Complexity: Simply Wonderful!

Featured

Chaos theory has been an immensely popular field of study that took roots in the 1960’s and ’70’s. At its heart, it appeals to the seeker in all of us, putting a limit on how much, in general, we can hope to predict. Why are some things easy to foretell, and why others are difficult, like, famously, the weather. The field also gave rise to some terms that captured public imagination – for instance Emergent behaviour, when the whole is more than the sum of its parts, Strange Attractors, when systems are driven to inevitable but unpredictable outcomes, or the Butterfly Effect, alerting us all, regardless of discipline, to the fact that small changes could have potentially huge consequences.

The impact of this field on almost all areas of science (and the social sciences) had been significant, so the 2021 Nobel prize in Physics, awarded to Parisi, Manabe and Hasselman for their work in the area of complexity, was entirely fitting, bringing much needed public recognition of the value of studying complex systems. By then, one might have argued, it was a no-brainer, given that climate change was beginning to occupy a major part of public discourse, and mankind had begun, belatedly, to realize that we were all in it together, that the climate was related to to the atmosphere, but also to physics, to chemistry, to geology, to economics, to agriculture and to society at large. Solving problems of this nature requires one to embrace the complexity, so to speak – the interrelatedness of these issues is intrinsic to the problem, and it cannot be simplified away.

When the IIT Madras magazine Shastra asked me if I would review Parisi’s book that had been published recently by Penguin, it was therefore another no-brainer. Having read a fair amount by Giorgio Parisi – mostly his scientific papers, truth be told – I was looking forward to learning what he might say in a book that he explicitly wrote for the public at large, In a Flight of Starlings: The Wonders of Complex Systems (Giorgio Parisi, Penguin Press, New York, 2023).

The review appears in the latest Shastra, and much of it is reproduced below, thanks to the kindness of the editors.

15

In 2021, Giorgio Parisi was awarded the Nobel Prize in physics for the discovery of the interplay of disorder and fluctuations in physical systems from atomic to planetary scales. He shared the prize with Syukuro Manabe and Klaus Hasselmann who earned the prize for the physical modelling of Earth’s climate, quantifying variability and reliably predicting global warming.

This was, arguably, the first award of a Nobel prize in the area of complexity: in their citation, the Foundation starts off by asserting, “Our world is full of complex systems characterised by randomness and disorder.”, and then adding (for Manabe and Hasselman) “One complex system of vital importance to humankind is Earth’s climate”. For Parisi, they said “[…] Giorgio Parisi discovered hidden patterns in disordered complex materials. His discoveries are among the most important contributions to the theory of complex systems.”Given that nonlinear science, complexity, disorder and order have all been areas of my own research interest, I am biased to be sure, but this is a book that is well worth the read. Parisi has, arguably, been one of the most influential (and cerebral) statistical physicists in the world, and the award of the Nobel comes at the tail of a long list of honours, of which there will likely be more.

The different areas that bear his imprimatur are many, the KPZ equation (for Kardar-Parisi-Zhang) in the study of aggregation, the Altarelli-Parisi equations in particle physics, the introduction of the multifractal formalism in turbulence, spin glasses, or the quantitative study of the murmuration of starlings (which feature in the title of the book and its cover).

feynman

This set of essays conveys the essence of Parisi’s discoveries to the general reader, for the most part quite successfully. Each fairly short, the chapters in the book fall into three categories. The general theme of complexity in physical systems, namely order and disorder in nature – phase transitions, emergence, spin glasses, flocking behaviour – is discussed in a few. Some personal history, on what it was like to be doing physics in Italy in the 1960’s and ’70’s is shared in a couple of chapters, and in some others Parisi shares his rumination on the nature of creativity, on metaphors in science, and a final word on the meaning of it all, why we do science in the first place. (Spoiler alert: we do it for fun! Parisi quotes Richard Feynman, “Science is like sex: sometimes something useful comes out, but that is not the reason we are doing it.”)

Reading a formal scientific paper by Parisi has never been very easy – there are tools and techniques that he pulls out of unfamiliar hats – but it has often been rewarding. This book is probably the only one that Parisi has written for a lay audience, and is based in part on a set of interviews with the science communicator Anna Parisi (who is not related to him). First published in Italian in 2021 as “In un volo di storni” the book was translated by Simon Carnell, and has an easy conversational style, keeping the frank matter-of-factness of Parisi as he describes how he developed ideas, partly by analogy, partly by metaphor, but mostly but hard work and application.

70

Parisi’s style is very direct. He describes the problems that interest him scientifically without either exaggerating their importance or dumbing down the language. For instance, the murmuration of starlings is a captivating sight, but there are deep questions that this phenomenon poses: is there a leader or is the behaviour self-organised, how is information transmitted through the flock, and so on. As he describes the work that he and his group carried out, one has a sense of the process of discovery, of what it takes to study such phenomena, and enough detail is shared to make one also feel like a participant and not just a bystander. The same goes for the chapters on phase transitions and spin glasses. The problems are articulated clearly, the history of the area is described in sufficient detail, and the explanations come with a number of diagrams but no equations (keeping Stephen Hawking’s warning in mind no doubt, that for every equation his readership would halve!). The basic ideas of the replica trick or of scale invariance are conveyed in simple language, and without obfuscation.

Although this works admirably in making the physics of complex systems very understandable, one might argue that the more valuable part of the book is the more reflective and analytical set of essays on the value of doing science. These essays were written prior to 2021, and it is clear that Parisi has long occupied an important and influential position as a public intellectual and as a voice for Italian science. Parisi was elected to the American Philosophical Society in 2013, and that seems entirely fitting for a person who has done much to reveal the commonality between diverse areas in physics (or nature), namely the idea of universality. He also speaks unselfconsciously about the powerful role of incubation and intuition in scientific discovery, about the role of the subconscious and “nonverbal” thinking. Some of these ideas are presented in an exploratory and tentative manner, but this only makes the ideas more approachable and adds to the charm.

Science advocacy is another running theme through the essays. Starting off, Parisi makes his position clear, that it is “essential that the public have a fundamental understanding of the practice of science” since “Our generation is on a road fraught with dangers. It is as if we were driving at night: the sciences are our headlights, but it is the responsibility of the driver to not leave the road and to take into account that the headlights have a limited range. Understanding the limitations of the sciences is as important as (or maybe even more than) believing in the value of science: Parisi is an articulate and honest spokesman for best practices in this domain. “Science needs to be defended not just for its practical aspects but for its cultural value” he says in the final pages, emphasising a point that is all too often ignored: it is not just for its value in making life better or in facilitating technological advances, but because it is integral to our modernity.

440px-E._M._S._Namboodiripad

This is something that, as it happens, I have felt strongly about for a while, and have in fact written about here in this blog, after I had attend a meeting in memory of EMS Namboodripad in Thrissur, Kerala in 2018. This talk appeared in the Indian Academy of Sciences online journal Dialogue: Science, Scientists, and Society as Science in the Public Sphere: Dissemination, Discussion, and Dialogue and eventually as a perspective in the Economic and Political Weekly in 2020 (vol. 60: pages 33–36) titled Science in the public sphere: Obligation and responsibility.

Continue reading “Complexity: Simply Wonderful!”

Ship of Fools

Narrenschiff_(1549)
A German woodcut from 1549 depiction of the ship of fools. From Wikimedia.

With all the discussion going on about the efficacy (or not) of online education, one should acknowledge that for over a decade now the holy triumvirate of Wikipedia, YouTube, and Google have been educating many of us online. May not be for degrees and such (although truth be told, for that too I suppose). So what the Big W tells me about the title of this post is that “the ship of fools is an allegory, originating from Book VI of Plato’s Republic, about a ship with a dysfunctional crew. The allegory is intended to represent the problems of governance prevailing in a political system not based on expert knowledge, such as democracies.” This cuts very close to the bone: although we nominally are a democracy, the lack of reliance on expert knowledge is telling on how we are managing the COVID-19 crisis.

As the pandemic progresses, the situation is gradually becoming pandemonic. At times like this, one’s thoughts turn to the arts, with Messiaen’s Quartet for the End of Time providing an ideal background score. (Link courtesy the Big Y, of course.)  My mind, admittedly frayed, has been pushed into incomprehension over the past few months as steadily extending lockdowns gnaw at my innards, the angst eating out my soul. Never before have I felt as alienated from the everyday workings of the country as I have felt over the past few weeks, as the separation between us, the people, and them, those who take and enforce decisions on us. Governments are, we like to believe, for us, by us, and of us, but I am more convinced that this one is really against us, it cannot understand the people.

c1_3580750
From The Bangkok Post

How does one count the many ways in which we and our beloved country – for which we are now beyond weeping – have been brutalized?   In how many acts of commission and omission have we been made complicit, as we watch our own backs and feel there, but for the grace of God, go I.

…….

I have stayed at this point in the post for days now, unable to write and move forward since it all seems so pointless. Every day the numbers rise, every day there is no direction. Every day dawns with a new policy which has no substance and which is effectively reversed by night. And every day it seems that there are problems on so many fronts, they simply don’t have a clue as to what to do.

The strange part is we are intellectually self-reliant and have been for quite some time, long before the call to this new self-reliance.  The raw material that can steer us through the pandemic is here: it has been built up over the past several decades. One only has to have the humility to listen to experts. We have, and it is no secret, the finest minds on the planet in myriad fields that are of value in our response to this public health crisis – mathematical analysis, modelling, statistics, economics, sociology, epidemiology, virology, genomics, medicine, engineering, you name it. But these minds work best when their research is funded consistently over the years, not just in spurts. And when they are given full independence, to say what they honestly believe. Drugs may be repurposed from one disease to another, but academics cannot be repurposed in the same way.  

perfectstUnfortunately, therefore, it does seem that we are riding this storm in a ship of fools. And on this ship, many hands do what they imagine that the captain demands and tell him what he likes to hear.

That is simply no way to steer through this mess. The waters are very, very choppy: regrettably for us, this is a perfect storm.  We will come through this, I imagine, but we will emerge enervated, weakened beyond recognition. I hope, though, that we will not either forget, or forgive.

The Idgah of Kharera

I am reminded today of Constantine Cavafy‘s poem In the same space, written in 1929 (and in the  translation of  Rae Dalven)

The surroundings of the house, centres, neighbourhoods
Which I see and where I walk; for years and years.

I have created you in joy and in sorrows:
Out of so many circumstances, out of so many things. 

You have become all feeling for me.

Today is an anniversary that I mark in some measure of solitude, with some sadness and with some sense of incredulity that it has been so long already.  The imagery of Cavafy perfectly reflects the way I feel, in remembering the joy a little bit more, and in remembering the sorrows just that much less.

Idgah_of_Kharehra
The Idgah, from Wikimedia Commons

I am fortunate to live in the proximity of the Idgah of Kharera, in my eyes a symbol of resurrection and hope, and in these times it can take on a meaning well beyond its own intentions. Each day, when I look upon it, each day (and I say this with no sense of exaggeration) it brings me some measure of calm. And in this solitude, I learn that the word that better describes how I feel is sukoon,Untitled.

I am fortunate to have the Idgah so close at hand, with a palash tree to my left, a huge neem to my right, red flowers strewn on the ground ungathered on my left, and the yellow leaves of the neem amassing on the right. Through the day the birds come and go. Hornbills, woodpeckers, parakeets, peafowl, tree-pies, bulbuls, mynahs, and of course, the ubiquitous pigeons. This Idgah is a living monument. 

IMG-20200123-WA0002.jpgThe park is also integral to the community that lives around here in Hauz Khas and that I now live in. “Which I see and where I walk”, if not for years and years, for long enough to identify. Morning walks, taking the dogs out, just coming to get some quiet… a lifeline for the regulars that frequent it on a daily basis. Closed to the public from sunset to sunrise, the monument is illuminated until 11 pm even now. And on full moon nights, it can be quite breathtaking. My bedroom window view is truly blessed.

Entry No 116 in INTACH’s Delhi The built Heritage  will tell you that this Idgah was built around the year 1404 by Mallu Iqbal Khan, who effectively ruled Delhi along with Nasiruddin Mahmud, the last of the Tuglaq dynasty.  The description is prosaic: The mosque consists of a west wall, battlemented and containing 11 mihrab recesses. It is originally terminated with circular bastions. At the north of the central mihrab there is a high pulpit reached by 13 steps, below which is an arched opening, a feature common in all Idgahs. 

Timur_defeats_the_sultan_of_Delhi

Built at a time when Delhi had been destroyed six centuries ago, an inscription on the south bastion of the Idgah mentions the desolate condition of Delhi after the invasion of Timur, who defeated the army of Sultan Nasir-ud-Din Mahmud Shah Tughluq and Mallu Iqbal on 17 December 1398.

The description of the battle is quite comical, though the outcome was not. Timur’s camels, loaded with hay that had been set alight, ran towards the Tuglaqs’ armoured elephants, causing a rout. That must have been quite a sight, and one can imagine the confusion and the pandemonium. In the event, Timur took Delhi in triumph. The Encyclopaedia of Islam says that Delhi was sacked, left in ruins, and Timur executed about a lakh of the inhabitants. In the aftermath, Mallu Iqbal Khan became the de facto ruler of Delhi till he was killed in battle in 1405, just after the Idgah was built, apparently in the region where Timur had camped outside the city of Siri.

IMG_20180924_090025121

But this is not about the history of Delhi, or even about the history of what feels like my Idgah.  During the two days I spent in Samarkand a couple of years ago, I had enough time to go around and see Timur’s footprint everywhere, still firm and unchanged. He was a great tactician, and his destruction of Delhi – one of the greatest cities in the world at that time – was so drastic, history tells us it took a century for the city to recover.

So too this pandemic shares that ruthlessness of purpose, and the virus is in its own way, a great tactician. Absolutely egalitarian, it has destroyed, if not the city, the smugness of our modernity. It has altered the fabric of our society, exposing our faultlines, exposing our prejudices and our privileges. There was a battle, and we lost it at the start. There was pandemonium in the way that so many people fled the city. There was chaos and there is confusion.

When we come back, and history tells us that we indeed will, one can only hope that there has been a radical change in the way we think, the way we live and clean, and the way we interact. I hope though, that it will not change our empathy, or the way we remember, or  the way we love. It will take some time to recover, but recover we shall.

The Natural Effectiveness of Mathematics in the Biological Sciences

download
Wigner, E. P., The unreasonable effectiveness of mathematics in the natural sciences. Commun. Pure Appl. Math., 1960, 13. 

Some years ago, too many for my reckoning, I was invited to contribute an article for a special issue of the journal Current Science (Bangalore), on the use of mathematics in  different scientific disciplines. I had occasion to read the article again after some 15 years, mainly to cannibalize it for a talk I had to give yesterday, I should confess. Some embarrassment is inevitable on reading something one has written some time ago (I have almost never looked at my Ph D thesis, for example) but I thought that some of it could be shared, so here is an abbreviated essay where I have not removed all the dated bits… The title, of course, acknowledges a great thinker and physicist, Eugene Wigner.

An increasingly quantitative approach within the biological sciences has been accompanied by a greater degree of mathematical sophistication. However, there is a need for new paradigms within which to treat an array of biological phenomena such as life, development, evolution or cognition. Topics such as game theory, chaos theory and complexity studies are now commonly used in biology, if not yet as analytic tools, as frameworks within which some biological processes can be understood. In addition, there have been great advances in unravelling the mechanism of biological processes from the fundamental cellular level upwards that have also required the input of very advanced methods of mathematical analysis. These range from the combinatorics needed in genome sequencing, to the complex transforms needed for image reconstruction in tomography. In this essay, I discuss some of these applications, and also whether there is any framework other than mathematics within which the human mind can comprehend natural phenomena.

It is a commonplace that in recent years the biological sciences have gradually become more quantitative. Far from being the last refuge of the nonmathematical but scientifically inclined, the modern biological sciences require familiarity with a barrage of sophisticated mathematical and statistical techniques.

By now the role of statistics in biology is traditional, and has been historically derived from the need to systematize a large body of variable data. The relation has been two- sided: biological systems have provided a wealth of information for statisticians and have driven the development of many measures, particularly for determining significance, as in the χ2 or Student’s-t tests. Indeed, Galton’s biometrical laboratory was instrumental in collecting and tabulating a plethora of biological measurements, and these and similar data formed the testing ground for a number of statistical theories.

The role of mathematics in biology is more recent. The phenomenal developments in experimental techniques that have helped to make biology more quantitative have necessitated the applications of a number of different mathematical tools. There have been unexpected and frequently serendipitous applications of techniques developed earlier and in a different context. The widespread use of dynamic programming techniques in computational biology, of stochastic context-free grammars in RNA folding, hidden Markov models for biological feature recognition in DNA sequence analysis, or the theory of games for evolutionary studies are some instances of existing methods finding new arenas for their application. There have also been the mathematics and the mathematical techniques that derive inspiration from biology. The logistic mapping, the discrete dynamical system that is so central to chaos theory, arose first in a model of population dynamics. Attempts to model the human mind have led to the burgeoning field of artificial neural networks, while the theory of evolution finds a direct application in the genetic algorithm for optimisation.

Mathematics is about identifying patterns and learning from them. Much of biology is still most easily described as phenomena. The underlying patterns that appear are nebulous, so extracting a set of rules or laws from the huge body of observations has not always been easy. Or always possible since some experiments (like evolution) are unrepeatable, and separating the essential from the inessential can be very difficult. Detail is somewhat more important in the life sciences: often it has been said that the only law in biology is that to every ‘law’, there is an exception. This makes generalizations difficult: biological systems are more like unhappy families. With the exception of natural selection, there are no clearly established universal laws in biology.

This is, of course, in sharp contrast to the more quantitative physical sciences where the unreasonable effectiveness of mathematics has often been commented upon. It might be held that these observations, coming as they do in the twentieth century, comment on a science that has already had about three centuries of development. The earlier stages of the fields that we now call physics or chemistry were also very poorly described by mathematics—there was no general picture beyond a set of apparently unrelated observations, and it required the genius of a Mendeleev, of a Faraday or Maxwell or Einstein to identify the underlying patterns and expose the mathematical structure that lay under some aspects of these fields. This structure made much of the modern physical sciences possible, and led to some of the most accurate verifications of the laws of physics. As predictive theories, relativity and quantum electrodynamics are unparalleled and have achieved astonishing accuracy. In a more complex setting, the seemingly infinite possibilities of organic chemical reactions have found organizational structure in the Woodward–Hoffman rules that combine an elementary quantum mechanics with notions of graph theory to make precise, semiquantitative predictions of the outcome of a large class of chemical reactions. What will it take to similarly systematize biology? Or to rephrase the question, what will the analogous grand theories in biology be?

The inevitable applications of mathematics are those that are a carry-over from the more quantitative physical sciences. As in the other natural sciences, more refined experiments have spearheaded some of these changes. The ability to probe phenomena at finer and finer scales reduces some aspects of biology to chemistry and physics, which makes it necessary to borrow the mathematics that applies there, often without modification. For instance, tomographic techniques rely on a complicated set of mathematical transforms for image reconstruction. These may be largely unknown to the working biologist who uses NMR imaging, but are a crucial component of the methodology, nonetheless. Similarly, the genome revolution was catalysed by the shotgun sequencing strategy which itself relied on sophisticated mathematics and probability theory to ensure that it would work. Several of the problems in computational biology arose (or at least were made more immediate, and their resolution more pressing) by the very rapid increase in experimental power.

The other sort of application of mathematics is, for want of a better descriptor, a systems approach, namely that which is not predicated by the reductionist approach to biology but instead by a need to describe the behaviour of a biological entity in toto.

Even the simplest living organisms appear to be complex, in way that is currently poorly described and poorly understood, and much as one would like, it is not possible to describe in all totality the behaviour of a living organism in the same way as one can the behaviour of, say, a complex material. The promise that there could be mathematical models that capture the essence of this complexity has been held out in the past few decades by several developments, including that of inexpensive computational power which has made possible the study of more realistic models of biological systems. Theoretical developments—cellular automata, chaos theory, neural networks, self-organization— have provided simple mathematical models that seem to capture one or the other aspect of what we understand as ‘complexity’, which itself is an imprecise term. There is one class of applications of mathematical or physical models to biology which attempt to adapt an existing technique to a problem, while another aims to develop the methods that a given problem needs. Each of these approaches have their own value and appeal. In the next sections of this article, I discuss some of the ways in which they have found application in the study of biological systems.

Richard_Hamming
Hamming, R. W., The unreasonable effectiveness of mathematics, Am. Math. Monthly, 1980, 87.

The resonance of the title with those of the well-known essays by Wigner and Hamming is deliberate, as is the dissonance. There are applications in the physical sciences where knowledge of the underlying mathematics can provide very accurate predictions. Comparable situations in the biological sciences may not arise, in part because it may be unnecessary, and in part because biological systems are inherently unpredictable since they are so fundamentally complex. The demands, as it were, that are made of mathematics in the life and physical sciences are very distinct, and therefore, it is very reasonable that the mathematics that finds application in the two areas can also be very different.

Is there any framework other than mathematics within which we can systematize any knowledge? Recent advances in cognitive studies, as well as information that is now coming from the analysis of genomes and genes, suggest that several aspects of human behaviour is instinctual (or ‘hardwired’). That mathematical reasoning is an instinct that we are endowed with is a distinct possibility, and therefore, it may not be given to us (as a species) to comprehend our world in any other manner. This point of view, that it is very natural that we should use mathematics to understand any science, is explored below.

In the last few years there has been a veritable explosion in the study of complex systems. The concept of complexity is itself poorly defined (‘the more complex something is, the more you can talk about it’ ), and as has been pointed out by others, ‘If a concept is not well-defined, it can be abused.’ Nevertheless, there is some unity in what studies of complexity aim to uncover.

A common feature of many complex systems is that they are composed of many interconnected and interacting subunits. Many systems, natural as well as constructed, are, in this sense complex. Examples that are frequently cited apart from those involving living organisms such as ecologies or societies, are the human brain, turbulent flows, market economies or the traffic. A second feature of complex systems is that they are capable of adaptation and organization, and these properties are a consequence of the interconnection and interactions of the subunits. The mathematics of complex systems would thus appear a natural candidate for application to biology. The drawback is that there is, at present, no unifying framework for the study of complex systems although there are some promising leads offered by studies of dynamical systems, cellular automata and random networks.

That the description of phenomena at one level may be inadequate or irrelevant at another has been noted for a long time. Thus the electronic structure of atoms can be understood quite adequately without reference to quarks, and is itself irrelevant, for the most part, when dealing with the thermodynamics of the material of which the atoms are constituents. Schrödinger, in a chapter of his very influential book (Schrödinger, E., What is Life?, Cambridge University Press, Cambridge, 1967) entitled ‘Is Life Based on the Laws of Physics’, observed that with regard to ‘the structure of living matter, that we must be prepared to finding it working in a manner that cannot be reduced to the ordinary laws of physics’. He further contrasted the laws of physics and chemistry, most of which apply in a statistical sense, to biological phenomena, which, even though they involve large numbers of atoms and molecules, nevertheless have nothing of the uncertainty associated with individual properties of the constituent atoms. Indeed, given a radioactive atom, he says, ‘it’s probable lifetime is much less certain than that of a healthy sparrow’.

But even at a given level, it frequently happens that the properties of a system cannot be simply inferred from those of its constituents. The feature of emergence, namely the existence of properties that are characteristic of the entire system but which are not those of the units, is a common feature of systems that are termed complex.

Distinction should be drawn between the complex and the complicated, though this boundary is itself poorly defined. For instance, it is not clear whether or not in order to be deemed complex, a system requires an involved algorithm (or set of instructions). The algorithmic complexity, defined in terms of the length of the (abstract) program that is required is of limited utility in characterizing most systems

starlingAttempts to decode the principles that govern the manner in which new properties emerge—for example the creation of a thought or an idea, from the firing of millions of neurons in the brain, or the cause of a crash in the stock market from the exit poll predictions in distant electoral constituencies—require new approaches. The principles themselves need not necessarily be profound. A simple example of this is provided by a study of flocking behaviour in bird flight. A purely ‘local’ rule: each bird adopts the average direction and speed of all its neighbours within distance R, say, is enough to ensure that an entire group adopts a common velocity and moves in unison. This behaviour depends on the density of birds as well as the size of R relative to the size of a bird in flight. If R is the size of a bird, then each bird flies on its own path, regardless of its neighbours: there is no flock. However, as R increases to a few times that of the bird, depending on the density, there can be a phase transition, an abrupt change from a random state to one of ordered, coherent, flight. And such a system can adapt rapidly: we have all seen flocks navigate effortlessly through cities, avoiding tall buildings, and weaving their way through the urban landscape at high speed.

boulezBut there are other aspects of complexity. A (western) orchestra, for instance, consisting as it does of several musicians, requires an elaborate set of rules so that the output is the music that the composer intended: a set of music sheets with the detailed score, a proper setting wherein the orchestra can perform, a specific placement of the different musical instruments, and above all, strict obedience to the conductor who controls what is played and when. To term this a complex system would not surprise anyone, but there is a sense in which such a system is not: it cannot adapt. Should the audience demand another piece of music, or music of another genre, an orchestra which has not prepared for it would be helpless and could not perform. Although the procedure for creating the orchestra is undoubtedly complicated, the result is tuned to a single output (or limited set of outputs). There is, of course, emergence: a single tuba could hardly carry a tune, but in concert, the entire orchestra creates the symphony.

Models like this illustrate some of the features that complex systems studies aim to capture: adaptability, emergence and self-organization, all from a set of elementary rules. The emphasis on elementary is deliberate. Most phenomena we see as complex have no obvious underlying conductor, no watchmaker, blind or not who has implemented this as part of a grand design (Dawkins, R., The Blind Watchmaker, Norton, New York, 1996). Therefore, in the past few decades, considerable effort has gone into understanding ‘simple’ systems that give rise to complex behaviour.

logistic‘Simple mathematical models with very complicated dynamics’, a review article published in 1976 (May, R. M., Simple mathematical models with very complicated dynamics. Nature, 1976, 261, 459) was responsible in great measure for the phenomenal growth in the study of chaotic dynamics. In this article—which remains one of the most accessible introductions to chaos theory— May showed that the simplest nonlinear iterative dynamical systems could have orbits that were as unpredictable as a coin-toss experiment. The thrust of much work in the past few decades has been to establish that complex temporal behaviour can result from simple nonlinear dynamical models. Likewise, complex spatial organization can result from relatively simple sets of local rules. Taken together, this would suggest that it might be possible to obtain relatively simple mathematical models that can capture the complex spatiotemporal behaviour of biological systems. 

A number of recent ambitious programs (eCell, A multiple algorithm, multiple timescale simulation software environment, http://www.e-cell.org) intend to study cellular dynamics, metabolism and pathways in totality, entirely in silico. Since the elementary biochemical processes are, by and large, well-understood from a chemical kinetics viewpoint, and in some cases the details of metabolic pathways have also been explored, entire genomes have been sequenced and the genes are known, at least for simple organisms, the attempt is to integrate all this information to have a working computational model of a cell. By including ideas from network theory and chemical kinetics, the global organization of the metabolic pathway in E. coli has been studied computationally. This required the analysis of 739 chemical reactions involving 537 metabolites and was possible for so well-studied an organism, and the model was also able to make predictions that could be experimentally tested. The sheer size of the dynamical system is indicative of the type of complexity that even the simplest biological organisms possess; that it is even possible for us to contemplate and carry out studies of this magnitude is indicative of the analytic tools that we are in a position to deploy to understand this complexity.  

In recent years, there has been considerable debate, and an emerging viewpoint, that the human species has an instinct for language. Champions of this school of thought are Chomsky, and most notably, Steven Pinker who has written extensively and accessibly on the issue

pinker
Pinker, S., The Language Instinct, Morrow, New York, 1994

The argument is elaborate but compelling. It is difficult to summarize the entire line of reasoning that was presented in The Language Instinct, but one of the key features is that language is not a cultural invention of our species (like democracy, say), but is hard-wired into our genome. Like the elephant’s trunk or the giraffe’s neck, language is a biological adaptation to communicate information and is unique to our species.

Humans are born endowed with the ability for language, and this ability enables us to learn any specific language, or indeed to create one if needed. Starting with the work of Chomsky in the 1950s, linguists and cognitive scientists have done much to understand the universal mental grammar that we all possess. (The use of stochastic context-free grammars in addressing the problem of RNA folding is one instance of the remarkable applicability of mathematics in biology.) At the same time, however, our thought processes are not language dependent: we do not think in English or Tamil or Hindi, but in some separate and distinct language of thought termed ‘mentalese’.

Language facilitates (and greatly enriches) communication between humans. Many other species do have sophisticated communication abilities—dolphins use sonar, bees dance to guide their hive mates to nectar sources, all birds and animals call to alarm and to attract, ants use pheromones to keep their nestmates in line, etc.—and all species need to have some communication between individuals, at least for propagation. However, none of these alternate instances matches anything like the communication provided by human language.

It is not easy to separate nature from nurture, as endless debates have confirmed, but one method for determining whether or not some aspect of human behaviour is innate is to study cultures that are widely spaced geographically, and at different stages of social development. Such cross-cultural studies can help to identify those aspects of our behaviour that are a consequence of environment, and those that are a consequence of heredity. The anthropologist Donald Brown (Brown, D., Human Universals, Temple University Press, Philadelphia, 1991) has attempted to identify human ‘universals’, a set of behavioural traits that are common to all tribes on the planet.

All of us share several traits beyond possessing language. As a species we have innumerable taboos relating to sex. Some of these, like incest avoidance, appear as innate genetic wisdom, but there are other common traits that are more surprising. Every culture, from the Inuit to the Jarawa, indulges in baby talk. And everybody dreams. Every tribe however ‘primitive’, has a sense of metaphor, a sense of time, and a world view. Language is only one (although perhaps the most striking) of human universals. Other universals that appear on the extensive list in his book, and which are more germane to the argument I make below, are conjectural reasoning, ordering as cognitive pattern (continua), logical notions, numerals (counting; at least ‘one’, ‘two’ and ‘many’) and interpolation.

The last few mentioned human universals all relate to a set of essentially mathematical abilities. The basic nature of enumeration, of counting, of having a sense of numbers is central to a sense of mathematics and brings to mind Kronecker’s assertion, ‘God made the integers, all else is the work of man’. The ability to interpolate, to have a sense of a continuum (more on this below), also contribute to a sense of mathematics, and lead to the question: Analogous to language, do humans possess a mathematics instinct?

21_TH_SESHANHENRI_POINCARE
Poincaré, H., Mathematics and Science: Last Essays,
Dover, New York, 1963.

Writing a century ago, Poincaré had an inkling that this might be the case. ‘… we possess the capacity to construct a physical and mathematical continuum; and this capacity exists in us before any experience because, without it, experience properly speaking would be impossible and would be reduced to brute sensations, unsuitable for any organization;…’ The added emphasis is mine; the observations are from the concluding paragraph of his essay, ‘Why space has three dimensions’.

If mathematics is an instinct, then it could have evolved like any other trait. Indeed, it could have co-evolved with language, and that is an argument that Keith Devlin has made recently (Devlin, K., The Maths Gene, Wiedenfeld and Nicolson, London, 2000).

At some level, mathematics is about finding patterns and generalizing them and about perceiving structures and extending them. Devlin suggests that the ability for mathematics resides in our ability for language. Similar abstractions are necessary in both contexts. The concept of the number three, for example, is unrelated either to the written or spoken word three, or the symbol 3 or even the more suggestive alternate, III. Mathematical thought proceeds in its version of mentalese.

An innate mathematical sense need not translate into universal mathematical sophistication, just as an innate language sense does not translate into universal poetic ability. But the thesis that we have it in the genes begs the question of whether mathematical ability confers evolutionary advantage, namely, is the human race selected by a sense for mathematics?

download
Wilson, E. O., Consilience: The Unity of Knowledge, A. A. Knopf,
New York, 1998.

To know the answer to this requires more information and knowledge than we have at present. Our understanding of what constitutes human nature in all its complexity is at the most basic level. The sociobiologist E. O. Wilson has been at the vanguard of a multidisciplinary effort toward consilience, gathering a coherent and holistic view of current knowledge which is not subdivided in subdisciplinary approaches. This may eventually be one of the grand theories in biology, but its resolution is well in the future. We need to learn more about ourselves.

Traditionally, any sense of understanding physical phenomena has been based on having the requisite mathematical substructure, and this tradition traces backward from the present, via Einstein, Maxwell and Newton, to Archimedes and surely beyond.

Such practice has not, in large measure, been the case in biology. The view that I have put forth above ascribes this in part to the stage of development that the discipline finds itself in at this point in time, and in part, to the manner in which biological knowledge integrates mathematical analysis. The complexity of most biological systems, the competing effects that give rise to organization, and the dynamical instabilities that underlie essentially all processes make the system fundamentally unpredictable, all require that the role played by mathematics in the biological sciences is of necessity very different from that in the physical sciences.

Serendipity can only occasionally provide a ready-made solution to an existing problem whereby one or the other already developed mathematical method can find application in biology. Just as, for example, the research of Poincaré in the area of dynamical systems gave birth to topology, the study of complex biological systems may require the creation of new mathematical tools, techniques, and possibly new disciplines.

F1.largeOur instincts for language and mathematics, consequences of our particular evolutionary history, are unique endowments. While they have greatly facilitated human development, it is also worth considering that there are modes of thought that may be denied to us, as Hamming has observed , similar to our inability to perceive some wavelengths of light or to taste certain flavours. ‘Evolution, so far, may possibly have blocked us from being able to think in some directions; there could be unthinkable thoughts.’ In this sense, it is impossible for us to think non-mathematically, and therefore there is no framework other than mathematics that can confer us with a sense of understanding of any area of inquiry.

In biology, as Dobzhansky’s famous statement goes, nothing makes sense except in the light of evolution. To adapt this aphorism, even in biology nothing can really make sense to us except in the light of mathematics. The required mathematics, though, may not all be uncovered yet.

1917. It was a Very Good Year.

thWhat Einstein said of Mahatma Gandhi, that generations to come will scarcely believe that such a one as this ever, in flesh and blood, walked upon this earth, is more than applicable to Einstein himself. From 2005 – declared by the United Nations as the World Year of Physics, to celebrate the centenary of Einstein’s annus mirabilis – onward, there have been many occasions to mark one hundred years of one or the other incredible contribution of his.  For Einstein, many years were very good indeed.

UntitledWhat makes 1917 special in some ways is the appearance of three other papers, each unrelated to the other (as the three of 1905) and which altered the fields that they touched upon. It would well have been called another annus mirabilis, had not 1905 already happened.

Early on in Volume 6 of the Collected Papers of Albert Einstein titled The Berlin Years: Writings, 1914-1917, his “Inaugural Lecture”, delivered upon his election to the Prussian Academy of Sciences, he alludes to the soon-to-be-presented papers on his General Theory of Relativity, saying “We have determined that inductive physics has questions for deductive physics and vice versa; and eliciting the answers will require the application of our utmost efforts. May we, by means of united efforts, soon succeed in advancing toward conclusive progress.” That conclusive progress was to appear in the journal Annalen der Physik in May 1916, entitled “Die Grundlage der Relativitätstheorie” (On the Theory of General Relativity). But that was 1916.

The first of 1917’s three gems was On the Quantum Theory of Radiation, wherein he  came up with stimulated emission and laid the foundations of laser physics. The second was Cosmological Considerations in the General Theory of Relativity, in which he set the foundations of modern cosmology.

UntitledAnd the third. On May 11, 1917, Einstein presented a paper to the German Physical Society,  and this was published on the 30th of the same month, in the journal Deutsche Physikalische Gesellschaft. Verhandlungen,  19, 82-92 (1917). Titled On the Quantum Theorem of Sommerfeld and Epstein, this paper essentially anticipated Hamiltonian chaos and its implications for quantum mechanics, the field of Quantum Chaos. Considering that it was written before wave mechanics and the Schrödinger equation, this paper is remarkable, also because Einstein seems to have explicitly understood the quantum implications of classical nonintegrability.

torus.png

As he puts it, if one examines any volume element in configuration space, any given orbit can pass through that region infinitely many times, either (a) with a (few) well defined values of the momentum – as on a torus – or (b) “there are infinitely many [values of the momentum] at the location under consideration“. In other words, small changes in the positions can correspond to very different momenta. Another way of saying something similar had to wait for Ed Lorenz in the 1960’s, who termed this as sensitive dependence on initial conditions, or what we term today as classical chaos.

The quantum condition he derives (and which now goes by the name of Einstein-Brilloiun-Keller-Maslov or EBKM quantization) uses the classical invariants identified by Poincaré, and Einstein goes on to give, in his view, the proper quantum conditions (11) that correspond to integrals along independent paths on n-dimensional tori (as in the figure above).

Untitled

But being Einstein, he “notices immediately that type b) excludes the quantum  condition we formulated” earlier in the paper: this is the insight that was to lead half a century later, to the beginnings of the field of quantum chaos, the knowledge that there were classical motions for which quantum conditions could be stated, and those for which it was not possible, at least not in the same way. More can be read about “Einstein’s Unknown Insight and the Problem of Quantizing Chaos” in Douglas Stone’s article in Physics Today in 2005.

I first came across this paper of Einstein’s in 1978 or 1979 as a postdoc when I was struggling through Arnold’s text on Classical Mechanics and working on semiclassical mechanics. It may (OK, does) not count among Einstein’s greatest works, but arguably it is the one that has had the greatest impact on my own research. Other works of Einstein have had a much bigger impact on all our lives, of course, but this one is a paper whose centenary I’d like to mark.

Untitled 2And the other two as well. 2017 is the centenary of another annus mirabilis, a smaller one than 2005 perhaps, but enormous by any other standards.