Skip to main content
PLOS Computational Biology logoLink to PLOS Computational Biology
. 2009 May 8;5(5):e1000380. doi: 10.1371/journal.pcbi.1000380

Pairwise Maximum Entropy Models for Studying Large Biological Systems: When They Can Work and When They Can't

Yasser Roudi 1,2, Sheila Nirenberg 2, Peter E Latham 1,*
Editor: Olaf Sporns3
PMCID: PMC2674569  PMID: 19424487

Abstract

One of the most critical problems we face in the study of biological systems is building accurate statistical descriptions of them. This problem has been particularly challenging because biological systems typically contain large numbers of interacting elements, which precludes the use of standard brute force approaches. Recently, though, several groups have reported that there may be an alternate strategy. The reports show that reliable statistical models can be built without knowledge of all the interactions in a system; instead, pairwise interactions can suffice. These findings, however, are based on the analysis of small subsystems. Here, we ask whether the observations will generalize to systems of realistic size, that is, whether pairwise models will provide reliable descriptions of true biological systems. Our results show that, in most cases, they will not. The reason is that there is a crossover in the predictive power of pairwise models: If the size of the subsystem is below the crossover point, then the results have no predictive power for large systems. If the size is above the crossover point, then the results may have predictive power. This work thus provides a general framework for determining the extent to which pairwise models can be used to predict the behavior of large biological systems. Applied to neural data, the size of most systems studied so far is below the crossover point.

Author Summary

Biological systems are exceedingly complicated: They consist of a large number of elements, those elements interact in nonlinear and highly unpredictable ways, and collective interactions typically play a critical role. It would seem surprising, then, that one could build a quantitative description of biological systems based only on knowledge of how pairs of elements interact. Yet, that is what a number of studies have found. Those studies, however, focused on relatively small systems. Here, we ask the question: Do their conclusions extend to large systems? We show that the answer depends on the size of the system relative to a crossover point: Below the crossover point the results on the small system have no predictive power for large systems; above the crossover point the results on the small system may have predictive power. Moreover, the crossover point can be computed analytically. This work thus provides a general framework for determining the extent to which pairwise models can be used to predict the behavior of large biological systems. It also provides a useful heuristic for designing experiments: If one is interested in understanding truly large systems via pairwise interactions, then make sure that the system one studies is above the crossover point.

Introduction

Many fundamental questions in biology are naturally treated in a probabilistic setting. For instance, deciphering the neural code requires knowledge of the probability of observing patterns of activity in response to stimuli [1]; determining which features of a protein are important for correct folding requires knowledge of the probability that a particular sequence of amino acids folds naturally [2],[3]; and determining the patterns of foraging of animals and their social and individual behavior requires knowledge of the distribution of food and species over both space and time [4][6].

Constructing these probability distributions is, however, hard. There are several reasons for this: i) biological systems are composed of large numbers of elements, and so can exhibit a huge number of configurations—in fact, an exponentially large number, ii) the elements typically interact with each other, making it impossible to view the system as a collection of independent entities, and iii) because of technological considerations, the descriptions of biological systems have to be built from very little data. For example, with current technology in neuroscience, we can record simultaneously from only about 100 neurons out of approximately 100 billion in the human brain. So, not only are we faced with the problem of estimating probability distributions in high dimensional spaces, we must do this based on a small fraction of the neurons in the network.

Despite these apparent difficulties, recent work has suggested that the situation may be less bleak than it seems, and that an accurate statistical description of systems can be achieved without having to examine all possible configurations [2], [3], [7][11]. One merely has to measure the probability distribution over pairs of elements and use those to build the full distribution. These “pairwise models” potentially offer a fundamental simplification, as the number of pairs is quadratic in the number of elements, not exponential. However, support for the efficacy of pairwise models has, necessarily, come from relatively small subsystems—small enough that the true probability distribution could be measured experimentally [7][9],[11]. While these studies have provided a key first step, a critical question remains: will the results from the analysis of these small subsystems extrapolate to large ones? That is, if a pairwise model predicts the probability distribution for a subset of the elements in a system, will it also predict the probability distribution for the whole system? Here we find that, for a biologically relevant class of systems, this question can be answered quantitatively and, importantly, generically—independent of many of the details of the biological system under consideration. And the answer is, generally, “no.” In this paper, we explain, both analytically and with simulations, why this is the case.

Results

The extrapolation problem

To gain intuition into the extrapolation problem, let us consider a specific example: neuronal spike trains. Fig. 1A shows a typical spike train for a small population of neurons. Although the raw spike times provide a complete description, they are not a useful representation, as they are too high-dimensional. Therefore, we divide time into bins and re-represent the spike train as 0 s and 1 s: 0 if there is no spike in a bin; 1 otherwise (Fig. 1B) [7][9],[11]. For now we assume that the bins are independent (an assumption whose validity we discuss below, and in more detail in the section “Is there anything wrong with using small time bins?”). The problem, then, is to find Inline graphic where Inline graphic is a binary variable indicating no spike (Inline graphic) or one or more spikes (Inline graphic) on neuron Inline graphic. Since this, too, is a high dimensional problem (though less so than the original spike time representation), suppose that we instead construct a pairwise approximation to Inline graphic, which we denote Inline graphic, for a population of size Inline graphic. (The pairwise model derives its name from the fact that it has the same mean and pairwise correlations as the true model; see Eq. (15).) Our question, then, is: if Inline graphic is close to Inline graphic for small Inline graphic, what can we say about how close the two distributions are for large Inline graphic?

Figure 1. Transforming spike trains to spike count.

Figure 1

(A) Spike rasters. Tick marks indicate spike times; different rows correspond to different neurons. The horizontal dashed lines are the bin boundaries. (B) Spike count in each bin. In this example the bins are small enough that there is at most one spike per bin, but this is not necessary—one could use bigger bins and have larger spike counts.

To answer this question quantitatively, we need a measure of distance. The measure we use, denoted Inline graphic, is defined in Eq. (3) below, but all we need to know about it for now is that if Inline graphic then Inline graphic, and if Inline graphic is near one then Inline graphic is far from Inline graphic. In terms of Inline graphic, our main results are as follows. First, for small Inline graphic, in what we call the perturbative regime, Inline graphic is proportional to Inline graphic. In other words, as the population size increases, the pairwise model becomes a worse and worse approximation to the true distribution. Second, this behavior is entirely generic: for small Inline graphic, Inline graphic increases linearly, no matter what the true distribution is. This is illustrated schematically in Fig. 2, which shows the generic behavior of Inline graphic. The solid red part of the curve is the perturbative regime, where Inline graphic is a linearly increasing function of Inline graphic; the dashed curves show possible behavior beyond the perturbative regime.

Figure 2. Cartoon illustrating the dependence of Inline graphic on Inline graphic.

Figure 2

For small Inline graphic there is always a perturbative regime in which Inline graphic increases linearly with Inline graphic (solid red line). When Inline graphic becomes large, Inline graphic may continue increasing with Inline graphic (red and black dashed lines) or it may plateau (cyan dashed line), depending on Inline graphic. The observation that Inline graphic increases linearly with Inline graphic does not, therefore, provide much, if any information about the large Inline graphic behavior.

These results have an important corollary: if one does an experiment and finds that Inline graphic is increasing linearly with Inline graphic, then one has no information at all about the true distribution. The flip side of this is more encouraging: if one can measure the true distribution for sufficiently large Inline graphic that Inline graphic saturates, as for the dashed blue line in Fig. 2, then there is a chance that extrapolation to large Inline graphic is meaningful. The implications for the interpretation of experiments is, therefore, that one can gain information about large Inline graphic behavior only if one can analyze data past the perturbative regime.

Under what conditions is a subsystem in the perturbative regime? The answer turns out to be simple: the size of the system, Inline graphic, times the average probability of observing a spike in a bin, must be small compared to 1. For example, if the average probability is 1/100, then a system will be in the perturbative regime if the number of neurons is small compared to 100. This last observation would seem to be good news: if we divide the spike trains into sufficiently small time bins and ignore temporal correlations, then we can model the data very well with a pairwise distribution. The problem with this, though, is that temporal correlations can be ignored only when time bins are large compared to the autocorrelation time. This leads to a kind of catch-22: pairwise models are guaranteed to work well (in the sense that they describe spike trains in which temporal correlations are ignored) if one uses small time bins, but small time bins is the one regime where ignoring temporal correlations is not a valid approximation.

In the next several sections we quantify the qualitative picture presented above: we write down an explicit expression for Inline graphic, explain why it increases linearly with Inline graphic when Inline graphic is small, and provide additional tests, besides assessing the linearity of Inline graphic, to determine whether or not one is in the perturbative regime.

Quantifying how well the pairwise model explains the data

A natural measure of the distance between Inline graphic and Inline graphic is the Kullback-Leibler (KL) divergence [12], denoted Inline graphic and defined as

graphic file with name pcbi.1000380.e054.jpg (1)

The KL divergence is zero if the two distributions are equal; otherwise it is nonzero.

Although the KL divergence is a very natural measure, it is not easy to interpret (except, of course, when it is exactly zero). That is because a nonzero KL divergence tells us that Inline graphic, but it does not give us any real handle on how good, or bad, the pairwise model really is. To make sense of the KL divergence, we need something to compare it to. A reasonable reference quantity, used by a number of authors [7][9], is the KL divergence between the true distribution and the independent one, the latter denoted Inline graphic. The independent distribution, as its name suggests, is a distribution in which the variables are taken to be independent,

graphic file with name pcbi.1000380.e057.jpg (2)

where Inline graphic is the distribution of the response of the Inline graphic neuron, Inline graphic. With this choice for a comparison, we define a normalized distance measure—a measure of how well the pairwise model explains the data—as

graphic file with name pcbi.1000380.e061.jpg (3)

Note that the denominator in this expression, Inline graphic, is usually referred to as the multi-information [7],[13],[14].

The quantity Inline graphic lies between 0 and 1, and measures how well a pairwise model does relative to an independent model. If it is 0, the pairwise model is equal to the true model (Inline graphic); if it is near 1, the pairwise model offers little improvement over the independent model; and if it is exactly 1, the pairwise model is equal to the independent model (Inline graphic), and so offers no improvement.

How do we attach intuitive meaning to the two divergences Inline graphic and Inline graphic? For the latter, we use the fact that, as is easy to show,

graphic file with name pcbi.1000380.e068.jpg (4)

where Inline graphic and Inline graphic are the entropies [15],[16] of Inline graphic and Inline graphic, respectively, defined, as usual, to be Inline graphic. For the former, we use the definition of the KL divergence to write

graphic file with name pcbi.1000380.e074.jpg (5)

The quantity Inline graphic has the flavor of an entropy, although it is a true entropy only when Inline graphic is maximum entropy as well as pairwise (see Eq. (6) below). For other pairwise distributions, all we need to know is that Inline graphic lies between Inline graphic and Inline graphic. A plot illustrating the relationship between Inline graphic, the two entropies Inline graphic and Inline graphic, and the entropy-like quantity Inline graphic, is shown in Fig. 3.

Figure 3. Schematic plot of Inline graphic (black line), Inline graphic (cyan line) and Inline graphic (red line).

Figure 3

The better the pairwise model, the closer Inline graphic is to Inline graphic. This is reflected in the normalized distance measure, Inline graphic, which is the distance between the red and cyan lines divided by the distance between the red and black lines.

Note that for pairwise maximum entropy models (or maximum entropy models for short), Inline graphic has a particularly simple interpretation, since in this case Inline graphic really is an entropy. Using Inline graphic to denote the pairwise entropy of a maximum entropy model, for this case we have

graphic file with name pcbi.1000380.e093.jpg (6)

as is easy to see by inserting Eqs. (4) and (5) into (3). This expression has been used previously by a number of authors [7],[9].

Inline graphic in the perturbative regime

The extrapolation problem discussed above is the problem of determining Inline graphic in the large Inline graphic limit. This is hard to do in general, but there is a perturbative regime in which it is possible. The small parameter that defines this regime is the average number of spikes produced by the whole population of neurons in each time bin. It is given quantitatively by Inline graphic where Inline graphic is the bin size and Inline graphic the average firing rate,

graphic file with name pcbi.1000380.e100.jpg (7)

with Inline graphic the firing rate of neuron Inline graphic.

The first step in the perturbation expansion is to compute the two quantities that make up Inline graphic: Inline graphic and Inline graphic. As we show in the section “Perturbative Expansion” (Methods), these are given by

graphic file with name pcbi.1000380.e106.jpg (8a)
graphic file with name pcbi.1000380.e107.jpg (8b)

where

graphic file with name pcbi.1000380.e108.jpg (9a)
graphic file with name pcbi.1000380.e109.jpg (9b)

Here and in what follows we use Inline graphic to denote terms that are proportional to Inline graphic in the limit Inline graphic. The Inline graphic in Eq. (9a) has been noted previously [7], although the authors did not compute the prefactor, Inline graphic.

The prefactors Inline graphic and Inline graphic, which are given explicitly in Eqs. (42) and (44), depend on the low order statistics of the spike trains: Inline graphic depends on the second order normalized correlation coefficients, Inline graphic depends on the second and third order normalized correlation coefficients (the normalized correlation coefficients are defined in Eq. (16) below), and both depend on the firing rates of the individual cells. The details of that dependence, however, are not important for now; what is important is that Inline graphic and Inline graphic are independent of Inline graphic and Inline graphic (at least on average; see next section).

Inserting Eq. (8) into Eq. (3) (into the definition of Inline graphic) and using Eq. (9), we arrive at our main result,

graphic file with name pcbi.1000380.e124.jpg (10a)
graphic file with name pcbi.1000380.e125.jpg (10b)

Note that in the regime Inline graphic, Inline graphic is necessarily small. This explains why, in an analytic study of non-pairwise model in which Inline graphic was small, Shlens et al. found that Inline graphic was rarely greater than 0.1 [8].

We refer to quantities with a superscript zero as “zeroth order.” Note that, via Eqs. (4) and (5), we can also define zeroth order entropies,

graphic file with name pcbi.1000380.e130.jpg (11a)
graphic file with name pcbi.1000380.e131.jpg (11b)

These quantities are important primarily because differences between them and the actual entropies indicate a breakdown of the perturbation expansion (see in particular Fig. 4 below).

Figure 4. Cartoon showing extrapolations of the zeroth order KL divergences and entropies (see Eqs. (9) and (11)).

Figure 4

These extrapolations illustrate why the two natural quantities derived from them, Inline graphic and Inline graphic, occur beyond the point at which the extrapolation is meaningful. (A) Extrapolations on a log-log scale. Black: Inline graphic; green: Inline graphic; cyan: Inline graphic. The red points are the data. The points Inline graphic and Inline graphic label the intersections of the two extrapolations with the independent entropy, Inline graphic. (B) Extrapolation of the entropies rather than the KL divergences, plotted on a linear-linear scale. The data, again shown in red, is barely visible in the lower left hand corner. Black: Inline graphic; solid orange: Inline graphic; solid maroon: Inline graphic. The dashed orange and maroon lines are the extrapolations of the true entropy and true pairwise “entropy”, respectively.

Assuming, as discussed in the next section, that Inline graphic and Inline graphic are approximately independent of Inline graphic, Inline graphic, and Inline graphic, Eq. (10) tells us that Inline graphic scales linearly with Inline graphic in the perturbative regime—the regime in which Inline graphic. The key observation about this scaling is that it is independent of the details of the true distribution, Inline graphic. This has a very important consequence, one that has major implications for experimental data: if one does an experiment with small Inline graphic and finds that Inline graphic is proportional to Inline graphic, then the system is, with very high probability, in the perturbative regime, and one does not know whether Inline graphic will remain close to Inline graphic as Inline graphic increases. What this means in practical terms is that if one wants to know whether a particular pairwise model is a good one for large systems, it is necessary to consider values of Inline graphic that are significantly greater than Inline graphic, where

graphic file with name pcbi.1000380.e160.jpg (12)

We interpret Inline graphic as the value at which there is a crossover in the behavior of the pairwise model. Specifically, if Inline graphic, the system is in the perturbative regime and the pairwise model is not informative about the large Inline graphic behavior, whereas if Inline graphic, the system is in a regime in which it may be possible to make inferences about the behavior of the full system.

The prefactors, Inline graphic and Inline graphic

As we show in Methods (see in particular Eqs. (42) and (44)), the prefactors Inline graphic and Inline graphic depend on which neurons out of the full population are used. Consequently, these quantities fluctuate around their true values (in the sense that different subpopulations produce different values of Inline graphic and Inline graphic), where “true” refers to an average over all possible Inline graphic sub-populations. Here we assume that the Inline graphic neurons are chosen randomly from the full population, so any set of Inline graphic neurons provides unbiased estimates of Inline graphic and Inline graphic. In our simulations, the fluctuations were small, as indicated by the small error bars on the blue points in Fig. 5. However, in general the size of the fluctuations is determined by the range of firing rates and correlation coefficients, with larger ranges producing larger fluctuations.

Figure 5. The Inline graphic dependence of the KL divergences and the normalized distance measure, Inline graphic.

Figure 5

Data was generated from a third order model, as explained in the section “Generating synthetic data” (Methods), and fit to pairwise maximum entropy models and independent models. All data points correspond to averages over marginalizations of the true distribution (see text for details). The red points were computed directly using Eqs. (1), (3) and (4); the blue points are the zeroth order estimates, Inline graphic, Inline graphic, and Inline graphic, in rows 1, 2 and 3, respectively. The three columns correspond to Inline graphic, 0.029, and 0.037, from left to right. (A, B, C) (Inline graphic). Predictions from the perturbative expansion are in good agreement with the measurements up to Inline graphic, indicating that the data is in the perturbative regime. (D, E, F) (Inline graphic). Predictions from the perturbative expansion are in good agreement with the measurements up to Inline graphic, indicating that the data is only partially in the perturbative regime. (G, H, I) (Inline graphic). Predictions from the perturbative expansion are not in good agreement with the measurements, even for small Inline graphic, indicating that the data is outside the perturbative regime.

Because Inline graphic does not affect the mean values of Inline graphic and Inline graphic, it is reasonable to think of these quantities—or at least their true values—as being independent of Inline graphic. They are also independent of Inline graphic, again modulo fluctuations. Finally, as we show in the section “Bin size and the correlation coefficients” (Methods), Inline graphic and Inline graphic are independent of Inline graphic in the limit that Inline graphic is small compared to the width of the temporal correlations among neurons. We will assume this limit applies here. In sum, then, to first approximation, Inline graphic and Inline graphic are independent of our three important quantities: Inline graphic, Inline graphic, and Inline graphic. Thus, we treat them as effectively constant throughout our analysis.

The dangers of extrapolation

Although the behavior of Inline graphic in the perturbative regime does not tell us much about its behavior at large Inline graphic, it is possible that other quantities that can be calculated in the perturbative regime, Inline graphic, Inline graphic, and Inline graphic (the last one exactly), are informative, as others have suggested [7]. Here we show that this is not the case—they also are uninformative.

The easiest way to relate the perturbative regime to the large Inline graphic regime is to ignore the corrections in Eqs. (8a) and (8b), extrapolate the expressions for the zeroth order terms, and ask what their large Inline graphic behavior tells us. Generic versions of these extrapolations, plotted on a log-log scale, are shown in Fig. 4A, along with a plot of the independent entropy, Inline graphic (which is necessarily linear in Inline graphic). The first thing we notice about the extrapolations is that they do not, technically, have a large Inline graphic behavior: one terminates at the point labeled Inline graphic, which is where Inline graphic (and thus, via Eq. (0a), Inline graphic; continuing the extrapolation implies negative true zeroth order entropy); the other at the point labeled Inline graphic, which is where Inline graphic (and thus, via Eq. (5) and the fact that Inline graphic, Inline graphic).

Despite the fact that the extrapolations end abruptly, they still might provide information about the large Inline graphic regime. For example, Inline graphic and/or Inline graphic might be values of Inline graphic at which something interesting happens. To see if this is the case, in Fig. 4B we plot the naive extrapolations of Inline graphic and Inline graphic (that is, the zeroth order quantities given in Eq. (11), Inline graphic and Inline graphic), on a linear-linear plot, along with Inline graphic. This plot contains no new information compared to Fig. 4A, but it does elucidate the meaning of the extrapolations. Perhaps its most striking feature is that the naive extrapolation of Inline graphic has a decreasing portion. As is easy to show mathematically, entropy cannot decrease with Inline graphic (intuitively, that is because observing one additional neuron cannot decrease the entropy of previously observed neurons). Thus, Inline graphic, which occurs well beyond the point where the naive extrapolation of Inline graphic is decreasing, has essentially no meaning, something that has been pointed out previously by Bethge and Berens [10]. The other potentially important value of Inline graphic is Inline graphic. This, though, suffers from a similar problem: when Inline graphic, Inline graphic is negative.

How do the naively extrapolated entropies—the solid lines in Fig. 4B—compare to the actual entropies? To answer this, in Fig. 4B we show the true behavior of Inline graphic and Inline graphic versus Inline graphic (dashed lines). Note that Inline graphic is asymptotically linear in Inline graphic, even though the neurons are correlated, a fact that forces Inline graphic to be linear in Inline graphic, as it is sandwiched between Inline graphic and Inline graphic. (The asymptotically linear behavior of Inline graphic is typical, even in highly correlated systems. Although this is not always appreciated, it is easy to show; see the section “The behavior of the true entropy in the large N limit,” Methods.) Comparing the dashed and solid lines, we see that the naively extrapolated and true entropies, and thus the naively extrapolated and true values of Inline graphic, have extremely different behavior. This further suggests that there is very little connection between the perturbative and large Inline graphic regimes.

In fact, these observations follow directly from the fact that Inline graphic and Inline graphic depend only on correlation coefficients up to third order (see Eqs. (42) and (44)) whereas the large Inline graphic behavior depends on correlations at all orders. Thus, since Inline graphic and Inline graphic tell us very little, if anything, about higher order correlations, it is not surprising that they tell us very little about the behavior of Inline graphic in the large Inline graphic limit.

Numerical simulations

To check that our perturbation expansions, Eqs. (8–10), are correct, and to investigate the regime in which they are valid, we performed numerical simulations. We generated, from synthetic data, a set of true distributions, computed the true distance measures, Inline graphic, Inline graphic, and Inline graphic numerically, and compared them to the zeroth order ones, Inline graphic, Inline graphic, and Inline graphic. If the perturbation expansion is valid, then the true values should be close to the zeroth order values whenever Inline graphic is small. The results are shown in Fig. 5, and that is, indeed, what we observed. Before discussing that figure, though, we explain our procedure for constructing true distributions.

The set of true distributions we used were generated from a third order model (so named because it includes up to third order interactions). This model has the form

graphic file with name pcbi.1000380.e262.jpg (13)

where Inline graphic is a normalization constant, chosen to ensure that the probability distribution sums to 1, and the sums over Inline graphic, Inline graphic and Inline graphic run from 1 to Inline graphic. The parameters Inline graphic and Inline graphic were chosen by sampling from distributions (see the section “Generating synthetic data,” Methods), which allowed us to generate many different true distributions. In all of our simulations we calculate the relevant quantities directly from Eq. (13) . Consequently, we do not have to worry about issues of finite data, as would be the case in realistic experiments.

For a particular simulation (corresponding to a column in Fig. 5), we generated a true distribution with Inline graphic, randomly chose 5 neurons, and marginalized over them. This gave us a 10-neuron true distribution. True distributions with Inline graphic were constructed by marginalizing over additional neurons within our 10-neuron population. To achieve a representative sample, we considered all possible marginalizations (of which there are 10 choose Inline graphic, or Inline graphic). The results in Fig. 5 are averages over these marginalizations.

For neural data, the most commonly used pairwise model is the maximum entropy model. Therefore, we use that one here. To emphasize the maximum entropy nature of this model, we replace the label “pair” that we have been using so far with “maxent.” The maximum entropy distribution has the form

graphic file with name pcbi.1000380.e274.jpg (14)

Fitting this distribution requires that we choose the Inline graphic and Inline graphic so that the first and second moments match those of the true distribution. Quantitatively, these conditions are

graphic file with name pcbi.1000380.e277.jpg (15a)
graphic file with name pcbi.1000380.e278.jpg (15b)

where the angle brackets, Inline graphic and Inline graphic, represent averages with respect to Inline graphic and Inline graphic, respectively. Once we have Inline graphic and Inline graphic that satisfy Eq. (15), we calculate the KL divergences, Eqs. (1) and (4), and use those to compute Inline graphic.

The results are shown in Fig. 5. The rows correspond to our three quantities of interest: Inline graphic, Inline graphic, and Inline graphic (top to bottom). The columns correspond to different values of Inline graphic, with the smallest Inline graphic on the left and the largest on the right. Red circles are the true values of these quantities; blue ones are the zeroth order predictions from Eqs. (9) and (10b).

As suggested by our perturbation analysis, the smaller the value of Inline graphic, the larger the value of Inline graphic for which agreement between the true and zeroth order values is good. Our simulations corroborate this: the left column of Fig. 5 has Inline graphic, and agreement is almost perfect out to Inline graphic; the middle column has Inline graphic, and agreement is almost perfect out to Inline graphic; and the right column has Inline graphic, and agreement is not good for any value of Inline graphic. Note that the perturbation expansion breaks down for values of Inline graphic well below Inline graphic (defined in Eq.(12)): in the middle column of Fig. 5 it breaks down when Inline graphic, and in the right column it breaks down when Inline graphic. This is not, however, especially surprising, as the perturbation expansion is guaranteed to be valid only if Inline graphic.

These results validate the perturbation expansions in Eqs. (8) and (10), and show that those expansions provide sensible predictions—at least for some parameters. They also suggest a natural way to assess the significance of one's data: plot Inline graphic, Inline graphic, and Inline graphic versus Inline graphic, and look for agreement with the predictions of the perturbation expansion. If agreement is good, as in the left column of Fig. 5, then one is in the perturbative regime, and one knows very little about the true distribution. If, on the other hand, agreement is bad, as in the right column, then one is out of the perturbative regime, and it may be possible to extract meaningful information about the relationship between the true and pairwise models.

That said, the qualifier “at least for some parameters” is an important one. This is because the perturbation expansion is essentially an expansion that depends on the normalized correlation coefficients, and there is an underlying assumption that they don't exhibit pathological behavior. The Inline graphic order normalized correlation coefficient for the distribution Inline graphic, denoted Inline graphic, is written

graphic file with name pcbi.1000380.e311.jpg (16)

A potentially problematic feature of the correlation coefficients is that the denominator is a product over mean activities. If the mean activities are small, the denominator can become very small, leading to very large correlation coefficients. Although our perturbation expansion is always valid for sufficiently small time bins (because the correlation coefficients eventually becomes independent of bin size; see the section “Bin size and the correlation coeffcients,” Methods), “sufficiently small” can depend in detail on the parameters. For instance, at the maximum population size tested (Inline graphic) and for the true distributions that had Inline graphic, the absolute error of the prediction had a median of approximately 16%. However, about 11% of the runs had errors larger than 60%. Thus, the exact size of the small parameter at which the perturbative expansion breaks down can depend on the details of the true distribution.

External fields and pairwise couplings have a simple dependence on firing rates and correlation coefficients in the perturbative regime

Estimation of the KL divergences and Inline graphic from real data can be hard, in the sense that it takes a large amount of data for them to converge to their true values. In addition, as discussed above, in the section “The prefactors gind and gpair”, there are fluctuations in Inline graphic associated with finite subsampling of the full population of neurons. Those fluctuations tend to keep Inline graphic from being purely linear, as can seen, for example, in the blue points in Fig. 5F and 5I. We therefore provide a second set of relationships that can be used to determine whether or not a particular data set is in the perturbative regime. These relationships are between the parameters of the maximum entropy model, the Inline graphic and Inline graphic, and the mean activity and normalized second order correlation coefficient (the latter defined in Eq. (19) below).

Since the quantity Inline graphic plays a central role in our analysis, we replace it with a single parameter, which we denote Inline graphic,

graphic file with name pcbi.1000380.e321.jpg (17)

In terms of this parameter, we find (using the same perturbative approach that led us to Eqs. (8–10); see the section “External fields, pairwise couplings and moments,” Methods), that

graphic file with name pcbi.1000380.e322.jpg (18a)
graphic file with name pcbi.1000380.e323.jpg (18b)

where Inline graphic, the normalized second order correlation coefficient, is defined in Eq. (16) with Inline graphic; it is given explicitly by

graphic file with name pcbi.1000380.e326.jpg (19)

(We don't need a superscript on Inline graphic or a subscript on the angle brackets because the first and second moments are the same under the true and pairwise distributions.) Equation (18a) can be reconstructed from the low firing rate limit of analysis carried out by Sessak and Monasson [17], as can the first three terms in the expansion of the log in Eq. (18b).

Equation (18) tells us that the Inline graphic of the Inline graphic and Inline graphic, the external fields and pairwise couplings, is very weak. In Fig. 6 we confirm this through numerical simulations. Equation (18b) also provides additional information—it gives us a functional relationship between the pairwise couplings and the normalized pairwise correlations function, Inline graphic. In Fig. 7A–C we plot the pairwise couplings, Inline graphic, versus the normalized pairwise correlation coefficient, Inline graphic (blue dots), along with the prediction from Eq. (18b) (black line). Consistent with our predictions, the data in Fig. 7A–C essentially follows a line—the line given by Eq. (18b).

Figure 6. The true external fields and pairwise interactions compared with the predictions of the perturbation expansion.

Figure 6

The top row shows the true external fields, Inline graphic, versus those predicted from Eq. (18a), and the bottom row shows the true pairwise interaction, Inline graphic, versus those predicted from Eq. (18b). Values of Inline graphic ranging from 5 to 10 are shown, with different colors corresponding to different Inline graphic. For each value of Inline graphic, data is shown for 45 realization of the true distribution. Insets show the Inline graphic of the mean external fields (top) and mean pairwise interactions (bottom). The three columns correspond exactly to the columns in Fig. 5. (A, B) (Inline graphic). There is a very good match between the true and predicted values of both external fields and pairwise interactions. (C, D) (Inline graphic). Even though Inline graphic has increased, the match is still good. (E, F) (Inline graphic). The true and predicted external fields and pairwise interactions do not match as well as the cases shown in (A, B, C, D). There is also now a stronger Inline graphic in the mean external fields compared to (A) and (B). The Inline graphic of the pairwise interactions in (F) is weaker than that of the external fields, but still notably stronger than the ones in (B) and (D). This indicates that the perturbative expansion is starting to break down.

Figure 7. The relation between pairwise couplings and pairwise correlations.

Figure 7

This figure shows that there is a simple relation between Inline graphic and Inline graphic, but not between Inline graphic and Inline graphic. (A, C, E) Inline graphic versus the normalized coefficients, Inline graphic (blue points), along with the predicted relationship, via Eq. (18b) (black line). (B, D, F) Inline graphic versus the Pearson correlation coefficients, Inline graphic, Eq. (26) (blue points). The three columns correspond exactly to the columns in Fig. 5 from left to right; that is, Inline graphic for (A, B), Inline graphic for (C, D), and Inline graphic for (E, F). The prediction in the top row (black line) matches the data well, even in the rightmost column.

A relationship between the pairwise couplings and the correlations coefficients has been sought previously, but for the more standard Pearson correlation coefficient [7],[9],[11]. Our analysis explains why it was not found. The Pearson correlation coefficient, denoted Inline graphic, is given by

graphic file with name pcbi.1000380.e358.jpg (20)

In the small Inline graphic limit—the limit of interest—the right hand side of Eq. (20) is approximately equal to Inline graphic. Because Inline graphic depends on the external fields, Inline graphic and Inline graphic (see Eq. (18a)) and there is a one-to-one relationship between Inline graphic and Inline graphic (Eq. (18b)), there can't be a one-to-one relationship between Inline graphic and Inline graphic. We verify the lack of a relationship in Fig. 7D and 7E, where we again plot Inline graphic, but this time versus the standard correlation coefficient, Inline graphic. As predicted, the data in Fig. 7D and 7E is scattered over two dimensions. This suggests that Inline graphic, not Inline graphic, is the natural measure of the correlation between two neurons when they have a binary representation, something that has also been suggested by Amari based on information-geometric arguments [18].

Note that the lack of a simple relationship between the pairwise couplings and the standard correlation coefficient has been a major motivation in building maximum entropy models [7],[11]. This is for good reason: if there is a simple relationship, knowing the Inline graphic adds essentially nothing. Thus, plotting Inline graphic versus Inline graphic (but not Inline graphic) is an important test of one's data, and if the two quantities fall on the curve predicted by Eq. (18b), the maximum entropy model is adding very little information, if any.

As an aside, we should point out that the Inline graphic is a function of the variables used to represent the firing patterns. Here we use 0 for no spike and 1 for one or more spikes, but another, possibly more common, representation, derived from the Ising model and used in a number of studies [7],[9],[11], is to use −1 and +1 rather than 0 and 1. This amounts to making the change of variables Inline graphic. In terms of Inline graphic, the maximum entropy model has the form Inline graphic where Inline graphic and Inline graphic are given by

graphic file with name pcbi.1000380.e382.jpg (21a)
graphic file with name pcbi.1000380.e383.jpg (21b)

The second term on the right side of Eq. (21a) is proportional to Inline graphic, which means the external fields in the Ising representation acquire a linear Inline graphic that was not present in our 0/1 representation. The two studies that reported the Inline graphic of the external fields [7],[9] used this representation, and, as predicted by our analysis, the external fields in those studies had a component that was linear in Inline graphic.

Is there anything wrong with using small time bins?

An outcome of our perturbative approach is that our normalized distance measure, Inline graphic, is linear in bin size (see Eq. (10b)). This suggests that one could make the pairwise model look better and better simply by making the bin size smaller and smaller. Is there anything wrong with this? The answer is yes, for reasons discussed above (see the the section “The extrapolation problem”); here we emphasize and expand on this issue, as it is an important one for making sense of experimental results.

The problem arises because what we have been calling the “true” distribution is not really the true distribution of spike trains. It is the distribution assuming independent time bins, an assumption that becomes worse and worse as we make the bins smaller and smaller. (We use this potentially confusing nomenclature primarily because all studies of neuronal data carried out so far have assumed temporal independence, and compared the pairwise distribution to the temporally independent—but still correlated across neurons—distribution [7][9],[11]. In addition, the correct name “true under the assumption of temporal independence,” is unwieldy.) Here we quantify how much worse. In particular, we show that if one uses time bins that are small compared to the characteristic correlation time in the spike trains, the pairwise model will not provide a good description of the data. Essentially, we show that, when the time bins are too small, the error one makes in ignoring temporal correlations is larger than the error one makes in ignoring correlations across neurons.

As usual, we divide time into bins of size Inline graphic. However, because we are dropping the independence assumption, we use Inline graphic, rather than Inline graphic, to denote the response in bin Inline graphic. The full probability distribution over all time bins is denoted Inline graphic. Here Inline graphic is the number of bins; it is equal to Inline graphic for spike trains of length Inline graphic. If time bins are approximately independent and the distribution of Inline graphic is the same for all Inline graphic (an assumption we make for convenience only, but do not need; see the section “Extending the normalized distance measure to the time domain,” Methods), we can write

graphic file with name pcbi.1000380.e399.jpg (22)

Furthermore, if the pairwise model is a good one, we have

graphic file with name pcbi.1000380.e400.jpg (23)

Combining Eqs. (22) and Eq. (23) then gives us an especially simple expression for the full probability distribution: Inline graphic.

The problem with small time bins lies in Eq. (22): the right hand side is a good approximation to the true distribution when the time bins are large compared to the spike train correlation time, but it is a bad approximation when the time bins are small (because adjacent time bins become highly correlated). To quantify how bad, we compare the error one makes assuming independence across time to the error one makes assuming independence across neurons. The ratio of those two errors, denoted Inline graphic, is given by

graphic file with name pcbi.1000380.e403.jpg (24)

It is relatively easy to compute Inline graphic in the limit of small time bins (see the section “Extending the normalized distance measure to the time domain,” Methods), and we find that

graphic file with name pcbi.1000380.e405.jpg (25)

As expected, this reduces to our old result, Inline graphic, when there is only one time bin (Inline graphic). When Inline graphic is larger than 1, however, the second term is always at least one, and for small bin size, the third term is much larger than one. Consequently, if we use bins that are small compared to the temporal correlation time of the spike trains, the pairwise model will do a very bad job describing the full, temporally correlated spike trains.

Discussion

Probability distributions over the configurations of biological systems are extremely important quantities. However, because of the large number of interacting elements comprising such systems, these distributions can almost never be determined directly from experimental data. Using parametric models to approximate the true distribution is the only existing alternative. While such models are promising, they are typically applied only to small subsystems, not the full system. This raises the question: are they good models of the full system?

We answered this question for a class of parametric models known as pairwise models. We focused on a particular application, neuronal spike trains, and our main result is as follows: if one were to record spikes from multiple neurons, use sufficiently small time bins and a sufficiently small number of cells, and assume temporal independence, then a pairwise model will almost always succeed in matching the true (but temporally independent) distribution—whether or not it would match the true (but still temporally independent) distribution for large time bins or a large number of cells. In other words, pairwise models in the “sufficiently small” regime, what we refer to as the perturbative regime, have almost no predictive value for what will happen with large populations. This makes extrapolation from small to large systems dangerous.

This observation is important because pairwise models, and in particular pairwise maximum entropy models, have recently attracted a great deal of attention: they have been applied to salamander and guinea pig retinas [7], primate retina [8], primate cortex [9], cultured cortical networks [7], and cat visual cortex [11]. These studies have mainly operated close to the perturbative regime. For example, Schneidman et al. [7] had Inline graphic (for the data set described in their Fig. 5), Tang et al. [9] had Inline graphic to 0.4 (depending on the preparation), and Yu et al. [11] had Inline graphic. For these studies, then, it would be hard to justify extrapolating to large populations.

The study by Shlens et al. [8], on the other hand, might be more amenable to extrapolation. This is because spatially localized visual patterns were used to stimulate retinal ganglion cells, for which a nearest neighbor maximum entropy models provided a good fit to their data. (Nearest neighbor means Inline graphic is zero unless neuron Inline graphic and neuron Inline graphic are adjacent.) Our analysis still applies, but, since all but the nearest neighbor correlations are zero, many of the terms that make up Inline graphic and Inline graphic vanish (see Eqs. (42) and (44)). Consequently, the small parameter in the perturbative expansion becomes Inline graphic (rather than Inline graphic), where Inline graphic is the number of nearest neighbors. Since Inline graphic is fixed, independent of the population size, the small parameter will not change as the population size increases. Thus, Shlens et al.may have tapped into the large population behavior even though they considered only a few cells at a time in their analysis. Indeed, they have recently extended their analysis to more than 100 neurons, and they still find that nearest neighbor maximum entropy models provide very good fits to the data [19].

Time bins and population size

That the pairwise model is always good if Inline graphic is sufficiently small has strong implications: if we want to build a good model for a particular Inline graphic, we can simply choose a bin size that is small compared to Inline graphic. However, one of the assumptions in all pairwise models used on neural data is that bins at different times are independent. This produces a tension between small time bins and temporal independence: small time bins essentially ensure that a pairwise model will provide a close approximation to a model with independent bins, but they make adjacent bins highly correlated. Large time bins come with no such assurance, but they make adjacent bins independent. Unfortunately, this tension is often unresolvable in large populations, in the sense that pairwise models are assured to work only up to populations of size Inline graphic where τ corr is the typical correlation time. Given that Inline graphic is at least several Hz, for experimental paradigms in which the correlation time is more than a few hundred ms, Inline graphic is about one, and this assurance does not apply to even moderately sized populations of neurons.

These observations are especially relevant for studies that use small time bins to model spike trains driven by natural stimuli. This is because the long correlation times inherent in natural stimuli are passed on to the spike trains, so the assumption of independence across time (which is required for the independence assumption to be valid) breaks badly. Knowing that these models are successful in describing spike trains under the independence assumption, then, does not tell us whether they will be successful in describing full, temporally correlated, spike trains. Note that for studies that use stimuli with short correlation times (e.g., non-natural stimuli such as white noise), the temporal correlations in the spike trains are likely to be short, and using small time bins may be perfectly valid.

The only study that has investigated the issue of temporal correlations in maximum entropy models does indeed support the above picture [9]: for the parameters used in that study (Inline graphic to 0.4), the pairwise maximum entropy model provided a good fit to the data (Inline graphic was typically smaller than 0.1), but it did not do a good job modeling the temporal structure of the spike trains.

Other systems—Protein folding

As mentioned in the Introduction, in addition to the studies on neuronal data, studies on protein folding have also emphasized the role of pairwise interactions [2],[3]. Briefly, proteins consist of strings of amino acids, and a major question in structural biology is: what is the probability distribution of amino acid strings in naturally folding proteins? One way to answer this is to approximate the full probability distribution of naturally folding proteins from knowledge of single-site and pairwise distributions. One can show that there is a perturbative regime for proteins as well. This can be readily seen using the celebrated HP protein model [20], where a protein is composed of only two types of amino acids. If, at each site, one amino acid type is preferred and occurs with high probability, say Inline graphic with Inline graphic, then a protein of length shorter than Inline graphic will be in the perturbative regime, and, therefore, a good match between the true distribution and the pairwise distribution for such a protein is virtually guaranteed.

Fortunately, the properties of real proteins generally prevent this from happening: at the majority of sites in a protein, the distribution of amino acids is not sharply peaked around one amino acid. Even for those sites that are sharply peaked (the evolutionarily-conserved sites), the probability of the most likely amino acid, Inline graphic, rarely exceeds 90% [21],[22]. This puts proteins consisting of only a few amino acids out of the perturbative regime, and puts longer proteins—the ones usually studied using pairwise models—well out of it.

This difference is fundamental: because many of the studies that have been carried out on neural data were in the perturbative regime, the conclusions of those studies—specifically, the conclusion that pairwise models provide accurate descriptions of large populations of neurons—is not yet supported. This is not the case for the protein studies, because they are not in the perturbative regime. Thus, the evidence that pairwise models provide accurate descriptions of protein folding remain strong and exceedingly promising.

Open questions

In our analysis, we sidestepped two issues of practical importance: finite sampling and alternative measures for assessing the quality of the pairwise model. These issues are beyond the scope of this paper, but in our view, they are natural next steps in the analysis of pairwise models. Below we briefly expand on them.

Finite sampling refers to the fact that in any real experiment, one has access to only a finite amount of data, and so does not know the true probability distribution of the spike trains. In our analysis, however, we assumed that one did have full knowledge of the true probability distribution. Since a good estimate of the probability distribution is crucial for assessing whether the pairwise model can be extrapolated to large populations, it is important to study how such estimates are affected by finite data. Future work is needed to address this issue, and to find ways to overcome data limitation—for example, by finding efficient methods for removing the finite data bias that affects information theoretic quantities such as the Kullback-Leibler divergence.

There are always many possible ways to assess the quality of a model. Our choice of Inline graphic was motivated by two considerations: it is based on the Kullback-Leibler divergence, which is a standard measure of “distance” between probability distributions, and it is a widely used measure in the field [7][10]. It suffers, however, from a number of shortcomings. In particular, Inline graphic can be small even when the pairwise model assigns very different probabilities to many of the configurations of the system. It would, therefore, be important to study the quality of pairwise models using other measures.

Methods

The behavior of the true entropy in the large Inline graphic limit

To understand how the true entropy behaves in the large Inline graphic limit, it is useful to express the difference of the entropies as a mutual information. Using Inline graphic to denote the true entropy of Inline graphic neurons and Inline graphic to denote the mutual information between one neuron and the other Inline graphic neurons in a population of size Inline graphic, we have

graphic file with name pcbi.1000380.e442.jpg (26)

If knowing the activity of Inline graphic neurons does not fully constrain the firing of neuron Inline graphic, then the single neuron entropy, Inline graphic, will exceed the mutual information, Inline graphic, and the entropy will be an increasing function of Inline graphic. For the entropy to be linear in Inline graphic, all we need is that the mutual information saturates with Inline graphic. Because of synaptic noise, this is a reasonable assumption for networks of neurons: even if we observed all the spikes from all the neurons, there would still be residual noise associated with synaptic failures, jitter in release time, variability in the amount of neurotransmitter released, stochastic channel dynamics, etc. Consequently, in the large Inline graphic limit, we may replace Inline graphic by its average, denoted Inline graphic. Also replacing Inline graphic by its average, denoted Inline graphic, we see that for large Inline graphic, the difference between Inline graphic and Inline graphic approaches a constant. Specifically,

graphic file with name pcbi.1000380.e458.jpg (27)

where this expression is valid in the large Inline graphic limit and the corrections are sublinear in Inline graphic.

Perturbative expansion

Our main quantitative result, given in Eqs. (8–10), is that the KL divergence between the true distribution and both the independent and pairwise distributions can be computed perturbatively as an expansion in powers of Inline graphic in the limit Inline graphic. Here we carry out this expansion, and derive explicit expressions for the quantities Inline graphic and Inline graphic.

To simplify our notation, here we use Inline graphic for the true distribution. The critical step in computing the KL divergences perturbatively is to use the Sarmanov-Lancaster expansion [23][28] for Inline graphic,

graphic file with name pcbi.1000380.e467.jpg (28)

where

graphic file with name pcbi.1000380.e468.jpg (29a)
graphic file with name pcbi.1000380.e469.jpg (29b)
graphic file with name pcbi.1000380.e470.jpg (29c)
graphic file with name pcbi.1000380.e471.jpg (29d)

This expansion has a number of important, but not immediately obvious, properties. First, as can be shown by a direct calculation,

graphic file with name pcbi.1000380.e472.jpg (30)

where the subscripts Inline graphic and Inline graphic indicate an average with respect to Inline graphic and Inline graphic, respectively. This has an immediate corollary,

graphic file with name pcbi.1000380.e477.jpg

This last relationship is important, because it tells us that if a product of Inline graphic contains any terms linear in one of the Inline graphic, the whole product averages to zero under the independent distribution. This can be used to show that

graphic file with name pcbi.1000380.e480.jpg (31)

from which it follows that

graphic file with name pcbi.1000380.e481.jpg

Thus, Inline graphic is properly normalized. Finally, a slightly more involved calculations provides us with a relationship between the parameters of the model and the moments: for Inline graphic,

graphic file with name pcbi.1000380.e484.jpg (32a)
graphic file with name pcbi.1000380.e485.jpg (32b)

Virtually identical expressions hold for higher order moments. It is this last set of relationships that make the Sarmanov-Lancaster expansion so useful.

Note that Eqs. (32a) and (32b), along with the expression for the normalized correlation coefficients given in Eq. (16), imply that

graphic file with name pcbi.1000380.e486.jpg (33a)
graphic file with name pcbi.1000380.e487.jpg (33b)

These identities will be extremely useful for simplifying expressions later on.

Because the moments are so closely related to the parameters of the distribution, moment matching is especially convenient: to construct a distribution whose moments match those of Inline graphic up to some order, one simply needs to ensure that the parameters of that distribution, Inline graphic, Inline graphic, Inline graphic, etc., are identical to those of the true distributions up to the order of interest. In particular, let us write down a new distribution, Inline graphic,

graphic file with name pcbi.1000380.e493.jpg (34a)
graphic file with name pcbi.1000380.e494.jpg (34b)

We can recover the independent distribution by letting Inline graphic, and we can recover the pairwise distribution by letting Inline graphic. This allows us to compute Inline graphic in the general case, and then either set Inline graphic to zero or set Inline graphic to Inline graphic.

Expressions analogous to those in Eqs. (31–33) exist for averages with respect to Inline graphic; the only difference is that Inline graphic is replaced by Inline graphic.

The KL divergence in the Sarmanov-Lancaster representation

Using Eqs. (28) and (34a) and a small amount of algebra, the KL divergence between Inline graphic and Inline graphic may be written

graphic file with name pcbi.1000380.e506.jpg (35)

where

graphic file with name pcbi.1000380.e507.jpg (36)

To derive Eq. (35), we used the fact that Inline graphic (see Eq. (31)). The extra term Inline graphic was included to ensure that Inline graphic and its first derivatives vanish at Inline graphic, something that greatly simplifies our analysis later on.

Our approach is to Taylor expand the right hand side of Eq. (35) around Inline graphic, compute each term, and then sum the whole series (we do not assume that either Inline graphic or Inline graphic is small). Using Inline graphic to denote the coefficients of the Taylor series, we have

graphic file with name pcbi.1000380.e516.jpg (37)

Note that because Inline graphic and its first derivatives vanish at Inline graphic, all terms in this sum have Inline graphic.

Because both Inline graphic and Inline graphic are themselves sums, an exact calculation of the terms in Eq. (37) would be difficult. However, as we show below, in the section “Averages of powers of ξp and ξq” (see in particular Eqs. (52) and (54)), they can be computed as perturbation expansions in powers of Inline graphic, and the result is

graphic file with name pcbi.1000380.e523.jpg (38)

where Inline graphic and Inline graphic are given by

graphic file with name pcbi.1000380.e526.jpg (39)

Inline graphic. The last equality in Eq. (39) follows from a small amount of algebra and the definition of the correlation coefficients given in Eq. (16). Equation (38) is valid only when Inline graphic, which is the case of interest to us (since the Taylor expansion of Inline graphic has only terms with Inline graphic).

The important point about Eq. (38) is that the Inline graphic and Inline graphic dependence follows that of the original Taylor expansion. Thus, when we insert this equation back into Eq. (37), we recover our original function—all we have to do is interchange the sums. For example, consider inserting the first term in Eq. (38) into Eq. (37),

graphic file with name pcbi.1000380.e533.jpg

Performing the same set of manipulations on all of Eq. (38) leads to

graphic file with name pcbi.1000380.e534.jpg (40)

This expression is true in general (except for some technical considerations; see the section “Averages of powers of ξp and ξq”); to restrict it to the KL divergences of interest we set Inline graphic to Inline graphic and Inline graphic to either Inline graphic or Inline graphic. In the first case (Inline graphic set to Inline graphic), Inline graphic, which implies that Inline graphic, and thus Inline graphic. Because Inline graphic has a quadratic minimum at Inline graphic, when Inline graphic, the second two terms on the right hand side of Eq. (40) are Inline graphic. We thus have, to lowest nonvanishing order in Inline graphic,

graphic file with name pcbi.1000380.e550.jpg (41)

with the Inline graphic correction coming from the last sum in Eq. (40). Defining

graphic file with name pcbi.1000380.e552.jpg (42)

where, recall Inline graphic, and inserting Eq. (42) into Eq. (41), we recover Eq. (8a).

In the second case (Inline graphic set to Inline graphic), the first and second moments of Inline graphic and Inline graphic are equal. This implies, using Eq. (32), that Inline graphic, and thus Inline graphic. Because Inline graphic (see Eq. (36)), the first three terms on the right hand side of Eq. (40)—those involving Inline graphic and Inline graphic but not Inline graphic—vanish. The next order term does not vanish, and yields

graphic file with name pcbi.1000380.e564.jpg (43)

Defining

graphic file with name pcbi.1000380.e565.jpg (44)

and inserting this expression into Eq. (43), we recover Eq. (8b).

External fields, pairwise couplings and moments

In this section we derive, to leading order in Inline graphic, expressions relating the external fields and pairwise couplings of the maximum entropy model, Inline graphic and Inline graphic, to the first and second moments; these are the expressions reported in Eq. (18). The calculation proceeds along the same lines as in the previous section. There is, though, one extra step associated with the fact that the quadratic term in the maximum entropy distribution given in Eq. (14) is proportional to Inline graphic, not Inline graphic. However, to lowest order in Inline graphic, this doesn't matter. That's because

graphic file with name pcbi.1000380.e572.jpg

where Inline graphic is defined as in Eq. (29d) except with Inline graphic replaced by Inline graphic, and we used the fact that Inline graphic. The second term introduces a correction to the external fields, Inline graphic. However, the correction is Inline graphic, so we drop it. We should keep in mind, though, that our final expression for Inline graphic will have corrections of this order.

Using Eq. (14), but with Inline graphic replaced by Inline graphic where it appears with Inline graphic, we may write (after a small amount of algebra)

graphic file with name pcbi.1000380.e583.jpg (45)

where Inline graphic is the same as the function Inline graphic defined in Eq. (29a) except that Inline graphic is replaced by Inline graphic, the subscript “ind” indicates, as usual, an average with respect to Inline graphic, and the two functions Inline graphic and Inline graphic are defined by

graphic file with name pcbi.1000380.e591.jpg (46)

and

graphic file with name pcbi.1000380.e592.jpg (47)

Given this setup, we can use Eqs. (55) and (56) below to compute the moments under the maximum entropy model. That's because both Inline graphic and its first derivative vanish at Inline graphic, which are the two conditions required for these equations to be valid. Using also the fact that Inline graphic, Eqs. (55) and (56) imply that

graphic file with name pcbi.1000380.e596.jpg (48a)
graphic file with name pcbi.1000380.e597.jpg (48b)
graphic file with name pcbi.1000380.e598.jpg (48c)

where the first term in Eq. (48b) came from Eq. (29d) with Inline graphic replaced by Inline graphic, the term “Inline graphic” in Eq. (48c) came from Inline graphic, and for the second two expressions we used the fact that, to lowest order in Inline graphic, the denominator in Eq. (45) is equal to 1.

Finally, using Eq. (19) for the normalized correlation coefficient, dropping the subscript “maxent” (since the first and second moments are the same under the maxent and true distributions), and inverting Eqs. (48b) and (48c) to express the external fields and coupling coefficients in terms of the first and second moments, we arrive at Eq (18).

Averages of powers of Inline graphic and Inline graphic

Here we compute Inline graphic, which, as can be seen in Eq. (37), is the key quantity in our perturbation expansion. Our starting point is to (formally) expand the sums that make up Inline graphic and Inline graphic (see Eqs. (29b) and (34b)), which yields

graphic file with name pcbi.1000380.e609.jpg (49)

The sum over Inline graphic is a sum over all possible configurations of the Inline graphic. The coefficient Inline graphic are complicated functions of the Inline graphic, etc. Computing these functions is straightforward, although somewhat tedious, especially when Inline graphic is large; below we compute them only for Inline graphic and 3. The reason Inline graphic starts at 2 is that Inline graphic; see Eq. (37).

We first show that all terms with superscript Inline graphic are Inline graphic. To do this, we note that, because the right hand side of Eq. (49) is an average with respect to the independent distribution, the average of the product is the product of the averages,

graphic file with name pcbi.1000380.e620.jpg (50)

Then, using the fact that Inline graphic with probability Inline graphic and Inline graphic with probability Inline graphic (see Eq. (29c)), we have

graphic file with name pcbi.1000380.e625.jpg (51)

The significance of this expression is that, for Inline graphic, Inline graphic, independent of Inline graphic. Consequently, if all the Inline graphic in Eq. (50) are greater than 1, then the right hand side is Inline graphic. This shows that, as promised above, the superscript Inline graphic labels the order of the terms.

As we saw in the section “The KL divergence in the Sarmanov-Lancaster representation”, we need to go to third order in Inline graphic, which means we need to compute the terms on the right hand side of Eq. (49) with Inline graphic and 3. Let us start with Inline graphic, which picks out only those terms with two unique indices. Examining the expressions for Inline graphic and Inline graphic given in Eqs. (29b) and (34b), we see that we must keep only terms involving Inline graphic, since Inline graphic has three indices, and higher order terms have more. Thus, the Inline graphic contribution to the average in Eq. (49), which we denote Inline graphic, is given by

graphic file with name pcbi.1000380.e641.jpg

Pulling Inline graphic and Inline graphic out of the averages, using Eq. (33a) to eliminate Inline graphic and Inline graphic in favor of Inline graphic and Inline graphic, and applying Eq. (51) (while throwing away some of the terms in that equation that are higher than second order in Inline graphic), the above expression may be written

graphic file with name pcbi.1000380.e649.jpg (52)

Note that we were not quite consistent in our ordering with respect to Inline graphic, in the sense that we kept some higher order terms and not others. We did this so that we could use Inline graphic rather than Inline graphic, as the former is directly observable.

For Inline graphic the calculation is more involved, but not substantially so. Including terms with exactly three unique indices in the sum on the right hand side of Eq. (49) gives us

graphic file with name pcbi.1000380.e654.jpg (53)

This expression is not quite correct, since some of the terms contain only two unique indices—these are the terms proportional to Inline graphic—whereas it should contain only terms with exactly three unique indices. Fortunately, this turns out not to matter, for reasons we discuss at the end of the section.

To perform the averages in Eq. (53), we would need to use multinomial expansions, and then average over the resulting powers of Inline graphic. For the latter, we can work to lowest order in Inline graphic, which means we only take the first term in Eq. (51). This amounts to replacing every Inline graphic with Inline graphic (and similarly for Inline graphic and Inline graphic), and in addition multiplying the whole expression by an overall factor of Inline graphic. For example, if Inline graphic and Inline graphic, one of the terms in the multinomial expansion is Inline graphic. This average would yield, using Eq. (51) and considering only the lowest order term, Inline graphic.

This procedure also is not quite correct, since terms with only one factor of Inline graphic, which average to zero, are replaced with Inline graphic. This also turns out not to matter; again, we discuss why at the end of the section.

We can, then, go ahead and use the above “replace blindly” algorithm. Note that the factors of Inline graphic, Inline graphic and Inline graphic turn Inline graphic and Inline graphic into normalized correlation coefficients (see Eq. (33)), which considerably simplifies our equations. Using also Eq. (39) for Inline graphic, Eq. (53) becomes

graphic file with name pcbi.1000380.e675.jpg (54)

We can now combine Eqs. (52) and (54), and insert them into Eq. (49). This gives us the first two terms in the perturbative expansion of Inline graphic; the result is written down in Eq. (38) above.

Why can we ignore the overcounting associated with terms in which an index appears exactly zero or one times? We clearly can't do this in general, because for such terms, replacing Inline graphic with Inline graphic fails—either because the terms didn't exist in the first place (when one of the indices never appeared) or because they averaged to zero (when an index appeared exactly once). In our case, however, such terms do not appear in the Taylor expansion. To see why, note first of all that, because of the form of Inline graphic, its Taylor expansion can be written Inline graphic where Inline graphic is finite at Inline graphic (see Eq. (36)). Consequently, the original Taylor expansion of Inline graphic, Eq. (37), should contain a factor of Inline graphic; i.e.,

graphic file with name pcbi.1000380.e685.jpg

where the Inline graphic are the coefficients of the Taylor expansion of Inline graphic. The factor Inline graphic, when expanded, has the form

graphic file with name pcbi.1000380.e689.jpg

As we saw in the previous section, we are interested in the third order term only to compute Inline graphic, for which Inline graphic. Therefore, the above multiplicative factor reduces to Inline graphic. It is that last factor of Inline graphic that is important, since it guarantees that for every term in the Taylor expansion, all indices appear at least twice. Therefore, although Eq. (53) is not true in general, it is valid for our analysis.

We end this section by pointing out that there is a very simple procedure for computing averages to second order in Inline graphic. Consider a function Inline graphic that has a minimum at Inline graphic and also obeys Inline graphic. Then, based on the above analysis, we have

graphic file with name pcbi.1000380.e698.jpg (55)

Two easy corollaries of this are: for Inline graphic and Inline graphic positive integers,

graphic file with name pcbi.1000380.e701.jpg (56a)
graphic file with name pcbi.1000380.e702.jpg (56b)

where the sum in Eq. (56a) runs over Inline graphic only, and we used the fact that both Inline graphic and Inline graphic are symmetric with respect to the interchange of Inline graphic and Inline graphic.

Generating synthetic data

As can be seen in Eq. (13), the synthetic data depends on three sets of parameters: Inline graphic, and Inline graphic. Here we describe how they were generated.

To generate the Inline graphic, we draw a set of firing rates, Inline graphic, from an exponential distribution with mean 0.02 (recall that Inline graphic, which we set to 15, is the number of neurons in our base distribution). From this we chose the external field according to Eq. (18a),

graphic file with name pcbi.1000380.e713.jpg

In the perturbative regime, a distribution generated with these values of the external fields has firing rates approximately equal to the Inline graphic; see Eq. (18a) and Fig. 6.

To generate the Inline graphic and Inline graphic, we drew them from Gaussian distributions with means equal to 0.05 and 0.02 and standard deviations of 0.8 and 0.5, respectively. Using non-zero values for Inline graphic means that the true distribution is not pairwise.

Bin size and the correlation coefficients

One of our main claims is that Inline graphic is linear in bin size, Inline graphic. This is true, however, only if Inline graphic and Inline graphic are independent of Inline graphic, as can be seen from Eq. (10b). In this section we show that independence is satisfied if Inline graphic is smaller than the typical correlation time of the responses. For Inline graphic larger than such correlation times, Inline graphic and Inline graphic do depend on Inline graphic, and Inline graphic is no longer linear in Inline graphic. Note, though, that the correlation time is always finite, so there will always be a bin size below which the linear relationship, Inline graphic, is guaranteed.

Examining Eqs. (42) and (44), we see that Inline graphic and Inline graphic depend on the normalized correlation coefficients, Inline graphic and Inline graphic (we drop superscripts, since our discussion will be generic). Thus, to understand how Inline graphic and Inline graphic depend on bin size, we need to understand how the normalized correlation coefficients depend on bin size. To do that, we express them in terms of standard cross-correlograms, as the cross-correlograms contain, in a very natural way, information about the temporal timescales in the spike train.

We start with the second order correlation coefficient, since it is simplest. The corresponding cross-correlogram, which we denote Inline graphic, is given by

graphic file with name pcbi.1000380.e738.jpg (57)

where Inline graphic is the time of the k th spike on neuron Inline graphic (and similarly for Inline graphic), and Inline graphic is the Dirac Inline graphic. The normalization in Eq. (57) is slightly non-standard—more typical is to divide by something with units of firing rate (Inline graphic, Inline graphic or Inline graphic), to give units of spikes/s. The normalization we use is convenient, however, because in the limit of large Inline graphic, Inline graphic approaches one.

It is slightly tedious, but otherwise straightforward, to show that when Inline graphic is sufficiently small that only one spike can occur in a time bin, Inline graphic is related to Inline graphic via

graphic file with name pcbi.1000380.e752.jpg (58)

The (unimportant) factor Inline graphic comes from the fact that the spikes occur at random locations within a bin.

Equation (58) has a simple interpretation: Inline graphic is the average height of the central peak of the cross-correlogram relative to baseline. How strongly Inline graphic depends on Inline graphic is thus determined by the shape of the cross-correlogram. If it is smooth, then Inline graphic approaches a constant as Inline graphic becomes small. If, on the other hand, there is a sharp peak at Inline graphic, then Inline graphic for small Inline graphic, so long as Inline graphic is larger than the width of the peak. (The factor of Inline graphic included in the scaling is approximate; it is a placeholder for an effective firing rate that depends on the indices Inline graphic and Inline graphic. It is, however, sufficiently accurate for our purposes.) A similar relationship exists between the third order correlogram and the correlation coefficient. Thus, Inline graphic is also independent of Inline graphic in the small Inline graphic limit, whereas if the central peak is sharp it scales as Inline graphic.

The upshot of this analysis is that the shape of the cross-correlogram has a profound effect on the correlation coefficients and, therefore, on Inline graphic. What is the shape in real networks? The answer typically depends on the physical distance between cells. If two neurons are close, they are likely to receive common input and thus exhibit a narrow central peak in their cross-correlogram. Just how narrow depends on the area. Early in the sensory pathways, such as retina [29][31] and LGN [32], peaks can be very narrow—on the order of milliseconds. Deeper into cortex, however, peaks tend to broaden, to at least tens of milliseconds [33],[34]. If, on the other hand, the neurons are far apart, they are less likely to receive common input. In this case, the correlations come from external stimuli, so the central peak tends to have a characteristic width given by the temporal correlation time of the stimulus, typically 100 s of milliseconds.

Although clearly both kinds of cross-correlograms exist in any single population of neurons, it is convenient to analyze them separately. We have already considered networks in which the cross-correlograms were broad and perfectly flat, so that the correlation coefficients were strictly independent of bin size. We can also consider the opposite extreme: networks in which the the cross-correlograms (both second and higher order) among nearby neurons exhibit sharp peaks while those among distant neurons are uniformly equal to 1. In this regime, the correlation coefficients depend on Inline graphic: as discussed above, the second order ones scale as Inline graphic and the third as Inline graphic. This means that the arguments of Inline graphic and Inline graphic are large (see Eqs. (42) and (44)). From the definition of Inline graphic in Eq. (36), in this regime both are approximately linear in their arguments (ignoring log corrections). Consequently, Inline graphic and Inline graphic. This implies that Inline graphic and Inline graphic scale as Inline graphic and Inline graphic, respectively, and so Inline graphic, independent of Inline graphic. Thus, if the bin size is large compared to the correlation time, Inline graphic will be approximately independent of bin size.

Extending the normalized distance measure to the time domain

In this section we derive the expression for Inline graphic given in Eq. (25). Our starting point is its definition, Eq. (24). It is convenient to define Inline graphic to be a concatenation of the responses in Inline graphic time bins,

graphic file with name pcbi.1000380.e789.jpg (59)

where, as in the section “Is there anything wrong with using small time bins?”, the superscript labels time, so Inline graphic is the full, temporally correlated, distribution.

With this definition, we may write the numerator in Eq. (24) as

graphic file with name pcbi.1000380.e791.jpg (60)

where Inline graphic is the entropy of Inline graphic, the last sum follows from a marginalization over all but one element of Inline graphic, and Inline graphic is the true distribution at time Inline graphic (unlike in the section “Is there anything wrong with using small time bins?”, here we do not assume that the true distribution is the same in all time bins). Note that Inline graphic is independent of time, since it is computed from a time average of the true distribution. That time average, which we call Inline graphic, is given in terms of Inline graphic as

graphic file with name pcbi.1000380.e800.jpg

Inserting this definition into Eq. (60) eliminates the sum over Inline graphic, and replaces it with Inline graphic. For simplicity we consider the maximum entropy pairwise model. In this case, because Inline graphic is in the exponential family, and the first and second moments are the same under the true and maximum entropy distributions, we can replace Inline graphic with Inline graphic. Consequently, Eq. (60) becomes

graphic file with name pcbi.1000380.e806.jpg

This gives us the numerator in the expression for Inline graphic (Eq. (24)); using Eq. (4) to write Inline graphic, the full expression for Inline graphic becomes

graphic file with name pcbi.1000380.e810.jpg (61)

where we added and subtracted Inline graphic to the numerator.

The first term on the right hand side of Eq. (61) we recognize, from Eq. (6), as Inline graphic. To cast the second into a reasonable form, we define Inline graphic to be the entropy of the distribution that retains the temporal correlations within each neuron but is independent across neurons. Then, adding and subtracting this quantity to the numerator in Eq. (61), and also adding and subtracting Inline graphic, we have

graphic file with name pcbi.1000380.e815.jpg (62)

The key observation is that if Inline graphic, then

graphic file with name pcbi.1000380.e817.jpg

Comparing this with Eqs. (8a) and (9a), we see that Inline graphic is a factor of Inline graphic times larger than Inline graphic. We thus have

graphic file with name pcbi.1000380.e821.jpg (63)

Again assuming Inline graphic, and defining Inline graphic Inline graphic, the last term in this expression may be written

graphic file with name pcbi.1000380.e825.jpg (64)

Inserting this into Eq. (63) and using Eqs. (4), (8a) and (9a) yields Eq. (25).

We have assumed here that Inline graphic; what happens when Inline graphic, or larger? To answer this, we rewrite Eq. (61) as

graphic file with name pcbi.1000380.e828.jpg (65)

We argue that in general, as Inline graphic increases, Inline graphic becomes increasingly different from Inline graphic, since the former was derived under the assumption that the responses at different time bins were independent. Thus, Eq. (25) should be considered a lower bound on Inline graphic.

Footnotes

The authors have declared that no competing interests exist.

YR and PEL were supported by the Gatsby Charitable Foundation (http://www.gatsby.org.uk) and by the US National Institute of Mental Health grant R01 MH62447. SN was supported by the US National Eye Institute grant R01 EY12978. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

References

  • 1.Rieke F, Warland D, de Ruyter van Steveninck R, Bialek W. Spikes: exploring the neural code. Cambridge, MA: MIT Press; 1997. [Google Scholar]
  • 2.Russ W, Lowery D, Mishra P, Yaffe M, Ranganathan R. Natural-like function in artificial WW domains. Nature. 2005;437:579–583. doi: 10.1038/nature03990. [DOI] [PubMed] [Google Scholar]
  • 3.Socolich M, Lockless S, Russ W, Lee H, Gardner K, et al. Evolutionary information for specifying a protein fold. Nature. 2005;437:512–518. doi: 10.1038/nature03991. [DOI] [PubMed] [Google Scholar]
  • 4.Oates J. Food distribution and foraging behavior. In: Smuts B, Cheney D, Seyfarth R, Wrangham R, Struhsaker T, editors. Primate societies. Chicago: University of Chicago Press; 1987. pp. 197–209. [Google Scholar]
  • 5.Wrangham R. Evolution of social structure. In: Smuts B, Cheney D, Seyfarth R, Wrangham R, Struhsaker T, editors. Primate societies. Chicago: University of Chicago Press; 1987. pp. 282–298. [Google Scholar]
  • 6.Eisenberg J, Muckenhirn N, Rundran R. The relation between ecology a social structure in primates. Science. 1972;176:863–874. doi: 10.1126/science.176.4037.863. [DOI] [PubMed] [Google Scholar]
  • 7.Schneidman E, Berry M, Segev R, Bialek W. Weak pairwise correlations imply strongly correlated network states in a neural population. Nature. 2006;440:1007–1012. doi: 10.1038/nature04701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Shlens J, Field G, Gauthier J, Grivich M, Petrusca D, et al. The structure of multi-neuron firing patterns in primate retina. J Neurosci. 2006;26:8254–8266. doi: 10.1523/JNEUROSCI.1282-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Tang A, Jackson D, Hobbs J, Chen W, Smith J, et al. A maximum entropy model applied to spatial and temporal correlations from cortical networks in vitro. J Neurosci. 2008;28:505–518. doi: 10.1523/JNEUROSCI.3359-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Bethge M, Berens P. Near-maximum entropy models for binary neural representations of natural images. In: Platt J, Koller D, Singer Y, Roweis S, editors. Advances in Neural Information Processing Systems 20. Cambridge, MA: MIT Press; 2008. pp. 97–104. [Google Scholar]
  • 11.Yu S, Huang D, Singer W, Nikolic D. A small world of neuronal synchrony. Cereb Cortex. 2008;18:2891–2901. doi: 10.1093/cercor/bhn047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Kullback S, Leibler R. On information and sufficiency. Ann Math Stat. 1951;22:79–86. [Google Scholar]
  • 13.Friedman N, Mosenzon O, Slonim N, Tishby N. Proc. of Uncertainty in Artificial Intelligence (UAI-17) San Mateo, CA: Morgan Kaufmann Publishers; 2001. Multivariate information bottleneck. pp. 152–161. [Google Scholar]
  • 14.Slonim N, Friedman N, Tishby N. Multivariate information bottleneck. Neural Comput. 2006;18:1739–1789. doi: 10.1162/neco.2006.18.8.1739. [DOI] [PubMed] [Google Scholar]
  • 15.Shannon C, Weaver W. The mathematical theory of communication. Urbana, Illinois: University of Illinois Press; 1949. [Google Scholar]
  • 16.Cover T, Thomas J. Elements of information theory. New York, NY: John Wiley & Sons; 1991. [Google Scholar]
  • 17.Sessak V, Monasson R. Small-correlation expansions for the inverse ising problem. J Phys A. 2009;42:055001. [Google Scholar]
  • 18.Amari S. Measure of correlation orthogonal to changing in firing rate. Neural Comput. 2009;21:960–972. doi: 10.1162/neco.2008.03-08-729. [DOI] [PubMed] [Google Scholar]
  • 19.Shlens J, Field G, Gauthier J, Greschner M, Sher A, et al. Spatial organization of large-scale concerted activity in the primate retina. J Neurosci. In Press 2009 [Google Scholar]
  • 20.Dill K. Theory for the folding and stability of globular proteins. Biochemistry. 1985;24:1501–1509. doi: 10.1021/bi00327a032. [DOI] [PubMed] [Google Scholar]
  • 21.Lockless S, Ranganathan R. Evolutionarily conserved pathways of energetic connectivity in protein families. Science. 1999;286:295–299. doi: 10.1126/science.286.5438.295. [DOI] [PubMed] [Google Scholar]
  • 22.Vargas-Madrazo E, Lara-Ochoa F, Jiménez-Montaño M. A skewed distribution of amino acids at recognition sites of the hypervariable region of immunoglobulins. J Mol Evol. 1994;38:100–104. doi: 10.1007/BF00175497. [DOI] [PubMed] [Google Scholar]
  • 23.Sarmanov O. Maximum correlation coeffcient (nonsymmetric case). Selected Translations in Mathematical Statistics and Probability. Amer. Math. Soc. Volume 2. 1962. pp. 207–210.
  • 24.Sarmanov O. Maximum correlation coefficient (nonsymmetric case). Selected Translations in Mathematical Statistics and Probability. Amer. Math. Soc. Volume 4. 1963. pp. 271–275.
  • 25.Lancaster H. The structure of bivariate distributions. Ann Math Stat. 1958;29:719–736. [Google Scholar]
  • 26.Lancaster H. Correlation and complete dependence of random variables. Ann Math Stat. 1963;34:1315–1321. [Google Scholar]
  • 27.Bahadur R. A representation of the joint distribution of responses to n dichotomous items. In: Solomon H, editor. Studies in Item Analysis and Prediction. Stanford University Press; 1961. pp. 158–168. [Google Scholar]
  • 28.Johnson D, Goodman I. Inferring the capacity of the vector Poisson channel with a Bernoulli model. Network. 2008;19:13–33. doi: 10.1080/09548980701656798. [DOI] [PubMed] [Google Scholar]
  • 29.Mastronarde D. Correlated firing of cat retinal ganglion cells. I. spontaneously active inputs to X- and Y-cells. J Neurophysiol. 1983;49:303–324. doi: 10.1152/jn.1983.49.2.303. [DOI] [PubMed] [Google Scholar]
  • 30.DeVries S. Correlated firing in rabbit retinal ganglion cell. J Neurophysiol. 1999;81:908–920. doi: 10.1152/jn.1999.81.2.908. [DOI] [PubMed] [Google Scholar]
  • 31.Nirenberg S, Carcieri S, Jacobs A, Latham P. Retinal ganglion cells act largely as independent encoders. Nature. 2001;411:698–701. doi: 10.1038/35079612. [DOI] [PubMed] [Google Scholar]
  • 32.Dan Y, Alonso J, Usrey W, Reid R. Coding of visual information by precisely correlated spikes in the lateral geniculate nucleus. Nat Neurosci. 1998;1:501–507. doi: 10.1038/2217. [DOI] [PubMed] [Google Scholar]
  • 33.Ts'o D, Gilbert C, Wiesel T. Relationships between horizontal interactions and functional architecture in cat striate cortex as revealed by cross-correlation analysis. J Neurosci. 1986;6:1160–1170. doi: 10.1523/JNEUROSCI.06-04-01160.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Nelson J, Salin P, Munk M, Arzi M, Bullier J. Spatial and temporal coherence in cortico-cortical connections: a cross-correlation study in areas 17 and 18 in the cat. Vis Neurosci. 1992;9:21–37. doi: 10.1017/s0952523800006349. [DOI] [PubMed] [Google Scholar]

Articles from PLoS Computational Biology are provided here courtesy of PLOS

RESOURCES