenerated. 5 data points, randomly selected from the initial function, serve as input to the firstgeneration. The agent at generation 1 describes these data points with an interpolation function. Next,5 data points are randomly selected from generation 1’s function to serve as input for the nextgeneration, and so on. The function used to describe 5 data points becomes less complex as it isiterated, and will probably stabilize as a linear function.

Unfortunately, in this interpolation model, as with many other learning algorithms, itis difficult to assess exactly what its biases are. What is needed, then, is a model withan explicitly-coded learning bias, so that different outcomes of iterated learning canbe attributed to specific manipulations of the bias. Fortunately, Bayesian statisticsprovides this framework.

Page 10: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

10

1.2 Bayesian Iterated Learning Models

For the readers who are unfamiliar with Bayesian statistics, I will introduce this topicwith a practical example:

Picture yourself walking down a street in Amsterdam. Someone bikes past you andyou catch a half-second clip of their voice. What language were they speaking? Tocome to a conclusion, a Bayesian-rational person would take into account threethings. First, the candidate languages. For simplicity’s sake let’s just say you havethree hypotheses: Dutch, Arabic, and English. Second, the data: a nice velar fricative.Third, the prior probability: what is the chance someone in that neighborhood wouldbe speaking any of those three languages? The likelihood that Dutch or Arabic wouldproduce a velar fricative is astronomically higher than for English. However, if youare anywhere near the tourist information center, you may just as well conclude thanan English-speaker was clearing their throat. Likewise, knowing if you were in aDutch or a Moroccan neighborhood would break the tie in the data likelihood of thevelar fricative. Additionally, the prior knowledge each person brings to an inductiveproblem can be different. If you happened to be one of those people still in line at thetourist information center, you might think that everyone in Amsterdam speaks Dutch,and therefore you would probably classify most people as Dutch-speakers at firstsound-byte.

By combining your knowledge of the data’s likelihood with the prior probability ofeach hypothesis, you will come to a solution. This solution is the posterior probabilityof each hypothesis now, after you have finished reasoning. Last, you will select youranswer in light of these posterior probabilities, choosing the hypothesis with thehighest posterior probability, if you’re smart.

Thus, the components of a Bayesian inference algorithm are:1) The hypotheses and the data likelihoods which accompany them2) The prior probability of each hypothesis: the bias3) The posterior probability of each hypothesis

As you can see, this is no longer a problem specific to language learning. Theinvestigation of iterated learning in terms of Bayesian agents brings the question ofwhich adds more, biases or cultural transmission, to a new, abstract level. Griffiths &Kalish (2005) were the first to use a Bayesian ILM to address this debate and theyfound that the outcome of iterated learning was completely determined by the priorprobability of each hypothesis. Here, this outcome is represented by the proportionthat each hypothesis was chosen over the course of the ILM when run to infinity.Clearly, this outcome of iterated learning must be determined analytically. Thisresulting distribution of hypotheses choices constitutes a stationary distribution, whichrepresents the outcome of iterated learning (Nowak et al., 2001).

The Griffiths and Kalish result showed that the stationary distribution over hypothesesexactly mirrored the prior probabilities of those hypotheses, regardless of specificprior distributions or other parameter manipulations. In particular, manipulating thebottleneck parameter had no effect whatsoever on the stationary distribution. Withthis, they determined that cultural transmission does not make an independentcontribution to the outcome of iterated learning and it is merely a vehicle which

Page 11: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

11

reveals the inductive bias of the learners. However, this result doesn’t make muchsense given that the bottleneck effect is robust in the many previous ILM simulations.

To counter this claim, Kirby et al. (2007) showed that this result was a consequenceof the particular hypothesis choice strategy that was implemented; sampling. Anagent that samples randomly chooses a hypothesis, weighted by the posteriorprobability of each hypothesis. This is known as probability matching in thepsychological literature. Conversely, Kirby et al. showed that a Bayesian ILM doesnot converge to the prior when agents are maximizers, who always choose thehypothesis with the highest posterior probability. Thus, the main question seemed tobe whether humans are maximizers or samplers. So, Smith & Kirby (2008) extendedtheir model to include biological evolution and showed that the maximizing strategyis evolutionarily stable over sampling. They concluded that natural selection favorsagents whose behavior can be affected by cultural transmission, so that culturaltransmission is the primary determiner of linguistic structure. They also asserted thatreal human behavior probably lies somewhere on a continuum between maximizingand sampling, and should be subject to a more fine-grained analysis.

At first glance, the initial Griffiths & Kalish results could be understood as confirminglinguistic nativism; that the ultimate structure of language is determined by our innatebiases and nothing else. However, the prior probability in the Bayesian model doesnot correspond only to the learner’s innate bias. In this simplistic model of acognitive agent, the prior represents all properties of the inductive task besides thedata itself. Therefore, the prior is everything the agent brings with it to the task; itsinnate biases, its learned biases, previous domain-specific experience, and even itsaffective state at the moment of induction.

In light of their own findings, Griffiths & Kalish also propose that ILMs using humansubjects can serve as a tool for revealing inductive biases, especially in cases whereresearchers have little a priori knowledge about what these biases might be (2006).They support their claim in two different experimental tasks where the associatedinductive biases are well-established by previous psychological experimentation. Inboth of these experiments, one in category learning (2006) and another in functionlearning (2007), the known inductive bias was revealed through iterated learning.However, this method should not be understood as a way to reveal innate biases, forthe same reason the prior, as characterized by Bayesian induction, should not be seenas representing only the innate bias. The biases which are revealed by human ILMsare likely to be task-specific, variable with training, and could be subject to primingand context manipulation.

In this thesis, I will construct my own Bayesian ILM in order to investigate thedifferent claims about how biases determine the outcome of iterated learning and whatcultural transmission adds to this outcome. Using the insights I gain from themodeling work, I will inform a hypothesis about human iterated learning behavior andtest this hypothesis within an experimental, function learning ILM with humansubjects.

Chapter 2 presents my implementation of a Bayesian ILM. This model willinvestigate the differential behaviors of maximizers and samplers under identicalconditions, including how each responds to particular parameter manipulations

Page 12: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

12

regarding biases, data likelihoods, population size, and heterogeneity. Chapter 3 willpresent an experimental ILM with human subjects. First, it will introduce the humansubjects ILM framework and describe one such experiment from Kalish et al. (2007).Three small experiments will be presented. One which replicates the originalexperiment of Kalish et al. (2007), one which contains a novel manipulation tosubject’s perception of the task, and one which tests a population of subjects with ahigh degree of mathematical training, and therefore, arguably different biases.Finally, Chapter 4 provides a general conclusion.

Lastly, I would like to add a note about the methodology used in the model analyses.Since my educational background is in Anthropology and Cognitive Science, I havechosen to approach this model with an empirical, rather than an analytical, standpoint. By empirically dissecting this model, I am able to provide some deeperinsights into the inner dynamics of the Bayesian ILM than some analytical dissectsallow. Some of the dynamics I have chosen to explore are simply invisible tomathematical descriptions that focus on the limits of model behavior and thecumulative end states of iterated learning when extrapolated to infinity. With thismethodology, I will attempt to draw a more complete picture of the mechanismswhich drive the model’s behavior. Many aspects of the model I will describe in thefollowing chapter certainly have straightforward analytical solutions which I have notentertained, however my goal here is to set forth a bridge between empirical researchon iterated learning systems and their analytical description. Hopefully, this thesiswill be equally informative for cognitive scientists and mathematicians alike, whomay want to continue the work I set forth here.

Page 13: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

13

Chapter 2

A Bayesian Iterated Learning Model

2.1 Introduction

In this chapter, I will describe the implementation and results of my own model ofiterated learning with Bayesian agents. Here, two models are constructed; one withagents who choose their hypothesis by sampling and one with agents who choose bymaximizing. Agents use Bayesian inference to produce and induce from data, whichis passed between agents across discrete, serially-organized generations. A variety ofparameter settings and their effect on the model’s behavior will be investigated. Thisinvestigation both replicates recent Bayesian ILM results and addresses newhypotheses regarding population size and heterogeneity.

Section 2.2, Model Description, will outline the components and structure of themodel and describe the parameters which will be manipulated. Section 2.3, ModelAnalyses, will describe the analytical tools commonly used in the existing BayesianILM literature, to assess model behavior. Here, I will also justify the use of severalapproximations for these solutions, which are obtained from the model simulations.These experimentally-obtained assessment tools will serve as the basis for thisresearch’s model analyses. Section 2.4, Model Results, will describe both the samplerand maximizer model’s behavior for a number of parameter manipulations regardingthe prior and likelihoods, the bottleneck effect, population size, and heterogeneity ofpriors. In conclusion, section 2.5 will provide a general discussion of the modelingresults.

2.2 Model Description

In this section I will outline the components of a Bayesian ILM and describe how theyare implemented in this model. The implementation and simulations were all carriedout in Matlab and the model’s code (Appendix A) was developed jointly with JelleZuidema.

2.2.1 The HypothesesIn this simulation, agents are considered to have a small set of hypotheses about thestate of the world, and each of these hypotheses assign different likelihoods to each ofa small set of observations that the agent can make about the state of the world. Thesehypotheses could represent, for instance, different languages that generate a set ofutterances, or different functions that describe a set of data points. However, theexact nature of the hypotheses is left under specified, in order to investigate the

Page 14: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

14

general dynamics inherent to Bayesian iterated learning. Thus, the basic properties ofthe model might be generalizable to a variety of systems where information isculturally transmitted, such as language and function learning, where Bayesianinference serves as a good approximation of the learning mechanism involved. In thismodel, the hypotheses are set at the beginning of each simulation and all agents havethis set of specified hypotheses. For simplicity of analysis, each hypothesis iscompletely defined by the likelihoods it assigns to each observation. Additionally, thenumber of hypotheses will be restricted to three, and each called H1, H2, and H3(Figure 2.1). Also, any particular combination of these three hypotheses will bereferred to as the “hypotheses structure.”

2.2.2 The DataThe observations that the agent can make about the state of the world will be referredto as data points. These will also be restricted to three and called d1, d2, and d3(Figure 2.1). The information that the agents pass between each other is a vector ofone or more of these three data points. As will be described in section 2.2.4, thenumber of data points in this vector defines the “transmission bottleneck.”

Figure 2.1Graph of hypotheses [.6 .3 .1; .2 .6 .2; .1 .3 .6] 1 and example prior vector [.7 .2 .1]. Each hypothesis’shape is entirely determined by the likelihoods it assigns to each data point.

2.2.3 The PriorThe prior probability of each hypothesis is stored in a 3-unit vector, where each entrylists the prior of each hypothesis. The shorthand for and example prior is [.7 .2 .1],showing the prior of H1, H2, and H3 respectively. The difference between thehighest and lowest probability create the bias strength. In this example, the biasstrongly favors H1. These probabilities of the prior vector sum to one, indicating thatthese are the only three hypotheses which can generate or account for the data.

1Hypotheses structure will be written in the shorthand above. The first set of three entries are thelikelihoods of data points 1, 2, and 3 according to H1. The next two sets correspond to H2 and H3.

Page 15: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

15

2.2.4 The Social StructureIn this model, agents are defined by the process of Bayesian induction from data,hypothesis choice, and data generation. Agents are organized into discretegenerations of one or more. When each generation consists of 1 agent, the simulationcan be characterized as a Markov chain and is identical to previous ILMs where oneadult transmits data to one child. When each generation consists of x agents > 1, eachagent will output an equal number of data points into the data vector, and this entirevector will serve as the input to each of the agents in the next generation.

2.2.5 Bayesian InferenceAgents both induce from data and produce data according to the likelihood values oftheir hypotheses. The particular likelihood values of one hypothesis determines thecomposition of the data string it is likely to produce. For example, H1 will produced1 70% of the time, d2 20% of the time, and d3 10% of the time. Therefore, acharacteristic, 10-sample data string for each hypothesis in figure 2.1 might look like:H1 _ [1 1 1 2 1 2 2 1 3 1]H2 _ [1 2 2 3 2 1 2 2 2 3]H3 _ [3 3 1 2 2 3 3 3 3 2]

When faced with a data string, such as one above, agents use Bayesian inference todecide which hypothesis was most likely to have produced it. Thus, agents useBayes’ Rule (eq. 2.1) to compute the probability that each hypothesis generated thedata string:

Equation 2.1

Here, P(h|d) denotes the posterior probability that a hypotheses could have generatedthe data in question. This is the outcome of Bayesian induction and is calculated foreach hypothesis. P(d|h) is the likelihood value of the data under the hypotheses inquestion. The data likelihood values for each hypothesis are defined by thehypothesis structure (Figure 2.1). P(h) is the prior probability of a hypothesis. P(d) isthe probability of the data averaged over all hypotheses.

Page 16: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

16

This method of calculating the posterior yields the normalized product of thelikelihood and prior and is equivalent to Bayes’ rule, above.

2.2.6 Data ProductionThe next step is for the agent to output a new data string. First, a hypothesis is chosenaccording to the posterior probabilities. Second, the data are generated from thechosen hypothesis.

Hypothesis choice - Maximizing vs. Sampling:There are a variety of ways in which the hypothesis could be chosen, however in thisstudy I will investigate two cognitively-grounded strategies: maximizing andsampling. Both of these strategies choose between hypotheses according to theirposterior probabilities. The maximizer simply chooses the hypothesis with thehighest posterior probability. But in the event there is a tie among hypotheses for thehighest posterior value, the maximizer randomly chooses between them. The samplerchooses one hypothesis randomly, but weighed by the posterior probabilities.

Example Posterior VectorH1 H2 H3

0.12 0.27 0.61

Table 2.1

According to the posterior values in table 2.1, the maximizer will choose H3. Thesampler will have a 12% chance of choosing H1, a 27% chance of choosing H2, and a61% of choosing H3. These different strategies are implemented separately, creating

Example calculation as implemented in the program:

hypotheses = [.6 .3 .1; .2 .6 .2; .1 .3 .6], prior = [.7 .2 .1], data string = [2 3 3]

1) Calculate likelihoodFor data point = 2, the corresponding likelihood values under each hypothesis H1, H2,H3 = [.3 .6 .3]. For data point = 3, the likelihoods are [.1 .2 .6]. Assumingindependence, the likelihoods of each element in the data string can be multiplied, toyield the likelihood of the string. Instead of multiplying the probabilities, the log ofthe likelihoods are added, to make it easier to deal with small numbers. Therefore, thelog likelihood of the data string [2 3 3], is calculated as:log [.3 .6 .3] + log [.1 .2 .6] + log [.1 .2 .6] = [-5.8091 -3.7297 -2.2256].

2) Calculate posteriorposterior = exp ( log prior + log likelihood )

First, the log of the prior vector is added to the log likelihood vector of the data string:exp ( [-0.3567 -1.6094 -2.3026] + [-5.8091 -3.7297 -2.2256] ) = [0.0021 0.00480.0108] Last, the posterior vector is normalized, to obtain a probability of 1 that oneof the three hypotheses generated the data. This yields: [0.1186 0.2712 0.6102].This posterior means that H3 is most likely (61%) to have generated the data string.The next likely is H2 at 27% and the least likely is H1 at 12%.

Page 17: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

17

two Bayesian iterated learning models which differ only in the respect of hypothesischoice. This leads to characteristic differences in the dynamics of each model, whichwill be addressed in the Analysis section.

Data choice:Data is generated from the chosen hypothesis according to the likelihood values ofthat hypothesis. Assuming the agent has chosen H3, each data point in the outputstring will be randomly generated, but weighted according to the likelihood of eachdata point under H3. Therefore, given the likelihood values of H3 = [.1 .3 .6], datapoint 1 has a 10% chance of being generated, data point 2 a 30% chance, and datapoint 3 a 60% chance. The next 3-value data string might look something like this: [32 3].

2.2.7 IterationCultural transmission is modeled by using each generation’s output data string as thenext generation’s input data string. All agents in one generation produce the samenumber of data samples, which are all concatenated into the output data string for thatgeneration. The likelihood of a data string is invariant to the order of the data samplesit contains. Each agent has no way of knowing the number of agents which producedthe data string or which data came from which agent. Additionally, each generationhas an identical composition of agents as the generation before it.

2.2.8 Model ParametersA variety of parameters can be manipulated to investigate the dynamics of the system.These manipulations will be used to compare and contrast the dynamics specific tothe Maximizer (MAP – maximum a posteriori) and Sampler models. The followingmanipulations that will be investigated in the present research are as follows:

1) The prior.2) Homogeneity and heterogeneity of the agents’ priors. Each agent in a

population greater than 1 can be assigned a different set of priors. This is theonly parameter which can be manipulated heterogeneously in the population.The remaining manipulations below hold for all agents in the simulation.

3) The hypotheses structure (the likelihood values of each hypothesis).4) The bottleneck. How many data samples each generation produces.5) Population size. Usually kept to 1 in the homogenous simulations and 2 in the

heterogeneous simulations.

Page 18: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

18

2.3 Model Analyses

2.3.1 Overview of Assessment MethodsEach model has a unique dynamical fingerprint. Understanding why two modelswork differently is understanding how their dynamics differ. Each parametermanipulation can potentially change the dynamics of the model, and depending on theproperties of the model, certain manipulations can change the dynamics in a different,but systematic way. Therefore, in order to characterize each model’s dynamicalfingerprint, we are looking for features that are invariant to specific parametermanipulations as well as changes in the dynamics that can be causally attributed tospecific changes in parameter settings.

A concrete representation of a “dynamical fingerprint” can be obtained byconstructing a transition matrix, or Q matrix (Nowak et al., 2001), for each model.This matrix gives the probabilities that each hypotheses will lead to itself or any otherhypotheses in the next generation. In essence, all probable trajectories that an ILMmight take are wrapped up in this matrix. From the Q matrix, we can also derive thestationary distribution, which is the stable outcome of iterated learning (Griffiths &Kalish, 2005; Kirby et al., 2007).

In the following sections, both the Q matrix and stationary distribution will beexplained in detail, for readers who may be unfamiliar with these terms. Additionally,I will justify the use of certain experimental approximations of these two analyticaltools. These approximation heuristics are readily obtainable from iterated learningsimulations and are especially valuable when the computational requirements of theanalytical solutions is high or simply not feasible.

The next section will walk through the analytical calculation of a couple Q matrices,as applied to the iterated learning model. Because the Q matrix defines the model’sdynamics, it is important to note, during the calculation process, how each modelcomponent comes into play. These seemingly minute details will have importantconsequences for understanding the mechanism behind the dynamics in later analyses.

2.3.2 Analytical and Experimental Q Matrix CalculationsIf the agent in one generation has hypothesis 1, then what’s the probability that theagent in the next generation will have hypothesis 1, 2, or 3? These probabilities aredisplayed in the transition matrix (or Q matrix). In the example Q matrix below(Table 2.2), when an (parent) agent in one generation produces data from H1, then theprobability that that data will lead the (child) agent of the next generation to chooseH1 is 80%. Since parent H1 can produce data that best supports H2 or H3, then“miscommunications” occur, leading the child to induce H2 or H3 each 10% of thetime.

Page 19: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

19

Example Q MatrixQ matrix child

H1 H2 H3

H1 0.8 0.1 0.1parent H2 0.1 0.8 0.1

H3 0.1 0.1 0.8

Table 2.2

Analytical Q matrix for Sampler with bottleneck of 1:The following will show the analytical calculation of the Q matrix for a Samplermodel with a bottleneck of 1 data sample per generation. All calculations in thissection will use the following prior and data likelihood values:

Data Likelihoods

Priors data 1 data 2 data 3

hypothesis 1 0.7 hypothesis 1 0.8 0.1 0.1hypothesis 2 0.2 hypothesis 2 0.1 0.8 0.1hypothesis 3 0.1 hypothesis 3 0.1 0.1 0.8

Table 2.3Prior and data likelihood values used for all calculations in section 2.2.2

Beginning with cell (H1, H1), we want to know how often parent H1 will produceeach possible data string, and how often each of those data strings will lead to thechild choosing H1. For a bottleneck of 1, there are just three possible data strings: [1][2] and [3]. As defined by the data likelihood values of each hypothesis, a parent withH1 will produce d1 with p = 0.8, d2 with p = 0.1, and d3 with p = 0.1. Next, theprobability that the child will choose H1 from each of the three data points is definedby the child’s computed posteriors (table 2.4) and their hypothesis choice strategy,sampling. All posterior probabilities are computed with Bayes’ rule as outlined insection 2.1.4.

Posterior ValuesH1 H2 H3

[1] 0.9492 0.0339 0.0169[2] 0.2917 0.6667 0.0417[3] 0.4118 0.1176 0.4706

Table 2.4

Therefore, when the sampler receives data string [1], it will choose H1 with p = .95,H2 with p = .03, and H3 with p = 0.02. To find out how often parent H1 will lead tochild H1, we must multiply the probability that each data string leads to child H1 bythe likelihood of that data string being generated by parent H1. Thus, the probabilityof parent H1 leading to child H1 is (0.9492*0.8) + (0.2917*0.1) + (0.4118*0.1) =0.8297

Page 20: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

20

Q matrix child

H1 H2 H3

H1 0.8297 0.1056 0.0648parent H2 0.3695 0.5485 0.0821

H3 0.4535 0.1641 0.3823

Table 2.5Analytically-calculated Q-matrix for Sampler with bottleneck of 1

Q matrix child

H1 H2 H3

H1 0.8231 0.1102 0.0667

parent H2 0.3685 0.5422 0.0893

H3 0.4539 0.1665 0.3796

Table 2.6Experimentally-calculated Q-matrix for Sampler with bottleneck of 1

Experimental Q matrix for Sampler with bottleneck of 1:For comparison, table 2.6 shows an experimentally-calculated Q matrix for the sameprior and likelihood values. The experimental calculation was obtained from themodel by setting the parent to one hypothesis, allowing it to generate a 1-sample datastring, and simply tallying how many times the child arrived at each hypothesis over10,000 runs. As evidenced in this comparison, and other trial calculations, thismethod of experimentally calculating the Q matrix reliably approximates theanalytical solution.

Analytical Q matrix for MAP with bottleneck of 1:All the steps above for the Sampler are the same for the Maximizer (MAP) except forthe way the posteriors enter the equation. As opposed to the samplers, which choosetheir hypotheses with the probability of each hypotheses posterior probability, theMAP simply chooses the hypothesis with the highest posterior probability. Goingback to the posteriors (Table 2.4), data string [1] will always lead to H1, [2] willalways lead to H2, and [3] will always lead to H3. So, multiplying the probabilitythat the parent produces each data string times the probability it will be induced undereach hypotheses, simply yields the data likelihoods as defined by each hypothesis(Table 2.7).

Q matrix child

H1 H2 H3

H1 0.8 0.1 0.1parent H2 0.1 0.8 0.1

H3 0.1 0.1 0.8

Table 2.7Analytically-calculated Q-matrix for MAP with bottleneck of 1

Page 21: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

21

Q matrix child

H1 H2 H3

H1 0.8062 0.0948 0.099parent H2 0.1016 0.7983 0.1001

H3 0.1038 0.0954 0.8008

Table 2.8Experimentally-calculated Q-matrix for MAP with bottleneck of 1

Experimental Q matrix for MAP with bottleneck of 1:Again, for comparison, Table 2.8 shows that the experimentally-calculated Q matrixclosely approximates the analytical Q matrix.

Q matrix calculations for Sampler with bottleneck of 2:As more data samples are allowed, computing the analytical solution becomes quitecumbersome. This is because the data likelihood and posteriors of all possible datastrings must be calculated. For a bottleneck of 2, there are 6 (order-independent) datastrings. Below are the new data likelihoods (Table 2.9) and the posterior values(Table 2.10) for every possible data string.

Data LikelihoodsH1 H2 H3

[1 1] 0.64 0.01 0.01[2 2] 0.01 0.64 0.01[3 3] 0.01 0.01 0.64

[1 2] or [2 1] 0.16 0.16 0.02[1 3] or [3 1] 0.16 0.02 0.16[2 3] or [3 2] 0.02 0.16 0.16

sum 1 1 1

Table 2.9

Posterior ValuesH1 H2 H3

[1 1] 0.9933 0.0044 0.0022[2 2] 0.0515 0.9412 0.0074[3 3] 0.0959 0.0274 0.8767

[1 2] or [2 1] 0.7671 0.2192 0.0137[1 3] or [3 1] 0.8485 0.0303 0.1212[2 3] or [3 2] 0.2258 0.5161 0.2581

Table 2.10

These new likelihoods (Table 2.9) are obtained by multiplying the likelihood valuesof the data points in question, as defined by each hypothesis (Table 2.3). Forexample, H1 produces data point 1 with p=.8, so producing it twice has theprobability of p=.64. All probabilities in each column sum to one because they coverall possible data strings.

Page 22: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

22

Again, the posterior values (Table 2.10) are computed with Bayes’ rule. When theSampler receives string [1 1], it will choose H1 with a 99% probability, and H2 andH3 with less than 0.5% probability each. To get each entry of the Q-matrix, all thelikelihoods of each string being induced under each hypothesis must be multiplied bythe likelihoods that each string is produced at all and then these values are summed.So, for parent H1 going to child H1, this value is the sum of the likelihood that eachstring is produced by parent H1 times the probability it will be induced as child H1:(.64*.9933)+(.01*.0515)+(.01*.0959)+(.16*.7671)+(.16*.8485)+(.02*.2258) = .9002Table 2.11 shows the analytical Q matrix and Table 2.12 shows the experimental Qmatrix for comparison.

Q matrix child

H1 H2 H3

H1 0.9002 0.0627 0.037parent H2 0.2197 0.7209 0.0594

H3 0.2591 0.1188 0.6221

Table 2.11Analytically-calculated Q-matrix for Sampler with bottleneck of 2

Q matrix child

H1 H2 H3

H1 0.9004 0.0635 0.0361parent H2 0.2218 0.7178 0.0604

H3 0.2588 0.1166 0.6246

Table 2.12Experimentally-calculated Q-matrix for Sampler with bottleneck of 2

Q matrix calculations for MAP with bottleneck of 2:Again, calculations for the MAP differ from the sampler in terms of hypothesischoice. To obtain the analytical Q matrix, only the likelihoods of strings with themaximum posterior will be summed under each hypothesis. These are the values inbold in Table 2.10. Here, strings [1 1], [1 2], [1 3] will always lead to H1. Strings [22], [2 3] will always lead to H2, and string [3 3] will always lead to H3. Therefore,the probability that the data from H1 will lead to H1 in the next generation is .64 (for[1 1]) + .16 (for [1 2]) + .16 (for [1 3]) = .96. Table 2.13 is the resulting analytical Qmatrix and Table 2.14 is an experimentally-calculated Q matrix for comparison.

Q matrix child

H1 H2 H3

H1 0.96 0.03 0.01parent H2 0.19 0.80 0.01

H3 0.19 0.17 0.64

Table 2.13Analytically-calculated Q-matrix for MAP with bottleneck of 2

Page 23: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

23

Q matrix child

H1 H2 H3

H1 0.9600 0.0289 0.0111parent H2 0.1900 0.8018 0.0092

H3 0.1900 0.1674 0.6439

Table 2.14Experimentally-calculated Q-matrix for MAP with bottleneck of 2

2.3.3 Analytical and Experimental Stationary Distribution CalculationsThe Q matrix summarizes the potential for transition dynamics in the system which itdescribes. But what can this dynamical fingerprint tell us about the outcome ofiterated learning? If an ILM simulation could be run for an infinite amount of time,the relative frequency of each chosen hypothesis would settle into a particulardistribution that is determined entirely by the Q matrix. This distribution is known asthe stationary distribution and serves as an idealized shorthand for the “outcome ofiterated learning.” As demonstrated by Griffiths & Kalish (2005) and Kirby et al.(2007), the stationary distribution is proportional to the first eigenvector of the Qmatrix. Therefore, the stationary distribution is easily determined for each model, bynormalizing the first eigenvector of its analytically-calculated Q matrix.

In an experimental run, the relative frequency of all chosen hypotheses are alsoentirely determined by the Q matrix, but because a run contains a finite number oftransitions, it represents one actual trajectory of transitions, from a larger set ofprobable trajectories under that Q matrix. However, when a large number oftransitions can be recorded in a simulation (by setting the number of generationssufficiently high), then a tally of the actual hypotheses chosen by the agents over thecourse of the simulation closely approximates the analytical stationary distribution.Below are the stationary distributions for each of the analytical Q matrices from theprevious section (Table 2.15). For comparison, next to each is the normalizedhypothesis history of a corresponding simulation run of 10,000 generations. Thenormalized hypothesis history is a reliable, experimental approximation of thestationary distribution.

Stationary Distribution Approximations

Table 2.15Normalized Hypothesis History approximates the analytical stationary distribution for both the Samplerand MAP model. The posterior mean is only a reliable approximation for the Sampler model. S1 =Sampler with bottleneck of 1, S2 = Sampler with bottleneck of 2, M1 = MAP with bottleneck of 1, M2= MAP with bottleneck of 2.

Page 24: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

24

Additionally, for the Sampler only, the average of all agents’ posterior values serve asa good approximation for the stationary distribution. This is because the hypothesesare chosen according to the exact proportions of the posterior vector. For the MAP,posterior mean can not be used as an approximation heuristic. MAP dynamics are nottied to the exact values of the posterior, because agents only respond to the maximum.Table 2.15 shows the posterior mean of the same simulation runs.

2.3.4 SummaryThe Bayesian ILM of the present research can be used to experimentally determinethe internal dynamics and associated stationary distribution of both Sampler and MAPmodels, and over a wide variety of parameter combinations. Determining the Qmatrices and stationary distributions through experimental calculations and simulationheuristics provide a good alternative to computing the analytical solutions, whichbecomes increasingly cumbersome as the bottleneck or population size or increases.Additionally, the simulations will allow the investigation of certain parametercombinations, such as multi-agent populations with heterogeneous biases, which donot have straightforward analytical solutions.

2.4 Model Results

This section will describe the differences between the MAP and Sampler given theparameter manipulations described earlier. First it will cover replicated aspects ofprevious Bayesian ILMs. Last, it will address new findings for multi-agentpopulations with heterogeneous and homogeneous biases.

2.4.1 Basic Sampler Behavior of 1-agent, 1-sample simulationsGriffiths & Kalish (2005) showed that the stationary distribution of the Sampleralways mirrors the prior. This was confirmed in my model for a 1-agent population.Over all combinations of priors and hypotheses structures tested, the Sampler model’sstationary distribution mirrored the prior. However, this was not the case for multi-agent populations, and will be addressed in section 2.4.4.

2.4.2 Basic MAP Behavior of 1-agent, 1-sample simulationsKalish et al. (2007) find that the MAP’s dynamics are effected by the prior, datalikelihoods (aka: hypothesis structure), and noise. However, it is not understoodexactly how the likelihoods affect the dynamics. Because my model does notinvestigate the effect of noise on the model’s behavior, it is more readily apparentwhich aspects of the dynamics are due to the prior and which are due to thehypothesis structure. The following explanations of MAP behavior in terms ofhypotheses structure and bias influence are novel and were informed by simulationswith the present model.

From the Q matrix calculations in the previous section, it is clear that the Q matrixvalues of a 1-sample simulation are the data likelihood values for each hypothesis.This leads to consistent patterns in the stationary distribution for particular types ofhypotheses structures. Overall, the hypotheses structures investigated in this modelcan be broken down into two main categories; canonical and asymmetrical.Canonical hypotheses structures are ones where each hypothesis is defined by thesame set of data likelihood values, but shifted so that each hypothesis’ peak is over a

Page 25: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

25

different data point. Examples of canonical hypotheses are in Table 2.16, a-e. Withina canonical hypotheses structure, each hypothesis has identical probabilities oftransitioning to every other hypothesis and therefore, when there is no prior bias, eachhypothesis is equally represented in the stationary distribution.

Asymmetrical hypotheses structure occurs when each of the hypotheses are notcomposed of the same values, and therefore have more complex transitionprobabilities among themselves. Examples of asymmetrical hypotheses are in Table2.16, f-j. Figure 2.1 is also an asymmetrical hypotheses structure. The stationarydistributions of this category of hypotheses are difficult to predict, however I haveidentified some general trends in the dynamics. Though, these trends may only holdfor this particular model’s implementation, with an equal number of hypotheses asdata points. The first concerns their relative peak height . The hypothesis with thehighest peak likelihood value will be represented with the highest proportion in thestationary distribution. Likewise, the hypothesis with the lowest peak will berepresented the least. The second concerns their relative overlap. When allhypotheses have peaks with equal likelihood values, but one has higher extremelikelihoods than the two hypotheses, as does H2 in example f, it will be representedwith the greatest proportion in the stationary distribution. These relationshipsregarding hypothesis overlap and relative likelihood values probably havestraightforward analytical solutions and are open points for further analyses.

Hypotheses Structure Effect on Stationary Distribution

Table 2.16Differences in normalized hypothesis history for the two categories of hypotheses; canonical andasymmetrical. All results above were calculated with an unbiased prior.

Because these relationships have to do with the entire hypotheses structure, the effectthat one hypothesis’ likelihoods has on the stationary distribution always depends onits context, which is the other two hypotheses. This makes for a difficult analysis.Figure 2.2 shows the manipulation of just one hypothesis, H2, in 4 different contexts,and with an unbiased prior. Here, H2’s peak is slowly raised from likelihood value0.33 (flat/no peak) to 0.9, as shown on the x-axis. The context hypotheses structuresare displayed in the columns of graphs at the sides (these graphs display the

Page 26: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

26

hypotheses structure as introduced in Figure 2.1). The left column shows a snapshotof the hypotheses structure in order for lines a-d at x = 0.3. The right column showsthe structures at x = 0.9. For a, H1 and H3’s peaks = 0.33 (flat), b peaks = 0.4, cpeaks = 0.6, and d peaks = 0.8. The y-axis shows the proportion of H2 in thenormalized hypothesis history. It is clear to see that raising the peak of H2, raises itsproportion in the hypothesis history. However, the higher the context hypotheses, thelower the proportion of H2. Additionally, the gray line at y = 1/3 marks the pointwhere all hypotheses are level in the hypothesis history. All hypotheses structuresfound at the intersection with this line are the canonical forms; where the H2 peak andcontext peaks are the same height.

Figure 2.2Proportion of H2 in the MAP stationary distribution as a function of H2’s hypothesis peak in 4different hypotheses structures. Peaks of context hypotheses H1 & H3 in a = 0.33 (flat), b = 0.4, c =0.6, d = 0.8. Prior is unbiased.

The picture becomes even more complex when a bias is introduced. Figure 2.3 showsthe same center graph as above, but for 3 different prior biases in favor of H2. Theunderlying dynamics remain the same, but the bias adds an additional layer ofcomplexity. When the maximum prior probability is higher than the maximumlikelihood value, the hypothesis which the bias favors becomes 100% represented inthe stationary distribution, meaning this is the only hypothesis which an agent is ableto choose. This is because the posteriors of all data strings will be maximum underthe hypothesis which the bias favors. When the maximum prior probability is equalto the maximum likelihood value (indicated by the stars), the H2’s proportion in the

Page 27: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

27

stationary distribution is raised considerably. But when the maximum priorprobability is less than the maximum likelihood value, there is no change to thestationary distribution. Therefore, no manipulation to the bias, when in this range,will affect the stationary distribution. To summarize Figure 2.3, the MAP hypothesisstructure plays a considerable role in shaping the system’s dynamics, but when theprior is high enough, these dynamics are overridden by the bias and all agents choosethe hypothesis that has the highest prior probability.

Figure 2.3Hypotheses structure effect on the MAP stationary distribution, with added effects from prior biases.

Table 2.17 directly visualizes this threshold for line c (hypotheses = [.6 .2 .2; .2 .6 .2;.2 .2 .6] of the middle graph in Figure 2.3. Here, the posterior values are given for allpossible data strings [1], [2], and [3]. The location of the maximum posterior values(in bold) are what determine the MAP hypothesis choice. Across this threshold, theseposterior maxima make a shift, thus shifting the outcome of iterated learning for thismodel. When the prior value is anywhere lower than the H2 peak, as in prior = [.205.59 .205], the dynamics remain completely determined by the hypotheses structure.However, nudging the prior up to [.2 .6 .2], which is the same level of the H2 peak,the posteriors move to favor H2 because the MAP is now faced with 2 maximumposterior values for 2 of the data strings and will choose them each 50% of the time.Finally, as soon as the prior bias for H2 exceeds the H2 likelihood peak, as in [.195.61 .195], all posterior maxima are located under H2. At this point, all agents in thesimulation will choose H2 for all possible data strings.

Posterior values under different priorsdata Prior = [.205 .59 .205] Prior = [.2 .6 .2] Prior = [.195 .61 .195]string H1 H2 H3 H1 H2 H3 H1 H2 H3

[1] 0.44 0.42 0.15 0.43 0.43 0.14 0.42 0.44 0.14[2] 0.09 0.81 0.09 0.09 0.82 0.09 0.09 0.82 0.09[3] 0.15 0.42 0.44 0.14 0.43 0.43 0.14 0.44 0.42

Table 2.17Hypotheses = [.6 .2 .2; .2 .6 .2; .2 .2 .6]

Page 28: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

28

Basic Sampler vs. Maximizer Conclusion:For the Sampler model, the most salient determiner of the dynamics is the prior.Although the transitions in the Q matrix are not trivially determined, the stationarydistribution derived from the Q matrix exactly mirrors the prior, despite manipulationsto the hypotheses structure. The MAP model’s dynamics, on the other hand, are mostsaliently determined by the data likelihood values. For 1-agent, 1-sample simulations,the Q matrix exactly mirrors the data likelihood values as defined by each hypothesis,and standard calculus should be able to predict the stationary distribution. When thehypotheses structure is canonical, then the probability of an agent choosing any givenhypothesis in the stationary distribution is equal. When the hypotheses structure is ofvarious, asymmetrical combinations, the stationary distribution reflects each of themdifferently. A prior bias adds yet more to the MAP dynamics, but only when it isstronger than the likelihood values.

2.4.3 The Bottleneck EffectThe number of data points that are transmitted between generations constitute thelearning bottleneck. The bottleneck size, therefore, equals the number of datasamples in the data string. Varying the bottleneck size directly effects thetransmission dynamics. When the bottleneck is large, there is a much higherprobability that the proportion of data samples in the data string faithfully reflects thelikelihoods of hypothesis it was generated from. This leads to greater fidelity oftransmission; where each generation usually chooses the same hypothesis as thegeneration before it. When very little data is transmitted over each generation,transmission fidelity is much lower, yielding many transitions between hypothesischoices within the simulation run. Transmission fidelity is directly visible in thediagonal axis of the Q matrix. A high probability of each hypothesis leading to itselfequals high transmission fidelity and a lower number of transitions in the simulationrun. As the bottleneck increases, transmission fidelity increases until it reaches 100%and Q matrix diagonals are all equal to 1. Depending on the strength of the bias andthe distinctiveness of the hypothesis peaks, this increase occurs at different speeds(Figure 2.4). However, this rate does not seem to be affected by hypothesis choicestrategy. All models will eventually reach 100% transmission fidelity at a certainbottleneck size.

In Figure 2.4, the transmission fidelity index used here is the average of the values onthe diagonal of the Q matrix. This indicates the probability, for any randomly-chosenhypothesis, that the child will choose the same hypothesis. When the index reaches 1,this means that all diagonal values in the Q matrix are 1. In this case,miscommunication is impossible and every generation will have the hypothesis of theprevious generation. Here, the outcome of iterated learning will be solely determinedby the initial data. Therefore, the hypothesis which the initial data best supports willbe the hypothesis that all generations will choose.

Page 29: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

29

Figure 2.4Increase in transmission fidelity is slower for models with weak biases and likelihoods.Behavior is more determined more by these factors than by hypothesis choice strategy.

Strong: prior = [.7 .2 .1] and hypotheses = [.8 .1 .1; .1 .8 .1; .1 .1 .8]Weak: prior = unbiased and hypothesis = [.4 .3 .3; .3 .4 .3; .3 .3 .4]

For a finite number of generations, all simulations will appear to display completetransmission fidelity when the bottleneck is wide enough. This will occur when theprobability of miscommunications (non-diagonal cell values) make it unlikely thatthey will appear within the given number of generations. For example, if oneparticular miscommunication has a probability of 0.01, it will usually not occur in asimulation of with less than 100 generations, but it likely to occur several times in asimulation of 10,000 generations.

For infinite generations, on the other hand, complete transmission will never occur aslong as the hypotheses overlap and there exists some probability of transitioning fromone hypothesis to another. But for finite runs, the practical appearance of completetransmission fidelity is determined by the combination of the prior and hypothesesstructure. When hypotheses have small overlap and a strongly-biased prior, less datasamples are needed to unequivocally indicate which hypothesis distribution they weregenerated from. In this case, complete transmission fidelity will occur at smallerbottleneck sizes (Figure 2.4, “MAP strong” and “Sampler strong”). However, forhypotheses with more overlap and weaker biases, complete transmission fidelity willoccur at larger bottleneck sizes (Figure 2.4, “MAP weak” and “Sampler weak”).

Page 30: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

30

Bottleneck effect differences between MAP and Sampler:For both the MAP and Sampler, transmission fidelity increases as the bottleneckwidens. The Sampler’s stationary distribution continues to mirror the prior, over allbottleneck sizes and priors tested. This confirms that the bottleneck has no effect onthe outcome of iterated learning for the Sampler model. However, it does effect theinternal dynamics of transmission and may well have an effect on the outcome ofiterated learning over finite time spans. The MAP’s stationary distribution, on theother hand, continues to be affected both by the likelihoods and priors, but changesnon-monotonically as the bottleneck widens. Though the MAP’s transmission fidelitysteadily increases, the dynamics reflected by the stationary distribution aresurprisingly unstable (Figure 2.5). Interestingly, this instability only occurs withasymmetrical hypotheses structures, where the slightest asymmetry leads to wildlydifferent stationary distributions for each bottleneck size. For canonical hypothesesstructures, all hypotheses continue to be equally represented in the stationarydistribution. Unfortunately, the cause of this strange behavior has not been obtained.

Figure 2.5MAP posteriors non-monotonically vary as bottleneck widens. The y-axis shows the proportion of H2in the experimentally-calculated stationary distribution (normalized hypothesis history). Prior =unbiased, Canonical Hypotheses = [.6 .2 .2; .2 .6 .2; .2 .2 .6], Asymmetrical Hypotheses = [.6 .3 .1; .2.6 .2; .1 .3 .6]

Bottleneck and data variance issues:Because these simulations are confined to a finite number of generations, theexperimentally-derived stationary distributions are less reliable under largerbottlenecks. This is directly due to the increase in transmission fidelity. Under smallbottlenecks, the high number of transitions in the simulation ensure that the resultingdistribution in the normalized hypothesis history reflects the true stationarydistribution. When transmission fidelity increases, the variation between simulationruns also increases, and thus more generations (or multiple runs) are needed to obtaina reliable approximation of the stationary distribution. If the simulation could be runan infinite number of generations, then the normalized hypothesis history would be

Page 31: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

31

the stationary distribution, and transmission fidelity would have no effect. Howeverthis is impossible. For all of the data in the present report, all simulations were runfor 10,000 generations. At this setting, variance begins to become a problem with Qmatrix and stationary distribution approximation around a bottleneck of 6-10. Abovethis level, multiple runs must be averaged to gain a more complete picture of themodel’s dynamics.

2.4.4 Population SizeWhen the population consists of multiple agents, which dynamics found in the single-agent models hold, and which do not? And what is the outcome of iterated learningwhen this population has heterogeneous biases? The remaining sections will answer,in terms of this model, these new questions regarding population size andheterogeneity.

In this model, each agent in a multi-agent population sees the same data string,separately calculates their posterior values, chooses their own hypothesis, andgenerates their own data. The data from each agent of the same generation are thenconcatenated into one unified data string, which is given to the next generation astheir input. When the population parameter is set to any number x, all generationshave x population members. When the number of data samples is set to y, each agentin the population produces y number of data samples, yielding a bottleneck size ofx*y.

The multi-agent and single-agent configurations differ in one respect: the data stringthat is passed between generations is not stochastically generated from one unifiedagent, but from many. This has different consequences for the MAP and the Samplermodels. For a homogeneous, multi-agent MAP model, the behavior of all agents inthe population is identical. Because all agents receive the same data string and haveidentical priors and hypotheses, the posterior of all agents will be the same (and this isalso the case for the Samplers). However, all the MAP agents will choose the samehypothesis (Table 2.18), because this choice is based on the maximum value of theiridentical posteriors. The only exception to two MAP agents choosing differenthypotheses based on the same data string is when there are multiple maximum valuesin their posterior. In this case, they each choose one of the maximum valuehypotheses randomly, with equal weight. This situation generally only arises whenthere is no bias in the prior values (to help diversify the posterior values). Aside fromthis exception, multiple MAP agents producing y samples, is equivalent to one MAPagent producing x*y samples (Table 2.18). Therefore, MAP dynamics due topopulation size are identical to the dynamics due to the bottleneck (see section 2.4.3).However, due to the implementation of the multi-agent model, where all agentsproduce equal an equal number of data samples, only even-numbered bottleneck sizescan be investigated for population sizes greater than 1. Therefore, the non-monotonicvariance in the MAP model (referring back to figure 2.5) is less apparent in thesecases.

Page 32: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

32

Normalized Hypothesis HistoryMAP

H1 H2 H3samples = 4 0.6352 0.2488 0.1160

population = 4 0.6304 0.2561 0.1135

SamplerH1 H2 H3

samples = 4 0.6977 0.1918 0.1105

population = 4 0.8104 0.1254 0.0642

Table 2.18Population size does not add new dynamics for MAP, but for Samplers it does – the stationarydistribution no longer mirrors the prior. Prior [.7 .2 .1], hypotheses [.8 .1 .1; .1 .8 .1; .1 .1 .8], 10,000generations.

Hypotheses Choice of Multi-agent Sampler vs. MAP Sampler MAP

H1 H2 H3 H1 H2 H3

Hypotheses agent 1 7609 1579 812 8234 1496 270Chosen agent 2 7674 1570 756 8234 1496 270

Table 2.19MAP agents choose the same hypothesis, whereas Samplers do not.

Prior [.7 .2 .1], hypotheses [.8 .1 .1; .1 .8 .1; .1 .1 .8], 10,000 generations.

For a homogeneous, multi-agent Sampler model, the dynamics are markedly different.Because samplers choose their hypotheses weighted by their posteriors, ahomogenous population will not choose the same hypotheses each generation (Table2.19). Therefore, the data samples do not come from the same set of likelihoodvalues. This has interesting implications concerning the perfect Bayesian rationalityof the agents. In the case of the MAP, the agents have all the possible sets oflikelihoods that the data could be generated from, already given to them as theirhypotheses. When a string of data is generated from a set of likelihoods which theagents are not explicitly given, then they are not longer perfect Bayesian reasoners.This is exactly the case with a multi-population of Samplers. When a data string ingenerated from 2 different hypotheses, these probabilities do not conform to thelikelihoods as defined by any of their hypotheses. The result is, for a multi-populationof Samplers, the stationary distribution no longer mirrors the prior (Figure 2.6).

Kalish et al. (2007) mathematically show that their single-agent results can begeneralized to multi-agent populations, where the stationary distribution will continueto mirror the prior. However, this proof would require, in practice, that eachSampling agent is given a new set of hypotheses, for each corresponding populationsize, where each hypothesis represents the combined likelihood set for each possiblecombination of hypotheses that the agents of the population may have whenoutputting into the data string. Although perfect Bayesian rationality is a simpleassumption for mathematical analyses of ILMs, the practicality of maintaining thisassumption is dubious for actual model implementations, let alone for actual humans.

Page 33: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

33

Figure 2.6The MAP model’s stationary distribution is invariant to population size. For Samplers, population sizedoes affect the dynamics and the stationary distribution no longer mirrors the prior. Stationarydistributions for populations 1 and 2, for MAP and Sampler models with: prior [.7 .2 .1], hypotheses [.8.1 .1; .1 .8 .1; .1 .1 .8], 10,000 generations.

Additionally, some systematic variance was observed for the multi-agent Samplermodel in regard to manipulations of the likelihood structure. Figure 2.7 shows thatthe stationary distribution mirrors the prior less and less as the hypotheses structurebecomes strongly peaked and the prior more biased. However, for a combination ofrelatively flat hypotheses and weakly biased priors, the stationary distribution stillmirrors the prior. Additionally, increasing the population size systematicallyamplifies the effect of the likelihoods on the Sampler’s stationary distribution (Figure2.8).

Figure 2.8 shows that the stationary distribution reflects hypotheses structure in theabsence of a prior bias. For the canonical hypotheses structure a, the stationarydistribution remains flat despite changes in population size. This is similar to theMAP behavior given canonical hypotheses under different bottleneck sizes. Also likethe MAP model, the Sampler is differentially sensitive to asymmetrical hypothesesstructures, however the relationships are in the opposite direction. Here, the highestpeaked hypothesis is the lowest in proportion in the Sampler’s stationary distributionand the lowest peaked hypothesis is the most represented.

Page 34: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

34

Figure 2.7Strong biases and peaked hypotheses lead the sampler away from converging to the prior.. a = prior [.3.3 .3] b = prior [.6 .2 .2] c = prior [.7 .2 .1] d = prior [.8 .1 .1], Population = 2.

Figure 2.8Population size amplifies Sampler sensitivity to hypotheses structure. a = hypotheses [.8 .1 .1; .1 .8 .1;.1 .1 .8] b = hypotheses [.4 .3 .3; .1 .8 .1; .3 .3 .4] c = hypotheses [.8 .1 .1; .3 .4 .3; .1 .1 .8]. Prior =unbiased. Population sizes 1 to 5.

Page 35: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

35

2.4.5 HeterogeneityA heterogeneous ILM was implemented by taking a multi-agent model and assigningdifferent prior vectors to each of the agents. This model, therefore, is the mostcomplex of all models constructed. For this reason, only 2-agent heterogeneouspopulations will be used as examples in this section.

The main result is that heterogeneous agents’ hypotheses choices converge as they areallowed to share more and more data, despite having fixed and different priors fromeach other (Figure 2.10 and 2.12). This conforms to the general tradeoff between thelikelihoods and prior in Bayesian induction; the more data that is seen, the less theeffect of the prior on the posterior distribution over hypotheses. Because the behaviorof both models is based on the posterior values, increasing the amount of data whichthe agents share produces increasingly similar posterior values, despite differences inagents’ priors. In the following analyses, convergence is measured by the Euclideandistance between each agent’s normalized hypotheses history vector.

Since agents’ behavior is converging, the natural question is, to what? To structurethis question more, I decided to investigate whether or not the converged behavior ofa heterogeneous ILM (where agent x and agent y each differ in their prior bias) is justan average of the behavior of one homogeneous run with agent x and one with agenty. It turns out, this is not a simple question. First, it is difficult to determine exactlywhat the true average of agent x and agent y’s behavior is, due to the variation amongruns inherent in the simulation. Additionally, as discussed previously, the stationarydistribution of a particular model changes as a function of bottleneck or populationsize. Therefore, we cannot just average the stationary distribution of agent x andagent y for comparison to a 2-agent simulation composed of agent x and agent y.Instead, we should match for the number of samples in the data string. For theSampler, it is established that manipulating bottleneck size does not effect thestationary distribution: it will continue to mirror the prior. However, manipulating thepopulation size slightly effects the stationary distribution away from mirroring theprior. Although, this effect isn’t noticeable at a population of 2 for relatively flathypotheses and a mildly biased prior. Therefore, the average behavior for Sampleragent x and Sampler agent y can be done with some confidence – by just averagingthe priors of the two agents – but only when hypotheses are relatively flat and the biasis weak.

Heterogeneous Sampler behavior:So, the question for the heterogeneous Sampler model is: Does the convergedbehavior of a 2-agent heterogeneous Sampler model come to the average of theirpriors? The answer appears to be yes. Figure 2.9 shows the convergence of a 2-agent, heterogeneous Sampler ILM and Figure 2.10 shows the difference (measuredin Euclidean distance) between each agent’s hypothesis history, for the data in Figure2.9. For clarity, Figure 2.9 does not plot the entire stationary distribution, but onlyH1’s proportion in the stationary distribution. The gray line indicates what H1 shouldbe if the convergence reflects a trivial average of individual agent behavior. Theheterogeneous behavior seems to converge to this trivial average. But this is difficultto tell with certainty because, by the time the agents’ behavior converges completely,the variance between runs (due to the increasing bottleneck size, see section 2.4.3) istoo high to determine what the true convergence values are.

Page 36: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

36

Looking at the last reliable run from Figure 2.9, at bottleneck 16, the normalizedhypotheses history of each agent are displayed in Table 2.20. Here, it is clear that theconvergence behavior is settling around the average of both agent’s priors.

Average of ConvergingBehavior

H1 H2 H3

agent 1 0.43 0.21 0.36agent 2 0.37 0.21 0.42 Average of both Priorsaverage 0.40 0.21 0.39 0.4 0.2 0.4

Table 2.20Hypotheses History of each agent at bottleneck = 16, from Figure 2.9, and the average of their priors.

Figure 2.9The behavior of agents with different priors converges when they are allowed to share more and moredata. Variation of individual runs is still a problem over large bottleneck sizes. Population = 2, Prioragent 1 = [.6 .2 .2], Prior agent 2 = [.2 .2 .6], Hypotheses = [.6 .2 .2; .2 .6 .2; .2 .2 .6].

Page 37: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

37

Figure 2.10Euclidean distance between the agents’ hypotheses histories, of which H1 is graphedin Figure 2.9.

In section 2.4.4 I discussed some parameter combinations, for the multi-agentSampler, which did not result in a stationary distribution that mirrored the prior.Namely, strongly-peaked hypotheses structures plus strong biases. Some of thesecombinations were also analyzed in the heterogeneous model, however, the normalvariance of each run subsumed the difference between the averaged priors and theaverage outcome of the multi-agent runs which did not mirror the prior. Therefore, itis impossible to determine a difference in convergence when comparing these twoconditions.

Heterogeneous MAP behavior:As I’ve demonstrated in previous sections, the dynamics of the MAP model are morecomplex than that of the Sampler, and the heterogeneous models are no exception.The MAP model behaves very differently over an increasing bottleneck depending onwhether the hypotheses structure is canonical or asymmetrical (refer back to Figure2.5). Due to the unexplained, non-monotonic variance over bottleneck size of MAPmodels with asymmetrical hypotheses, I will restrict the heterogeneous MAP analysesto canonical hypotheses structures.

To determine whether the MAP model convergence of agent x and agent y is a trivialaverage of each of agent x’s and agent y’s normal stationary distribution, we need toknow what agent x and agent y’s normal behavior is. For the MAP, there is nodifference in dynamics whether the models are matched for population size orbottleneck. Therefore, the stationary distribution of a 1-agent x, bottleneck = 2simulation and the stationary distribution of a 1-agent y, bottleneck = 2 simulation canbe averaged to represent a trivial convergence state. For all the MAP models testedhere, their convergence state does not conform to this average exactly. Figure 2.11shows convergence in MAP hypotheses choice behavior, again just for H1’sproportion in the stationary distribution. The analytically-determined stationary

Page 38: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

38

distribution for single-agent, bottleneck = 2 model for agent 1 is [.83 .15 .03] andagent 2 is [.03 .15 .83]. The average of these vectors is [.43 .15 .43]. However, thehypothesis history of the heterogeneous model does not settle upon this average. Dueto the variation over runs, it is difficult to determine what the actual convergence stateis. However, it seems that the heterogeneous model is converging somewhere off ofthis average, rather then homing in on it as the Sampler model did.

Figure 2.12 shows some additional simulations, where each graph shows a simulationwith a different set of agent priors. These also seems to converge somewhere off ofthe trivial average. Recall that MAP stationary distributions are differentiallysensitive to whether the maximum prior value is higher or lower than the maximumhypotheses peak value (refer back to Figure 2.3). This differential sensitivity is alsoconfirmed in the heterogeneous MAP model. In figure 2.12 and 2.13, models (a) and(b) have a bias which exceeds the maximum likelihood values of the hypothesisstructure, whereas the bias in models (c) and (d) do not. The convergence behaviorfor these two sets of models are qualitatively different.

Figure 2.11MAP model shows signs of converging to something other than the average behavior of appropriatelymatched single-agent, homogeneous models (represented by the horizontal line). Hypotheses = [.8 .1.1; .1 .8 .1; .1 .1 .8], prior agent 1 = [.7 .2 .1] and prior agent 2 = [.1 .2 .7]

Page 39: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

39

Figure 2.12MAP convergence for 4 simulations, each where agents have a different set of priors. a-d: Hypotheses= [.6 .2 .2; .2 .6 .2; .2 .2 .6], a: prior agent 1 = [.4 .3 .3] and prior agent 2 = [.3 .3 .4], b: prior agent 1 =[.59 .205 .205] and prior agent 2 = [.205 .205 .59], c: prior agent 1 = [.6 .2 .2] and prior agent 2 = [.2 .2.6], d: prior agent 1 = [.8 .1 .1] and prior agent 2 = [.1 .1 .8]

Figure 2.13Euclidean distance between the agents’ hypotheses histories, of which H1 is graphed in Figure 2.12.

Page 40: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

40

Summary Heterogeneous ILM:For both the MAP and Sampler models, the behavior of 2 agents with heterogeneousbiases converges as a function of the bottleneck size. The more agent’s share eachother’s data, the more they choose the same hypotheses as each other. At very largebottlenecks, agents choose the exact same hypothesis throughout the simulation,showing that a sufficiently strong data likelihood value can override agent’sdifferences in prior biases. However, the inherent variance of a finite simulation overhigh bottlenecks, makes it impossible to obtain an exact distribution of hypothesesthat is being converged to. It appears that the Sampler model tends towardconvergence to a trivial average of the agents’ priors. The MAP model seems toconverge slightly below this level. However, overall the behavior of the Sampler andthe MAP in the heterogeneous model are qualitatively very similar. I’ve probablyshied away from addressing the true complexity of convergence by limiting myself to2-agent, bottleneck = 2 simulations, canonical hypotheses structures, and symmetricalprior sets. Undoubtedly, such simulations will yield more variance in behavioraccording to more fine-grained categories of parameter settings. However, thesesimulations are only a first step in addressing bias heterogeneity in a Bayesian ILM.

2.5 Model Discussion

This Bayesian ILM both replicated the general properties of previous existingBayesian ILMs and provided new results regarding multi-agent populations and biasheterogeneity. The replications are that a single-agent Sampler model’s stationarydistribution always mirrors the prior bias of the agent (Griffiths & Kalish, 2005), anda MAP model’s stationary distribution is determined both by the prior and the datalikelihood values (Kalish et al., 2007). Also, for a range of parameters, the prior hasno effect on the MAP stationary distribution (Smith & Kirby, 2008?), but above thisthreshold, iterated learning amplifies this bias (Kirby et al., 2007). Additionally, astrong bottleneck effect was observed, with the general effect of increasingtransmission fidelity of both the Sampler and MAP models (Kalish et al., 2007). Lastto note, the normalized history of all hypotheses choices over the course of thesimulation yielded the same solution as analytically calculating the Q matrix, as inNowak et al. 2001. All of these replications attest to this particular implementation asa valid Bayesian iterated learning model.

Throughout the replication work, new insights into the role of data likelihoods forboth the MAP and Sampler were obtained. The focus of all previous research withBayesian ILMs is the on prior and how its manipulations affect the stationarydistribution. Kalish et al. (2007) manipulated the degree of hypotheses overlap, aswell as noise level, but it appears their hypotheses correspond to what I call thecanonical form in my analyses. Otherwise, the rest of the literature does notmanipulate the data likelihoods of their model. My analyses of both canonical andasymmetrical hypotheses structure, and in the absence of noise, sheds new light onhow the likelihoods effect the stationary distribution, by way of determining theposterior values, which determine the transition probabilities of the Q matrix, whichyields a particular stationary distribution. There is a great deal of complexity inherentin the nature of the hypotheses overlap, where different hypotheses structures can beshown to determine the outcome of iterated learning just as much as the prior biasdoes. These complexities are also strongly influenced by the bottleneck, showing that

Page 41: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

41

hypotheses overlap is responsive to the pressures of cultural transmission. However,in the case of asymmetrical hypotheses, this sensitivity to the bottleneck issurprisingly unstable.

The results of Smith & Kirby (2008?) demonstrate that the MAP strategy ofhypothesis choice is evolutionarily stable over that of samplers. However, the MAPparameters for which this result was proven, it seems, were derived from a canonicalhypotheses structure and for the range of priors which are unaffected by bias strength(refer back to Figure 2.3). The result is consistent behavior of the Smith & Kirby’sMAP model over the bias values they selected. My results show that this is a subsetof MAP behavior and that unstable behavior is easily obtained for the rightrelationship between hypotheses and priors. So, perhaps MAP would not be theevolutionary stable strategy in all cases. Or additionally, perhaps it can be shown thata certain range of MAP parameters is evolutionary stable over other MAP parametersets. Knowledge of this kind would help guide the right choice of MAP parametersets to use for iterated learning simulations, rather than convention (i.e. assuming oneun-manipulated set of canonical hypotheses).

Another novel result was obtained from a manipulation in population size. By justincreasing the population size to 2, the Sampler model’s stationary distribution doesnot strictly mirror the prior. The result is that the hypothesis with the highest priorbecomes amplified in the stationary distribution, and some sensitivity to the likelihoodstructure emerges. This result is due to the fact that Samplers, according to thismodel’s multi-agent implementation, can no longer be classified as perfect Bayesianreasoners (refer back to section 2.4.4).

The population manipulation was also informative in terms of this thesis’ ultimatequestion; what does cultural transmission add? Since population size, in part, definesthe social structure and transmission dynamics of an ILM, any manipulation topopulation size that affects the stationary distribution can be taken as evidence forcultural transmission “adding” something. Referring back to Figure 2.6, it is clearthat cultural transmission adds additional dynamics in the case of the Sampler, but notin the MAP. Interestingly enough, the existing literature claims the reverse: culturaltransmission adds nothing to the Sampler model, but does to the MAP model.

In summary, this chapter has shown that the particular dynamics which werepreviously thought to differentiate the Sampler and MAP models, may not be as clearcut as they previously seemed. It is clear that the parameters which encodemanipulations to the cultural transmission system (bottleneck and population size)affect both the Sampler and MAP models. It is also clear that non-convergence to theprior cannot be taken as evidence against the Sampling strategy or support for MAPstrategy.

One suggestion for future Bayesian ILM research would be to investigate modelswhere agents are no longer perfect Bayesian reasoners, by giving agentsheterogeneous hypotheses structures. If each generation of agents do not have theexact same hypotheses structures as the previous generation, then agents will not beable to calculate the optimal posterior probabilities over hypotheses. It is certainly thecase that humans are not perfect Bayesian reasoners, because we do not havecomplete knowledge of the exact likelihoods involved in the processes of our

Page 42: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

42

environment, but rather we learn these probabilities and construct our ownhypotheses, imperfectly, through experience. Another suggestion would be to explorehypothesis choice strategies that are a mix between Sampling and MAP behavior. Itis more likely that human behavior can be better approximated by a strategy that lieson the continuum between sampling and maximizing, rather than one at either of theseextremes. Both of these suggestions should yield results which further inform usabout the outcome of iterated learning in human populations.

Page 43: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

43

Chapter 3

A Function Learning ILMwith Human Subjects

3.1 Introduction

Through the modeling work in the previous chapter, we have seen that if learners areperfectly Bayesian-rational, have perfect knowledge of the distributions from whichdata could be drawn, and sample from the posterior, then iterated learning willconverge to the prior. If any of these conditions do not hold, then the prior,likelihoods, and selection strategy all influence the outcome of iterated learning.

Ultimately, we are interested in explanations for the structure of natural language andhow cultural transmission mechanistically translates the properties of individuallearners into the properties of human language. If we take this model seriously, thenwe would expect that human learning biases can be read off of the universalproperties of human language. However, if we would like to make this claim and takean observed universal as evidence for a specific learning bias, then we must be surethat the likelihood structure and selection strategy are of the required kind. Likewise,if we want to predict what sort of universal will arise from a particular bias, then weneed to know the state of the other parameters involved in order to make such apredictions.

Therefore, the ideal way to proceed is to study subjects where one or more of theserequirements is known, so that the others can be inferred. Such a parameter analysisis straightforward within a computational model, however, it is difficult to knowwhich combinations and ranges of parameters approximate language induction andtransmission in an actual population of human learners. Clearly, by experimentallyconstructing an iterated learning model with human subjects, the uncertaintiesregarding the appropriateness of the learning algorithm are circumvented.

Experimental ILMs with human subjects show promise as a powerful framework fortesting predictions of both computational models and psychological studies regardinglearning biases and the cultural transmission of language. Over the past couple years,some initial explorations into this framework have been made. In verticaltransmission models, where information is transmitted serially, from one generation tothe next, Kirby et al. (2008) demonstrated in human learners, the emergence ofregularization, increased learnability, and compositionality due to a transmissionbottleneck. Flaherty (2008) also demonstrated a learnability increase with children,however these language did not become regular. As for horizontal transmissionmodels, where individuals repeatedly interact and negotiate a communication system,

Page 44: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

44

Galantucci (2005) showed the emergence of a communication system where thecommunication channel was undefined and Scott-Phillips (2008) showed theemergence of a symbolic communication system, even when communicative intentwas not pre-established between subjects. Additionally, Kalish et al. (2007)demonstrated regularization and learnability in a vertical ILM with human subjects,but in the domain of function learning; not a communication system.

The present research will take up the experimental ILM in function learning, as putforward by Kalish et al. (2007). This is a simple paradigm with established resultsregarding the role of human learning biases and has been successfully modeled with aBayesian ILM. Because these task-specific learning biases are more or less known,this task offers an ideal setting for studying the role of the data likelihoods in apopulation of human learners. In this chapter, I will present a replication of Kalish etal. (2007) iterated function learning experiment with human subjects. Additionally, Iwill present a novel, second condition which tests if human subjects display abehavioral difference to a manipulation in their perceived reliability of the trainingdata.

In both the original experiment and the present replication, subjects attempt to learnand then reproduce the underlying relationship between the lengths of two differentbars over a series of trials (see Figure 3.2 for an example screen shot). The twolengths constitute (x,y) pairs, so this underlying relationship can be described as afunction that relates these two data points over the complete set of (x,y) pairs. The xvalue serves as the subject’s stimuli and is encoded by the length of a horizontal, bluebar. Each trial, a new stimuli bar length appears and subjects indicate thecorresponding y value by adjusting the height of a vertical, red response bar. Duringthe training phase, a feedback bar is presented alongside the subject’s response bar,showing the correct response bar height for the target (x,y) pair. In the testing phase,this feedback bar does not appear and the subject’s responses are recorded as the newy-values for the corresponding stimuli of each trial. At the end of the testing trials, anew set of (x,y) pairs has been obtained, and reflects what the subject inferred aboutthe relationship behind the data in their training set. This new set of (x,y) pairs thenserves as the training set for the next subject in the iterated learning chain.

With this experiment, Kalish et al. demonstrated that iterated learning reliably lead tohuman behavior that was consistent with the known bias for this task, within only afew generations. Their study consisted of 4 conditions, each containing 8 chains ofnine generations. Each condition was defined by the initial function which was usedto generate the (x,y) pairs of the first generation’s training set. These four functionsare positive linear, negative linear, u-shape, and random (x,y) pairs. Figure 3.1 showsrepresentative chains from each condition, taken from Kalish et al. (2007).

Page 45: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

45

Figure 3.1Results from Kalish et al. (2007) iterated function learning experiment with human subjects. Shownare 5 representative chains from the 4 conditions, where the initial function is: (A) positive linear, (B)negative linear, (C) U-shaped or (D) & (E) random. Each set of axes plot the testing phase responsesof each subject, which was trained on the data to the left of it. Regardless of the initial data, iteratedlearning converges to the positive linear function, with the highest inductive bias, and occasionally tothe negative linear function, with the next highest bias.

Regardless of the initial data, the subjects in the iterated learning chains converged toone of the a prior preferred solutions: a positive linear function with positive slope(with the highest bias), or one with negative slope (with the second-highest bias).Kalish et al. attest that these bias rankings are established in previous psychologicalstudies on function learning, which show that linear functions with positive sloperequire the least training to learn (Brehmer, 1971 & 1974) and are consistent withsubjects’ initial responses (Busemeyer et al., 1997).

For all chains in all conditions, this study reports convergence to one of these twofunctions, with no exceptions. Assumedly, this also means that no intermediate, semi-stable states were obtained on the route to convergence. The consistency of thisresult, over so many subjects, is somewhat surprising, because there are other,conceivably easy, ways to solve this task, such as dividing the apparently-complexfunction into simpler sub-functions in order to approximate the whole (resulting in adiscontinuous function), responding discretely to the continuous stimuli (resulting in astep function), or by hedging bets in the testing phase if the subject was undecidedbetween two or more plausible underlying relationships. The last or these behaviors,bet hedging, seems to be evidenced in the original results. In Figure 3.1C, subjects 1-3 seem to be guessing between the positive and negative linear functions. Or perhaps,they inferred that both of these functions were generating their training data.However, the motivation behind subject’s behaviors could only be obtained from anexit questionnaire, and cannot be concluded from their testing data alone.

It is important to note that this specific function learning task was developed byKalish et al. (2004) to demonstrate knowledge partitioning and showed that subjectsreadily divided a complex function into multiple, simpler sub-functions. This studyprovides empirical evidence that individuals are capable of inferring a discontinuous

Page 46: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

46

function for this particular task paradigm. However, subjects were explicitly awarethat the underlying function was somewhat complex, and this might have elicitedadditional strategies. If subjects were instructed in Kalish et al. (2007) that therelationship was simple, generated by one rule, or continuous, then this may haveexcluded some of these other possible behaviors. It is also possible that such arestriction to subjects’ expectations of the possible, task-relevant hypotheses mighthave played into the demonstrated convergence to the prior, where other solutionscould be equally stable for human learners. Therefore, in the present replication, Iwas careful to not over specify the nature of the function, so not to add additionalexpectations toward a specific class of functions (such as continuous, generated by asingle rule, or based on one-to-one mappings). Leaving the nature of the functionambiguous establishes more potential for varied behavior. Additionally, anybehavioral differences regarding the manipulation of the likelihoods may not bevisible under a task in which subjects entertain a restricted set of hypotheses.

In the present study, the Kalish et al. experiment is replicated. The findings confirmthat the majority of chains converge to the known bias, the positive linear function.However, this replication also obtained semi-stable discontinuous functions, discreteresponses, and confirms through exit questionnaires, bet-hedging behavior. Thesebehaviors will be discussed further in the Results section. A second experiment(condition 2) contains a novel manipulation to subjects’ perceived reliability of theirtraining data in comparison to the replication experiment (condition 1). Thismanipulation was accomplished by an addition to the instructions, informingparticipants that some of the trials in the training phase would be random pairs, to adda small level of noise to the training, and the testing phase would test how well theylearned the underlying relationship. (See Appendix B for both instructions.) Only theinstructions differed between conditions; otherwise, the set up of the experimentremained exactly the same. In both conditions of the following experiment,instructions, methodology, and stimuli were kept as similar as possible to the originalstudy. Additionally, all chain initializations were set to a different set of random (x,y)pairs, corresponding to Kalish et al.’s 4th condition. Although Kalish et al. show thatthe initial data plays no role in the long run, this initialization was still used to rule outan initial bias in the data toward any specific function.

The chosen manipulation of subjects’ perceive role of the data was motivated by theinfluential role which the data likelihoods played in the model. If the data likelihoodshave a psychologically real correspondence within the process of human induction,then manipulating subjects’ perceived role of the data should also affect the dynamicsof convergence. However, it is important to note that model, of course, is at best adescription of human behavior. I make no claims to be manipulating subjects’ “datalikelihoods”, but instead manipulating how much the subjects might rely on the datawhen inferring the underlying relationship between the stimuli. However, accordingto the model, this manipulation would be best characterized by a change to thelikelihood values, where less informative data is represented by flatter hypothesesthan more informative data.

It is clear that human subjects are neither perfectly Bayesian-rational, nor do theyhave perfect knowledge of the distributions from which the data might be drawn.Additionally, their hypothesis choice strategy might fall on a continuum betweenmaximizing and sampling. According to the modeling work, violating these

Page 47: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

47

assumptions does not exclude the prior from determining the outcome of iteratedlearning, but it does suggest that the likelihoods will play a role in determining thisoutcome. Assuming that the Bayesian model is a good approximation of humaninductive inference, manipulating the role of the data should not affect which bias isconverged to (because the priors are not the target of manipulation), but the dynamicsof convergence themselves. For example, this could affect the rate of convergence, orthe proportion of chains which conform to each hypothesis. This manipulation choiceis supported by the known trade-off in Bayesian statistics between the influence of theprior vs. the influence of the data. Where data provides little information, the prior ismore influential, and vice a versa. It is hypothesized that in the condition where thedata are perceived to be less informative, chains will converge to the positive (oroccasionally negative) linear function quicker than in the replication condition.Additionally, it is predicted that behavior corresponding to a wider variety ofhypotheses will be obtained in condition 1 than in condition 2. And because bothconditions leave the possible hypotheses unspecified, more varied behavior will beobtained in both conditions than in the original study by Kalish et al. (2007).

3.2 Method

3.2.1 ParticipantsExperiments 1 & 2: Participants were solicited by email invitation from anestablished mailing list experiment participants. 56 respondents completed the onlineexperiment. Condition 1 consisted of 8 chains and 32 subjects and condition 2consisted of 7 chains and 20 subjects. 4 subjects from condition 1 were automaticallyexcluded due to contemporaneous login. Not all subjects completed the exitquestionnaire, so the age range and gender is not known for all participants.Experiment 3: 9 graduate students of Logic and Cognitive Science constituted 1chain. 3F/6M, aged 23-30. They took a computerized, offline version of theexperiment in person.

3.2.2 Apparatus and stimuliThe experiment was implemented as an online java applet by Federico Sangati. Theapplet displayed all trials and collected all results. On each trial, two bars encoded thex and y values of the function by their lengths. The stimulus was the x-value and waspresented as a horizontal blue bar in the upper left-hand corner of a 800 by 600 pixelapplet window. The stimulus bar was 20 pixels wide, ranging from length 5 pixels (x= 1) to 480 pixels (x =100). A response was made by rolling the mouse to adjust thelength of a vertical red response bar, in the lower right-hand side of the screen. Theresponse bar was 20 pixels wide, ranging from length 4 pixels (y = 1) to 400 pixels (y= 100). This proportion skew was copied from the original study as an additionalmeasure against building a bias toward linearity into the task interface. Thebackground was white and the maximum values of each bar were not marked (Figure3.1).

Page 48: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

48

Figure 3.2Example screen shot, with labels added. This is taken from the implementation of the presentexperiment, but is nearly identical to the original interface of Kalish et al. (2007).

At the beginning of a trial, the blue bar appeared. The response bar could be adjustedat will and no time constraint was imposed. When the subject had adjusted the bar tothe desired height, they pressed the return key to record their response. During thetraining phase of the experiment, a feedback bar displayed the target response afterthe return key was pressed. The feedback bar was presented 10 pixels to the right ofthe response bar and had the same width and possible range as the response bar. If theresponse was correct (within a 5 unit / 20 pixel range of the correct y value) thefeedback bar would appear in green and the screen remained as is for 1-second studyperiod. If the response was incorrect, the feedback bar would be shown in yellowuntil the subject readjusted the response bar to the exact height of the feedback barand pressed return again. This readjustment period looped until the correct answerwas recorded. Once recorded, the feedback bar would be shown in green and thescreen remained as is for a 2-second study period. Afterward, the next trial began.The testing phase was identical to the training phase, except no feedback in the formof the feedback bar was given. After the subject pressed the return key to record theirfirst response, the next trial began. The training and testing phases each consisted of50 trials.

3.2.3 ProcedureThe experiment had a training and a testing phase, each consisting of 50 trials. Thestimuli and responses of one subject’s testing phase served as the stimuli andfeedback bar for the next subject’s training phase. Thus, each subject can be referredto as a “generation” in an iterated learning chain. All chains were initialized on arandom training set: the training set of the first generation consisted of 50randomized (x,y) pairings. The length of the stimuli bar encoded the x value and thefeedback bar/response bar encoded the y value. The testing set for all generations,consisted of 50 x values; 25 selected randomly from the 50 training x values and 25selected randomly from the 50 unused x values. The subject’s responses in the testingphase were saved as the new set of 50 (x,y) pairs for the next generation’s training setand presented in random order. This was the only form of contact between

Page 49: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

49

participants and they were unaware that their data would be used for another test-taker.

3.2.4 Data Collection and AnalysesWhen a subject accessed the experiment link, they would be directed at random to oneof the chains, where the last recorded testing set of that chain would serve as theirtraining set. Due to this implementation, the number of generations of each chainvaries, and some chains did not receive enough subjects and remain un-converged.Additionally, some chains that showed convergence to the positive linear function forat least 2 generations were truncated by the experimenter. The original studydisplayed the robustness of this convergence state and it was preferred to initialize asmany chains as possible with the limited number of subjects. If two or more subjectswere directed to the same chain contemporaneously, both would complete the task,but only the first completed testing set would serve as the input to the next generation.IP addresses were also logged and each computer was blocked from running theexperiment twice. Lastly, the testing phase was followed by a brief questionnaire(Appendix B).

Because this is a relatively new function learning paradigm, it is unclear what the bestquantitative analyses might be. Kalish et al. (2007) mainly presented a qualitativecharacterization of their results. For their quantitative analysis, they computed thecorrelation of each function to the positive linear function, y = x. However, in thepresent experiment, not all chains converge to the positive linear function, andtherefore such a correlation measure would not be equally informative for all chains.The present results will also be characterized qualitatively and focus on a comparisonand contrast of the dynamics obtained here to those of the original study.

3.3 Results

Condition 1 constitutes a replication of Kalish et al. (2007). Figure 3.2 shows allchains collected for this condition. Of these 8 chains, 5 seem to converge to thepositive linear function. The remaining chains show no sign of definite convergence,though chain #1 seems to be headed toward negative linear. Of the converged chains,estimated convergence to the positive linear function occurred by an average of 2.4generations. Figure 3.3 shows all results from condition 2, where subjects where tolda small level of noise existed in the training set. Of these 7 chains, 3 seem toconverge to the positive linear function, and 1 to the negative linear function. Of theconverged chains, estimated convergence to the positive or negative linear functionoccurred by an average of 2 generations.

Looking at these results cumulatively, some behavior is evident which was notconclusively obtained in the original study. First, is the phenomenon of bet hedging,which can be taken as evidence that subjects are sampling. However, there is no wayto tell if they are in fact sampling from all possible hypotheses or just outputtingaccording to their most probable current hypotheses. Bet hedging is apparent incondition 1, chain 1, generation 1 (C1-1-1), C1-1-5, and C2-7-2. As obtained from 2exit questionnaires, subjects report that they guessed the rule is “red bar = eitherdouble or half of the blue bar” (y = 2x or _x) but that they weren’t sure so they

Page 50: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

50

Figure 3.2This figure shows the results from the 8 chains of condition 1. Each set of axes plot the testing phaseresponses of each subject, which was trained on the data to the left of it. The leftmost column showsthe random initialization that the first subject of each chain was trained on. 5 of these chains seem toconverge to a linear function with positive slope, and chain 1 seems to be headed toward a negativelinear function. Bet hedging and discontinuous functions also appear.

responded by choosing randomly according to both rules, in attempt to “get at leastsome answers correct.”

Second, is the appearance of discontinuous functions. C1-5-2, C1-6-3&4, C2-6-1,2&3 all appear to be categorizing certain lengths of the blue bar and applyingdifferent rules relating red bar size to each category. The strongest evidence of thisshown in C2-7-3&4. Here, this discontinuous function seems relatively stable andclearly originated in the bet-hedging behavior of generation 2.

Additionally, C2-6-1 reported a different continuous solution, which was graduallyincreasing the red bar when the blue bar increased, but keeping the red bar in its midrange. Some participants also guessed the relationship could be related to time-coursedependence (where one trial’s answer was dependent on the trial before it), but nonereported to output answers consistent with such a hypothesis. It is also clear, from theexit questionnaires, that when subjects had absolutely no idea of what the underlying

Page 51: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

51

Figure 3.3This figure shows the results from the 7 chains of condition 2. Each set of axes plot the testing phaseresponses of each subject, which was trained on the data to the left of it. The leftmost column showsthe random initialization that the first subject of each chain was trained on. 3 of these chains seem toconverge to a linear function with positive slope, and 1 to a linear function with negative slope. Chains6 & 7 show additional behaviors, maintained for 3 generations.

Figure 3.4This figure plots the results of Experiment 3, where subjects are graduate student of logic and cognitivescience and constitute one chain of iterations. The wider variation and higher fidelity of transmissionsuggest that these subjects entertained a wider variety of hypotheses than the subjects in the previousexperiments.

Page 52: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

52

relationship might be, they generated their testing answers by either simulatingrandom responses or by matching the red bar to the blue bar length, resulting in apositive linear function. The latter was the case for C1-7-1, C2-2-2, and C2-4-1.

One additional, small experiment (Figure 3.4) was conducted with graduate studentsin logic and cognitive science, who were told that this was a function learningexperiment. In general, these students displayed a lot of motivation to “correctlyfigure out” the underlying function. The resulting behavior of these subjects give riseto chain full of variation in hypotheses choice, but also

a higher degree of transmission fidelity between these more complex functions. It ispossible that these subjects entertained a wider variety of hypotheses, or even havedeveloped different learning biases for function learning tasks than the generalpopulation. It is also likely that these subjects took the data into account differentlythan subjects in the previous experiments, through higher attention levels or workingmemory capacity, explaining the increased transmission fidelity.

The subject in generation 1 reported bet hedging between y = 2x and y = _ x. Thisserved as a foundation for the next 2 subjects who inferred that the relationship wasgenerated by 2 rules, y = 2x and y = _ x. Subject 5 categorized the stimuli bar intotwo separate rules. Subject 6 found a continuous function based on a proportion ofthe stimulus bar’s remainder. In generation 7, the extreme values of the previoussubject are regularized as maximum response bar lengths for the smallest third ofstimulus bar sizes. Last, subjects 8 and 9 reported using the samecategorization and rules in the testing phase. It is doubtful that this chain willconverge to a positive linear function anytime soon, due to the saliency of the rulecoding x < 1/3 its maximum length = maximum y. It seems likely, however, that thisfunction could converge to a fully discrete function, which might be just as stable asthe positive linear function.

3.4 Discussion

These experiments elicited a wider variety of behavior in than has been previouslydemonstrated in an iterated function learning experiment. This is probably due to anunder specification of the nature of the function to be learned, resulting in subjectsentertaining, on average, more hypotheses than subjects in Kalish et al. (2007). Theseresults show that some aspect of human subject’s hypotheses can be reached bymanipulating their expectations of the processes which generate the data they see.The more expectations they have toward a particular class of functions (such ascontinuous, one-to-one mappings), the less they will entertain hypotheses which donot conform with those expectations. However, this result could very well beexplained by a change in subjects’ prior distribution over hypotheses between theoriginal and present experiments. And thus, serves as evidence that the distribution ofprior probabilities over hypotheses can be manipulated through instructions orcontext. Additionally, it is quite likely that the biases which subjects bring to aspecific task can be directly manipulated, perhaps by changing the domain orreasoning in which the task is couched. For example, if the two bars were said to

Page 53: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

53

represent the volume of two cups (linear) or the speed and stopping distance ofvehicles (exponential), the corresponding ranking of biases might be altered.

This variance in behavior obtained in the present experiments, coupled with a smallnumber of chains, makes it unclear whether there was a difference in convergence ratebetween conditions 1 and 2. Perhaps collecting more chains could provide a clearerpicture. There is some evidence, though, that the purported noise facilitated subjectsin choosing their first hunch. In particular, subject C2-1-3 reported entertaining awide variety of possible relationships to no success during the training phase, but then“remembered reading something about the noise added and I decided to stick to myoriginal idea…” which was inverse proportionality of the two bars.

Although individual variation in ILMs may obscure differences between experimentalmanipulations, these differences shouldn’t necessarily be controlled for, but insteadstudied in greater depth. These experiments demonstrated that experimental ILMsoffer a very incomplete picture of individual learning biases. Looking back at theresults, a wide variety of responses are obtained in the first generation of each chain,to initializations of the same class; random (x,y) pairs. Although we can see differentresponses from different people to similar stimuli, there is not way to tell, from suchan ILM, what range of responses to this stimuli any particular subject is capable of.

Despite an unclear difference between conditions, this experiment largely replicatesthe results of Kalish et al. (2007). This helps to establish this relatively new task as areliable paradigm for revealing the inductive biases of human function learning. AsKalish et al. assert, the experimental ILM may be a good tool for revealing humaninductive biases for tasks where they are unknown or where researches have few apriori hypotheses about what they might be. Additionally, the human subjects ILMmay lend itself well to the experimental manipulation of convergence patterns or thebiases themselves. Such manipulations could be useful in testing hypothesesinformed by computational modeling, as explained in this chapter’s introduction.

In general, an experimental ILM can inform us more about human iterated learningbehavior than a computational ILM and the relevant psychological studies combined.However, it will provide the strongest results when coupled with these othermethodologies. It is possible that certain psychological pre-tests could be conductedwith experimental ILM participants to ascertain subject-specific biases. Subjectscould be grouped into ILMs according to differences in biases and this could becorrelated with differences that are obtained in the dynamics or resultingcommunication systems in each ILM. It is also possible that the actual trajectorywhich a particular chain takes can be explained by the particular biases of theindividuals at each generation. Using experimental and computational ILMs incombination can help us to infer unknown behaviors (such as hypothesis choicestrategy) from observed behavior (such as sensitivity to data likelihoods). Becausethe models provide us with insights into what behaviors are typically associated withwhat parameter settings, we may be able to infer certain underlying relationships inthe human subject, when particular behaviors are obtained. Eventually, it may bepossible to read off the biases of individuals from the properties of the languages theydevelop in an experimental ILM. Until then, experimental ILMs will still serve asgood tools for revealing the general, shared biases in a population of human learners.

Page 54: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

54

Chapter 4

General Conclusion

Overall, the present research has demonstrated a wider variety of behavior than hasbeen previously obtained in computational and experimental iterated learning models,yielding new insights into the complex interplay between individual biases and thecultural transmission of language.

The debate over how much innate biases vs. cultural transmission determine theoutcome of iterated learning, seems to be somewhat reconciled. The presentmodeling results demonstrated that Samplers in a population larger than 1 do notconverge to the prior and are sensitive to manipulations in the data likelihoods. Thus,when an ILM does not converge to the prior, this can neither be taken as evidenceagainst the sampling strategy, nor for the MAP strategy. Additionally, both modelsare sensitive to dynamics imposed by cultural transmission; the MAP model issensitive to bottleneck size, and the Sampler model to population size. Thus, wecannot expect that either of these models will simply converge to the prior withinILMs that more realistically approximate human social systems.

Another important point which appears when comparing the computational model tothe experimental results, is a confusion over what “converging to the prior” means ineach case. In the analytical solutions to the computational model, converging to theprior means that the stationary distribution for a given model exactly mirrors the priorprobabilities over hypotheses. Therefore, in the stationary distribution, eachhypothesis is chosen with identical proportions to its assigned prior probability.Therefore, if we want to carry this term over to human populations, we actually needto know the distribution of prior probabilities, over all hypotheses in the population.Then, we must show that the population’s behavior corresponds to each of thesehypotheses with identical proportions. Unfortunately, due to the finiteness ofexperimental ILM runs, such a demonstration may be impossible to obtain.

What is actually obtained in my experiment and in Kalish et al. (2007) is subjectbehavior which usually corresponds to the function with the known highest bias, andoccasionally to the function with the known second-highest bias. This hardly isevidence for true convergence to the prior. In fact, this behavior would be expectedfrom maximizers and samplers alike, as well as mixed strategies in between, becausemaximizers also choose the hypothesis that corresponds to the highest prior themajority of the time. What is more compelling evidence for sampling behavior is thesubject-reported behavior of bet hedging, which was obtained in the experimentalILM. Though, on the other hand, other subjects also entertained many hypothesesduring the training phase, but only choose one of these with which to generate theirresponses during the testing phase. This behavior could equally be seen as

Page 55: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

55

maximizing behavior. However, this lack of behavioral distinction betweenhypothesis choice strategies does not discredit the experimental ILM as a powerfultool for revealing human inductive biases. As long as the most prominent behaviorscorrespond to the highest biases, these biases can be identified regardless ofassumptions about human hypotheses choice strategies.

The last important points regard the use of Bayesian inference as a model of humancognition. First, it may not be the case that the prior fully specifies the bias for theBayesian inference algorithm once its adapted into an agent within a culturaltransmission system. Additions to the Bayesian inference algorithm, such as ahypothesis choice strategy, must be implemented so that this algorithm can outputdata for other agents in the simulation. These additions are probably building inadditional biases to the agents behavior, as is apparent in the behavioral differencesbetween Samplers and Maximizers. This raises doubt to the claim that BayesianILMs are the solution to previous ILM confounds, where the learning algorithms hadimplicit and incomparable biases.

Second, we would certainly want to know how much the behavior of the modelschange when agents are no longer perfect Bayesian reasoners. Relaxing thisassumption could give a better account of human behavior in iterated learning. Itcould be quite interesting, for researchers in cognitive science, to investigatecomputational ILMs where agents are heterogeneous in respect to their hypothesesstructures, because an in-depth study of the hypotheses component of Bayesianinference could provide a formal framework to investigate the representationalconstraints of individual cognitive agents, and how they affect the transmission oflanguage.

The work presented in this thesis has attempted to synthesize the findings of acomputational ILM and an iterated learning experiment with human subjects. Ashopefully demonstrated in the present research, this combination of methodologiescan provide us with deeper insights into explaining the structure of human language,as rooted both in the biases of individual cognitive agents, and the system of culturaltransmission in which the interact.

Page 56: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

56

Appendix A

Bayesian ILM code for the MAP agentMatlab

%% Parameters:N_gen = 10000; % number of generationsN_pop = 2; % must enter same number of rows in prior matrix as N_popN_hyp = 3; % number of hypothesesN_sam = 1; % number of samplesN_sampop = N_sam*N_pop; % number of samples, totaled over agentsN_dat = 3; % number of data-values (assuming data-values range from 1 to N_dat)

%% Main program loop

%% initializeposteriorhistory = zeros(N_gen,N_pop,N_hyp);hypotheseshistory = zeros(N_gen,N_pop);hyphist = zeros(N_pop,N_hyp);hyphistnorm = zeros(N_pop,N_hyp);posterior = zeros(N_pop,N_hyp);agents_posterior = zeros(N_pop,N_hyp);agents_hypothesis = zeros(N_pop,1);data_each_agent = zeros(1,N_sam);posteriormean = zeros(N_pop,N_hyp);summary = zeros(N_pop,N_hyp);likelihood = zeros(1,N_hyp);prior = zeros(N_pop,N_hyp);prior = log ([.8 .1 .1; .1 .1 .8]); % must enter #rows=N_pophypotheses = log ([.6 .2 .2; .2 .6 .2; .2 .2 .6]);data = zeros(1,N_sampop); % data is a vector. each agent's output follows in chunksdata = random('unid',N_dat,[1,N_sampop]),

%% iteratefor generation=1:N_gen,

%calculate posterior for a=1:N_pop, likelihood = [0.0 0.0 0.0]; %resets likelihood to zeros, each loop for i=1:N_sampop, likelihood = likelihood +

transpose(hypotheses(:,data(1,i))); end; agents_posterior(a,:) = logBayesRule(prior(a,:),likelihood); end; agents_posterior; %matrix of all agents posteriors for this generation

posteriorhistory(generation,:,:) = agents_posterior; for a=1:N_pop, posteriormean(a,:) = sum(posteriorhistory(:,a,:)) ./ N_gen; end;

%Maximizer

%choose hypothesis for a=1:N_pop, %randomize order hypotheses are evaluated to be max or not, because %if there are multiple identical max values, max() always returns the firstone, biasing towards h1, then h2. maxtest = transpose(randsample(3,3)); [value,position] = max(agents_posterior(a,:)); if agents_posterior(a,maxtest(:,1)) == max(agents_posterior(a,:)); hstar = maxtest(:,1); elseif agents_posterior(a,maxtest(:,2)) == max(agents_posterior(a,:));

Page 57: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

57

hstar = maxtest(:,2); elseif agents_posterior(a,maxtest(:,3)) == max(agents_posterior(a,:)); hstar = maxtest(:,3); else 'I cant program'; end; agents_hypothesis(a,:) = hstar; hypotheseshistory(generation,a) = hstar; end; agents_hypothesis; %matrix of all agents' chosen hypothesis for this generation

%generate data data = []; for a=1:N_pop, data_each_agent =

randsample(1:N_dat,N_sam,true,exp(hypotheses(agents_hypothesis(a),:))); data = [data data_each_agent]; end;

end;

%create hyphistfor a=1:N_pop, for h=1:N_hyp, for g=1:N_gen, if hypotheseshistory(g,a) == h; hyphist(a,h) = (hyphist(a,h))+1; else hyphist(a,h) = hyphist(a,h); end; end; end;end;

for a=1:N_pop, hyphistnorm(a,:) = hyphist(a,:) ./ sum(hyphist(a,:));end;

Bayesian ILM code for the Sampler agent hypothesis choiceMatlab

%choose hypothesis agents_hypothesis = []; for a=1:N_pop, agents_hypothesis(a) = randsample(1:N_hyp,1,true,agents_posterior(a,:)); hypotheseshistory(generation,a) = agents_hypothesis(a); end; agents_hypothesis; %matrix of all agents' chosen hypothesis for this generation

%generate data data = []; for a=1:N_pop, data_each_agent =

randsample(1:N_dat,N_sam,true,exp(hypotheses(agents_hypothesis(a),:))); data = [data data_each_agent]; end;

Page 58: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

58

Appendix B

Experiment instructions for condition 1:

Instructions:

Thank you for your participation! This experiment consists of 2 parts. In the first part, you will be taught therelationship between the sizes of two different bars. In the second part, you will be tested to see how well youlearned this relationship.

Part 1 Instructions:

Part 1 will teach you the relationship between the sizes of a blue and a red bar. During this part of theexperiment, pay attention and try to learn this relationship the best you can.

Here’s what will happen in Part 1:

1. At the top of the screen, a blue bar will be shownat a particular size. Each trial, it will be a differentsize. Your job is to adjust the red bar (by rolling yourmouse) so that the red bar has the size you want. Youwill learn, over the course of Part 1, what the correctresponse is.

2. Once the red bar is the size you want, press thespace bar on your keyboard to record your response.

When your response is correct: When your response is not correct.

Page 59: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

59

If your response was correct, a green barwill appear to the right of the red bar for a1-second study period. (If your answer isvery close, but not exact, it will be accepted). Thenthe next trial will begin.

If your response was incorrect, a yellowbar will appear to the right of the red barand show you the correct answer. Youmust re-adjust the red bar to the exactheight of the yellow bar and press thespace bar. There will be a 2-second study period andthen the next trial will begin.

Part 1 consists of 50 trials. There is no time constraint.

Part 2 Instructions:

Part 2 will test how well you have learned the relationship between the blue and the red bar. This part isidentical to Part 1 except that the feedback bar will not appear. Once you record your response with thespace bar, the next trial will begin.

Try your best to indicate the correct size of the red bar during this part of the experiment!

Part 2 consists of 50 trials. There is no time constraint.

Good Luck!

Click here to run the experiment.

Page 60: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

60

Experiment instructions for condition 2:

Instructions:

Thank you for your participation! This experiment consists of 2 parts. In the first part, you will be taught therelationship between the sizes of two different bars. In the second part, you will be tested to see how well youlearned this relationship.

Part 1 Instructions:

Part 1 will teach you the relationship between the sizes of a blue and a red bar. For most of the trials in Part 1,the blue and red bar sizes will correspond to this relationship. However, in some trials the blue and red barsizes will be random – this adds a small level of noise. During this part of the experiment, pay attention and tryto learn the underlying relationship the best you can.

Here’s what will happen in Part 1:

1. At the top of the screen, a blue bar will be shownat a particular size. Each trial, it will be a differentsize. Your job is to adjust the red bar (by rolling yourmouse) so that the red bar has the size you want. Youwill learn, over the course of Part 1, what the correctresponse is.

2. Once the red bar is the size you want, press thespace bar on your keyboard to record your response.

When your response is correct: When your response is not correct.

If your response was correct, a green barwill appear to the right of the red bar for a1-second study period. (If your answer isvery close, but not exact, it will be accepted). Thenthe next trial will begin.

If your response was incorrect, a yellowbar will appear to the right of the red barand show you the correct answer. Youmust re-adjust the red bar to the exactheight of the yellow bar and press thespace bar. There will be a 2-second study period andthen the next trial will begin.

Page 61: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

61

Part 1 consists of 50 trials. There is no time constraint.

Part 2 Instructions:

Part 2 will test how well you have learned the underlying relationship between the blue and the red bar. Thispart is identical to Part 1 except that the feedback bar will not appear. Once you record your response withthe space bar, the next trial will begin.

Try your best to indicate the correct size of the red bar during this part of the experiment!

Part 2 consists of 50 trials. There is no time constraint.

Good Luck!

Click here to run the experiment.

Page 62: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

62

References

Bartlett, F. C. (1932). Remembering: A study in experimental and social psychology.Cambridge: Cambridge University Press.

Batali, J. (1998). Computational simulations of the emergence of grammar. In Hurford, J. R.,Studdert-Kennedy, M., Knight, C. (Eds.), Approaches to the Evolution of Language:Social and Cognitive Bases, pages 405-426. Cambridge: Cambridge University Press.

Brehmer, B. (1971). Subjects’ ability to use functional rules. Psychonomic Science, 24, 259-260.

Brehmer, B. (1974). Hypotheses about relations between scaled variables and in the learningof probabilistic inference tasks. Organizational Behavior & Human DecisionProcesses, 11, 1-27.

Brighton, H. (2002). Compositional Syntax From Cultural Transmission. Artificial Life, 8(1).Brighton, H. & Kirby, S. (2001). The survival of the smallest: stability conditions for the

cultural evolution of compositional language. In Kelemen, J. & Sosik, P. (Eds.),ECAL01, pages 592-601. Springer-Verlag.

Brighton, H., Smith, K., & Kirby, S. (2005). Language as an evolutionary system. Physics ofLife Reviews, 2, 177-226.

Busemeyer, J. R., Byun, E., DeLosh, E. L., & McDaniel, M. A. (1997). Learning functionalrelations based on experience with input-output pairs by humans and artificial neuralnetworks. In Lamberts, K. & Shanks, D. R. (Eds.), Knowledge, concepts, andcategories: Studies in cognition, pages 408-437. Cambridge: Cambridge, MA: MITPress.

Christiansen, M. & Kirby, S. (2003). Language Evolution: Consensus and controversies.Trends in Cognitive Science, 7(7), 300-307.

Cornish, H. (2006). Iterated Learning with Human Subjects: an Empirical Framework for theEmergence and Cultural Transmission of Language. Unpublished Masters thesis,School of Philosophy, University of Edinburgh, U.K.

Flaherty, M. & Kirby, S. (2008). Iterated language learning in children (abstract). In Smith, A.D. M., Smith, K., & Ferrer I Cancho, R. (Eds.), Proceedings of the 7th InternationalConference (EVOLANG7), pages 425-426. World Scientific.

Galantucci, B. (2005). An experimental study of the emergence of human communicationsystems. Cognitive Science, 29, 737-767.

Griffiths, T. L. and Kalish, M. L. (2005). A Bayesian view of language evolution by iteratedlearning. In Bara, B.G., Barsalou, L., and Bucciarelli, M. (Eds.), Proceedings of theTwenty-Seventh Annual Conference of the Cognitive Science Society, pages 827-832.Erlbaum, Mahwah, NJ.

Griffiths, T. L., Christian, B. R., & Kalish, M. L. (2006). Revealing Priors on CategoryStructures Through Iterated Learning. Proceedings of the 28th Annual Conference ofthe Cognitive Science Society.

Hare, M., & Elman, J. L. (1995). Learning and morphological change. Cognition, 56,61-98.

Hurford, J. R., (2000). Social transmission favors linguistic generalization. In Knight, C.,Studdert-Kennedy, M., Hurford, J. R. (Eds.), The Evolutionary Emergence of

Page 63: How learning biases and cultural transmission structure ...tuvalu.santafe.edu/~vanessa/homepage/Publications... · computational, Bayesian iterated learning model is constructed to

63

Language: Social Function and the Origins of Linguistic Form, pages 324-352.Cambridge: Cambridge University Press.

Kalish, M. L., Lewandowsky, S., & Kruschke, J. K. (2004). Psychological Review, 111(4),1072-1099.

Kalish, M. L., Griffiths T. L., & Lewandowsky, S. (2007). Iterated learning: Intergenerationalknowledge transfer reveals inductive biases. Psychonomic Bulletin & Review. 14(2),288-294.

Kirby, S. (1998). Language evolution without natural selection: From vocabulary to syntax ina population of learners. Unpublished manuscript.

Kirby, S. (1999). Function, Selection, and Innateness: the Emergence of LanguageUniversals. Oxford university Press.

Kirby, S. (2000). Syntax without Natural Selection: How compositionality emerges fromvocabulary in a population of learners. Unpublished manuscript.

Kirby, S. (2001). Spontaneous evolution of linguistic structure: An iterated learning model ofthe emergence of regularity and irregularity. IEEE Journal of EvolutionaryComputation, 5, 102-110.

Kirby, S., Dowman, M., & Griffiths, T. L. (2007). Innateness and culture in the evolution oflanguage. PNAS, 104(12), 5241-5245.

Kirby, S., Cornish, H., & Smith, K. (2008). Cumulative Cultural Evolution in the Laboratory:an experimental approach to the origins of structure in human language. Proceedingsof the National Academy of Sciences, 105(31), 10681-10686.

Kuhl, P. K. (2004). Early language acquisition: cracking the speech code. Nature ReviewsNeuroscience, 5(11), 831-843.

Lieberman, A. M., et al. (1967). Perception of the Speech Code. Psychological Review, 74,431-61.

Lieberman, E., Michel, J., Jackson, J., Tang, T., & Nowak, M. (2007). Quantifying theevolutionary dynamics of language. Nature, 449, 713-716.

Nowak, M. A., Komarova, N. L., & Niyogi, P. (2001). Evolution of universal grammar.Science, 291, 114-118.

Pinker, S. (1984). Language Learnability and Language Development. Cambridge, MA:Harvard University Press.

Scott-Phillips, T. C., Kirby, S., & Ritchie, G. R. S. (2008). Signalling signalhood and theemergence of communication (abstract). In Smith, A. D. M., Smith, K., & Ferrer ICancho, R. (Eds.), Proceedings of the 7th International Conference (EVOLANG7),pages 497-498. World Scientific.

Smith, K. (2002). The cultural evolution of communication in a population of neuralnetworks. Connectionism Science ,14, 65-84.

Smith, K. (2003). Learning biases and language evolution. In Kirby, S. (Ed.) LanguageEvolution and Computation (Proceedings of the Workshop on Language Evolutionand Computation, 15th European Summer School on Logic, Language andInformation).

Smith, K., & Kirby, S. (2008). Natural selection for communication favors the culturalevolution of linguistic structure. In Smith, A. D. M., Smith, K., & Ferrer I Cancho, R.(Eds.), Proceedings of the 7th International Conference (EVOLANG7), pages 283-290. World Scientific.

Vogt, P. (2003). Iterated learning and grounding: from holistic to compositionallanguages. Unpublished manuscript.

Weisbuch, G. (1991). Complex systems dynamics: an introduction to automatanetworks. Santa Fe Institute Studies in The Sciences of Complexity LectureNotes, vol. 2.

Zuidema, W. (2003). How the poverty of the stimulus argument solves the poverty of thestimulus argument. In Becker, S., Thrun, S., & Obermayer, K. (Eds.) Advances inNeural Processing Systems 15. Cambridge, MA: MIT Press.