Cite as in K.Pribram, ed, Origins: Brain and Self-Organization, Hillsdale NJ: Erlbaum, 1994, p.680-706.

 

Additional comments since publication appended to the end. (See the book for the actual figures.)

 

 

 

              The Brain as a Neurocontroller:

New Hypotheses and New Experimental Possibilities

 

 

                                                                                          Paul J. Werbos

                                                                    Room 675, National Science Foundation*

                                                                             Arlington, Virginia, USA 22230

                                                                                        pwerbos@nsf.gov

 

This paper will describe how a new body of mathematics -- initially motivated by neuroscience but developed in recent years through engineering applications -- can begin to yield a predictive, empirical understanding of the phenomenon of intelligence in the brain. The paper is mainly written for neuroscientists, or for engineers working with neuroscientists; it tries to describe crucial new experiments which need to be performed in order to test and refine this new understanding.

                The biggest single obstacle to the full use of mathematics in real neuroscience is the sheer difficulty of the relevant mathematics. The brain is far more complex than today's computers; therefore to understand it, one must use even more sophisticated mathematics than the average research engineer is familiar with. Because of this difficulty, a few "middle men" have presented oversimplified description of biology to the engineers, and oversimplified descriptions of the engineering to the biologists. These oversimplifications have often led to considerable misunderstanding and justified mistrust.

                Because of these communications problems, this paper will be written in an extremely informal style. It will consist mainly of the transcript of a one-hour talk, edited for readability, with a few critical updates inserted. The first section will explain the fundamental approach, and move directly to the "bottom line" -- to some specific areas where new experiments are badly needed. The next two sections will discuss the underlying theory and mathematics in more detail. The second section will discuss the issue of supervised learning, which can shed light on local circuits within the brain. The final section will discuss the major concepts of neurocontrol, which can shed light on the global organization which unifies these local circuits into a truly intelligent system.

                                                                                                       

 

INTRODUCTION AND OVERVIEW

 

Goals of This Talk

 

I really am grateful to speak for once to an audience that is said to have a lot of physiologists in it. I wish I had more chances to do this, because I think that some of the things that we've learned on the engineering side lead to some very interesting experimental possibilities on the physiological side; if we had more chances to talk to each other, we could learn a lot more about experiments which nobody is doing which could lead to some very exciting results in the future. That is what I would really like to talk about today. 

 

 _______________________________________________________________________________

*The views herein are those of the author, not those of NSF; however, as government work, it is in the public domain conditional upon proper citation. This is an updated version of a paper in Computational Neuroscience Symposium 1992, edited by M.Penna, S.Chittajalu and P.G. Madhavan, available from Madhavan at the Electrical Engineering Department at IUPUI in Indianapolis, Indiana.

 

 

 

                Now, because it is late in the day, I figured it might be useful for me to summarize everything I am going to say in one list, so that you can see that it is finite, anyway. I'm basically going to try to make four major points today:

 

(1) First, I'm going to argue that we can understand intelligence or the brain in the same kind of mathematical way  that we understand physics, as a real science. I'm not saying we're there yet, but I think it can be done.  

 

(2) Second, I'm going to argue that neurocontrol gives us new mathematics, which is the mathematics we need in  order to understand the brain mathematically. 

 

(3) I'm going to argue that neurocontrol has made enormous progress in the last few years, in terms of new  engineering applications, new mathematical designs and ideas, and new links to the brain.  Jim Bower has described this process as a kind of convergent evolution.  If you look at the simple‑minded neural nets you see in a lot of the neural net conferences, they don't have much connection to biology. But when you look at people who have to solve really difficult, hard engineering control problems, they're driven to some of the same  complexities we observe in the brain.  So I would argue concretely there are signs of convergent evolution. 

           

(4) Finally, most important, is that what we now have learned about what the brain might be doing suggests new opportunities for experiment. It suggests some surprising predictions. If the predictions are right, then you can  use experiments to surprise a lot of people and have fun changing the culture, and if they're wrong, you can surprise a lot of mathematicians and come up with some new computational principles that people think are impossible. So either way, it's really important. 

           

A caveat here is that as an NSF program director I'm not telling you that I've got a lot of money for this. In fact, I'm not allowed to spend money on things other than neuroengineering; my present budget is too small to allow anything more. I think that this is a very unfortunate situation, because if we're going to try to understand the human mind and human learning -- subjects of truly enormous importance -- then we have got to bring these things together; but right at the moment, there's essentially zero dollars available for the specific kind of two-way cooperation I'll be talking about today. I really wish somebody could fix that.  (As this book goes to press, The Biology Directorate at NSF is preparing a Collaborative Research Initiative which could help fill this vacuum; however, the exact role of Engineering in that initiative is not yet clear.)

                If this were an audience of policy people, or people who talk to their congressman, I would spend ninety percent of my time up here on items number one and two on my list.  I could spend a good hour on this  -- on the theory and the philosophy and all of that. If this were an engineering audience, I would talk about the applications and the designs; I have done that for about eight hours at a stretch.  But here I am going to try to jump ahead to the brain stuff, but this is a little risky.  You have to bear in mind that the kind of mathematics that's relevant to the brain is not the easy stuff.  The kind of math you can totally understand in twenty minutes‑‑that isn't what relates to the brain. The brain is a little more complicated, so I'm going to have to jump over some stuff and give some citations. 

 

Can we Understand the Brain Mathematically?:

Prospects for a Newtonian Revolution

 

Before I get going, though, I really do want to say a little bit about the generalities here. 

                I suspect that a lot of the people in neuroscience started out by wanting to understand the human mind.  They really wanted to understand something fundamental and important. But then they ran into a problem. Do you remember the old saying:"When you're up to your knees in alligators, it's hard to remember that your goal was to drain the swamp."?  All of us have that problem, from time to time. I suspect that a lot of neuroscientists discover, as time goes on, that the brain is so complex that they lose hope of figuring it out in their own lifetimes.  Some 

people have made a formal philosophy of that; they say, "look, the information content in any one brain is more complex than what I have spare neurons to understand, so by definition I cannot understand another brain, let alone everybody's brain." 

 

                But let us think about that idea a little more carefully.

                If you try to know all of the synapse strengths, the connections, the state of all the networks in somebody's brain, and the reverberatory dynamics -- then of course, that is too complicated to ever understand in your life. There is no way that all of those details can be fully known scientifically. There will always be lots and lots of islands of understanding, and those islands are useful.  We've seen good examples of studying connections even here today. But they don't tell you how intelligence works as a whole system. They're just little islands.  And that is very discouraging. 

                But think back, how did physicists solve this problem, how did physics become a science? Basically there was this guy Isaac Newton, and what did he do? Instead of trying to describe every physical object in the universe, physics gave up on that, and they said "let us try to understand instead the simple underlying dynamics which change all of that complicated stuff over time."  Maybe all of these complicated things are governed by something simple enough you can understand it.  In physics, "simple enough" meant a page of equations and a thousand pages of explanation -- not trivial, but understandable. 

                My argument is that the same kind of approach could work on the brain if you think of learning as the dynamics.  There is every reason to believe that underneath the complexity in the cerebral cortex and so on, there is a generalized, modular plasticity. Lashley has shown this, and I've heard of recent experiments where they've trained linguistic cortex to develop edge detectors just by wiring it up differently. It's very clear that there is a uniform, generalized modularity there in the interesting parts of the brain, which ought to be understandable if we focus on the learning, the plasticity[1]. Knowing the laws of learning would not immediately tell us a lot of specialized things about how we process specific sensory inputs in specific ways, but physicists have found that if you understand the underlying dynamic laws that control everything else, that's incredibly important later on when you try to do engineering. 

                So let us try to see if we can create ‑‑ I think we can create, in principle -- a Newtonian revolution, by focusing on the basic laws of learning in the high‑level, modular organs like the  cerebral cortex, the limbic lobes, the cerebellum, the olive, and so on. We won't ever understand the motor pools that way; they're like ad hoc preprocessors and postprocessors. But the really important stuff we can understand, in principle. 

                But of course you can't do that unless you have the right math.  Newton had derivatives. Well, we too have a new flavor of derivatives; that's what backpropagation is really all about, at its base [2], but backpropagation is only one small part of this large area that I'm calling neurocontrol. Years ago, there was great work by David Robinson comparing conventional control theory with what goes on in the vestibular system. His early work was a really great example of cross-disciplinary collaboration, really applying mathematical concepts in detail, but it did not really focus on the issue of learning (unlike his newer work). It really scares me that the time delay between developing the necessary mathematics and control theory in the engineering context and applying it to the eye tracking system may have been something like a 20‑40 year time delay.  Our new math -- neurocontrol -- is only about two years old, in the engineering world. I hope we don't have to wait 40 years before we start to apply it to the brain!  If we follow the normal course of government  funding and human inertia, we might well wait 40 years; however, if we work a little hard to do what's unnatural, the mathematics is available today, if we have the will to go ahead with it. 

                I'll talk more about these issues later on, when I discuss recent progress in neurocontrol.

 

Neuroengineering and Neuroscience:

What is the Basis for Collaboration?

 

Let me move ahead now to the first slide (Figure 1 on the next page).  This is again a generality slide.  Like everybody else here I'm arguing that we need interdisciplinary cooperation, but I'd like to say a little bit about where the problem is, because we need to do more than just say  interdisciplinary cooperation is needed; we need to have a concrete image in our heads of what it's about, or else we'll never be able to implement it. 

                A lot of people are excited because folks in the neuroscience side of the world studying the brain are now using neural network models.  They are building up the field of computational neuroscience, which still belongs on the left-hand side of Figure 1. In computational neuroscience,  we describe the brain by use of differential equations or other mathematical models, instead of just verbal anecdotes and whatnot.  That's exciting.  On the right-hand side of figure 1, in neuroengineering, we are using neural network systems to solve real‑world engineering problems; that's also very exciting.  

 

 

 

                                                                 Figure 1. An NSF Definition of Neuroengineering

               

                But the problem is this: what is the connection between the left and right sides of Figure 1?  Even in today's symposium, which is very interdisciplinary, it is pretty easy to classify most of the talks into who is doing computational neuroscience and who is doing engineering applications.  It's  like a gulf.  And what's the problem?          What's happening is that we're both using neural network models, but one group is using as its standard of validation: "Does the model fit the low‑level circuit and the empirical data down at the low‑level circuit?"  Maybe more than that.  But in engineering, the test is: "Does it work?"  So we have two different communities, based on two different standards of validation.  But in reality, the brain itself meets both tests.  The real circuits not only fit their own biological data, they also work in solving very complex control challenges.  Instead of having two communities, using two different standards of validation to inspire and to evaluate their work, we need to think of using both standards of validation together.  And that's how we can get feedback back and forth here. I won't elaborate on this today; this is just a matter of general principles. As I said before, unfortunately, my tiny bit of money is entirely on the neuroengineering side, and that's something that needs to be changed. 

 

Neural Nets and Neurocontrol:

Where Is the Right Mathematics?

 

A lot of people are worried that the artificial neural network (ANN) community, the engineering community, is itself caught in a kind of local minimum. It is true that 90% of the papers you see in a neural network conference these days talk about pattern recognition, and what are they actually doing?  Usually, they are doing pattern classification, using associative memory or other simple systems. Usually they are "training" ANNs to match databases which contain definite targets for what the output of the ANN should be, for every single example in the database.  There are lots of uses for this kind of task.  But that's not intelligence.  That's not consciousness, that's not what the mind does.  We humans are not just simple classification machines! This really ought to be obvious to anyone.

                This situation is kind of scary; you have to ask what is the relevance of that stuff?  Now, I'm not going to talk today about consciousness or the mind/body problem; if I'm brave at SMC on 

Monday I'll talk about that[3,4,5], but here I'm going to focus on physiology. 

                If we agree that neuroengineering has been caught in a kind of local minimum or intellectual rut, then what is the way to get out of that local minimum? If you forgive a pun, I will argue that we can get out of that local minimum by climbing out, by climbing up a ladder -- and here's the ladder (Figure 2, on the next page), the ladder of designs of neurocontrol. 

                Again, let me warn you that this is just a quick overview; I'll be giving you citations to more detailed information later on. There are many, many designs in this emerging area of neurocontrol, which I define as the use

 

 

                                                                 Figure 2. The Ladder of Designs in Neurocontrol

 

of well‑specified neural nets -- either natural or artificial, just mathematically‑defined neural nets -- to generate control outputs, which could be to motors, muscles, glands, stock transactions or whatever, but real actions in the real world. 

                In the neurocontrol field, we do have very simple designs, and these are the most popular. They're easy to do; they're a great start for people who want to get their students going, and start to build up software. This is the right place to start, but these designs do have very limited power, and they certainly are not like the brain. 

                In the middle level of the ladder, we have what I call the state‑of‑the‑art group, and I'd say there are about

four groups that are really in this category. It's curious that industry is here more than  academia; I don't know why. Are university people scared to do new things? I don't know, but these state‑of‑the‑art groups have mostly taken a couple of years to build up; maybe that's the problem, that you've got to keep your students around long enough, and build up modular software packages. After a couple of years of struggling, these groups have gotten real‑world applications just this year, of things that were only on paper two or three years ago.  And they have proven -- with really exciting, important applications -- that these more advanced, more brain‑like methods are far more powerful in solving real-world problems.  There are just incredibly important engineering problems that have been solved that are in the mill; again, today I won't talk about this a lot today, but I may make some reference to it.  

                After these methods on paper were used in real applications, it was a challenge to us theorists to move ahead of the applications people and come up with new methods to overcome the limits of the older ones, so that now on paper this year there are new methods which did not exist two or three years ago.  And now, on paper, it looks as if these designs and ideas really have the potential to achieve true brain‑like intelligence. So my bottom line is that at least on paper we now  have the math we need to understand real intelligence.  I'm not saying that these ideas are working yet on real systems, and that's what I try to pay people to do, to climb up this ladder with real engineering systems.

                By the way, I'm saying that the bottom level of the ladder is a good place to start, but when I fund people, the higher they can go up the ladder, the higher the probability of funding.  They may have to climb one step at a time, but they had better be moving upwards in a visible way. I'm trying to develop the engineering math that will be necessary to understand the brain.  I'm using engineering as a discipline to get the math we need for what's really interesting, which is the mind and the brain. 

 

Four New Empirical Possibilities: A Summary

 

Now before getting into the intricacies of neurocontrol, I would first like to give you my real bottom line. I would like to summarize four empirical areas where I think new experimental work could be really crucial. I will try to explain the reasoning behind all this in more detail later, but for now I will just give a summary:

 

(1) First,  I'm going to argue that some form of backpropagation -- not the simple three‑layer kind that most people have seen, but a more complicated, advanced form of backpropagation -- almost certainly must exist in the brain in order to explain some of the capabilities that we have observed there.  That in turn suggests that we have to look for some novel mechanisms, to carry information backwards both within and between cells. Between cells, it is now well-known that nitric oxide (NO) acts as a backwards transmitter. In addition, a group of researchers including Timothy Bill -- one of the important pioneers in Long-Term Potentiation (LTP) [6] -- has discovered a new presynaptic receptor intimately related to LTP[7]. (The group speculates that this receptor may be involved in adapting the nearby synapse, but there is no reason to believe that this is its only function.) Back in 1974, after I had developed the backpropagation algorithm, I speculated that the cytoskeleton might take care of the backwards flows within cells [2]; this still appears to be a viable possibility [5,8,9,10], but there is new evidence that the usual kinds of field effects in membranes could also be involved[11]. David Gardner has shown that such backwards mechanisms are crucial to learning even at the level of aplysia[12,13]. Nevertheless, all of this is only just a beginning. 

                There is a lot of engineering work needed in this area, both in theory and in instrumentation.  It's really frightening to me, when I look at how critical the cytoskeleton is in the nervous system (it is like half the nervous system!), to see that the amount of work that's been done understanding how the cytoskeleton relates computationally (or might relate) is negligible.  We don't yet know that it's relevant, but we don't yet know that it's irrelevant either. It is amazing to me that we can just sit back and ignore it and give it maybe ten thousand a year, when we're spending a billion dollars on the other half, when we don't know what it does is.  It's really frightening; we really need to be studying the cytoskeleton in any case, and backpropagation is just one of the things to look for when we do it.

                In looking for backpropagation, you don't necessarily have to look at the cytoskeletal level. There are other kinds of experiments you can do, where region A has a forward fiber to region B (e.g., A might be a part of the limbic zone and you move on to something like the motor cortex), and sometimes you can find that the plasticity in A seems to depend on what happens in B. It would be interesting to see if you could cut the fiber from A to B and then see if you lose the plasticity in A.  There's no way that could happen in a classical neuron model that's all feedforward and membrane‑driven, but if it does happen then that means that you can unhinge the neuron model. I have tried to persuade Karl Pribram to look into experiments like that, and his (informal, not for scientific publication) response was "I've already done it, I've already proven this." 

                Pribram's response was really very interesting to me. If you ask a lot of the middlemen between the neural network field and biology, they'll tell you that this is impossible; however, when I ask Pribram he says it's already been proven, that there is a backpropagation there.  I don't know whether to take his informal statements at face value yet; I think we need a lot closer collaboration to evaluate those experiments to see what they mean mathematically, but it's clear there is a lot to be done here. 

 

(2) True reverse engineering of hippocampal and other slices. In the talk by Sclabassi earlier today, we heard some very exciting things about the hippocampus. It was particularly fascinating to hear that the kind of learning you get from LTP clearly doesn't represent the real nonlinearity of the system.  I would speculate that appropriate slices through the brain can generate model systems that you can play with like artificial neural nets, where you can control the inputs and outputs. Why is it that when we do experiments in neural systems we try to always do them under natural conditions?  If we think that biological neural circuits are general purpose learning machines, then let's play with them! 

                Let's see if we can use a slice of neural tissue to learn to recognize an 

arbitrary pattern that hasn't been seen in nature.  Let's find out what are the capabilities.  Let's find out what the plasticity is in these more micro, more mathematical ways.   And I would speculate,  for example, that a slice through hippocampus and cerebral cortex that maintains those local recurrent links will have a better learning capability, in a sense I hope to have time to define, than  any of the Hebbian or backpropagation, feedforward nets that are in use today. 

                In other words, there are two classes of nets people are using a lot -- the classic Hebbian, the Grossbergian nets, and then maybe the multi‑layer perceptron (MLP) nets; I'm willing to bet that there are critical learning problems which I hope to talk about, which that kind of slice can solve better than any of the nets people now believe in, on the biological or the engineering side. Once you prove this, empirically, then I have some ideas for what is going on there, but the experiments are what's crucial for now. I think if you do the experiment, you will shake up a lot of people, and then they'll start thinking about those more powerful designs that we're just now starting to look at in engineering. 

                There is a whole lot to be done in this area. Once you have taken the first step -- demonstrating and describing plasticity on the slice -- you can then start looking for the learning mechanisms that underlie that plasticity.  So in a way this might be a good place to begin before getting into some of the harder issues I discussed earlier.

                Similar kinds of experiments could also be done in culture, if the right kinds of cells can be grown together in culture. Many biologists worry that cell cultures (and even slices) are very artificial. It can be dangerous to draw too many conclusions from what we see in culture, because the presence of other cell types and inputs in the brain could lead to very different kinds of behavior. Nevertheless, when groups of cells in culture do succeed in demonstrating certain kinds of engineering capabilities - such as the ability to learn to approximate mathematical functions more complex than those which Hebbian or MLP nets can learn -- then we probably can conclude that these cells possess these capabilities (or more) in nature, in the brain. There may be great value in figuring out what kinds of cells need to be present, as part of a culture, to generate what kinds of learning capabilities.

 

(3) A third area has to do with the inferior olive, which governs learning in the cerebellar system.  I am told by Pellionisz that Llinas and his group have observed plasticity in the inferior olive, which is crucially related to the cerebellum and lower‑level motor control.  I haven't looked at the experiments myself, but based on a very careful examination of the cerebellum, working jointly with Pellionisz  (not with tensor theory, working with Pellionisz on some new ideas), it is my conclusion  that something unusual is going on[14]. 

                There are two possibilities  ‑‑ or rather, I'm predicting one of two possibilities. Both of them are very surprising. First of all, before doing the experiment proper, the first stage is to replicate the phenomenon of plasticity in the olive. Then you have to cut one of two fibers, and show that cutting those fibers eliminates the plasticity; this would narrow down the plasticity to one of two possible mechanisms. (The two fibers are: (1) the climbing fibers; (2) the collateral fibers from the deep cerebellar nuclei and vestibular nucleus to the olive.)

                I hope that somebody can do this experiment soon. This may well be the most finite and do‑able thing on this whole list here.  So I really hope somebody looks at this. I have described that in more detail at the end of a recent paper[14]. This is an important experiment nobody has done‑‑I don't think it should be that hard.  And it is really critical to our next step in understanding what the cerebellum is doing.  

                After this talk was given, I found out that the first step -- of simply replicating plasticity in the olive -- itself a serious challenge. The original experiments by Llinas et al, reported in Science in 1975, are still highly controversial. Furthermore, there are certain learning tasks -- like those described by Richard Thompson -- which do not seem to elicit plasticity in the olive. (Just as most physical tasks only require the use of a few muscles, so too do many learning tasks exercise only a part of our learning abilities.) Hockberger and Alford at Northwestern University, working in communication with James Houk, in a small grant supported by NSF, have begun to explore learning mechanisms in olive cells, at a molecular level, in culture. It is predicted that olive cells can display plasticity in culture, but that it will be crucial to include enough other cell types (e.g. deep cerebellar nuclei cells) and to provide appropriate learning challenges through appropriate stimulation of the system; this may or may not be possible within the time-frame of this small start-up project. It is truly amazing that such a basic, important issue has been left unresolved for so many years. (Similar gaps exist in the present understanding of the nucleus basalis, another small but crucial piece of brain circuitry.)

                A more technical issue, crucial to working out the fine points of this system, is the ability of the cerebellum to learn time sequences and delays [14]. This ability clearly depends on certain short-term memory capabilities of Purkinje cells, but it is very tricky to design a circuit which reproduces such capabilities. (See chapter 13 of [15].) Tam, at the University of North Texas, has begun some learning experiments in cultures which might shed light on this issue[16]; again, however, the efforts so far are only a crude beginning.

 

(4) Fourth, there is room for more true reverse engineering of the cerebellar motor system. Suzuki, Kawato et al[17] have done a magnificent job in getting this area started, but a lot more needs to be done. Suzuki et al have basically shown that the lower motor system is doing optimal control, not adaptive control in the classic sense, and not translation between different kinds of coordinate systems, but optimal control.  I think that someone could play with that circuit a lot more than anyone has done so far. Suzuki et al, and Houk, think they know where the reward or utility functions are coming in from; if they are right, we could perturb these inputs and prove what the power is of this system in optimization, in adapting to new regimes. Again, we could play with the lower motor system, by perturbing its inputs to see what capabilities it has as a general‑purpose optimizer. 

 

                In brief, I have described four general areas where new kinds of experiments could be extremely useful. I don't know if I'm describing the tasks in exactly the right way. This is just an attempt to get the process started. I'm just a dumb engineer, as they often say.  But I think that something  needs to be done to get us moving into these new kinds of areas, and there is some theory behind the ideas above.

 

 

SUPERVISED LEARNING: RECENT ANN RESULTS

AND IMPLICATIONS FOR NEUROCONTROL AND BIOLOGY

 

Supervised Learning As A Neural Net Paradigm

 

A lot of people in neuroengineering get upset when I talk about control applications and control, because a lot of people in the artificial field really have this old idea (illustrated in Figure 3), that supervised learning is the same as neural network theory.  They think that neural network theory is the same as learning a map from an input vector X to a target vector Y, in hopes that in the future you'll be able to predict the right target vector.  And you go through training sets and you learn over and over again what this mapping is. 

 

 

                                                                   Figure 3. What Does Supervised Learning Do?

 

                In fact, if you look at the Granger‑Lynch model of the hippocampus (arguably the best existing model of the hippocampus as an associative memory) ‑‑ that's another form of supervised learning; it's just plain old pattern classification you're studying.  Supervised learning tends to be an all-pervasive paradigm, even for biologically motivated research. Many people tend to think that supervised learning is fundamental theory, and that anything else is just dirty applications.

 

Supervised Learning versus Neurocontrol

 

Supervised learning is certainly useful, and it may well exist in subsystems of the brain, but it turns out that for really powerful control systems, you have to do stuff that is a lot harder. 

                What you have to do is stuff like this (see Figure 4 on the next page). When I'm giving a tutorial on how to do real neurocontrol in engineering, it turns out that I have to spend an hour or two on each one of the three main boxes in Figure 4.  You do have supervised learning systems in these designs, but they're like little modules.  And then you have a big system, a neurocontrol system, that takes these lower level modules and integrates them and links them.  I often compare this situation to how we build computers: there's a lot of science to building the chip, but there's a lot of science to putting chips together to making a computer. Supervised learning is a general purpose concept, but neurocontrol is also general-purpose and fundamental; they simply address different general-purpose tasks.

                It turns out that the work that's been done in neurocontrol at these multiple levels of organization has parallels to the brain, at multiple levels of organization.  So the stuff we've learned down at the supervised learning level tends to be relevant to issues like what is the circuitry like within the cortex, within recurrent nets, or within the cerebellum, while the higher-level stuff is important when we try to figure out the organization that connects those systems.  So that means I should talk about both of these levels and explain them before I talk about the brain. So I should spend eight hours before it all becomes crystal clear.  Forgive me, it won't be quite as crystal clear as I like, because I don't have the eight hours.  

 

 

 

 

                                                               Figure 4. Four Task Areas Critical to Neurocontrol

 

Three Supervised Learning Modules Used Today in Neurocontrol

 

     First of all let me talk at the low level, at the supervised learning level.  Little nets that learn pattern recognition. What has been useful in engineering? 

     Basically, there are three kinds of networks that people really use in real‑world control 

applications:

 

(1) The most common is the multilayer perceptron (MLP). (See Figure 5 on the next page.))  Please don't call it a "backpropagation network"!  The MLP is only one special case of what you can adapt with backpropagation. Furthermore, the MLP is a lot older than what I did in 1974. Bernie Widrow or Rosenblatt are the guys that should take credit for the MLP design itself. 

 

 

                                                Figure 5. An Example of a Three-Layer Multilayer Perceptron (MLP)

 

                The MLPs are basically the McCulloch‑Pitts networks, the feedforward things. There have been some wonderful theorems about what they can do.  

 

(2) and (3) Almost as common are the CMAC and the RBF designs. (Figure 6) These networks are examples of "local" learning systems. We have already heard about these designs from Nick DeClaris; for example, CMAC was first proposed by Albus, for a PhD thesis under DeClaris. There are many other local learning systems discussed in neural net meetings, but DeClaris' students happened to hit on what was useful more than most students. There are also many modified versions of the CMAC and the RBF, which give improved performance in control applications[15].

 

 

 

 

 

                                                                  Figure 6. Structure of CMAC or RBF Network

 

                Basically, these local learning rules perform forecasting by association.  The MLP gives you a global model and is good for learning global functions, causal relations, etc.  The local systems are

more like forecasting by precedent. When you've got a new situation, you predict the result will be like what it was before when you had a similar situation.  It's an associative memory, and 

this is what the Granger‑Lynch stuff is, just another example of the same general principle. 

                So these are the things that most people use. These are feedforward nets, easy to implement and I'll show in a couple of charts how some people have used them. There are many other forms of associative memory, based on Hebbian learning, which have appeared in the biologically oriented literature; however, those kinds of nets are not used very often, for one reason or another. Most likely, people feel that the Hebbian nets now available are very similar in capability to the CMAC and RBF, because they are based on forecasting by precedent, while being harder to implement in real time. Why work harder to achieve the same capability? Another factor, however, is that people do tend to implement what is easiest first, not what is most powerful.

                Local learning systems can be adapted in a variety of ways -- through least squares, through backpropagation (i.e. derivative-based learning), etc. No matter how they are adapted, they tend to be faster to adapt than global MLPs, because they do not have to "undo" what was learned in a previous region of space when they explore a new region of space. They are usually set up so that different weights are active in different regions of space. They tend to require many more weights, but learn faster. There may be ways to combine the best advantages of global and local networks together in one system[18], but no one has implemented the right kind of hybrid as yet.

 

What Kinds of Functions Can Such Modules Represent?

 

So what are the capabilities of these different kind of networks? Well, there are a lot of theorems.  There's a guy named Andy Barron who has proven some beautiful theorems showing that those three‑layer neural nets like MLPs are much better than Taylor series, polynomial expansions, or local learning rules even, to approximate any smooth function.  It's beautiful. 

                In control applications, however, to really control an arm efficiently, sometimes you don't have a smooth function.  Then you've got a problem and it often doesn't work, and a lot of control applications three layers won't work.  So, a guy named Sontag at Rutgers has proven that it really can approximate the control functions you need if you've got two hidden layers. 

                But there's another problem. You can approximate any function, but you need to approximate it parsimoniously.  MLPs are as good as any other feedforward net, but in general there are some functions which you cannot approximate parsimoniously with any feedforward net. This means that you need enormous numbers of hidden layers to approximate them, and enormous numbers of hidden neurons. 

                Marvin Minsky, years ago, gave an example of this in his famous book Perceptrons. He described the "connectedness" problem, which sounds at first like a very typical character‑recognition, pixel‑type problems, but it turns out to be a little different. Imagine that you've got a grid of 50x50 input pixels and that they're all either white or black, and that you're trying to recognize a desired pattern.  What you're trying to do is to output a one if the blacks are all connected, and a zero if they're not.

                Now it turns out that Minsky showed that the number of hidden units required for this task is just enormous. As the number of pixels grows, it becomes astronomical; no feedforward net of any kind is going to do a good job.  But if you allow recurrence, recurrent feedback connections, what I call simultaneous recurrence (a special kind of feedback connection), then you can represent it  parsimoniously. 

                I would argue that the kind of language-processing problem that Jim Anderson described earlier (from Fodor) is a problem in this family, where a feedforward net can't represent it parsimoniously, but a simultaneous recurrent net can represent it parsimoniously.  And I've seen systems of that general sort, which so seem to work on that kind of problem, but I haven't studied Fodor's example in great detail. 

                It turns out that, if you want to deal with control problems like navigating a robot through a cluttered room where the clutter keeps moving and it's in a new position, you need to worry about finding a connected path.  So there are good arguments[15,18] that higher‑level intelligence has to use these kind of networks. 

                So that seems easy, but is it?  Well, first of all, what happens if you have parts of the brain that have to make decisions quickly? Simultaneous-recurrent nets take time to settle down‑‑what then?  Then you've got to have a feedforward net, and what happens then? 

 

Fast Feedforward Nets: The Cerebellum

 

     Well, Figure 7 (on the next page) shows an example of a feedforward net with two hidden layers that is good for fast, general motor control.  This comes from Nauta[1]; it is a diagram of the cerebellum. In the cerebellum, you start out with inputs along the mossy fibers (which some people would call an input "layer"). These inputs go to the granule cells, which operate as a first hidden layer, and you've got zillions of them. Some people say there are more granule cells than there are any other kind of cell in the brain, maybe ten times that number. That reinforces

 

the point‑‑you need a lot of hidden nodes if you try to do complex tasks with a (relatively local) feedforward net.  

     The next hidden layer in the cerebellum is the Purkinje cell layer. The output layer is basically the deep cerebellar layer and the vestibular nucleus (the FTN cells of the vestibular nucleus, to be precise); those two systems are basically together ‑‑ they're not right next to each other but they form one output layer, for functional purposes.  In summary, the cerebellum is not based on simultaneous-recurrence, which is slow but powerful. Strictly speaking, however, it is not just a static MLP, either. Above all, the Purkinje layer has some working memory capabilities[14], similar to well-known ANN designs. Such capabilities are tricky to adapt[15, chapter 13]; this, in turn, suggests that Purkinje cells might possibly be adapted by a combination of the well known olive-to-cerebellum mechanism, plus a local mechanism supporting the working memory effects. Also, because the Purkinje cells are large cells, and because continuity in output is very desirable at this level, it is conceivable that dendritic field effects, as described by Pribram, could occur within these cells[4,18].

 

 

 

 

                                                                       Figure 7. The cerebellum, from Nauta [1]

 

Slower, More Powerful Nets: The Cerebrum

 

But again, this kind of feedforward arrangement is not very good for the really higher‑level kinds of functions like finding a connected path or connectedness. There are some kinds of functions you just can't expect the cerebellum to solve, because they require a more sophisticated processing.  That leads to a prediction that somehow or other the higher levels, like the limbic lobes and the cerebral cortex, must form a kind of two‑level system, where the low level is settling down...but first let me describe Figure 8, which shows what a simultaneous recurrent net (SRN) is. 

 

 

                                                                    Figure 8. A Simultaneous Recurrent Network

 

                Mathematically, an engineer would say that you plug in the inputs into any old feedforward structure (the net labelled f in Figure 8). But then you take the outputs of f, and feed them back in as inputs to f, and see how that changes the outputs. You keep feeding the outputs back in as inputs, again and again, until the values of the outputs settle down to some kind of equilibrium, y. Now, to use a system like that... you have to plug in the inputs, wait through many cycles of the inner system f until the outputs settle down, and then that becomes just one cycle time of your bigger system.  Now, when I first saw the engineering of this[15], I said this can't have anything to do with the brain, because you need really fast cycles inside of a longer cycle; how could that be biologically plausible?  I knew that we needed all this to get intelligence in engineering, and I couldn't think of anything else; therefore, in the 1990 Decade of the Brain Symposium (sponsored by the National Federation for Brain Research and the INNS), I presented this from an Engineering point of view, without any notion of how the connection to the brain could be made.

                Later that same day, Walter Freeman presented his model of the hippocampus, which involved exactly the same kind of loops within loops as in my model! He showed how very close inner recurrent loops operate at a very high cycle time, embedded within a larger, slower theta rhythm. For the inner loop, he said that the basic calculation cycle time is like 400 hertz, versus about 4 hertz -- quite enough to implement Figure 8. (Some people quote 80 hertz -- based on Fourier analysis rather than cycle times -- but that still would be enough for some functionality in this design. When I checked with Pribram, he assured me that fast, 1ms. synapses allow such high-frequency computation.) VonderMalsburg has convinced me that such dual-loop effects are even more certain to occur in the neocortex, but the neocortex contains additional capabilities and complexity which may make it harder to work with at present. I would speculate that SRN capabilities are crucial both to binocular vision and to the image segmentation capabilities of neocortex.

                In brief, the biological data appears to fit the model beautifully.  Now, if you look 

at Granger and Lynch's model of the hippocampus, it doesn't do that.  In their model its like a feed forward, associative memory, and you only have an outer recurrence that's used to generate the 

associative memory.  So what is that inner loop doing?  Maybe the hippocampus is more powerful than associative memory. Maybe we need something more powerful than an associative memory to form emotions and make plans in our life, and maybe somebody can do an experiment proving it.  I hope so.  

                Parenthetically, it should be noted that SRNs -- unlike feedforward networks -- can have problems in settling down to a stable equilibrium. In engineering, one can use a "tension" term [15,18] to reduce the probability of instability, but the possibility cannot be totally eliminated. The tension parameter is a global parameter, with an interesting analogy to the global level of adrenalin in the bloodstream. Karl Pribram has pointed out that there is a strong analogy between this "tension" term and the "unpleasure" principle of Freud[19], which plays a key role in understanding the possibility of instability in human brains. In the human limbic system, Pribram's empirical discussion [20] suggests that the hippocampus is an SRN, acting mainly as a "hidden layer" of the limbic network, a network in which the amygdala is the ultimate or penultimate output layer. (This suggests a kind of crude analogy between the hippocampus and the cerebellar cortex.)

 

 

 

Other Forms of Hebbian Learning

 

That covers most of what I really want to say about supervised learning. Again, please excuse my glossing over the many, many details; each one of these topics can be discussed in much more detail, and is so discussed in the papers cited.

                For the sake of completeness, however, I should say a little about two forms of Hebbian learning which I did not mention above.

                Most people who work with Hebbian learning would argue that there are really two different kinds of Hebbian learning system which could be used on supervised learning problems. There are local associative memory systems, which I discussed above. But there are also global systems, which are generally linear, and require that inputs be decorrelated before they enter the supervised learning system. A lot of decorrelating networks have been designed for use with such nets. However, after discussing this matter with Pribram, I am convinced that this latter class of network is not relevant to systems like the human brain. Pribram and others have shown again and again that biological representation systems have a great deal of redundancy (e.g., like wavelets but with a 1.5 amplification factor instead of 2, etc., as in the Simmons talk today). One would expect such redundancy, in any system which also has to have a high degree of fault tolerance. This is inconsistent with the mathematical requirement of orthogonality. In addition, the limitation to the linear case is not encouraging, either.

     In 1992, I developed an alternative learning design which appears Hebbian in character, but has radically different properties[4,18]. It provides a mathematical representation of certain ideas by Pribram about dendritic field processing, which the talk today by Simmons provides strong empirical support for. It is closely linked to Chris Atkeson's experiments with locally-weighted regression, which has performed very well in robotics experiments at MIT. In retrospect, as I reconsider the issue of information flows around dendrites, I suspect that the design still needs to be revised, to account explicitly for the three-dimensional nature of local information flows, at least for biological modeling. In any case, the alternative design is still feedforward in terms of what it accomplishes; therefore, it might possibly be worth considering as a model of the innermost loop of the neocortex, but it does not obviate the need for simultaneous recurrence, and for the unusual kinds of nonHebbian feedback (as in Figure 8) required to adapt key parts of the neocortex and hippocampus -- if those systems are as powerful as I suspect.

 

Summary

 

In summary, I predict that the human brain contains some very complex circuitry, as required to solve some very complex adaptation problems. At present, most people would find it hard to believe that something that complicated is there, even though it does fit these new results of Freeman and so on. I think we need new experiments, based on living slices, to help get home to people that it's this kind of complexity that's in that system, and that the old models are simply not good enough.  So that's the end of supervised learning. 

 

 

THE HIGHER LEVEL OF ORGANIZATION: NEUROCONTROL

 

Now let me talk about neurocontrol.  This is a subject I've talked about for eight hours at a stretch, so I will have to cut out a lot of important material here today. First, I want to talk about why this is crucial to understanding intelligence. I'll skip over my slides on engineering application areas. I will talk a little bit about the kind of designs that engineers are using today, but only a little.  Mainly I will focus on the design concepts which relate directly to understanding the brain.

 

Why is Neurocontrol the Right Mathematics For Understanding the Brain?

 

This is a chart (Figure 9) that people look at and say, "I already know this."  But if people could understand the implications of what they already know, this world would be a different place. There are some implications in what we already know that people haven't thought through.  Now what I am going to talk about here is the reason why the human brain is a neurocontroller; let me give you the argument in a few stages. 

 

 

                                                                       Figure 9. The Brain As a Neurocontroller

 

                Step one: we know that the brain is an information processing system.  I would call it a computer, except that people will think of sequential machines.  But it's really a computer; it's an information processing system; its sole biological function is to be a computer.  And what does it compute?  It computes actions that control glands and muscles.  So the point is that the function of the brain as a whole system is to perform control. 

                Some people think of control as something that's only in the cerebellum, just to control finger movements. That's not true.  Nauta, in his classic text on neuroanatomy[1], stresses how you cannot separate what is the control system from the rest of the brain.  Now that doesn't mean that the whole brain does tracking or pursuit movements; no, it doesn't do that; it does a higher order kind of control, of course.  And you might use the term "sensorimotor control," if you will.  But the point is that the brain as a whole system has the function of calculating these things.  Everything in the brain is there to help it compute these outputs.  So the function of the brain is to do that; if you want to understand the brain as a whole system, you have to understand the mathematics of what it takes to build a controller that has these kind of control capabilities, which again go far beyond mere trajectory tracking. (Those who think of control as trajectory planning may still have troubles with this; I urge them to reconsider the definition of control, and recognize that the overall mathematical literature on control has always been far wider than this in scope.)

                Furthermore, you can't even understand a subsystem until you know how it fits into the whole system. Therefore, you can't even understand subsystems of the brain until you put them into this greater context, which is neurocontrol: the brain is a neurocontroller.  

 

Capabilities of the Brain As a Controller

 

So, next slide (Figure 10).  This slide shows what I regard as the most exciting and crucial capabilities of the brain as an intelligent controller.  I should have added an extra line here about learning in real time; it's just so obvious, but it's something we've got to keep in mind.

 

 

                                                           Figure 10. Capabilities of the Brain As a Neurocontroller

 

 

                The brain can control millions of actuators in parallel -- well, maybe only 900,000 ‑‑ it's the same principle, huge numbers. What about conventional controllers? Most control engineers regard one actuator as a normal problem and ten as a large problem.  Thus the brain has an incredible capability, very exciting to engineers.  It can handle nonlinearity and noise routinely, without being destabilized.  And above all, most critical, it includes what you might call a long‑term planning horizon.  The AI people would say the long-term planning capability is the real intelligence.  And the brain also has a high‑speed coordination capability through the cerebellum, basically.  

 

Brain Capabilities Versus ANN Capabilities

 

Now, how do these capabilities compare with anything we can conceive of in mathematics?  Is there any hope of understanding them?  Well, we presume there's a hope of understanding them, but is there a way that we can conceive of to understand them? 

                The next slide (Figure 11) provides a list of what's been done in Artificial Neural Networks in control. These are the basic kinds of capabilities that exist today.  I've read hundreds of papers on this topic, but they all boil down to this.  I've seen a lot of people try to wriggle out of my basic taxonomies, but these are basically the capabilities you've got.  You've got people using neural nets in subsystems in control; that's not really neurocontrol.  You've got people who have learned to clone experts, learned to copy a human movement.  You've got people who

 

                        NEURAL NETWORKS (ANNs) IN CONTROL:

 

                                    1. In Subsystems

 

                                    2. Copy Experts

                                                Supervised Control

 

                                    3. Follow Path, Setpoint, Ref. Model

                                                Direct Inverse Control

                                                Neural Adaptive Control

 

                                    4. Optimal Control Over Time

                                                Backpropagation of Utility (Direct)

                                                Adaptive Critics

 

                                                            Figure 11. Four Tasks Performed by ANN Controllers

                                                                   (Subsystem, Cloning, Tracking, Optimization)

 

do classical adaptive control, which is just a matter of trajectory tracking; for example, somebody tells you where to move your arm and you move it there. And then you've got systems that optimize over time. 

                Now, cloning or copying is definitely not what humans do.  I admit that we may imitate our  

parents a little bit, but they don't tell us exactly what to do; they don't give us a complete vector of actions, and that's what the cloning designs require. 

                Likewise, we're not simple trajectory followers. We don't have somebody who tells us where to move our arm; maybe we have subsystems -- maybe, maybe not -- but that isn't what the human brain as a whole does.  

                So, that really leaves us only one choice, which has to do with optimization over time.  Now, the notion of optimization over time is one people have taken seriously for centuries in studying human behavior. People have screamed at the idea for centuries as well, because it's clear that humans don't do a perfect job of optimization, but that's okay.  If, in engineering, you do the best possible job of optimization, your system still won't be perfect. You don't have to worry about designing a system so perfect that it's implausible as a model of the human brain; in fact, that's the last thing you have to worry about, okay?  There are a lot of imperfections in the engineering optimization systems quite close to those you see in biological systems. 

                Furthermore, the general concept of reinforcement learning is very pervasive in animal behavior. There's a guy named Harry Klopf who has recently shown that if you take some very simple optimization networks, they regenerate not only what's called Skinnerian conditioning (in other words, conditioning by reward and punishment); just incidentally, they replicate Pavlovian conditioning as well. So this concept really is powerful enough to cover the basic things we observe. 

                In summary, my argument here is that the optimizing designs are the only form of neural  networks we know of, the only form of mathematics we can conceive of, which is really relevant to understanding brain-like intelligence. It seems to me that it's got to be what's going on in the brain.

                If you look again at Figure 11, you will see that there are only two useful ways of doing the optimization over time (for large-scale problems).  There is a thing called backpropagation of utility: I proposed this approach in my Ph.D. thesis in 1974; I invented it, so I have a vested  interest in it, but I'll still tell you it's not biological.  Flat out, it's not biological.  I can explain why at length, if anybody's interested. I've written it up[2,15]; it's not biological.  There's only one thing that's biological, and that's what I call adaptive critics, and I'm going to say that the brain is an adaptive critic system.  I've got to explain what that means now. 

 

ANN Control Designs Versus Brain Capabilities

 

Table 1 is a matrix comparing these designs against the capabilities of the brain. For the most part, this table simply reinforces what I just said, that the simple cloning stuff doesn't involve planning, that tracking

 

                                                     Table 1. Matrix of ANN Designs Versus Key Brain Capabilities

 

 

 

 

 

 

 

 

 

 

 

trajectories doesn't involve planning either, and that backpropagating utility is not brain‑like because it can't handle noise and it's not so good for real time. But if you look at the table, you can see that it subdivides the world of adaptive critics a little further.

                There was an old class of adaptive critics developed by Barto, Sutton and Anderson in 1983. It was a very useful first step in popularizing the idea. It can handle noise and optimization 

over time and real time learning. A lot of people are having fun with it. It's good for a lot of engineering applications, but it won't handle many motors.  And people have to understand it's very limited; it can't handle large problems.  And when I say large, I mean like ten variables in an aircraft control problem.  I mean that this is a really heavy limitation and that's been proven empirically now. (Tesauro has shown that a challenging AI problem -- playing backgammon -- yields very well to this design; that problem involves many theoretically possible states, but still involves a limited number of discrete action choices. Haykin's new book reviews the engineering experience -- which fits my summary here -- and White and Sofge, among others, have reported similar experience[15].)

                On paper, in 1987 (actually before '87, in 1977), I proposed an alternative class of adaptive critics which combine backpropagation with adaptive critics.  And let's call that Advanced Adaptive Critics. Combining backpropagation with adaptive critics in a fundamental way, not in a superficial way. And on paper, it looked to me that this approach solved the problem of handling many variables at once.  And, as I said, I published that in 1977 as a solution to the slow convergence problem of the simpler critics.  So that was actually before the Barto‑Sutton-Anderson paper on the same kind of area.  

                In 1987 there were no working examples, but some people heard me present this in mid‑1989 and now there are four working examples, at least, and two that are pretty close to real, and two  of the examples are in large uses.  There is one company that is a spinoff from McDonnell‑Douglas (Neurodyne[15]) that has used these things to solve composite materials manufacturing problems that could never be solved before by any kind of neural network method, where the two nets would not work because they couldn't learn fast enough in real time. They have also used it in the control of an F‑15 (providing real-time adaptation to aircraft damage in two seconds), to the control of a prototype of critical systems in the National Aerospace Plane, and elsewhere.

So there are actually many applications in one of these groups. (See [19] for some further applications.)

                So the bottom line is this: we now have real‑world examples of advanced adaptive critics, and they have indeed proven much more powerful.  But we have other designs that nobody's tried yet, that are on the books, and we're just waiting to go up the ladder another step or two.  So there has been proof that this kind of design works.  

 

Where to Get Mathematical/Implementation Details on Neurocontrol Designs

 

There is no time today to describe all the ins-and-outs of the different adaptive critic designs, let alone all the other designs which have proven useful in some engineering applications. Today I will mainly focus on a few points about adaptive critics which I think are biologically relevant.  But I can give you some citations that tell you how to implement these designs.

                In this field, it is really critical to get the right citations.  Frankly, there are a lot of papers and books out on neurocontrol or on recurrent networks which are wrong or obsolete; there are a lot more than you might expect. I have seen a lot of very bright people having a very hard time adapting recurrent networks, for example, based on believing some things they read by world-famous people, who should have known better. So I really want to stress these citations.  

                Until December of last year (1991) I spent at least a good year working with people in industry and academia, and on my own, to come up with a complete statement of where we are in terms of the state‑of‑the‑art designs and the new ideas. The results are in a new book called the Handbook of Intelligent Control, which finally came out in September 1992[15]. This book is the up‑to‑date, complete source on what is known in neurocontrol today.  It contains an authoritative statement not only of my own views, but of many other key workers in field, along with truly real-world benchmark problems and applications. (See [19] for further applications and information on patents.)

                Now, there is still a problem here. Because this book is a complete handbook, it is not an elementary textbook. In order to get a basic introduction that goes a bit beyond what I am saying here today, there are two conceptual overviews I would recommend: [22], and [19] or [23].  If you're a good enough engineer, the Handbook might even be straightforward enough by itself, but if you find it a little hard going, these introductions may be important.  

                I have a friend who's begun to implement some of the very hard architectures in the Handbook, and he says that it's been very useful for him to refer to the tutorial in [22].  He says that [22] is the easiest‑to‑read thing that I have ever written, and he says that this tutorial is very, very important in really implementing neurocontrol in a hardcore engineering way. 

                So reference [22] is for mathematical details; the other references -- [19] or [23] -- are for 

conceptual details, and applications, and excitement, and that kind of stuff. They are good for policy people. You need [22] only if you really want to implement neurocontrol and understand the math in detail.  Reference [23] has the advantage of being shorter than [19], and being more easily available; reference [19] has the advantage of containing new information about how to combine neural nets and fuzzy logic, as well as some basic background on ANNs. 

 

Neurocontrol Designs to be Discussed

 

Now, aside from these references, I am not going to talk at all about those forms of neurocontrol which are relevant only to engineering. I won't talk about cloning. There are neat applications of cloning in controlling a model of the National Aerospace Plane, but I won't talk about them today. I won't talk about neural adaptive control, even though it's painful, because those designs do raise some issues of some relevance to biology. Within the realm of optimization, I will not talk about direct methods like the backpropagation of utility, even though there are important applications like improving efficiency and reducing pollution in chemical plants[15]. I will talk briefly about direct inverse control -- an approach to solving tracking problems, because some people believe that approach is relevant to low-level motor control in biological organisms. Then I will talk about adaptive critics -- a more relevant family of designs -- in more detail.

 

 

Direct Inverse Control (A Simple Approach to Trajectory Following)

 

Direct inverse control I've got to talk about because there are a lot of people who tacitly assume that this is what the brain does.

                I was really glad to hear Dave Robinson talk at this workshop about how people think the brain is mapping from coordinate system A to coordinate system B, and about why this is a fundamentally misleading assumption. I agree with Dave very strongly that that is a bad way to describe what goes on in the cerebellum.  And why? 

                People assume that the cerebellum is doing something like this (Figure 12). Suppose that you're trying to get a robot arm to follow a trajectory in physical space. You've got spatial coordinates X1 and X2. You control the

 

                   Figure 12. Direct Inverse Control

 

thetas (joint angles). Somebody else is going to give you the X's, and you've got to figure out what are the thetas that get you to the right X's.  Physically, you know that the position (X1,X2) is a function of the thetas.  So if this function is invertible, then the thetas are a function of the X's, and what you can do is this: you can flail the arm around, get data on the thetas and the X's, real data, and then learn the mapping from the 's to the thetas.       That's the basic idea. People like Kuperstein have taken this kind of approach, and people like Miller have done it, too. Kuperstein was an important pioneer in getting this approach started, but his reported error statistics (circa 3 percent) were too high to be useful in a practical way in robotics. Even in biology, we know that more accurate tracking is accomplished. I am puzzled why some people in neural nets go back so often to that early work, now that Kuperstein himself has moved on to other problems.

                Now, Miller has shown that you can fix up the basic error statistics[15,17], but we still know that the human arm has many degrees of freedom.  We know there are a lot of degrees of freedom in the human arm, more than

there are in space; thus we know this mapping is not invertible for the human arm. So what is going on here?

     Suzuki, Kawato and Uno have done a lot to explain this situation. They have lots of experiments proving that there is actually an optimization going on in this motor control system of the human arm. The details of that are in many places; for example, we have an MIT press book[17] where Kawato talks about that, and their experiments are unbelievable.  I know that Hogan from MIT has  recently questioned their conclusions, but neutral observers like Massone at Northwestern have looked over this controversy; it is my understanding that the empirical evidence is

overwhelming that Suzuki et al are right on this particular point.

                So we have a real optimizer down there in the cerebellum; we don't have the direct inverse coordinate-mapping kind of stuff; and that leads to some very interesting possibilities. Somebody should do some experiments to see how flexible and plastic that system is as an optimizer.  Can they change what it optimizes? Can they use it to solve interesting optimization problems?  And then the question is: how does it do it biologically?  How does the cerebellum act as an optimizer?  

                By the way, there are a lot of engineering applications of direct inverse control. For example, it has been used to control a simulator of the space shuttle main arm; you may be seeing the results on television in a few years, if all works as expected, but I don't have time to get into those kinds of details today. The bottom line is that direct inverse control has its uses, but is not relevant to the adaptive part of the human brain.

 

Adaptive Critics: A Way To Approximate Dynamic Programming

 

Now let's talk about adaptive critics.

                One way of defining adaptive critics is to call them systems to approximate dynamic programming.  When I give this definition to engineers, they say, "Oh, you're talking about something real. I thought it was garbage. You mentioned Freud and animal learning, so I assumed this can't work."  But actually, you know that animal learning really does work; it is a powerful information processing system. Still, it is legitimate to look at it as an approximation to dynamic  programming; that is one legitimate point of view. 

                In engineering, it is well‑known that (for good mathematical reasons) there is only one set of techniques that is capable of finding the optimal strategy of action in a general, noisy nonlinear environment over time.  There's only one that can do that in a general case, and that's called dynamic programming.  

                Figure 13 illustrates the basic idea of dynamic programming.

                The way that dynamic programming works is that you give the system a utility function U or a performance function or primary reinforcement ‑‑ there are a thousand names for it; in other words, you tell it what you want to optimize over time, over the future.  You also give it a stochastic model of the environment. Then, what dynamic programming does is that it comes out with something called the J function -- at least that's what Bryson and Ho

 

 

                                                           Figure 13. Inputs and Outputs of Dynamic Programming

 

call it. (To be more precise, in dynamic programming you solve an equation called the Bellman equation, in order to find this function J.) I like to call this function J a "strategic utility" function. It's just another utility function; it looks similar to U, in many ways.

                The basic theorem of dynamic programming is this: if you maximize J in the short term, that will give you the strategy that optimizes U in the long term.  So dynamic programming translates  a difficult problem in planning or optimization over time into a problem in short term optimization.  That is the essence of it.  

 

                And then the next question is: if there is only one exact way to do this, why don't we use dynamic programming for everything?  In engineering and in biology, why don't we use it for everything, if it's the only exact and perfect thing?  Well the reason is simple: it's too expensive.  It may be the minimum-cost method, but the minimum cost is astronomical even when you have just a few variables. You can't do it exactly.  A corollary of that is nobody will ever come up with a neural network that plays a perfect game of chess.  So the next time a computer scientist tells you "gee, perfectly adapting a neural net is an NP‑hard problem; therefore this isn't real and

 

 

we've got to give up on it," ignore the computer scientist.  The goal is not to play a perfect game of chess; that isn't what the brain does; and that isn't what our artificial systems can do, because it can't be done.  That's one reason why the systems aren't perfectly optimal: it's not possible in a real‑ world engineering sense. 

     So what can you do?  For a real‑world, general purpose system that tries to optimize in the real world, we have to come up with a general purpose approximation to dynamic programming. That's what adaptive critics are.  More precisely, adaptive critic systems are systems which contain networks whose job is to approximate either the J function of dynamic programming or its derivatives or something very close to it.  That's what it is; you can call this "approximate dynamic programming" if you've got to sell it to a boss that doesn't like neural nets, and you'll be completely honest.  You can call it adaptive critics among neural network people, or -- if you're talking to animal psychologists --  you can call it reinforcement learning, although that tends to understate what it's good for. Those are all legitimate names for what I am talking about here.  

 

Intuitive Meaning of U and J

 

Now let me give you a little intuition about the meaning of all this, because once again, I'm saying the human brain is an adaptive critic system.  So I am claiming we have a network in our brain that approximates the J function.                  Table 2 gives us some intuition of what the J function is.  If you're playing chess, the ultimate goal, at least in computer chess, is to win and not to lose.  That's the intrinsic utility, U.  But there's a famous rule of thumb, that a queen is worth nine points, a rook is worth five, etc.; people use that rule of thumb to see if they are making progress in the game. Beginners play to maximize points.  Sophisticated people learn that holding the center is also worth something.  And there are studies  of Bobby-Fischer type people which argue that

 

                                                                                 Table 2. Examples of U and J

 

 

they don't really see ahead twenty moves -- even though they love to talk that way ‑‑ but that what they really do is to perform a very complex recurrent, strategic assessment to understand how well they're doing.  And they really look ahead one move, but they do a complicated, strategic assessment of their options one move ahead.  

                In humanistic psychology you could think in terms of pleasure and pain, and hope and fear. In typical animal psychology, there's primary reinforcement and secondary reinforcement. This U/J concept occurs all over.  I don't think this concept is new to me; it's been hard‑wired into our brains for years.  But maybe we're just making it mathematical for the first time. 

                So when I say that the human brain is an adaptive critic, what does this theory really mean? In common-sense terms, all I'm saying is that we're governed by our hopes and fears, and these phenomena of hope and fear are irreducible things built into the human brain. I think that this is plausible. I don't think that it's a weird AI kind of theory, that hopes and fears are the fundamental grammar or representation wired into a big part of our intelligence.  

 

The Barto/Sutton/Anderson Design

 

Figure 14 illustrates the 2-Net design I mentioned earlier. Because I don't have much time left today, let me explain very briefly the main reason why this design can be slow. The problem is that you've just one global reinforcement

 

 

                                                     Figure 14. 2-Net Adaptive Critic of Barto, Sutton and Anderson

 

signal broadcasting to an action net.  If you only have one action, then you know which action was wrong.  But if

you've got a hundred actions, and you made a mistake, then you don't know which one to change, in what direction.  You don't have a sense of cause and effect; without cause and effect, it's kind of hard to learn. The ideas in this design may still be useful as part of a larger hybrid, but, by itself, this design could never describe a large-scale system like the brain.  So now let's move on a little bit. 

 

The Backpropagated Adaptive Critic: From Freud to Engineering

 

Figure 15 illustrates an idea I proposed in a journal article in 1977.  Let me stress that this is only one of many advanced adaptive critic designs; you've got to go to the Handbook to get the complete list; this is only one. (I picked this one because it's the one case where there was a typo in the Handbook; the version here is correct.)   

 

 

                                                                 Figure 15. BAC (Version Which Uses J Output)

 

                This is where the backpropagation algorithm really came from, originally, in unpublished reports I wrote in 1970-72. Later I learned how to simplify the idea of backpropagation, so people could copy it and reproduce it.  That really happened in 1981[24], when I started playing the simple perceptron games with backpropagation and publishing that stuff, but Figure 15 is where I really came from in the first place. 

                And where did this idea come from? Believe it or not, I developed this idea as a mathematical translation of an idea from Freud. That's where backpropagation started. Freud had this idea. Freud was interested in neural nets. He had to make money later on, and he regretted that he had to make money doing stupid things, okay?  That's on the record.  He started out going to medical school, studying physiology, and what he really wanted to do was to build a neural network theory of the human mind. He felt that he had developed a valid theory, and he came back to that theory later in life.  

                His model began with the idea that human behavior is governed by emotions. Does that sound weird?  Not if you're a human, but sometimes I almost wonder who is and who isn't, when I see some of the theories floating around these days. At any rate, Freud had the idea that there was something called cathexis or psychic energy or emotional charge attached to things he called objects.  According to his theory, people first of all learn cause-and-effect associations; for example, they may learn that "object" A is associated with "object" B at a later time.  And his theory was that there is a backwards flow of emotional energy. If A causes B, and B has emotional energy, then the emotional energy will flow from B back to A.  And if A causes B to an extent W, then the backwards flow of emotional energy from B back to A will be proportional to the forwards rate. That really is backpropagation.  That really is backpropagation, and my argument is that you have to have that.  I cannot conceive of a way of using cause and effect infor­mation without doing what Freud said.  If A causes B, then you have to find a way to credit A for B, directly.  You have to exploit the fact that you know that A causes B to the extent W.  So your flow of emotional energy has to use that number W that repre­sents the forward association; you can't get out of that.  I see no mathematical way to get out of it, and I've looked at a million attempts, by Grossberg and others; there's no way to get out of that.  If you want to build a powerful system, you need a backwards flow. 

 

                Now, in mathematical terms, I can now give you a very different interpretation of Figure 15. In Figure 15, I am properly and dutifully following dynamic programming. (I figured this out later after I had the design.) What I'm really doing is exactly what dynamic pro­gramming tells me to do, which is to pick an action vector u so as to maximize J. I maximize J directly and intelligently by calculating the derivatives of J with respect to action, any using those derivatives.  This is not error  backpropagation.  This is backpropagation as a way of calcu­lating derivatives; I work back the derivatives with a special chain rule, the chain rule for ordered derivatives.  I work backwards the deriva­tives of J with respect to U.  And this is easy to implement, although you've got to read the Handbook to get the details, since I don't have the time here today. 

 

                So the bottom line is this: because of this argument of Freud's, which is just about inescapable mathematically, nobody has found an alternative to this general approach. I would predict that there must be this kind of mechanism somewhere in the brain.  I can't see any way you could do it otherwise; I haven't seen a contrary model that could work on that scale. And, as I mentioned before, my paper on the cytoskeleton[5] and other papers by Dayhoff and others describe plausible mechanisms that could implement it.  It is very plau­sible.

                On the engineering side, there are already a number of applications of Advanced Adaptive Critics (which I define as adaptive critic systems which adapt an Action network based at least in part on estimated derivatives of J with respect to u). There is an application to the continuous production of high-quality composite materials, which have a potential market worth many billions of dollars, et cetera. There is an F15 application I mentioned before. Again, I have no time for the details.

                In the future, we might get into earth orbit at a much lower cost if we can solve certain control problems; they turn out to be optimal control problems, which classical control cannot handle very well. Again, dynamic programming -- the appropriate classical method -- is too expensive, but adaptive critics are not. The National Aerospace Plane office has recently given contracts to people to use adaptive critics, to solve optimiza­tion problems that can't be solved any other way.  There is reason to believe we cannot reduce the cost of earth orbit without using this stuff, and it's being implemented today. A benchmark version of this control challenge is in [15].

 

Overall Architecture of the Brain

 

How does Figure 15 relate to the brain?  Figure 16 on the next page is from a paper I published in 1987[25]. (That paper was the first to inform the Barto-Sutton-Anderson group about the dynamic interpretation here, and about my prior work; whatever its failings, it had a major influence on later developments.).  This figure presented a very early approximation of what I think is going on.

 

 

 

                                           Figure 16. A First-Pass Neurocontrol Interpretation of the Brain (From [25])

 

                I have argued that the hypothalamus (and maybe the epithalamus) are computing a built‑in utility function. The hypothalamus is not the greatest center of plasticity, but it's a powerful source of primary reinforcement. I'm arguing that the primary function of the limbic lobes is as a critic network; we certainly know that there's secondary reinforcement in the limbic nodes, no surprise  there. (See, for example, the classic work by Papez and by James Olds.) 

                I'm also arguing that the cerebral cortex includes the function of Model network; in other words, understanding cause-and-effect relations is its primary function. In engineering terms, that means that the cortex is performing system identification, which includes filtering or working memory (short-term memory) as a secondary function. This fits the recent studies demonstrating working memory capabilities in the temporal cortex (Goldman-Rakic) and in the frontal cortex, and the work by Barry Richmond demonstrating the relevant kinds of lagged recurrent effects even in visual cortex. (Because of plasticity studies, it would hard to imagine that a truly fundamental capability like working memory could be limited to any one architectonic area of the neocortex.)  

 

Architecture of the Olive/Cerebellum System

 

Finally, in closing, I would like to come back to the cerebellum, which is the Action network within a complete multi-net lower-level control system. I would like to come back to the key question which I left open earlier in this section: "How could the lower motor system perform optimization? How could it be an adaptive critic system?" Here, I will summarize some thoughts from [14].

                Houk has done a lot of work showing that the inferior olive sends training signals which adapt the Purkinje cells, in a way that looks like an adaptive critic arrangement. So this is consistent with the idea that the cerebellum is an Action net, and that the olive is a Critic directly adapting it.  What Houk doesn't talk about is plasticity in the output layer of the network, which is a deep cerebellar array (plus FTN cells of the vestibular nucleus). Lisberger produced a flow chart of the FTN system for an excellent paper in Science which nevertheless confused some people in the neural net community, who apparently thought that the chart was a complete wiring diagram (even though Lisberger himself discussed other connections in the text); this has led to some nonempirical Hebbian models so unrealistic and so incomplete that the authors should not be mentioned.

 

 

 

 

 

 

 

 

                                                    Figure 17. Flow Chart of the FTN System From David Robinson

                                               IO=Inferior Olive; cf=Climbing Fiber;Pc=Purkinje Cell;gc=Granule Cell

 

                Figure 17 is a better flow chart borrowed from Dave Robinson, who has a new paper that will explain this better. The bottom line is that there are cells of the vestibular nucleus (and presumably in the deep nuclei) that also get climbing fiber input from the olive, and there's good evidence that they train the output layer as well as the Purkinje layer.  It turns out, if you look at this arrangement mathemati­cally, this arrangement turns out to be equivalent to backpropagating through the action network, just like that adaptive critic architecture I was describing.  And it's neat,  because it's an electronic way of doing it that I wouldn't have thought of. The unique many-to-one architecture in the cerebel­lum makes this mathematically a valid implementation of backpropagation if you have the right training signal (a derivative signal) for the output layer. (Strictly speaking, the Purkinje-to-deep connectivity is more like 850-to-35, according to Pellionisz, rather than many-to-one in a precise sense; however, this is no problem, if we assume that the 35 related deep cells are representing a common variable, using multiple channels to permit greater precision.) 

                But how do you get this training signal from the olive?  If you read things like Houk and Barto,  they begin to be a little incoherent when it comes explaining how the olive learns to give a training signal specific to a given action variable.  Now it turns out, on the neurocontrol side, that there is one and only one class of working design (we now know) that yields output in a powerful way that you can use to train individual action nodes.  And it's this weird thing here (in Figure 18)...and you have this particular kind of critic network that outputs training sig­nals to an action network, and what do these things represent?  They represent the derivatives of J with respect to the individual, specific action variables.

 

 

 

 

                                                              Figure 18. Example of an ANN Neurocontrol Design

 

                Now, I don't think the cerebellum is doing this, I've matched Figure 18 to the cerebellar circuit and it doesn't fit. But it is very clear to me from what is going on here that those olive signals must represent the derivative of J with respect to u. Is there any way that that could happen?  

                Well, it turns out the cerebellum suggests a new design different from what we've had before, that is still under­standable in terms of the same mathematics. The basic idea is that the olive output has to be trained in one of two ways:

 

either (1) there's a local target that tells each climbing fiber what the derivatives of J with respect to u are. In this arrangement, the point is that the olive is antic­ipating a slower system. The need for speed explains why it is good to have an olive; that's why you don't just use the target itself to train the cerebellum.

       (2)  The other hypothesis is that a lot of complicated learning is going on inside the olive  (perhaps like Figure  18, for example). But you can't do that unless you have an efference copy of the action vector u.  

                So there are only two possible hypotheses.  One is that there is some kind of backpropagation along the climbing fi­bers -- either in the climbing fibers or along them -- so that somehow a target which is local to the action region gets back to train the olive.  That's one possibility; it's pretty crazy, but you could cut the fiber and find out.  The other possibility is that the fibers that go from the deep nuclei back to the olive provide a complete efference copy which is used in training the olive.  If that is true, then you cut those fibers and eliminate plasticity in the olive (or modify it substantially).  

                You can do the experiment. I can't. I hope that you do. Thank you.

 

References

 

[1] W.Nauta and M.Feirtag, Fundamental Neuroanatomy. W.H.Freeman, 1986.

[2] P.Werbos, The Roots of Backpropagation: From Ordered Derivatives to Neural Networks and Political  Forecasting, Wiley, 1993.

[3] P.Werbos, "Neural networks and the human mind: new mathematics fits humanistic insight," in Proc. Conf. SMC,     IEEE, 1992. An updated version is in [2].

[4] P.Werbos, "Quantum theory and neural systems: alternative approaches and a new design," in K.Pribram, ed.,

 

 

 

    Rethinking Neural Networks: Quantum Fields an Biological Evidence, INNS Press/Erlbaum, 1993.

[5] P.Werbos, "The cytoskeleton: why it may be crucial to human learning and neurocontrol," Nanobiology, Vol.1,      No.1, 1992.

[6] M.Baudry and J.Davis, eds, Long-Term Potentiation: A Debate of Current Issues, MIT Press, 1991.

[7] T.Smirnova, S.Laroche, M.Errington, A,Hicks, T.Bliss and J.Mallet, "Transsynaptic expression of a

    presynaptic glutamate receptor during hippocampal long-term potentiation," Science, Vol. 262, p.430-436,

    October 15, 1993.

[8] J.Dayhoff, S.Hameroff, R.Lahoz-Beltra and C.Swenborg,"Intracellular mechanisms in neuronal learning: adaptive  models", IJCNN92 Proceedings, IEEE, 1992.

[9] F.Fukamauchi, C.Hough and D.Chuang, "m2- and m3-Muscarinic Acetylcholine receptor mRNAs have different  responses to microtubule-affecting drugs," Molecular and Cellular Neurosciences, Vol. 2, p.315-319, 1991.

[10] C.Hough, F.Fukamauchi and D.Chuang, "Regulation of b-adrenergic receptor mRNA in rat C6 glioma cells is  sensitive to the state of microtubule assembly," Journal of Neurochemistry, Vol. 62, No.1, 1994.

[11] G.Stuart and B.Sakmann, "Active propagation of somatic action potentials into neocortical pyramidal cell dendrites," Nature, Vol. 367, No. 6458, p.69-72, Jan. 6, 1994.

[12] D.Gardner, "Backpropagation and neuromorphic plausibility," WCNN93 Proceedings, INNS Press/Erlbaum, 1993.

[13] D.Gardner, The Neurobiology of Neural Networks, MIT Press, 1993.

[14] P.Werbos and A.Pellionisz, "Neurocontrol and neurobiology," In IJCNN Proceedings, IEEE, 1992.

[15] D.White and D.Sofge,eds, Handbook of Intelligent Control: Neural, Fuzzy and Adaptive

   Approaches. Van Nostrand, 1992.

[16] D.C.Tam, personal communication and "A new conditional correlation statistics for detecting spatio-temporally correlated firing patterns in a biological neuronal network," WCNN93 Proceedings, INNS Press/Erlbaum, 1993.

[17] W.Miller,R.Sutton and P.Werbos,eds, Neural Networks for Control. MIT Press, 1990.

[18] P.Werbos, "Supervised learning," WCNN93 Proceedings, INNS Press/Erlbaum, 1993.

[19] P.Werbos, "Elastic fuzzy logic: a better fit to neurocontrol and true intelligence," Journal of

     Intelligent and Fuzzy Systems, Vol. 1, No. 4, 1993.

[20] K.Pribram and M.Gill, Freud's Project Reassessed, Basic Books, 1976.

[21] K.Pribram,"Familiarity and novelty: the contributions of the limbic forebrain to valuation and the

     processing of relevance," in D.Levine and S,Leven, eds, Motivation, Emotion and Goal Direction in

     Neural Networks, Erlbaum, 1992.

[22] P.Werbos, "Backpropagation through time: what it does and how to do it," Proceedings of IEEE, October 1990  issue. A slightly updated version is in [2].

[23] P.Werbos, "Neurocontrol: where it is going and why it is crucial," in I. Aleksander and J.Taylor, eds,

     Artificial Neural Networks II. North Holland, 1992. An updated version is in [2].

[24] P.Werbos, "Applications of advances in nonlinear sensitivity analysis," in R.Drenick and F.Kozin, eds,

     Systems Modeling and Optimization: Proc. of the 1981 IFIP Conf., Springer-Verlag, 1982. Reprinted in [2].

[25] P.Werbos, "Building and understanding adaptive systems: a statistical/numerical approach to factory automation  and brain research, IEEE Trans. SMC, Jan.-Feb. 1987.

 

 

 

 

Additional Points After Publication

 

1. The exploratory project by Houk, Cockberger and Alford was not successful. They were unable even to attempt the search for plasticity in olive cells, because they were unable to create a viable culture of these cells. Following up on the collaborative work of Houk and Barto, assuming a Barto-Sutton-Anderson Critic model, they felt it was most critical to have a co-culture of olive cells with spinal cells (to supply pain inputs, to provide the U input necessary to the training of the Critic.). I suggested that cerebellar cells (preferably deep) be added as well, because the more sophisticated critic models I claim to be essential would also requirte such cells. I have argued that the simple Barto-Sutton-Anderson design simply does not make sense in this application. (This, despite the fact that I published the same simple critic adaptation rule in 1977, 6 years prior to the Bart-Sutton-Anderson paper; in that 1977 paper, however, I pointed out the weaknesses of that method when dealing with highly multivariate systems. By engineering standards, the cerebellum is incredibly multivariate.) Nevertheless, they never even attempted a co-culture including cerebellar cells.

    In their work, however, Houk and Hockberger did cite a successful co-culture of olive and cerebellum cells by J. Mariani of Paris. Clearly, a replication of that work would be the first step to going ahead and testing this crucial first issue of demonstrating plasticity in olive cells. This, as a secondary matter,would rule out at least some versions of the feedback-error-learning model proposed by Kawato for this system.

  

2. Since this was published, I have had occasion to look morte carefully at the work of Gardner, underl;ying some of the published claims. Gardner’s book takes pains to describe evidence regarding reverse synapses. Some of the more recent work cited in this paper actually makes the point much more firm; however, the additional details in Gardner’s book are of at least some interest. Howqever, Garner does not address at all the more central issue of complete retrograde flows, requiring backwards transmission within cells. One of the papers cited here does that somewhat (re retrograde electrical flows), and the paper by Dayhoff et al in IJCNN92 deal with it much more clearly than anyone else for the option of microtubule effects. In the invertebrate literature, however, Mpitsos has pointed out to me an amazing rich litterature on both classical and Palovian conditioning, which -- so far as I can tell as yet -- does still suggest that one could find true backpropagation (of utility) even in very simple organisms.

It should be possible to prove something analogous to the Bel;l’s Theorem in physics, pointing towards specific experiments which unambiguously are inconsistent with the ordinary postulated “forwards” flows of information in the learning process. There are truly exciting possibilities here which are only just now coming into focus.

 

3. Naturally there has been great progress in implementing some of these designs in engineering since this paper was published.  A more up-to-date understanding of the entire neurocontrol situation may be found in chapter F1.10 of the Handbook of Neural Computation from Oxford U. Press, Fiesler et al eds, 1995. Some recent results in more brain-like adaptive critic architectures are reviewed in my paper in 1994 Workshop NN/FS/ES/VR, edit by M.Padgett of Auburn U. (who may have advance copies), published in 1995 by SPIE of Bellingham, Washington.

A more specific discussion for engineers of how they can play a critical role in engineering-biology collaboration, and how, is in my paper in the proceedings of the 1994 Yale Workshop on Adaptive and Learning Systems, available from the editor, Prof. K.S. Narendra at the Electrical Engineering Department of Yale University.

 

4. Grossberg has pointed out that certain limitations in the Klopf architecture would make it intrinsically unable to account for more complex examples of classical conditioning. He argues that an “expectations” subsystem is essential. I fully agree with Grossberg’s position. The more recent Critic designs cited here fully account for that concern. In particular, in postulating that the thalamo-cortical system is adapted primarily as a system identification or expectations system, I am hypothesizing that a rather large chunk of the brain is dedicated to that task. Nevertheless, it is reassuring that a very simplified version of this class of model does perform well relative to more classical models, for the kinds of experiments that have been the focus of animal behavior modelling in the past.