Categories

Architecture, orientation and mathematics

Recently, I became intrigued with the discussions of topology that I found among architects and historians of architecture. I saw a few familiar threads running through these discussions – like the emergence and self-organizing principles of biology, together with the view that mathematics was not, primarily, a tool but more a point of view.

I was introduced to the term bioconstructivism in a 2012 paper by John Shannon Hendrix: “Topological Theory in Bioconstructivism.”

Bioconstructivism involves the engagement in architecture of generative models from nature. This is in the tradition of natura naturans in architecture, which is the imitation of the forming principles of nature, as opposed to natura naturata, the direct imitation or mimesis of the forms. According to Plotinus in the Enneads, it is the purpose of all the arts to not just present a “bare reproduction of the thing seen,” the natura naturata, but to “go back to the Ideas from which Nature itself derives”

…An important element of Bioconstructivism is autopoiesis or self-generation, taking advantage of digital modeling and computer programs to imitate the capacity for organisms in nature to organize themselves, or for unorganized or fluid material to consolidate itself, based on the inner active principle of the organism, an “essential force” or “formative drive” which contradicts the mechanistic theories of Galileo, Descartes and Newton. The monad of Leibniz, for example, can self-generate in “integrals” from pre-existing sets of variables, resulting in “continuous multiplicity.”

But I must admit that, from this paper, I didn’t get a clear sense for how topology was used or understood in either architectural design or critique. However, in a 2007 essay on this very topic (on the website RZ-A), I found the following:

Toni Kotnik (the author of ‘The Topology of Type’) believes that the only goal for introducing topology to architecture has been ‘to overcome dialectical strategies of homogeneity or heterogeneity, which dominated the architectural discussion throughout the 20th century’ but rather than giving a real answer to these debates ‘a superficial practice has been established in which every non-linear deformation of the usually used canon of forms gets classified as topological design both by the architect and by the critics.’ (17) He also thinks that the current architecture is not really based on topology but ‘on differentiable dynamical systems and popular spin-offs like chaos theory and fractal geometry’ and suggests that ‘a topological approach to architecture should not be seen as a form-generating tool but as an abstract form of thinking to structure sensorial and rational perceptions in a spatial way’.

I searched for “The Topology of Type,” but found instead, Kotnik’s piece “…there is geometry in architecture.” And here, I was excited to see, a provocative discussion of mathematics and architecture, broader and more interesting than the more specific discussions of various applications of topological ideas.

Kotnik begins with this:

Within contemporary architectural design the form – rule relationship is often understood as the application of geometric rules in a generative process of form-finding, that is rules are a logico-algebraic text out of which architectural form emerges through the manipulation of data. By looking at the etymological roots of mathematics another reading of geometry can be uncovered that relates geometry back to bodily experience and the question of spatial orientation. This enables the re-introduction of the body into contemporary discourse of digital architecture.

Kotnik makes the point that, historically, geometry has been viewed as something that architects use (or consume), not something that they produce. The introduction of digital computing in architecture, however, opens the door to “the emergence of architectural form out of the manipulation of data” which, Kotnik observes, could introduce scientific thinking and methodology into the design process. He moves, then, into the discussion of mathematics that I so much enjoyed, beginning with the etymology of the word mathematics.

…mathematics has its roots in the Greek ta mathemata, which means what can be learned where learning, mathesis, is about the recognition of the unchanged, the stable, of the Being in a world of constant Becoming.

Kotnik argues that mathematics is “the human search for patterns as a means of orientation,” and so “in its original meaning, mathematics is about relating the body with the world around. As such, mathemata is about orientation.”   (emphasis added)

Kotnik briefly surveys the developments in mathematics within cultural histories. Euclidean geometry is seen as “an act of physical orientation based on the creation of an intellectual structure,” shaping the individual perception of the world. He calls attention to Riemann’s reformulation of the foundation of geometry, his concept of manifold, and the influence of philosopher Johann Friedrich Herbart is noted. Herbart understood that the variety of perceptions people have, as they move through space, undergo a “graded fusion,” that “glues” individual perceptions together to form a geometric image. Related to the argument I made in Cognition, brains and Riemann, Kotnik suggests

Riemann’s idea of manifolds renders Herbart’s psychological concepts more concrete and precise for mathematical application.

…Euclid’s geometry and its universalized form in Riemann’s manifolds are examples of how mathematical concepts can be created by formalizing the act of orientation.

This discussion of mathematics and architecture rests on the idea that mathematics is not primarily a tool, but almost an articulation of sensation and its consequences. Mathematics, perhaps, can be seen as providing alternative ways to organize sensation or, more to the point, as alternative ways to understand what we see.

I’ll conclude with this last bit from Kotnik:

Measuring space is a central activity for structuring our surroundings and an important orientation element in the human environment. However, this environment must not be understood in a limited way as the purely physical environment. It must be seen more broadly and comprehensively as a multilayered perceptual and experiential space. This expanded understanding of space and the incorporation of the subject into space transforms measurement into an act of individual orientation. The importance of such bodily measuring of space,of geometrein, is not only justified by philosophers like Heidegger, Merleau-Ponty or Deleuze but also by developments in contemporary neuroscience. What emerges is an understanding of architecture as a primarily emotional and perceptual experience grounded in biological values that have to be put forward in the design process.

The mathematical nature of self-locating

A 2011 TED talk in London was brought to my attention recently. The speaker, Neil Burgess  from University College London, spoke on the topic, “How your brain tells you where you are.” Burgess investigates the role of the hippocampus in spatial navigation and episodic memory. In the talk he describes the function of what are called place cells, boundary-detection cells, and grid cells. I wrote (also in 2011) about a Science Daily report on studies dedicated to understanding how hippocampal neurons represent the temporal organization of experience, in particular, how they bridge the gaps between events that are not continuous. But here I want to focus more on how the brain constructs spatial experience.

Electrophysiological investigations have identified neurons that encode the spatial location and orientation of an animal. Among these are those that have come to be called place cells, head direction cells, and grid cells. Burgess also talked about boundary detection cells in his 2011 TED talk, which seem to be coordinated with head direction. I was particularly struck by the clarity of some of the data images he presented. He showed his audience images produced by the firing of boundary detection cells in response to boundaries in the environment of a rat. In one of them we could see cell firings in response to one of the walls in the rat’s environment. In another, created after a second wall had been added to the environment, the firing was duplicated with respect the added wall. Burgess also presented the image of cells that fired when the rat was about midway between the walls, and one could see directly that when the rat’s box was expanded, the firing locations expanded. These boundaries needn’t be rectangular walls. They can be the drop at the edge of a table or the circular wall of a circular box.

Grid cells are creating representations a little differently, in a quasi-mathematical way, as their name suggests. Burgess tells about the rats again:

Now grid cells are found, again, on the inputs to the hippocampus, and they’re a bit like place cells. But now as the rat explores around, each individual cell fires in a whole array of different locations which are laid out across the environment in an amazingly regular triangular grid…So together, it’s as if the rat can put a virtual grid of firing locations across its environment — a bit like the latitude and longitude lines that you’d find on a map, but using triangles. And as it moves around, the electrical activity can pass from one of these cells to the next cell to keep track of where it is, so that it can use its own movements to know where it is in its environment.

Both boundary detection cells and grid cells reflect a sensory perception of the environment. But neurons also encode movement from proprioceptive information that can be used to measure the body’s displacement as we move (path integration). This is not movement defined by the environmental changes that occur, but from the body’s sensations of itself.
In a more recent paper Burgess and co-author C. Burgess  describe an interesting test of the errors that can be produced by the iterative neural processing of self-motion.

This process, known as path integration or dead reckoning, requires the animal to update its representation of self-location based on the cumulative estimate of the distance and direction it has traveled. It can be shown that an animal is utilising path integration by introducing a known error into its representation of direction or distance: in the case of the gerbils, if they are rotated prior to the return leg of the journey, and this is done slowly so that the vestibular system does not detect the motion, then the animals head towards the nest with an angular error equal to the amount they were rotated by.

So when we try to find our way back to something, like where we parked the car, we likely use boundary-detecting cells to remember distances and directions to buildings and boundaries, but we also remember the path we took, represented by the firing of grid cells and path integration. The interaction of these things seem to contribute to the pattern of neural firings that become associated with a particular place, a cognitive map of that place, formed by what are called place cells. There is a nice discussion of the history of the study of place cells which includes a number of images at BrainFacts.org. There the point is made that the ‘cognitive map’ defined by place cells is a ‘relation among neurons,’ not among points in space.

In brief, we can think of the “map” of a session in terms of space (the spatial relations of firing fields) and time (the tendency for pairs of cells to fire together or not). Since the speed of rats is restricted, these are essentially equivalent. An important concept is that the map is entirely in the brain. In this description, a map is defined by the relation among hippocampal neurons, not by the relationships between neurons and the environment. The linkage to the environment is critical, but does not define the map.
The temporal relations are important for two reasons. First, neurons in the brain do not know about space directly, but they know about time. Neurons can code the timing relations of the neurons that project to it, but not the spatial relations. In other words, within the brain, the map is a timing map that encodes the temporal overlap between cell pairs.

 

There are a few interesting things going on here. No doubt the grid cell idea and the vector-like measures of displacement that are encoded when we move around, trigger memories of mathematics. They are like our mathematical analyses of the 3-dimensional space of our experience. Place cells, on the other hand, are like another level of abstraction. They seem to have more in common with coding and non-spatial analyses, even though we don’t seem to know how they do what they do. The neurons that fire to represent a particular location have no spatial relationships among themselves. Neighboring place cells do not indicate neighboring environmental areas. And while correlated to sensory input, they are part of a non-sensory system. It is the integration of this system with the neural representations of boundaries, direction, and distance (among other things) that create our spacial awareness. Certainly sensory information is being subject to some kind of transformation. Spatial relations are translated into what look like purely temporal ones (the timing of neuron firing). The non-sensory system then stores a coded representation of a sensory one. Here again we see, not the mathematical modeling of brain processes but more their mathematical nature.

What we see when….

I recently listened to Krys Boyd’s interview with Peter Mendelsund, author of the new book What We See When We Read,  on North Texas’ public radio. Mendelsund is an award-winning book jacket designer. The interview had the effect of connecting his thoughts about reading to thoughts that I have had about mathematics. It wasn’t immediately obvious, even to me, why. But I think I’m beginning to understand.

An excerpt from the book was published in the Paris Review. This excerpt focuses on the incompleteness of the visual images that our minds create when we are reading, despite the fact that we experience them as clear or vivid. Mendelsund quotes William Gass who wrote on the character of Mr. Cashmore from Henry James’s The Awkward Age:

We can imagine any number of other sentences about Mr. Cashmore added … now the question is: what is Mr. Cashmore? Here is the answer I shall give: Mr. Cashmore is (1) a noise, (2) a proper name, (3) a complex system of ideas, (4) a controlling perception, (5) an instrument of verbal organization, (6) a pretended mode of referring, and (7) a source of verbal energy.

The quote is from the book Fiction and the Figures of Life, a collection of essays first published in 1979.  Following Gass a little further we find these remarks:

But Mr. Cashmore is not a person. He is not an object of perception, and nothing whatever that is appropriate to persons can be correctly said of him. There is no path from idea to sense (this is Descartes’ argument in reverse), and no amount of careful elaboration of Mr. Cashmore’s single eyeglass, his upper lip or jauntiness is going to enable us to see him.

Mendelsund adds this:

It is how characters behave, in relation to everyone and everything in their fictional, delineated world, that ultimately matters…

Though we may think of characters as visible, they are more like a set of rules that determines a particular outcome. A character’s physical attributes may be ornamental, but their features can also contribute to their meaning.

(What is the difference between seeing and understanding?)

He follows this with a very mathematical looking statement where the characters (along with some physical attributes), as well as particular events and their cultural environment, are represented by letters. Their interaction is somehow formalized in symbol.

These are all words that have been used with respect to mathematics – “not an object of perception,” “behavior that matters only in relation,” “a set if rules that determines a particular outcome…”

Mendelsund occasionally uses mathematical ideas to describe some of what may be happening in the reading (and the writing) of a story. There are the maps of novels, the graphs and contours of plot, the vectors in Kafka’s vision of New York City. And these observations:

Anna can be described as several discrete points (her hands are small; her hair is dark and curly) or through a function (Anna is graceful)

If we don’t have pictures in our minds when we read, then it is the interaction of ideas – the intermingling of abstract relationships – that catalyzes feeling in us readers. This sounds like a fairly unenjoyable experience, but, in truth, this is also what happens when we listen to music. This relational, nonrepresentational calculus is where some of the deepest beauty in art is found. Not in mental pictures of things but i the play of elements…

…But we don’t see “meaning.” Not is the way that we see apples or horses…

Words are like arrows – they are something and they also point toward something.

Any text can be seen as communication through words (symbol), that can be aided by pictures, but that only lightly relies on them. The reader builds an internally consistent world, grounded mainly in concepts, whose structure is communicated in symbol. And both structure and meaning are never fully completed. This certainly sounds a lot like mathematics. But more striking about Mendelsund’s work in particular, is his making direct use of his experience to explore profound philosophical questions.  What happens when we read tells us something about ourselves.

The world, as we read it, is made of fragments. Discontinuous points – discrete and dispersed.

(So are we.  So too our coworkers; spouses; parents; children; friends..)

We know ourselves and those around us by our reading of them, by the epithets we have given them, by their metaphors, synechdoches, metonymies.  Even those we love most in the world.  We read them in their fragments and substitutions.

The world for us is a work in progress.  And what we understand of it we understand by cobbling these pieces together – synthesizing them over time.

It is the synthesis that we know.  (It is all we know.)

And all the while we are committed to believing in the totality – the fiction of seeing.

…Authors are curators of experience.

…reading mirrors the procedure by which we acquaint ourselves with the world. It is not that our narratives necessarily tell us something true about the world (though they might), but rather that the practice of reading feels like, and is like, consciousness itself; imperfect; partial; hazy; co-creative.

Writers reduce when they write, and readers reduce when they read. The brain itself is built to reduce, replace, emblemize…Verisimilitude is not only a false idol, but also an unattainable goal. So we reduce. And it is not without reverence that we reduce. This is how we apprehend the world. This is what humans do.

Picturing stories is making reductions. Through reductions, we create meaning.

There is significant overlap here with how I see the doing and the making of mathematics.  Mathematics is the making of meaning through reduction and synthesis.  Emerging from some adjustment in the direction of the mind’s eye, mathematics mirrors, in another way, how we are acquainted with the world. It finds meaning that opens up other parts of that world, a bit more for us.  And it tells us something about the nature of vision and understanding itself. Mathematics will not be fully embraced by our culture until we see this – until we recognize its own living nature.

 

What mathematics can make of our intuition

The CogSci 2014 Proceedings have been posted and there are a number of links to interesting papers.

Here are some math-related investigations:

A neural network model of learning mathematical equivalence

The Psychophysics of Algebra Expertise:  Mathematics Perceptual Learning Interventions Produce Durable Encoding Changes

Two Plus Three is Five:  Discovering Efficient Addition Strategies without Metacognition

Modeling probability knowledge and choice in decisions from experience

Simplicity and Goodness-of-fit in Explanation:  The Case of Intuitive Curve-Fitting

Cutting In Line:  Discontinuities in the Use of Large Numbers by Adults

Applying Math onto Mechanism:  Investigating the Relationship Between Mechanistic and Mathematical Understanding

Pierced by the number line: Integers are associated with back-to-front sagittal space

Equations Are Effects: Using Causal Contrasts to Support Algebra Learning

One of the presentations I attended is represented by the paper:
Are Fractions Natural Numbers Too? This study challenges the argument that human cortical structures are ill-suited for processing fractions, a view which has been used to justify the well-documented difficulty that many children have with learning fractions.

Such accounts argue that the cognitive system for processing number, the approximate number system (ANS), is fundamentally designed to deal with discrete numerosities that map onto whole number values. Therefore, according to innate constraints theorists, fractions and rational number concepts are difficult because they lack an intuitive basis and must instead be built from systems originally developed to support whole number understanding.

…Emerging data from developmental psychology and neuroscience suggest that an intuitive (perhaps native) perceptually based cognitive system for grounding fraction knowledge may indeed exist. This cognitive system seems to represent and process amodal magnitudes of non- symbolic ratios (such as the relative length of two lines).

This particular study is fairly well-focused, however.  Researchers aim to demonstrate a link between our sensitivity to non-symbolic ratios and the acquired understanding of magnitudes represented by symbolic fractions.   Given this focus, the study looked at individual responses to “cross-format comparisons of various fractional values (i.e. ratios composed of dots or circles vs. traditional fraction symbols).  For example, a given symbolic ratio was given with a numerical numerator and a numerical denominator.  The dot stimulus would show an array of dots in the numerator, of a certain quantity, and another array, of a different quantity, in the demoninator.  The circle stimulus showed a blackened disc of a certain area in the numerator and another, of a different area, in the denominator.  With a reasonable amount of care taken in their analysis, the authors concluded that they had found evidence of “flexible and accurate processing of non-symbolic fractional magnitudes in ways similar to ANS processing of discrete numerosities.

Considered in concert with other recent findings, our evidence suggests that humans may have an intuitive “sense” of ratio magnitudes that may be as compatible with our cortical machinery as is the “sense” of natural number. Just as the ANS allows us to perceive the magnitudes of discrete numerosities, this ratio sense provides humans with an intuitive feel for non-integer magnitudes.

An important consequence of this kind of evidence is their suggestion that the widespread difficulty with fractions may be the result teaching fractions incorrectly –  with partitioning or sharing ideas that use counting skills and whole number magnitudes instead of encouraging the use of what may be our intuitive ratio processing system.  This point was driven home for me when I looked at the circle representations of ratios that were used in the study.  I found them very effective, very readable.

This view is certainly consistent with the proposal that a mental representation of continuous magnitudes predates discrete counting numbers  (as with Gallistel, et al).   But I also think that this initiative points to something likely to be important in cognitive science as well as math education.   My own hunch is that an intuitive sense of ratio is likely grounded in continuous magnitudes, like length and area or perhaps even in tactile sensations of measure, like in cooking, as was suggested by one of the paper’s authors.  And I think it plays some role in the long debate over the relationship between discrete and continuous numbers that can be seen in the history of mathematics.  One could argue that the ancient Greek’s rigorous distinction between number and magnitude contributed to their remarkable development of geometric ideas.  With the number concept isolated away from the geometric idea of magnitude, perhaps their geometric efforts were liberated, allowing a focused elaboration on that ‘intuitive sense of ratio,’ extending it, permitting manifold and deep results.  Understanding this cultural event in the light of cognitive processes might inform our ideas about how mathematics emerges as well as how to communicate that development in mathematics education.

Embodiment and a Philosophy of Mathematics

Yesterday I gave a talk at a symposium at the 36th annual Cognitive Science Conference. The content of the talk was described this way in our symposium proposal:

Mathematics has been the subject of experimental studies in cognitive science that explore the sensory grounding of number and magnitude. But mathematics also provides conceptual schemes that can manage our comprehension of complex integrated neural activity, like Giulio Tononi’s qualia space. Visual processes, like stereopsis, may be said to be mathematical in character, and the brain is often described as performing computations on sensory data as it constructs the elements of our experience. Mathematician Yehuda Rav has argued that mathematics grows on the scaffolding of cognitive mechanisms that have become genetically fixed with human adaptation. Joselle Kehoe will present a philosophy of mathematics informed by the significance of selected studies in cognitive science and selected moments in the history of mathematics. It will be considered in the light of structural coupling – the embodiment concept of enaction introduced by Varela, Thompson and Rosch.

I had the opportunity to listen to a number of talks and was struck by the extent to which computer modeling governs investigative strategies.  These models are designed to mirror neural processing.  The models inevitably influence the formulation of new questions, which often provide a refreshing angle on the phenomenon being investigated.  But I began to wonder if modeling strategies might not also begin to create conceptual paths from which it may become difficult to exit.  It also became increasingly clear to me that modeling of this sort can capture mathematical relationships (like a particular differential equations) in such a way that the formal mathematical description of the relationships could appear to not be necessary.  The impulse from some to dismiss the need for formal representations could raise some interesting questions about the nature and the role of these representations.  They certainly help identify the paths that connect concepts within mathematics?  And these bridges between different branches of mathematics powerfully extend our understanding of both the physical and the conceptual.  While many of the things I heard discussed were not designed to talk about what mathematics is, they indirectly addressed exactly that question.

A brief summary of my own talk, with links to supporting content, is this:

A view, held by many in cognitive science, is that an innate sense of magnitude forms a non-symbolic real number system.  These are continuous magnitudes that arise from our experience of time, space, motion, etc.  This view is supported by the observation that pre-verbal humans and non-verbal animals do perform non-symbolic arithmetic operations on these magnitudes.  There are also studies that address how even plants are able to adjust their rate of starch consumption based on the amount of starch they have stored, and the length of time till dawn (when they can begin storing again).  Researchers have argued that language, somehow, picked out the discrete integers with which we learn to count (since language itself is a discreet system).  The evolution of the real number system in mathematics could then be seen as an investigation of the relationship between number and magnitude.  From the Greek separation of quantity (number) and magnitude (the ratios and proportions of continuous quantities like length and time), through the 18th century difficulties with infinitesimals in calculus, and the eventual definitive placement of irrational numbers on the real number line, mathematicians continued to test these relationships and their implications.  It is as if this work reflects a struggle to talk about a number system that existed before we were able to talk.

Brain processing itself seems to have a mathematical character – particular neurons in the visual system will only respond to particular orientations (a working abstraction), and optical illusions follow the pattern of statistical judgments (what we see is the most likely interpretation of the retinal image we have).  These probabilistic inferences are also used to model learning in a fairly general sense.

Riemann’s 1854 work on the foundations of geometry can be seen as an instance of a mathematician’s point of view being informed by ideas about perception.  We know this only because he acknowledged the influence of John Friedrich Herbart in his famous paper on the foundations of geometry.   Herbart held the view that space was not the thing that contained the objects around us, but rather a mental image constructed by any number of things seen in relation to each other (including time and color).  The combined influence of Gauss and Herbart moved Riemann’s thinking to propose that the concepts of measure and geometry could only have meaning if they began with the most general idea, a manifold, – any collection of elements that were related either discretely or continuously. This work is key to Relativity and points to both set theory and topology.

Mathematics has become increasing significant in the sciences.  David Deutsch’s recent work on Constructor theory can be used as an illustration of a mathematics (an algebra in this case) structured to produce, within itself, new statements that cannot be expressed in current physical theories but have new physical theoretical content.

The point of these observations is to suggest that mathematics has played the dual role of characterizing what we see as well as how we see it.  If we imagine that an effective body/brain would have a structure that somehow maps with, or couples with its world, then a very interesting question arises – how are we matched?  Mathematics may be in the best position to contribute an answer to this question.   It may be able to tell us something new about both ourselves, and about how the ‘thoughtful’ is found in nature.

In this way, mathematics is an inquiry like both art and science, looking at our experience and all that we see around us.  Perhaps it can be used to refresh our view of both.

 

Mathematics in the light of Maturana’s biology of cognition

As I have investigated all of the things in science and mathematics that get my attention, I have developed an impression of mathematics that, philosophically, seems most consistent with Humberto Maturana’s biology of language.  Maturana outlines his perspective in great detail in an essay by the same name that appeared in 1978 in the text Psychology and Biology of Language and Thought, edited by George Miller and Elizabeth Lenneberg.

Language must arise as a result of something else that does not require denotation for its establishment, but that gives rise to language with all its implications as a trivial necessary result. This fundamental process is ontogenic structural coupling, which results in the establishment of a consensual domain.

 

In a piece that appeared in Cybernetics and Human Knowing,  Maturana  says the following:

Living in language, doing all the things that we do in language, however abstract they may seem, does not violate our structural determinism in general, nor our condition as structure determined systems. As I observed our languaging behavior and the behavior of other animals, I realized that the central aspect of languaging was the flow in living together in recursive coordinations of behaviors or doings, and that notions of communication and symbolization are secondary to actually existing in language.

And also:

Language cannot be understood as a biological phenomenon if we do not take seriously our operation as structure determined systems. If we do not do so we remain trapped in the belief that language is a system of communication and thinking with representations (symbolizations) of an independent reality that contains us as its primary constitutive feature. And if we do not understand language as a biological phenomenon we shall remain in the mystery of self-consciousness through believing that this somehow reveals an intrinsic cosmic duality, and we shall not be able to understand ourselves as the self-conscious transitory beings that we are.

I see at least two important ideas that are needed to bear the weight of Maturana’s perspective.  The first of these is the dynamic interaction of unities:

…we human beings exist in structural coupling with other living and not living entities that compose the biosphere in the dimensions in which we are components of the biosphere, and we operate in language as our manner of being as we live in the present, in the flow of our interactions, in our domains of structural coupling.

The other is the shift away from questions about ‘being’ to an inquiry into ‘doing.’

As a result of this fundamental conceptual change, my central theme as a biologist (and philosopher) became the explanation of the experience of cognition rather than reality, because reality is an explanatory notion invented to explain the experience of cognition (see Maturana, 1980).

While there is no mention of mathematics in any of the Maturana pieces that I’ve looked at, it seems to me that the evidence is growing that mathematics happens, like language, and that we live in it, much like the way Maturana describes our “living in language.”   And I suspect that what we might not yet understand is the extent to which we share it, as it exists in manifold forms throughout nature.  Brain processing can look very mathematical (the abstractions in visual processing and learning, the Bayesian models of learning, etc.).  Foraging patterns, eye movement patterns and the patterns in how we search for words all look the same.  And, of course, our perception of the mathematical nature of the universe itself continues to enable an almost incomprehensible expansion of what we can know.

In a paper on How Humberto Maturana’s Biology of Cognition Can Revive the Language Sciences, Alexander Kravchenko takes note of some of the resistance to Maturana’s perspective but concludes with the following remarks:

One of the most important consequences of adopting the biology of language is the relational turn in approaching the mind/language problem. Much of what an organism does and experiences is centered not on the organism but on events in its relational experiential domain, one that crosses the boundary of skin and skull.  In its endeavor to answer the question “How does the brain compute the mind?” the neural theory of language overlooks the incoherence of the proposition that mind is a complex computational function of the brain. In the biology of cognition there is no such thing as “the mind” in the operation of the nervous system, and “the mind” is nothing but an explanatory notion: “language, self-consciousness and mindedness are different forms of existing in the relational domain in which a living being lives, not manners of operation of the nervous system” (Maturana, Mpodozis & Letelier 1995: 25).  (emphasis added)

 

It is this relational idea that can have a significant impact on a philosophy of mathematics,  as it will inevitably locate mathematics both in and around us.

The mathematics of common sense

I will be joining a few colleagues for a symposium at CogSci2014 and I’ve been gathering some notes for my talk.  The talk will focus on the impact of embodiment theories on a philosophy of mathematics.  As I looked again at some of the things I’ve chosen to highlight in my blogs, I came upon a talk given by Josh Tenenbaum, Professor in the Department of Brain and Cognitive Sciences at MIT.  I’ve read about aspects of his work before, but after listening to the talk he prepared for the Simons Foundation,  I became even more interested in the implications that his investigation of Bayesian models of cognition might have for mathematics.  One of the goals of the Simons Foundation talk was to highlight what Tenenbaum called “some of the new and deep math that has come out of the quest to understand intelligence.”  I  was particularly struck by his straightforward suggestion that statistics is not only intuitive, but that it may be part of our intuition.

The following remarks appear in the text introduction to the talk:

The mind and brain can be thought of as computational systems — but what kinds of computations do they carry out, and what kinds of mathematics can best characterize these computations?  The last sixty years have seen several prominent proposals: the mind/brain should be viewed as a logic engine, or a probability engine, or a high-dimensional vector processor, or a nonlinear dynamical system. Yet none of these proposals appears satisfying on its own. The most important lessons learned concern the central role of mathematics in bridging different perspectives and levels of analysis — different views of the mind, or how the mind and the brain relate — and the need to integrate traditionally disparate branches of mathematics and paradigms of computation in order to build these bridges.

…The recent development of probabilistic programs offers a way to combine the expressiveness of symbolic logic for representing abstract and composable knowledge with the capacity of probability theory to support useful inferences and decisions from incomplete and noisy data. Probabilistic programs let us build the first quantitatively predictive mathematical models of core capacities of human common-sense thinking: intuitive physics and intuitive psychology, or how people reason about the dynamics of objects and infer the mental states of others from their behavior.

There are a few themes in this talk that are worth noting.  There is certainly the suggestion that the brain’s computations are a kind of mathematics that happens within the body itself.  Tenenbaum was clear, however, about the extent to which brain processes are not understood.  No one yet knows how the brain actually learns, or how it translates symbols into meaning, nor how any of this can be understood with respect to the activity of neurons.  But he has been gathering evidence, on more than one front, that supports the idea that probability theory has a lot to say about “the things that the brain is good at” — like visual perceptions, learning the cause and effect relationships in the physical world,  understanding words and the meaning of actions…etc.

In one of the studies Tenenbaum discussed, Bayesian estimates that were computed from priors (defined by empirical statistics) were compared to the judgments that individuals made when asked to make predictions about those same variables.   Subjects in the study were asked to estimate things like the amount of money a movie will gross, a human life expectancy, or a movie’s run time.  For example, they might be asked, if you read about a movie that has made $60 million to date, how much will it make in total.  There were five groups of subjects, and each of the groups were given different numbers.  In our example one group may have been given $30 million, another $50 million, etc.  The prior for a parameter describes what is known, a priori, about the parameter being estimated.  In this case, what is known was gathered from empirical data.  And the distribution of priors for the various categories (movies runs, life expectancies, movie grosses) are qualitatively different from each other.  The data for one of the variables, for example, may produce a power distribution, while the data for another produces a Gaussian distribution.   You can apply Baye’s rule to each of these distributions to compute a posterior distribution – priors updated by experience or evidence.  The median of the posterior distribution provides what’s called a posterior predictive estimate.  In the studies Tenenbaum cites, applying Bayesian inference to the priors fits very well with the estimates people actually made, both quantitatively and qualitatively.
A recent paper entitled A tutorial introduction to Bayesian models of cognitive development, which Tenenbaum co-authored,  is a broad treatment of how and why Bayesian inference is used in probabilistic models of cognitive development. There, the point is made that the Bayesian framework is generative.  By generative is meant that the data observed has been generated by some underlying process.  And one of the values of Bayesian models is their flexibility.

…because a Bayesian model can be defined for any well-specified generative framework, inference can operate over any representation that can be specified by a generative process. This includes, among other possibilities, probability distributions in a space (appropriate for phonemes as clusters in phonetic space); directed graphical models (appropriate for causal reasoning); abstract structures including taxonomies (appropriate for some aspects of conceptual structure); objects as sets of features (appropriate for categorization and object understanding); word frequency counts (convenient for some types of semantic representation); grammars (appropriate for syntax); argument structure frames (appropriate for verb knowledge); Markov models (appropriate for action planning or part-of-speech tagging); and even logical rules (appropriate for some aspects of conceptual knowledge).
The representational flexibility of Bayesian models allows us to move beyond some of the traditional dichotomies that have shaped decades of research in cognitive development: structured knowledge vs. probabilistic learning (but not both), or innate structured knowledge vs. learned unstructured knowledge (but not the possibility of knowledge that is both learned and structured)p10

In his talk, Tenenbaum says that intelligence is about finding structure in data.   This is key, I think, to why mathematics has played so prominent a role in physics.  And, as Tenenbaum says, “it is the math, not a body of empirical phenomena that supports reduction and bridge-building.”  His talk for the Simons Foundation is aimed at highlighting the significance of mathematics in the study of cognition — from Bayesian inference to Bayesian networks (probabilistic graphical models) where arrows describe probabilistic dependencies and algorithms compute inferences.  His arguments lead to the development of probabilistic programming.  He also made a brief reference to mathematics being explored by some of his colleagues, that would serve to translate inferences into stochastic (random) circuits, suggesting a potential parallelism to the brain.

Neuroscience, Tenenbaum points out, is using the language of electrical engineering and common sense is at the heart of human intelligence.  He goes on to explain that the toolkit of graphical models is not enough to capture the causal processes underlying our intuitive reasoning about the physical world or our intuitive psychological/social reasoning.  To do this, one needs to define probabilities “not over a fixed number of variables, but over something much more like a program.”   Current work is focused on designing programs that can be run forward for prediction and backward for inference, explanation and learning.

You can find a nice account of recent work on probabilistic programing on Radar.

In Tenenbaum’s tutorial paper, he quoted Laplace (1816) who sort of summed things up when he said:

Probability theory is nothing but common sense reduced to calculation.

Beauty, Passion and Computation

I listened to a couple of interviews with Gregory Chaitin on the Closer to Truth website. They may have been part of TV episodes that I haven’t seen but I was actually invigorated by some of the things he said, and it made me want to share them.

One of the interviews (in two parts) is found under the heading Is Mathematics Eternal, and is contained within the topic What are the Deep Laws of Nature? Here, Chaitin seemed most intent on dispelling the standard view of mathematics, where mathematics is seen, perhaps, as the epitome of rationality, or as thought without feeling, or logical structure without any inherent meaning.  The discussion begins with a discussion of beauty.

We mathematicians are not machines,” he says, “we’re not calculating machines at all.  We’re human beings and we’re emotional. Why are we devoting our lives to mathematics?

His reply is essentially that mathematics seduces them, that mathematicians are passionate about mathematics.  He tells us that what many mathematicians are really looking for is beauty.  And here he uses a phrase that is likely filled with a number of thoughts worth exploring.  “It’s something sensual,” he says, “it’s the sensuality of ideas.”  The whole body is somehow participating.  And I believe that too, but how does one understand the sensuality of ideas?  What does it say about how mathematics is happening.  This sensuality may only be reached when one participates in mathematics at a certain depth.

Chaitin compared the effect that pure mathematics had on him as an adolescent boy, to the effect that women had on him.  “It was confused in my case,” he tells us.  The reason for the confusion, as he sees it, may be that beauty is in some way connected to the life force.  “One wants to create something beautiful, one wants to be illuminated by it.”  Chaitin uses the word illuminated more than once, and it is a particularly good choice of word.

When asked about what makes a proof beautiful, one of the things he said is, “at first it seems surprising and then it seems inevitable.  And you ask yourself, how come you didn’t see it sooner?” Even when speaking about a beautiful painting, Chaitin uses the phrase, “it illuminates you,” like a light coming out of the canvas.  He says the same of a beautiful proof, “it illuminates you.”  I think the image (in both cases), that something becomes lit, is a provocative image, referring to both light and understanding.  “A proof that isn’t illuminating is useless.”    The purpose of a proof, he explains, is not to prove something, it’s to give you understanding.  It should give you some intuition about why the result is right.

As the first segment nears its end, Chaitin brings up the mathematician Ramanujan.

I think Ramanujan is great because he contradicts everything that mathematics is supposed to be.

Chaitin tells us that Ramanujan believed that an equation is of value only if it expresses one of God’s thoughts. Ramanujan said of his own ideas that they were given to him, in his sleep,  by the goddess Namagiri.  Then Chaitin moves on to Cantor. “He is the most weird contradictory person.”  Cantor was trying to understand God.  He saw the contradictions inherent in his ideas but was not discouraged by them because, perhaps, his ideas were more important. Cantor was learning something about the infinite.  Chaitin also reminds us that Poincare referred to Cantor’s worked as a disease from which he hoped future generations recovered.  But Cantor’s work is foundational now.  Having been cleansed of some of its difficulties, and with Cantor’s more irrational inspiration hidden, it strongly influenced the course mathematics took in the 20th century.

Chaitin is making the argument that these peculiar aspects of Ramanujan’s and Cantor’s experiences are ignored because they contradict the way people expect to think about mathematics.

I prefer to take the other extreme.  I believe mathematics is based only on emotion and inspiration and it’s totally irrational and we don’t know where it comes from especially in cases where it’s an obsession.

Chaitin also recalled the work of Leonard Euler:

He created a lot of the mathematics that physicists and engineers use, but it was like a river, like a torrent of creative thought.  Every week he would do another wonderful paper.  He would give the whole train of thought and you would think when you read it, oh, I could do that.  But no.  Today a lot of his proofs are not considered rigorous.  But even though his proofs are not what today is considered a valid proof, he discovered all this mathematics.   Where did all those ideas come from?

In another segment, under the heading Is Information fundamental, Chaitin is addressing the question of whether information is  more fundamental to the universe than matter. The inspiration for these ideas may be the computer.  But what if the computer isn’t just a metaphor?  Chaitin asks.  A theory can be thought of as a computation.  You input your theory and the output is the physical universe, or mathematical theorems.  And when is a theory good?  This goes back to Leibniz, he says. A theory is good when it’s a compression, when what you put into the computer is simpler or smaller than what you get out.  Then you understand.  And that understanding can be mathematical or it can be physical.  And maybe the right way to think about the universe is that the universe is a computation  – computing its future state from its current state.  One can think of everything as a computation – understanding, the physical universe, DNA, as well as current technology.

Chaitin suggests that this provides a whole new way of looking at epistemology, which reach back to ideas presented by Leibniz after 300 years of development. We have reinvented an old question:  Is the universe built of matter or of mind?
I wrote a bit on Leibniz a couple of years ago.   I will revisit Leibniz and Chaitin for more on this.

Savants, neurons, and ants

Jason Padgett, author of the book Struck by Genius, appeared on CBS This Morning on April 24.  On May 5, livescience also did a piece on him and his book.  Padgett was assaulted in 2002 and suffered a severe concussion.  But, following this head injury, he acquired an extraordinary facility for seeing mathematics.  He is, as many say, an “acquired savant,” or prodigy in mathematics.
Padgett is quoted in the CBS article as saying:

I thought it was the pain medicine that they had given me that made me feel so strange….Things looked like individual picture frames coming in and clouds moving.  Instead of looking smooth, they looked like little tangent lines in a spiral.  Everything was discreet and chunky.

Padgett’s new skills are attributed, in part, to synesthesia, a condition where the senses blend – one might see music or hear color.  From what I’ve read so far, it seems to be his visual experience that is most changed.  But his visual experiences have motivated him to draw as well as do mathematics.  In fact, for him, the mathematics appears to be contained in the drawings (which, in itself, is noteworthy).

Livescience author Tanya Lewis reports it this way:

With Padgett’s new vision came an astounding mathematical drawing ability. He started sketching circles made of overlapping triangles, which helped him understand the concept of pi, the ratio of a circle’s circumference to its diameter. There’s no such thing as a perfect circle, he said, which he knows because he can always see the edges of a polygon that approximates the circle.

According to the livescience article, functional magnetic resonance images of Padgett’s brain showed significant activation in the left parietal cortex, although it’s not clear in the article what Padgett is doing when the activation occurs.  The parietal cortex is known to be an area where sensory information is integrated.  I considered the relevance of the parietal cortex to mathematics in my +plus article as well as in a post from a few years ago.

In a 2005 article Edward Hubbard, Manuela Piazza, Philippe Pinel and Stanislas Dehaene write on the interactions between number and space in the parietal cortex.  The authors conclude:

In the more distant future, it might become possible to study whether more advanced mathematical concepts that also relate numbers and space, such as Cartesian coordinates or the complex plane, rely on similar parietal brain circuitry. Our hypothesis is that those concepts, although they appear by cultural invention, were selected as useful mental tools because they fit well in the pre-existing architecture of our primate cerebral representations. In a nutshell, our brain organization both shapes and is shaped by the cultures in which we live.

But studies with Padgett went a bit further.  In her article Lewis reports:

…the fMRI only showed what areas were active in Padgett’s brain. In order to show these particular areas were causing the man’s synesthesia, Brogaard’s team used transcranial magnetic stimulation (TMS), which involves zapping the brain with a magnetic pulse that activates or inhibits a specific region. When they zapped the parts of Padgett’s parietal cortex that had shown the greatest activity in the fMRI scans, it made his synesthesia fade or disappear, according to a study published in August 2013 in the journal Neurocase.

Berit Brogaard,  who performed these studies, has shown in another study that “when neurons die, they release brain-signaling chemicals that can increase brain activity in surrounding areas.”  The increased activity may not last, but if it produces structural changes, then the brain-activity changes would persist.  Lewis doesn’t develop this idea but, for me, it calls to mind the kind of interaction ecologist Deborah Gordon described in her TED talk: What ants teach us about the brain, cancer and the internet. There Gordon describes how systems without a central control are directed by interaction.  Ants operate within a network created by their interaction.  For example, in situations where there may be a high cost to venturing out, the decision to forage for food can be a tricky one.  Ants seem to add up the stimulation they receive from other ants to know whether they should begin.  She compares the effectiveness of this interaction to neurons that add up the stimulation from other neurons to know whether to fire or not.

Author and psychiatrist Darold Treffert interviewed Padgett.  Padgett writes:

…he told me that these innate skills are, in his words, “factory-installed software” or “genetic” memory.  After interviewing me…..he also suggested that all of us have extraordinary skills just beneath the surface, much as birds innately know how to fly in a V-formation and fish know how to swim in a school.

While the experience of the savant is rare, and often coupled with disabilities of another kind,  I think these mysterious talents provide a unique opportunity to explore the enigma of how we come to know anything.  Perceiving very large quantities, in the absence of counting; having a sensory experience of numeric relationships; playing an instrument without ‘learning’ it, all these things happen.  Treffert recounts some of his encounters in a very nice piece which appeared as a Scientific American Guest Blog last July.   These remarkable abilities must be showing us something about the nature of the disciplines involved as well as the potential of brain functions.

In his book, Padgett says:

To me, a tree is more than its geometry, but geometry is also far more than most people realize.   I think it’s everything.

And in the CBS report he says:

It means that we all have this ability within us,” he said. “I had no prior training whatsoever, and, like, we’re all doing complicated math and physics. … If somebody throws a ball and you catch it, you calculated the gravitational field, everything, and you caught that ball. Your brain- if you wrote that as an equation, that’s a serious equation to write, but you could just catch the ball automatically.

Shakespeare, art, religion and mathematics

I recently considered the role that mathematics plays in bringing meaning, or perhaps even story, to our experience. Mathematics is often used to reveal the structure that can be found in large sets of data, or in any number of physical things that change over time,  or in the properties of the abstractions themselves.  Mathematics, then, sort of tells us what we’re looking at, within what would otherwise be an unwieldy or even meaningless data.   In my last post, I took note of the way that mathematics was used to communicate meaning in the structure of cathedral walls – in particular Lincoln Cathedral whose construction began at the end of the 11th century and continued through the Middle Ages. There I shared the suggestion that the geometry of the Cathedral was used to convey, or even to instruct on, the fundamental nature of our being.  It happens now that another New Scientist article, Shakespeare: poet, playwright, scientist broadened the issue for me.

The article was written by broadcaster and author Dan Falk, whose most recent book is The Science of Shakespeare. In the article, Falk considers, specifically, evidence in the play’s words that Shakespeare made use of astronomical observations, and explores how he may have known about them.

 

Galileo’s telescopic observations came toward the end of Shakespeare’s career, but Cymbeline, first performed in 1611, offers tantalizing hints that he was aware of the findings.  In the play’s final act, the hero, Posthumus, falls into a dream-like state, and the ghosts of four family members appear and move around him in a circle.  The ghosts cry out for Jupiter, the Roman god.  On hearing their pleas, he descends onto the stage.  So we have Jupiter and four ghosts moving in a circle.

 

But the book also addresses the emergence of the scientific world view more generally and, in doing so, takes note of the relationship between science and religion.  I recently saw a group of students who had gathered on the University of Texas, Dallas campus with posters  designed to provoke a debate about truth, science and religion.  When I saw them, it occurred to me that mathematics never comes up in these debates, despite the fact that it is so intimately tied to ancient, medieval and modern world views.  Reading the history of the relationship between science and religion, as Falk surveys it, encouraged me to pursue my own thoughts about whether an updated philosophy of mathematics might contribute to the science vs. religion debate. Early in his book, Falk makes the following observations:

…the rediscovery of classical texts, via Arabic translations, triggered a new wave of learning across Europe. Those works included the writings of Aristotle and Ptolemy…,as well as the geometry of Euclid, the medical writings of Galen, and much more…This wave of learning was closely linked to the activities of the Roman Catholic Church. The best medieval schools had been those associated with the monasteries and the great cathedrals. By the late Middle Ages these had also become centers of what we would now call science…There were also the universities, the earliest having been founded around 1200; these, too, functioned largely as religious institutions. The highest degree offered was in theology – though to obtain it, the student also had to master mathematics, logic, and natural philosophy.

The close connection between science and faith may seem strange to the modern reader, living at a time when Western society, and Western science in particular, has become a secular endeavor….The evolving relationship between science and religion is a large and complex subject…For one thing, religion was simply part of the fabric of society; all of the key figures of the Scientific Revolution were men of faith of one kind or another.

…As historian Paul Kocher puts it, early modern science “was more often cited as proving God’s existence than disproving it…And as Principe writes, the study of nature was seen as “an inherently religious activity…But the link between science and faith, as Principe stresses, is deeper than this. For the thinkers of early modern Europe, he writes, “the doctrines of Christianity were not personal choices. They had the status of natural or historical facts.”

Mathematics inspired many of the thoughts of an emerging scientific world-view, but it is also closely associated with the more mystical side of things. Kepler, as Falk points out, was obsessed with numerology, like the ancient Pythagoreans. His work relies on observations of the physical as well as the mystical. Quoting Allen Debus, Falk tells us:

His subtle mixture of mathematics and mysticism is “far removed from modern science, but it formed an essential ingredient of its birth.”

Mathematics is often at the beginning of things. The considerations of Shakespeare’s Hamlet explored in Falk’s article, and in his book, should remind us of the complexity of human thought.  Relationships among art, science, and religion are inherent.  They are all narratives that grow out of the same ground. Their coexistence, it would seem, is just a fact of our existence, and one of the keys to a deeper understanding of who we are.

If mathematics is, as I tend to see it, the mind itself building structure with the elements of thought, then mathematics’ development is like the development of another sense.  It is, after all, the structure we give to any sensory data that creates a meaningful thing perceived. And so it is with mathematics.  Mathematics, like story and name, brings meaning, not mechanics, to our experience.  It increasingly links what we see to what we think we know, as we try to reconcile our immediate experience with what’s ‘out there.’  The evolution of mathematical concepts allows us to probe deeper and deeper into the universe, as well as into our own nature.  And this is the subject of both religion and science.  So why is mathematics so consistently absent from our debates?