A verbatim copy of



COMPUTATION AND THE SINGLE NEURON

by Christof Koch

taken from NATURE, 16 January, 1997



annotated and with two appendicesa1   a2

by   M.Robert Showalter


Evidence and argument are presented here supporting the position that current neuroscience data cited by Koch and connected to Koch's article, combined with the zoom FFT EEG measurements of David Regan, strongly supports the Showalter-Kline (S-K) passive conduction theory(1). The S-K theory appears to permit memory, and to imply processing speeds and logical capacities that are much more brain-like than permitted by the current Kelvin-Rall conduction model. The annotation of Professor Koch's article here is believed to be a good way to show how S-K theory fits in with what is now known and thought about neural computation. I argue that Regan's data, combined with the data cited and interpreted by Koch, requires the sort of sharp resonance-like brain behavior that occurs under S-K, that cannot occur under Kelvin-Rall.

Over the past few decades, neural networks have provided the dominant framework for understanding how the brain implements the computations necessary for its survival. At the heart of these networks are simplified and static models of nerve cells. But neuroscience is undergoing a revolutiona0, and one consequence is that the picture of how neurons go about their business has altered.

To appreciate that, we need to start with a simple account of neuronal information processing. A typical neuron in the cerebral cortex, the proverbial gray matter, receives input from a few thousand neurons and, in turn, passes on messages to a few thousand other neurons (Fig. 1). These connections are hard wired in the sense that each connection is made by a dedicated wirea1, the axon, so that, unlike processors in a computer network, there is no competition for communication bandwidth.

Apart from the axons, there are three principal components of a neuron. The cell body (or soma); dendrites, which are short, branched extensions from the cell body, and which in the traditional view simply receive stimuli from other neurons and pass them on to the cell body without much further processing; and synapses, the specialized connections between two neurons.

Synapses are of two types, excitatory and inhibitorya2. An excitatory synapse will slightly depolarize the electrical potential across its target cell, while an inhibitory input will hyperpolarize the cell. If the membrane potential at the cell body exceeds a certain threshold value, the neuron generates a millisecond-long pulse, called and action potential or spike. (Fig.2, overleaf). Otherwise, it remains silent. The amount of synaptic input determines how fast the cell generates spikes; these spikes are in turn conveyed to the next target cells through the output axon. Information processing in the average human cortex would rely on the proper interconnections of about 4 x 1010 such neurons in a network of stupendous sizeb2.

In 1943, McCullough and Pitts showed that this (direct connection) view is at least plausible. They made a series of simplifications that have been with us ever since. Mathematically, each synapse is modelled by a single scalar weight, ranging from positive to negative depending on whether the synapse is excitatory or inhibitory. As a whole, the neuron is represented as a linear threshold element; that is, the contributions of all the synapses, multiplied by their synaptic weights, add linearly at the cell body. If the threshold is exceeded, the neuron generates a spike. McCollough and Pitts argued that, with the addition of memory, a sufficiently large number of these logical "neurons," wired together, can compute anything that can be computed on any digital computera3.

Significant as their (McCollough and Pitts) result was, it left some major questions unanswered. Most importantly, how is such a network set up in the absence of an external programmer? A network of these abstract neurons has to be programmed from the outside to do any particular job, but the brain assembles by itself. Clearly, the information needed to provide a unique specification of the 2 x 1014 synapses in human cerebral cortex cannot possibly be encoded in the genome. Moreover, a key feature of biological nervous systems that radically distinguishes them from present-day computers is their ability to learn from previous experience, presumably by adjusting synaptic weightsa4. Finally the properties of real neurons are much more elaborateb4 and variable (Fig 2b) than the simple and reliable units at the heart of McCullough and Pitts's neural networks and that of their progeny.

A possible mechanism addressing the need for self-programming was postulated a few years later by Donald Hebb: "When an axon on cell A is near enough to excite cell B and repeatedly or consistently takes part in firing it, some growth process or metabolic change takes place in one or both cells such that A's efficiency, as one of the cells firing B, is increased." Here, according to the principle known as synaptic plasticity, the synapse between neuron A and neuron B increases its "weight" if activity in A occurs at the same time as activity in Ba5

Much of the excitement in the neural network revolution of the 1980's was triggered by the discovery of learning rules for determining the synaptic weights using variants of Hebb's rule of synaptic plasticity. These rules allow the synaptic weights to be adjusted so that the network computes some sensible function of its input - it learns to recognize a face, to encode eye position or to predict the stock market, for example.

Underlying these developments was the view that memory was stored in the synaptic weights. Changing connection strength can make the network converge to a different stable equilibrium, equivalent to recalling a different memory (associative memory, ref 3). Experimentally, the best studied aspect of increasing connection strength is long-term potentiation, which can be categorized as an increase in synaptic weight lasting for days or even weeks. It is induced by simultaneous activity in the pre-and postsynaptic terminals (Fig.1), in agreement with Hebb's rule. Of more recent vintage is the discovery of a complementary process, a decrease in synaptic weight termed long-term depressiona6.

These neural-network models assumed that, to cope with the apparent lack of reliability of single cells, the brain makes use of a "firing rate" code. Here, only the average number of spikes within some suitable time window, say a fraction of a second, matters. The detailed pattern of spikes was thought to be largely irrelevant (Fig, 2b). This hypothesis was supported by experiments in which monkeys were given the task of discriminating a pattern of moving dots amid background noise. The monkey's performance could be statistically predicted by counting spikes in single neurons within visual cortex, implying that all of the information needed to perform the task is available from the average firing rate alone. Further precision could in principle be obtained by averaging the firing rates over a large number of neurons (a process known as population coding), without any need to invoke the precise timing of individual spikes as a source of informationa7.

A snapshot of the "standard view" of information processing in the brain, say as of 1984, would be based on simple, linear-threshold neurons using firing-rate code that waxes and wanes over hundreds of milliseconds to communicate information among neurons. Memory and computation are assumed to be expressed by the firing activity of large populations of neurons whose synaptic weights are appropriately set by synaptic learning algorithms. But since then neuroscience has undergone a phase of exponential growth, with progress in three separate areas - in the study of dendrites, of spikes and their timing, and of synaptic plasticity.


ACTIVE DENDRITES

It was Ramon y Cajal who, in the late nineteenth century, first revealed the intricacies and complexities of real neurons. He showed that a prototypical neuron, say a cortical pyramidal cell as shown in Figs 1 and 3, has an extended dendritic tree which is scores of times larger in surface area than the cell body. It has become clear that dendrites do much more than simply convey synaptic inputs to the cell body for linear summation. Indeed, if this is all they did, it is not obvious why dendrites would be needed at all; neurons could be spherical in shape and large enough to accommodate all the synaptic inputs directly onto their cell bodies. A few neurons do follow this geometrical arrangement but the vast majority are more like the cell shown in Fig. 3, with an extended dendritic tree. This is where many of the synaptic inputs to the cell are received, but much of the membrane area is still devoid of synapses; so the function of the elaborate structure cannot simply be to maximize the surface area for synaptic contact.

Dendrites have traditionally been treated as passive cables, surrounded by a membrane which can be modelled by a conductance in parallel with a capacitor. When synaptic input is applied, such an arrangement acts as a low-pass filter, removing the high frequencies but performing no other significant information processinga8. Dendrites with such passive membranes would not really perturb our view of neurons as linear threshold unitsb8.

But dendrite membranes are not passive - they contain voltage-dependent membrane conductances. These conductances are mediated by protein complexes that allow charged ions to flow across the membrane. As long ago as the 1950's, Hodgkin and Huxley showed how the transient charges in such conductances generate and shape the action potentiala9. But it was assumed that they are limited to the axon and the adjacent cell body. We now know that many dendrites of pyramidal cells are endowed with a relatively homogeneous distribution of sodium conductances as well as a diversity of calcium membrane conductancesb9.

What is the function of these active conductances? In a passive cable structure, synaptic input to the more distant regions of the dendritic tree (as the to of Fig.3) would quickly saturate, delivering only a paltry amount of electric current to the spike initiating zone far awaya10. But it turns out that synaptic input to this part of the tree is sufficient to elicit somatic action potentials, and the likely explanation (supported by computer models) is that calcium and potassium membrane conductances in the distant dendrites can selectively amplify this inputb10.



Voltage-dependent conductances can also subserve a specific nonlinear operation, multiplication. In a passive dendritic tree, the effect of two synaptic inputs is usually less than the sum of the two individual inputs; that is, they show saturation. This saturation effect can be minimized by spreading the synapses far apart. If, however, the dendritic tree contains sodium and calcium conductances, or if the synapses use a particular kind of receptor (the so-called NMDA receptor), the inputs can interact synergistically: now, the strongest response occurs if inputs from different neurons are located close to each other on a patch of dendritic membrane. Computer simulations show that such a neuron effectively performs a multiplication; that is, its firing rate is proportional to the product, rather than the sum, of its inputs. Multiplication is one of the most common operations carried out in the nervous system (for example, for estimating motion or the time-to-contact with an approaching stimulus), and it is tempting to speculate that this may be achieved through such processes at the cellular level.

A further development has been the realization that the distribution of calcium ions within dendrites may represent another crucial variable for processing and storing information. Calcium enters the dendrites through the voltage-gated channels, and this, along with its diffusion, buffering and release from intracellular stores, leads to rapid local modulations of calcium concentration within the dendritic tree. The concentration of calcium can, in turn, influence the membrane potential (through calcium-dependent membrane conductances) and - by binding to buffers and enzymes - turn local biochemical signalling pathways on or offa11.

It was also Ramon y Cajal who postulated the law of "dynamic polarization,' which stipulates that dendrites and cell bodies are the receptive areas for the synaptic input, and that the resulting output pulses are distributed unidirectionally along the axon to its targets. This assumes that action potentials travel only along axons: no signal was thought to travel outwards along the dendrites.

From work on brain slices, however, it seems that this is by no means the whole story. Single action potentials can propagate not only forwards from their initiation site along the axon, but also backwards into the dendritic tree (a phenomenon known as "antidromic spike invasion"). It remains unclear whether dendrites can initiate action potentials themselves. If they can, such dendritic spikes could support theoretical proposals that all-or-none logical operations occur in the dendritic tree. The next step will be to find out whether action potentials can propagate into the dendritic tree under more natural conditions - that is, using sensory stimuli in an intact animal.

Thus, it is now evident that the dendritic tree is far more complex than the linear cable models of yesteryear assumed. Dendrites provide the substrate for numerous nonlinear operations, and endow neurons with much greater information-processing capacity than was previously suspecteda12.

TIMING COUNTS

The second area in which our thinking has changed has to do with the role of time in neuronal processing. There are two main aspects to this issue - first, the relationship between the timing of an event in the external world and the timing of the representation of that event at the single-neuron level; second, the accuracy and importance of the relative timing of spikes between two or more neurons.

Regarding the first question, some animals can discriminate intervals of the order of a microsecond (for instance, to localize sounds), implying that the timing of sensory stimuli must be represented with similar precision in the braina13. But this usually involves highly specialized pathways, probably based on the average timing of spikes in a population of cellsb13. However, it is also possible to measure the precision with which individual cells track the timing of external events. For instance, certain cells in the monkey visual cortex are preferentially stimulated by moving stimuli, and these cells can modulate their firing rate with a precision of less than 10 ms (ref 16).

The second aspect of the timing issue is the extent to which the exact temporal arrangement of spikes - both within a single neuron and across several neurons - matters from information processing. In the past few years there has been a resurgence of signal processing and information-theoretical approaches to the nervous system. In consequence, we now know that individual neurons, such as motion-selective cells in the fly or single auditory inputs in the bullfrog, can encode between 1 and 3 bits of sensory information per strike, amounting to rates of up to 300 bits per second. This information seems to be encoded using changes in the instantaneous interspike interval between a handful of spikesa14. Such a temporal encoding mechanism is within 10-40 per cent of the theoretical maximum allowed by the spike train variability. This is quite remarkable because it implies that individual spikes in a single cell in the periphery can carry significant amounts of information, quite at odds with the idea that neurons are very unreliable and can only signal in the aggregate. At these rates, the optic nerve, which contains about one million fibers, would convey between on and a hundred million bits per second - compare this with a quad-speed CD-ROM drive, which transfers information at 4.8 million bits per second.

The relative timing of spikes in two or more neurons can also be remarkably precise. For instance, spikes in neighboring cells in the lateral geniculate nucleus, which lies midway between the retina and the visual cortex, are synchronized to within a millisecond. And within the cortex, pairs of cells may fire action potentials at predictable intervals that can be as long as 200 ms, two orders of magnitude longer than the delay due to a direct synaptic connection, with a precision of about 1 msa15. Such timing precision across populations of simultaneously firing neurons is believed to be a key element in neuronal strategies for encoding perceptual information in the sensory pathwaysb15.

If neurons care so much about the precise timing of spikes - this is, if information is indeed embodied in a temporal code - how, if at all, is it decoded by the target neurons? Do neurons act as coincidence detectors, able to detect the arrival time of incoming spikes at millisecond or better resolution? Or do they integrate more than a hundred or so relatively small inputs over many tens of milliseconds until the threshold for spike initiation is reached? (Ref 23, Fig 2a) These questions continue to be widely and hotly debateda16.

SYNAPTIC PLASTICITY

Back-propagating action potentials are puzzling if considered solely within the context of information processinga17. But they make a lot of sense is seen as "acknowledgement signals" for synaptic plasticity and learning. A Hebbian synapse is strengthened when pre and postsynaptic activity coincide. This can occur if the presynaptic spike coincides with the postsynaptic spike that is generated close to the cell body and spreads back along the dendritic tree to the synapse. A new and beautiful study shows that the order of the arrival time between the presynaptic spike and the back-propagated postsynaptic spike is critical for synaptic plasticity.

If the presynaptic spike precedes the postsynaptic spike, as should occur if the first participates in triggering the second, then long term potentiation occurs - that is, the synaptic weight increases. If, however, the order is reversed, the synaptic weight decreases. So sensitive is this sequence that a change of timing of as little as 10 ms either way can determine whether a synapse is potentiated or depresseda18. The purpose of this precision is presumably to enable the system to assign credit to those synapses that were actually responsible for generating the postsynaptic spike.



These (time sensitive) experiments come at a fortuitous time, for theoretical work has begun to incorporate asymmetric timing rules into neural network models. The attraction of doing so is that it allows the network to form associations over time, enabling it to learn sequences and predict events. It is therefore gratifying to find that synapses with the required properties do indeed exist within the brain.

In the past year there has also been the emergence of a new way of thinking about short-term plasticity, one that complements the view of long-term synaptic changes for memory storage. This has come about by joint experimental-theoretical work (refs 26, 27, see overleaf), suggesting that individual synapses rapidly adapt to the presynaptic firing rate, primarily signalling an increase or a decrease in their input. That is, synapses continuously adapt to their input, only signalling relative changes, which means that the system can respond in a highly sensitive manner to a constantly and widely varying external and internal environmenta19. This is entirely different from digital computers that enforce a strict segregation between memory (onboard cache, RAM, or disk) and computation. Indeed they are carefully designed to avoid adaptation and other usage dependent effects from occurring.

Interestingly, single-transistor learning synapses - based on the floating gate concept underlying erasable programmable ROM digital memory - have now been built in a standard CMOS manufacturing process. Like biological synapses, they can change their effective weight in a continuous manner while they carry out computations. Floating-gate synapses will greatly aid attempts to replicate the functionality of nervous systems by the appropriate design of neuromorphic silicon neurons using analog very-large-scale-integrated (VLSI) circuit fabrication technology.





HYBRID COMPUTER

Overall, then, current thinking about computation in the nervous system has the brain as a hybrid computer. Individual nerve cells convert the incoming streams of digital pulses into spatially distributed variables, the postsynaptic membrane potential and calcium redistribution. This transformation involves highly dynamic synapses that adapt to their inputsa20.

Information is then processed in the analog domain, using a number of linear and nonlinear operations (multiplication, saturation, amplification, thresholdinga21) implemented in the dendritic cable structure and augmented by voltage dependent membrane and synaptic conductances. The resulting signal is then converted back into digital pulses and conveyed to the following neuronsb21. The functional resolution of these pulses is in the millisecond range, with temporal synchrony across neurons likely to contribute to coding. Reliability could be achieved by pooling the responses of a small number (20-200) of neuronsc21.

And what of memory? It is everywhere (but can't be randomly accessed). It resides in the concentration of free calcium in dendrites and the cell body; in the presynaptic terminal; in the density and exact voltage-dependency of the various ionic conductances; and the density and configuration of specific proteins in the postsynaptic terminalsa22.

Only very little of this complexity is reflected in today's neural network literature. Indeed, we sorely require theoretical tools that deal with signal and information processing in cascades of such hybrid, analog-digital computational elements. We also need an experimental basis, coupled with novel unsupervised learning algorithms, to understand how the conductances of a neuron's cell body and dendritic membrane develop in time. Can some optimization principle be found to explain their spatial distribution?

As always, we are left with a feeling of awe for the amazing complexity found in nature. Loops within loops across many temporal and spatial scales. And one has the distinct feeling that we have not yet revealed every layer of the oniona23. Computation can also be implemented biochemically - raising the fascinating possibility that the elaborate regulatory network of proteins, second massagers and other signalling molecules in the neuron carry out specific computations not only at the cellular but also at the molecular level.

ARTICLE INSERT: THE ADAPTABLE SYNAPSEa24

In short-term synaptic depression, the postsynaptic response to a regular train of presynaptic spikes firing at a fixed frequency f gradually lessens. The response of the first spike might be large, but subsequent responses will be diminished until they reach a steady state (expressed in terms of A, the fractional reduction in postsynaptic effect). This is shown here (part a) for presynaptic spikes at 40 Hz frequency (data from ref 26).

This depression generally recovers within 0.1 to 0.5 s. For firing rates above 10 Hz, A is roughly inversely proportional to the firing frequency. In other words, within a few hundred milliseconds the synapse will have adapted to the presynaptic firing with a response roughly independent of the firing rate (due to the inverse relationship between A and f). If the synaptic


consequence, the transient change in the postsynaptic response will be proportional to the relative change in firing frequency. This is demonstrated in computer simulations in which the presynaptic firing rate of a couple of hundred such synapses converging onto a model neuron is increased fourfold (part b - the increases are from 25 to 100 Hz on the left and from 50 to 200 Hz on the right; data from ref. 27).

Even though the final input rate is twice as high on the right side as on the left, the firing rate of the neuron is roughly the same. This is because the fractional increase - relative to the background rate - is the same in both cases. This form of short-term depression in synaptic strength therefore enables synapses to respond to relative changes in firing rates rather than to absolute rates.

Appendix 1.   Resonance in neurons, as measured, and as calculated by S-K theory.

David Regan has measured brain magnetic fields (MEG) and scalp voltages (EEG) during evoked stimulation. Figs 9 and 10(5) show some of Regan's measurements using his zoom FFT technique.

a1

Fig. 9


The caption for this figure reads:

Figure 10 shows evidence of bandwidths narrower than bin widths of .0019 Hz. The 4F component is shown. Bandwidths of the peaks measured in Fig 9 may have been no wider than this.




Fig. 10



These sharp, high information content patterns, with stimuli frequency multiples organized roughly as shown, show that brain includes sharply resonant components, and shows that these components are coupled with very small lags and with small damping. The bandwidths Regan measures are too sharp by at least an order of magnitude to be produced by membrane channel activity.

The pyramidal cells in brain, interpreted according to the S-K theory, should behave in a manner that generates the kind of behavior that Regan measured. Conduction lines have very low distortion. Effective line inductances are far greater than those predicted by Kelvin-Rall - high enough so that neurons can be inductively coupled via the intercellular medium, which conducts millions of time faster than line conduction speed. The dendritic spines have the sharply resonant requirements Regan's data appears to require.

Resonance:

Resonance is logically interesting. Enormous resonant magnifications of tightly selected signals are possible. In this sense, resonant systems can function as highly selective amplifiers. This fact is a foundation of communication technology. Radio and television offer familiar examples of resonant selectivity. Radio and television receivers exist in an electromagnetic field consisting of a bewildering and undescribable variety of electromagnetic fluctuations. Reception occurs because the resonant receiver is selective for a specific frequency at a high degree of phase coherence. Signals off frequency are not significantly detected, and "signals" of random phase that are on frequency cancel rather than magnify in resonance. Radar receivers also operate on the principle of resonance. Other examples are our telephone system and cable television system, each organized so that a multiplicity of different signals can be carried in physically mixed form over the same conduits. These "mixed" signals can be separated and detected with negligible crosstalk by resonant means.

Electrical resonance can store up energy in an oscillation having a peak voltage Q times the oscillating voltage of the exciting disturbance. Resonant systems may all be described in wave propagation terms, and many can also be treated in lumped terms. The LRC oscillator common in differential equation textbooks is an example of a resonant system described in lumped terms.



The International Dictionary of Applied Mathematics explains inductance-resistance-capacitance (LRC) series resonance as follows, and describes behavior generally characteristic of resonance. The "coil" is a lumped inductance, the "condenser" is a lumped capacitance, and "j" is the square root of -1.

. . . In an a-c circuit containing inductance and capacitance in series ... the impedance is given by





L is the inductance, and C is the capacitance. It can be readily seen that at some frequency the terms in the bracket will cancel each other, and the impedance will equal the resistance alone. This condition, which gives a minimum impedance (and thus a maximum current for a fixed impressed voltage) and unity power factor is known as series resonance. Where the resistance is (relatively) small the current may become quite large. As the voltage drop across the condenser or coil is the product of the current and the impedance of that particular unit, it may also become very large. The condition of resonance may even give rise to a voltage across one of these units that is many times the voltage across the whole circuit, being, in fact, Q times the applied voltage for the condenser and nearly that for the coil. This is possible since the drops across the coil and condenser are nearly 180 degrees out of phase, and thus almost cancel one another, leaving a relatively small total voltage across the circuit . . .(6)



Fig 11 shows how the voltage oscillation stored in a resonant system grows when it is driven by a input signal at its resonant frequency. The growth shown depends on stimulus phase, a fact on which FM transmission depends. If, at some time, the input signal voltage shown were to shift phase 180 degrees, the resonant voltage would decrease as fast as it is shown increasing here.














The resonant amplification factor, Q, achieved after time to equilibrium(7), is:






For an LRC resonator, Q is




High Q's are prized in information processing systems, partly because bandwidth (the frequency difference between the half power points on a resonance curve) is inversely related to Q according to the formula:




For spines, Q's in the tens of thousands are possible.

Columns can also be sharply resonant. Columns (transmission lines) of 1/4 and 1/2 wavelength have been used as resonators in musical instruments for many centuries. More recently, column resonance has been used with precision in the radar and communication fields. Well terminated lengths of neural passage that are sharply open (short circuited) at both ends are resonant when there length is exactly 1/4 of wavelength. For column length lc:




and integer multiples of resonant frequency, omegao. A well terminated length of neural

passage that sharply closed on one end will be resonant at




and integer multiples of these frequencies.



For a neural line in the constant velocity regime:




As a 1/4 wave resonant column of length lc:

.


Q of the column resonator will be inverse with attenuation per









Column resonators are more powerful information handling devices than lumped resonators of the same Q, because they magnify and store repeating WAVEFORMS that fit as standing waves within them. (For this reason, a wind instrument or pipe organ wave form can be much more complicated than a sine wave.) In contrast, an LRC resonator stores a sine wave.

The brain appears to be a resonant system, and if it may be judged by the Q's it shows, a very capable one. Considering Regan's frequencies of 7-46 Hz, and setting bw at .0019 Hz, we calculate Q's of 3680 to 24,200 for the ensembles that represent frequency peaks. These are very high calculated ensemble Q's, higher than the Q's of even the best nonsuperconducting tuned circuits. Individual resonator Q's must be higher still. Regan's measurements give upper bounds on bandwidths. From neuroanatomy, the candidate resonant structure seems plain - the dendritic spines. There are about 1013 dendritic spines in brain, and, based on the values of effective inductance predicted here, the spines appear to be resonant structures with very high resonant amplification factors (Q's).





SPINE ANATOMY AND SPINE RESONANCE:

Fig 12(8) shows camera lucida drawings of a neuron body and proximal dendrites of a rat hippocampal (CA1) neuron showing common spine types: thin, mushroom shaped, and stubby. The "thin" type is the commonest in cortex (about 70%), and can be considered as both an LRC element and a column resonant element. In LRC mode, the bag and shaft have a lumped capacitance, and shaft inductance and resistance are considered in lumped form. In column mode, the shaft is a column open at both ends, with a termination correction.





Figure 13 shows an electrical model of a thin spine. In the scaled figure, 4/5 of spine capacitance is in the bag section.



The "spine" of Fig. 13 can be modelled as an LRC resonant system. Capacitance is the capacitance of the "bag" section, plus half the capacitance of the shaft section. The shaft has resistance of R, and an effective inductance of Ledelta x. The Le/R ratio is inverse with diameter. Different bag sizes for the same shaft size yield different LC products, and different resonant frequencies. Q, radian frequency, and bandwidth, for the LRC case are:








In the model, bandwidth is proportional to diameter. Let's arbitrarily chose a shaft diameter of .1 micron, shaft length of

.5 micron, interspine medium conductance of 110 ohm-cm, membrane capacitance of 1 microfarad/cm2, and zero membrane leakage, g. Holding these values, and varying bag size, yields the following relation between frequency and Q.

Q = o(910)

For Regan's measured frequency range of 7-45 hz (44-283 radians/sec,) Q's between 40,000 and 257,000 are estimated. Regan's data correspond to Q's about a decade smaller, between 3,680 and 24,000 over that same 7-45 Hz frequency range. This is an acceptable fit because:

Regan must have measured ensemble properties, not properties of single neural elements.

Regan's setup could have detected no tighter bandwidths than he did detect. and

Within the constraints of biological knowledge, we could have guessed other values of the parameters to come closer to Regan's values (or even to match them.)

Figs 14 and 15 below show steady-state magnification of a signal as a function of frequency calculated for the LRC spine model of Fig 13. The peak magnification factor is about 70,000. Note the sharpness of the magnification as a function of frequency.




The model spine of Fig 13 would also have a column resonance mode. Spine column resonant frequency will be approximately





and could estimate column resonant Q from equation () as




Electrical compliance of the bag would shift these resonant frequencies and Q's somewhat from the simple 1/4 wave column calculation set out above, but the correction would involve details that can be considered elsewhere.

Referring again to Regan's data, the brain has many (about 1013) spines. If spine resonant frequencies are widely distributed, and some reasonable fraction of the dendritic spines are in the high Q state, one would expect fixed frequency stimuli, such as Regan supplied, to yield the sort of excitation curves that Regan observed. Coupling of the spines would be via the very rapid conduction of the extracellular medium, not via conduction along dendrites or axons.

SPINE SWITCHING:

Spines are adapted for off switches. A single membrane channel can turn off spine resonance. This may be useful

because otherwise, voltage buildups sufficient to break down the dielectric of the spine bags, with consequent spine destruction, might occur(9) and

because binary switching is useful in information processing.





Suppose there is one membrane channel in the bag portion of the spine. If that channel is open, it acts as a shunt, damping voltage fluctuations that occur across it. There will be a shunt across the bag membrane. The spine will have a shunt limited Q, Qdamped. Let Rc be shunt channel resistance. If Qdamped<<Q, as it will be for a channel, we can say that:







The Qdamped << Q assumption makes sense for reasonable membrane channel values (between 4 and 400 picosiemens(10).) Opening of one channel will change a spine from a sharply resonant state, with a Q in the thousands, to a Q less than 10, a very wide bandwidth state. A single channel therefore acts as an on-off switch.



Closure:



Evidence and argument has been presented here supporting the position that current neuroscience data cited by Koch and connected to Koch's article, combined with the zoom FFT EEG measurements of David Regan, strongly supports the Showalter-Kline (S-K) passive conduction theory. That theory appears to permit memory, processing speeds and logical capacities that are much more brain-like than permitted by the current Kelvin-Rall conduction model. The annotation of Professor Koch's article was employed here as a good way to show how S-K theory fits in with what is now known and thought about neural computation. Regan's data, combined with the data cited and interpreted by Koch, requires the sort of sharp resonance-like brain behavior that occurs under S-K, that cannot occur under Kelvin-Rall.




Appendix 2: The S-K transmission equations:

The currently accepted passive neural conduction equations are the standard conduction equations of electrical engineering, usually written in a contracted form that discards the terms in L that are negligible in this equation.








Robert Showalter and Stephen Jay Kline have found that these equations lack crossterms because special restrictions on the use of dimensional parameters in coupled finite increment equations have not been understood(11)

. The crossterms, which can also be derived (within a scale constant) by standard linear algebra based circuit modelling, are negligible in most engineering applications. But these crossterms are very large in the neural context. The Showalter-Kline (S-K) equations are isomorphic to the standard conduction equations of electrical engineering and are written as follows.






For set values of resistance R, inductance L, membrane conductance per length G, and capacitance C, these equations have the solutions long used in electrical engineering. The hatted values are based on a notation adapted to crossproduct terms. In this notation, the dimensional coefficients are divided into separate real number parts (that carry n subscripts) and dimensional unit groups, as follows.




For wires,, the crossproduct terms are negligible, and the two kinds of equations are the same. But under neural conditions the crossproduct terms are LARGE. For instance, effective inductance is more than 1012 times what we now assume it to be. The S-K equation predicts two modes of behavior.



When G is high (some channels are open) behavior similar to that of the current model is predicted.

When G is low, transmission has very low dissipation, and the system is adapted to inductive coupling effects including resonance.





Under S-K, attenuation of waves per wavelength or per unit distance varies over a much larger range than occurs in the now-accepted theory. There is a low membrane conductance regime where attenuation of waves is small, and wave effects are predicted. However, as channels open attenuation increases enormously, and waves may be damped out in a few microns. Under the new model a neural passage can be either sharply "on" or sharply "off" depending on the degree of channel-controlled membrane conductance. In the high g regime, attenuation per wavelength values are qualitatively similar for the Kelvin-Rall and S-K.



Figures 2, and 3 plot unit wave amplitude after one wavelength (right axis) or damping exponent per wavelength (left axis) as a function of membrane conductance, g, for both Kelvin-Rall and S-K theory. The curves map functions that move rapidly - exponents are graphed in log-log coordinates.







Figure 2 plots calculated responses at the low frequency of 10 radians/second for neural process diameters ranging over five decades (from 1000 microns to .1 micron). For the 1000 and 100 micron cases the Kelvin-Rall and S-K curves are almost the same. Results for these large diameters and low frequencies are also nearly the same on an attenuation per unit length and a phase distortion basis. These conditions correspond to squid axon experiments that are famous tests of the Kelvin-Rall theory. However, for smaller diameters, attenuation according to S-K theory is much less than that according to Kelvin-Rall.

Figure 3 plots calculated responses at 10,000 radians/second (1591 Hz) for the same diameters plotted in Figure 2. Attenuation values are substantially less for the new theory than for the Kelvin-Rall theory even for the 1000 micron diameter case. For Kelvin-Rall, the value of the attenuation exponent per wavelength never falls below 2. This means that the maximum amplitude of a wave after 1 wavelength of propagation is .00187 (about 1/535th) of its initial value under Kelvin-Rall for any diameter neural process. It makes little sense to talk of "wave propagation" and no sense to talk about "resonant responses" under these conditions. In contrast, according to the new theory, as much as 99.995% of unit wave amplitude may remain after a single wavelength. Under these very different conditions, notions of wave propagation and resonance do make sense.

Fig. 5 plots conduction velocity versus frequency for a 1 micron dendrite or spine neck versus frequency in the case where membrane conductance, g, is approximately zero.




In the Kelvin-Rall theory, conduction velocity is proportional to the square root of frequency. In the S-K model, conduction velocity rapidly approaches an asymptote. Above a frequency threshold, conduction speed is almost constant. For large diameter neural processes, this threshold is so high that the velocity-frequency relation is similar for both theories. But for small neural processes, velocity is almost constant above quite low threshold frequencies. The following chart is based on a of 110 ohm-cm and a membrane capacitance of 1 microfarad/cm2. For a .1 micron dendrite, 99.99% of peak velocity occurs at a very low frequency (.0511 cycles/second.)


diameter                              Frequencies for the following fractions of
(microns)                                     Peak velocity (radians/second)

                                                                  95%         99%      99.99%
  1000                                                     1,320        3,160        32,120
    100                                                         132            316         3,212
      10                                                            13.2           31.6         321.2
        1                                                               1.32           3.16        32.12
         .1                                                               .132            .316        3.212





"The 99% velocity cutoff frequency according to the new model offers a good basis for comparing phase distortion predictions between Kelvin-Rall and the new theory. Phase distortion occurs when different frequency components of a signal move at different speeds. Phase distortion can rapidly degrade the information content of a signal. In Kelvin-Rall, phase distortions are enormous. However, for the new model, in the low G limit, a dendrite or axon will have a characteristic frequency 99 that has 99% of maximum propagation velocity. Above 99, propagation will be almost free of phase distortion. 99 correlates with diameter, conductivity, and membrane capacitance according to the relation



.

The S-K theory has new, interesting characteristics in its low g, low attenuation mode. This may be described as an "on" state, in contrast to the high g, high attenuation "off" state. In the "on" condition, with g very small, important relationships are simple, particularly for small neural diameters.

Components of important formulae are the "radical value" analogous to the radius in the argand diagram characterizing the transmission line, and the angle "" on that argand diagram. The radical for the low g limit model is




a neurally interesting characteristic emerges when this is factored as follows:




There is a term proportional to frequency (the fourth root of frequency4 ) The relative importance of this term grows rapidly as diameter, d, decreases. For dendrites and dendritic spines in brain, it is this frequency term that predominates. The radical can be approximated as




The rotational angle for the new theory is approximately




The quantity inside the arctangent above is strongly dependent on diameter scale, resistivity, and frequency.

Attenuation per unit distance is




In the range where resonance is of interest (above the 99% of maximum velocity frequency) the following approximations are useful:

For the wavelength :



Attenuation per wavelength (in the max velocity range) is approximated as follows:





Velocity is




For frequencies that yeild near-maximum velocity, sin() is almost 1, and



Impedance and Impedance Mismatch Effects:

Dendritic neural passages, which function as lossy transmission lines, have impedance in the electrical engineering sense. Impedance may be defined in the usual way, but with hatted values substituted for the unhatted ones.






Note that impedance has the units of lumped resistance - (for the same reason that coax and other wave carrying lines are specified in ohms, with the ohmic values far in excess of the static resistances of the lines.) The interface between two sections of transmission line generates reflections unless their impedances are matched. For a line with an impedance Zo terminated in another line (or lumped resistance) called the "load", and having an impedance Zl, the reflection coefficient Kl (the ratio of reflected to incident voltage at the load) will be




Reflection coefficient will vary from -1 to +1. When Zl>>>Zo then Kl is approximately 1; when Zl<<<Zo then Kl is approximately -1; when Zl = Zo then Kl is 0, and there is no reflection. Similar reflection rules and similar consequences are familiar to microscopists and students of acoustics.

Impedance jumps in neural passages due to changes in crossection or g are analogous to similar sharp impedance jumps that occur in wind musical instruments. Wind instruments work because they are abruptly switchable impedance mismatch devices driven by cyclic energy input means that adapt a fluctuating input flow to match air column load. A sharply defined quasi-steady resonance is established in the musical instrument's air column. Because the reflection coefficient at the discontinuity is not perfect, some of the sound from this column radiates to listeners of the music. Analogies between the acoustical function of wind instruments and neural lines are useful when considering two linked issues.

Consider a dendrite of 2 microns diameter, with a group of channels of one hundred 200 picosiemen channels over a 1 micron axial distance. Conditions are sigma=110 ohm-cm, capacitance/area=10-6 farads/cm2 ,gclosed channel=10-12 mhos/cm2. When the channels are closed, they do not effect conduction. Suppose that the channels open so that g goes from 10-12 to some much higher value. See Fig 6, which plots a change for reflection coefficient from 1.0 to less than .0025. Fig 6 was calculated for 10 radians/second, but values for much higher or substantially lower frequencies would be about the same. Note that the difference in reflection coefficient that occurs between 10-12 and 10-6 mhos/cm2 is small, but that very large changes in reflection coefficient are calculated for higher conductances. Opening a large number of channels in the side of a dendritic passage can change the channel from a transmission line to a sharply reflecting discontinuity. The physics is analogous to that which occurs when a clarinettist opens or covers a finger hole.

Abrupt changes in crossection are also calculated to produce analogous reflecting impedance mismatches.

Fig 7 shows reflection coefficient at a discontinuity between a smaller and a larger diameter. Calculated conditions are sigma= 50 ohm-cm, 10-6 farads/cm2, g=10-9, frequency=900 radians/sec, d1=1 micron. The shape and slope of this function is insensitive to changes in frequency, initial diameter, sigma, g and c. Abrupt changes in crossection can produce strongly reflecting impedance mismatches. When changes in crossection or impedance are gradual, they can occur with little or no reflection, for reasons exactly analogous to those that permit smooth transitions in the bells of wind instruments or the gradual impedance (refraction) transitions that can be arranged in optics and in waveguide practice. The S-K equations are switching, resonance adapted equations well adapted to the information processing that brains do.










NOTES AND REFERENCES (for the annotated Koch article and Appendices):

1. Regan, D. HUMAN BRAIN ELECTROPHYSIOLOGY: Evoked Potentials and Evoked Magnetic Fields in Science and Medicine Elsevior, New York, Amsterdam, London 1989 pp 100-115.

2. Sejnowski, T.J. "The year of the dendrite" SCIENCE, v.275, 10 January 1997 pp 178

3. Appendix 2, p. 6

4. Zecevic, D. "Multiple spike-initiation zones in single neurons revealed by voltage-sensitive dyes NATURE v.381 23 May 1996

pp 322-324.

5. Regan, op. cit. Fig 1.70A and Fig 1.70B, pp 106-107.

6. The International Dictionary of Applied Mathematics D.Van Nostrand Company, Princeton, Toronto, New York, London 1960

7. op. cit.

8. C.H. Horner Plasticity of the Dendritic Spine, Progress in Neurobiology, v.41, 1993 pp. 281-231, Fig 3, p 285.

9. I believe that the extensive destruction of spines and dendrites that occurs in severe epilepsy may happen in this way.

10. I.B.Levitan and L.K.Kaczmarek THE NEURON: Cell and Molecular Biology Oxford University Press, Oxford, New York, 1991, p. 65-66.

11. When the derivation of the conduction equation from a finite "physical" model is carefully done, a series of crossterms arise in addition to the terms in R, L, G, and C. This has long been known, but the crossterms have been thought to be infinitesimal. However, when the rule that crossproducts in dimensional parameters must be evaluated in intensive (point) form is accounted for, these crossterms are finite. Crossterms that are large enough to effect neural conduction (which happen to go as the inverse cube of diameter) are included in the equation. The new crossterm parameters can be combined with the old term parameters in a hat notation. For instance, the hatted value of R includes R and a crossterm that, like R, is associated with i. The hatted value of L includes L and a crossterm that, like L, is associated with di/dt.