The structure and components of our neural systems affect many func...
One of the main goals of this paper is to identify characteristics ...
The body can determine what it needs and through selective gene exp...
The study monitors synapse density over time in an area responsible...
To get an idea of the precision of the data, the team used a method...
We know that flipping a coin will either give us heads, or tails, a...
The computer algorithm and network has many parallels with the natu...
In Figure B: For the pruning algorithm, we start with a network tha...
How the pruning algorithm works: Pruning decisions were made locall...
Here, we are exploring why experimentally observed decreasing pruni...
Rate Comparisons: Increasing: Starts out slow, by eliminating a...
The more edges existing, the more energy is consumed. An optimal ne...
2s-patch (even)- A specific group of source nodes is connected to a...
The Erdős–Rényi (ER) random graph model (G(n,M) model) creates all ...
As q represents the likelihood of getting an S T edge, it may be i...
Can evolution really help construct the brain in such a way that th...
Scientists theorize that the success of decreasing rate pruning is ...
Much of the study of neurobiology is heavily focused on the regulat...
Decreasing pruning rates are also found in the brain, consistent wi...
The EPTA staining method allows for improved visualization of synap...
SVMs are classifiers- they are machine learning models that are tra...
Theoretically, this allows the system to remove edges that can make...
You may ask, why don't we just link the source node to the target n...
RESEARCH ARTICLE
Decreasing-Rate Pruning Optimizes the
Construction of Efficient and Robust
Distributed Networks
Saket Navlakha
1
, Alison L. Barth
2
*, Ziv Bar-Joseph
3
*
1 Center for Integrative Biology, The Salk Institute for Biological Studies, La Jolla, California, United States of
America, 2 Department of Biological Sciences, Center for the Neural Basis of Cognition, Carnegie Mellon
University, Pittsburgh, Pennsylvania, United States of America, 3 Lane Center for Computational Biology,
Machine Learning Department, Carnegie Mellon University, Pittsburgh, Pennsylvania, United States of
America,
* barth@cmu.edu (ALB); zivbj@cs.cmu.edu (ZBJ)
Abstract
Robust, efficient, and low-cost networks are advantageous in both biological and engi-
neered systems. During neural network development in the brain, synapses are massively
over-produced and then pruned-back over time. This strategy is not commonly used when
designing engineered networks, since adding connections that will soon be removed is con-
sidered wasteful. Here, we show th at for large distributed routing networks, network function
is markedly enhanced by hyper-connectivity followed by aggressive pruning and that the
global rate of pruning, a developmental parameter not previously studied by experimental-
ists, plays a crit ical role in optimizing network structure. We first used high-throughput
image analysis techniques to quantify the rate of pruning in the mammalian neocortex
across a broad developmental time window and found that the rate is decreasing over time.
Based on these results, we analyzed a model of computational routing networks and show
using both theoretical analysis and simulations that decreasing rates lead to more robust
and efficient networks compared to other rates. We also present an applicatio n of this strat-
egy to improve the distributed design of airline networks. Thus, inspiration from neural net-
work formation suggests effective ways to design distributed networks across several
domains.
Author Summary
During development of neural circuits in the brain, synapses are massively over-produced
and then pruned-back over time. This is a fundamental process that occurs in many brain
regions and organisms, yet, despite decades of study of this process, the rate of synapse
elimination, and how such rates affect the function and structure of networks, has not
been studied. We performed large-scale brain imaging experiments to quantify synapse
elimination rates in the developing mouse cortex and found that the rate is decreasing
over time (i.e. aggressive elimination occurs early, followed by a longer phase of slow
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 1 / 23
a11111
OPEN ACCESS
Citation: Navlakha S, Barth AL, Bar-Joseph Z (2015)
Decreasing-Rate Pruning Optimizes the Construction
of Efficient and Robust Distributed Networks. PLoS
Comput Biol 11(7): e1004347. doi:10.1371/journal.
pcbi.1004347
Editor: Lyle J. Graham, Université Paris Descartes,
Centre National de la Recherche Scientifique,
FRANCE
Received: December 12, 2014
Accepted: May 20, 2015
Published: July 28, 2015
Copyright: © 2015 Navlakha et al. This is an open
access article distributed under the terms of the
Creative Commons Attribution License, which permits
unrestricted use, distribution, and reproduction in any
medium, provided the original author and source are
credited.
Data Availability Statement: All images generated
to study developmental pruning in the mouse
somatosensory cortex are available in an online
database at: http://www.snl.salk.edu/~navlakha/
pruning_data/
Funding: This work was supported in part by the
National Institutes of Health award no. F32-
MH099784 to SN; by grants from the National
Institutes of Health award no. 0171-088 and the
McKnight Foundation to ALB; and by grants from the
McDonnell Foundation programme on Studying
Complex Systems and from the US National Science
elimination). We show that such rates optimize the efficiency and robustness of distrib-
uted routing networks under several models. We also present an application of this strat-
egy to improve the design of airline networks.
Introduction
Neural networks in the brain are formed during development using a pruning process that
includes expansive growth of synapses followed by activity-dependent elimination. In humans,
synaptic density peaks around age 2 and subsequently declines by 5060% in adulthood [14].
It has been hypothesized that synaptic pruning is important for experience-dependent selection
of the most appropriat e subset of connections [1, 5], and it occurs in many brain regions and
species [69]. This strategy substantially reduces the amount of genetic information required
to code for the trillions of connections made in the human brain [10]. Instead of instructing
precise connections, more general rules can be applied, which are then fine-tuned by activity-
dependent selection. Although the molecular and cellular mechanisms driving activity-depen-
dent pruning have been extensively investigated [1, 3, 4], global aspects of this highly-distrib-
uted process, including the rate at which synapses are pruned, the impact of these rates on
network function, and the contrast of pruning-versus growth-based strategies commonly used
in engineering to construct networks, has not been studied.
While the specific computations performed within neural and engineered networks may be
very different, at a broad level, both types of networks share many goals and constraints [11].
First, networks must propagate signals efficiently while also being robust to malfunctions (e.g.
spike propagation failures in neural networks [1214]; computer or link failures in communi-
cation networks [15]). Second, both types of networks must adapt connections based on pat-
terns of input activity [16]. Third, these factors must be optimized under the constraint of
distributed processing (without a centralized coordinator) [17, 18], and using low-cost solu-
tions that conserve important metabolic or physical resources (e.g. number of synapses or wir-
ing length in biological networks; energy consumption or battery-life in engineered networks)
[1921]. For example, on the Internet or power grid, requests can be highly dynamic and vari-
able over many time-scales and can lead to network congestion and failures if networks are
unable to adapt to such conditions [22, 23]. In wireless or mobile networks, broadcast ranges
(which determine network topology) need to be inferred in real-time based on the physical dis-
tribution of devices in order to optimize energy efficiency [24]. Although optimizing network
design is critical for such engineered systems across a wide range of applications, existing algo-
rithms used for this problem are not, to our knowledge, based on experience-based pruning, in
part because adding connections that will soon be eliminated is considered wasteful.
Here, we develop a computational approach informed by experimental data to show that
pruning-inspired algorithms can enhance the design of distrib uted routing networks. First, we
experimentally examined developmental pruning rates in the mouse somatosensory cortex, a
well-characterized anatomical structure in the mouse brain [25]. Using electron microscopy
imaging across 41 animals and 16 developmental time-points, coupled with unbiased and
high-throughput image analysis [26], we counted over 20,000 synapses and determined that
pruning rates are decreasing over time (i.e. early, rapid synapse elimination is followed by a
period of slower, gradual elimination). Next, to translate these observations to the computa-
tional domain, we developed a simulated environment for comparing algorithms for distrib-
uted network construction. We find that over-connection followed by pruning leads to
significant improvements in efficiency (routing distance in the network) and robustness
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 2 / 23
Foundation award nos. DBI-0965316 and DBI-
1356505 to ZBJ. The funders had no role in study
design, data collection and analysis, decision to
publish, or preparation of the manuscript.
Competing Interests: The authors have declared
that no competing interests exist.
(number of alternative routes between two nodes) compared to commonly-used methods that
add connections to initially-sparse networks. To determine if these results hold more generally,
we analyzed the theor etical basis of network construction by pruning and found that dec reas-
ing rates led to networks with near-optimal connectivity compared to other rates (increasing,
constant, etc.), which we also confirmed using simulations. Finally, we adapted a pruning-
based strategy to improve the design of airline networks using real traffic pattern data.
The novelty of our approach is two-fold. First, while synaptic pruning has been studied for
decades, previous analyses have determined that synaptic density peaks during early develop-
ment and is reduced by late adolescence and adulthood [69]. However, fine-scale measure-
ments to statistically establish the rate of synapse elimination have not been made. Second,
while substantial prior work linking neural and computational networks has focused on the
computation performed by neural networks [27, 28], our work focuses on the construction of
networks and provides a quantitative platform to compare different network construction pro-
cesses based on their cost, efficiency, and robustness. Our goals here are to model pruning
from an abstract, graph-theoretic perspective; we do not intend to capture all the requirements
of information processing in the brain, and instead focus on using pruning-inspired algorithms
for improving routing in distributed networks. Overall, our results suggest that computational
thinking can simultaneously lead to novel, testable biological hypotheses and new distributed
computing algorithms for designing better networks.
Results
Neural networks employ decreasing rates of synapse elimination
Many generative models have been propos ed to understand how networks evolve and develop
over time (e.g. preferential attachment [29], small-world models [30], duplication-divergence
[31, 32]), yet most of these models assume that the numb er of nodes and edges strictly grows
over time. Synaptic pruning, however, diverges from this strategy. To better understand how
pruning is implemented and whether it can be used to construct networks for broad routing
problems, we sought to measure this process experimentally. Although pruning is a well-estab-
lished neurodevelopmental phenomenon, previous experimental studies have primarily
focused on identifying the time period over which pruning begins and ends but have largely
ignored the dynamics in between these end-points [6, 9, 33], lacking crucial pruning rate infor-
mation that may be useful for using pruning-based strategies for building distributed networks.
To determine the rate of synapse loss in developing neural networks, we focused on a well-
characterized region of the neocortex, layer 4 of somatosensory cortex representing the D1
whisker (Fig 1A), where both thalamic inputs and recurrent circuitry are established in the first
two postnatal weeks [3436]. Because this region of primary sensory cortex does not receive
significant input from other cortical layers [37], measurements of synaptic pruning reflect the
maturation of an extant network, uncontaminated by the addition of synapses over the analysis
window. In addition, the somatotopic anatomy of the whisker (barrel) cortex insured that com-
parisons across different animals and time-points could be made for the identical small cortical
region (Fig 1B).
Changes in synaptic density over time were obtained from sampling 41 animals over 16
developmental time-points ranging from postnatal day 14 (P14) to P40 (Table 1). Over 20,000
synapses in nearly 10,000 images were identified using a synapse-enhancing reaction that spe-
cifically highlights synaptic contacts for electron microscopy [ 38, 39], coupled with unbiased
machine learning algorithms (Fig 1C; Materials and Methods)[26]. Consistent with prior esti-
mates that sampled only the peak and the end-point [9, 33], peak synaptic density occurred at
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 3 / 23
Fig 1. Experimental pipeline, image processing, and decreasing pruning rates. (A) Schematic of the somatotopic mapping of whiskers to neocortical
columns in the mouse somatosensory cortex. (B) Tangential sections from flattened brains preserve the structure of the barrel-field, enabling easy
identification and isolation of the D1 barrel in tissue from different ages. (C) A support vector machine (SVM) classifier is trained using manually labeled
examples of synapses and non-synapses in electron microscopy images. (D) Example images of synapses at three different time points corresponding to
peak synapse density (P19), and later drop-off (P24 and P32). Scale bar represents 500nm. (E) Developmental pruning rate (raw data, left; binned data,
right). Red lines show spline interpolations of the data points. Insets show that the majority of synapses are pruned during the first half of the developmental
pruning period.
doi:10.1371/journal.pcbi.1004347.g001
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 4 / 23
P19 and density declined steeply to mature levels three weeks later (Fig 1D and 1E). Synapse
density at P40 was similar to adult mice sampled at P65 (S4 Fig).
Pruning rates were decreasing over time, i.e. rapid elimination was followed by a slower
period of pruning. To determine the significance of this observation and to test whether only a
single sample or time-point was driving the rate, we used a leave-one-out cross-validation
strategy (Materials and Methods). First, the pruning period was divided into either 2 or 5
equally-spaced intervals over time from P19 to P40. Second, for each fold in the cross-valida-
tion, either one sample was left-out or one time-point was left-out. Third, a spline interpolation
curve was fit and was used to compute the percentage of synapses pruned across successive
intervals. When dividing the period into 2 intervals (P19P29, n = 18 animals and P29P39,
n = 18 animals), there was a significant decrease in the percentage of synapses pruned within
the first interval compared to the second interval (average percentage of synapses pruned from
P19 to P29: 39.99%; (standard deviation over cross-validation folds: 2.93); average percentage
of synapses pruned from P29 to P39: 10.87% (standard deviation: 4.56); P < 0.01, unpaired
2-sample t-test; Fig 1E). When dividing into 5 intervals, we also found a significant decrease in
percentage of synapses pruned within the first interval versus the second (27% versus 15%;
P < 0.01 unpaired 2-sample t-test) and similar decreases across the next two intervals (Fig 2).
The slight rise in pruning in the last interval (7%) may be due to the addition of layer-4-inner-
vating afferents from other brain areas [40] (indeed, we see a small rise in synapse density at
P33, followed by additional pruning; S6 Fig). Nonetheless, the majority of the pruning still
occurs during the first two intervals compared to the last three (P < 0.01), which is quantita-
tively indicative of a decreasing rate.
To further assess the reproducibility of these results, synapse density was adjusted for 3D
analysis, which also confirmed a decreasing rate of synapse elimination (S5 Fig). These data
indicate that neural networks are modified by aggressive pruning of connections, followed by a
later, slow phase of synaptic elimination.
Table 1. Overview of experiments and synapses detected.
Post-natal day (P) # Animals # Images # Synapses (Avg/Image) Precision Recall
14 2 269 500 (1.86) 89.00 50.00
17 3 400 964 (2.41) 92.40 50.00
19 3 739 2437 (3.30) 88.37 49.50
21 3 736 1984 (2.70) 96.57 49.70
22 2 397 1121 (2.83) 80.80 50.00
23 4 767 1913 (2.50) 92.75 49.88
24 1 136 418 (3.07) 96.50 49.50
26 3 523 1176 (2.25) 91.93 49.83
28 2 915 1779 (1.95) 95.70 49.45
30 3 642 1186 (1.85) 95.40 49.83
32 4 1480 2192 (1.48) 96.00 49.70
33 1 256 642 (2.51) 98.00 50.00
34 2 508 1235 (2.43) 88.80 49.50
36 3 763 1611 (2.11) 90.83 49.73
38 3 734 1397 (1.91) 92.63 49.80
40 2 489 803 (1.64) 94.10 49.65
16 41 9754 21355 92.38 49.76
doi:10.1371/journal.pcbi.1004347.t001
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 5 / 23
Pruning outperforms growing algorithms for constructing distributed
networks
Theoretical and practical approaches to engineered network construction typically begin by
constructing a basic, backbone network (e.g. a spanning-tree) and then adding connections
over time as needed [17]. Such a process is considered cost efficient since it does not introduce
new edges unless they are determined to improve routing efficiency or robustness. To quantita-
tively compare the differences between pruning and growing algorithm s, we formulated the
following optimization problem: Given n nodes and an online stream of source-target pairs of
nodes drawn from an a priori unknown distribution D (Fig 3A), design an efficient and robust
network with respect to D (Materials and Methods). Efficiency is measured in terms of the
average shortest-path routing distance between source-target pairs, and robustness is measure d
in terms of number of alternative source-target paths (Materials and Methods).
The distribution D represents an input-output signaling structure that the network needs to
learn during the training (developmental) phase of network construction. This situation occurs
in many computational scenarios. For example, wireless and sensor networks often rely on
information from the environment, which may be structured but unknown beforehand (e.g.
when monitoring river contamination or volcanic activity, some sensors may first detect
changes in the environment based on their physical location and then pass this information to
other downstream nodes for processing) [24]. Similarly, in peer-to-peer systems on the Inter-
net, some machines preferentially route information to other machines [41], and traffic pat-
terns may be unknown beforehand and only discovered in real-time. In the brain, such a
distribution may mimic the directional flow of information across two regions or populations
of neurons.
After training, the goal is to output an unweighted, directed graph with a fixed number of
edges B, representing a limit on available physical or metabolic resources. To evaluate the final
Fig 2. Finer analysis of decreasing synaptic pruning rates. The pruning period was divided into 5 intervals and the percentage of synapses pruned
across successive intervals is depicted by the red bars. Statistics were computed using a leave-out-one strategy on either individual samples from the raw
data (A) or on entire time-points using the binned data (B), where samples from a 2-day window were merged into the same time-point. Error bars indicate
standard deviation over the cross-validation folds. All successive points are significantly different (P < 0.01, two-sample t-test).
doi:10.1371/journal.pcbi.1004347.g002
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 6 / 23
Fig 3. Computational network model and comparison between pruning and growing. (A) Example distribution (2-patch) for source-target pairs. (B) The
pruning algorithm starts with an exuberant number of connections. Edges commonly used to route source-target messages are retained, whereas low-use
edges are iteratively pruned. (C) The growing algorithm begins with a spanning-tree and adds local shortcut edges along common source-target routes. (D)
The no-learning algorithm chooses random edges and does not attempt to learn connections based on the training data. (E+F) Learned networks were
evaluated by computing efficiency (E, the average shortest-path distance amongst test pairs) and robustness (F, the average number of short alternative
paths between a test source and target). Error bars indicate standard deviation over 3 simulation runs.
doi:10.1371/journal.pcbi.1004347.g003
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 7 / 23
network (test phase), additional pairs are drawn from the same distribution D, and efficiency
and robustness of the source-target routes is computed using the test pairs.
Importantly, decisions about edge maintenance, growth, or loss were local and distributed
(no central coordinator). The pruning algorithm begins with a dense network and tracks how
many times each edge is used along a source-target path. In other words, each edge locally
keeps track of how many times it has been used along a source-to-target path. Edges used
many times are by definition important (according to D); edges with low usage values are then
iteratively eliminated modeling a use it or lose it strategy [42, 43](Fig 3B). Initially, we
assumed elimination occurs at a constant rate, i.e. a constant percentage of existing edges are
removed in each interval (Materials and Methods). The growing algorithm first constructs a
spanning-tree on n nodes and iteratively adds local edges to shortcut common routes [44](Fig
3C). These algorithms were compared to a fixed global network (no-learning) that selects B
random directed edges (Fig 3D).
Simulations and analysis of final network structure revealed a marked difference in network
efficiency (lower values are better) and robustness (higher values are better) between the prun-
ing, growing, and no-learning algorithms. In sparsely connected netw orks (average of 2 con-
nections per node), pruning led to a 4.5-fold improvement in efficiency compared to growing
and 1.8-fold improvement compared to no-learning (Fig 3E; S8 Fig). In more densely con-
nected networks (average of 1020 connections per node), pruning still exhibited a significant
improvement in efficiency (S7 Fig). The no-learning algorithm does not tailor connectivity to
D and thus wastes 25% of edges connecting targets back to sources, which does not enhance
efficiency under the 2-patch distribution (Fig 3A). Remarkably, pruning-based networks
enhanced fault tolerance by more than 20-fold compared to growing-based networks, which
were particularly fragile due to strong reliance on the backbone spanning tree (Fig 3F).
Simulations confirm advantages of decreasing pruning rates
The pruning algorithm employed in the previous simulations used a constant rate of connec-
tion loss. Given our experimental results of decreasing pruning rates in neural networks, we
asked whether such rates could indeed lead to more efficient and robust networks in our simu-
lated environment. To address this question, the effects of three pruning rates (increasing,
decreasing, and constant) on network function were compare d (Materials and Methods).
Increasing rates start by eliminating few connections and then removing connections more
aggressively in later intervals. This is an intuitively appealing strategy since the network can
delay edge elimination decisions until more training data is collected. Decreasing rates initially
prune aggressively and then taper off over time, which forces earlier decision-making but pro-
vides more time for network stabilization.
Simulations show that the biologically-motivated decreasing rates indeed improve upon the
constant rate used previously and created the most efficient and robust networks (Fig 4A 4C).
In particular, for the sparsest networks, decreasing rates were 30% more efficient than increas-
ing rates (20% more efficient than constant rates) and exhibited similar gains in fault tolerance.
This was particularly surprising because efficiency and robustness are often optimized using
competing topological structures: e.g. while alternative paths enable fault tolerance, they do not
necessarily enhance efficiency. Further, fewer source-target pairs were unroutable (discon-
nected from each other) using decreasing rates than any other rate (Fig 4B ), which means that
these networks were overall better adapted to the activity patterns defined by the distribution
D. Performance of pruning algorithms was also qualitatively similar when starting with sparser
initial topologies, as opposed to cliques (S9 Fig).
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 8 / 23
Interestingly, decreasing rates also consume the least energy compared to the other rates in
terms of total number of edges maintained during the developmental period (S10 Fig), which
further supports their practical usage.
An alternative biologically-inspired model for building networks
Neurons likely cannot route signals via shortest paths in networks. To explore a more biologi-
cally plausible, yet still abstract, process for network construction, we developed a network-
flow-based model that performs a breadth-first search from the source node, which requires no
global shortest path computation (Materials and Met hods ).
Using this model, we see the identical ordering of performance amongst the three rates,
with decreasing rates leading to the most efficient and robust networks, followed by constant
and then increasing (Fig 5). While our original goal was not to model the full complexity of
neural circuits (e.g. using leaky integrat e-and-fire units, multiple cell types, etc.), this analysis
shows the generality of our biological findings and relevance of pruning rates on network
construction.
Comparing algorithms using additional source-target distributions
The previous results compared each network construction algorithm using the 2-patch distri-
bution (Fig 3A). This distribut ion is unidirectional with equal probability of sampling any
node within the source and target sets, respectively. Next, we compared each network design
algorithm using four additional input distributions. For the 2s-patch distribution (Fig 6A),
with probability x, a random source and target pair is drawn, but with probability 1x, a ran-
dom pair is drawn from amongst a smaller more active set of sources and targets. This distribu-
tion models recent evidence suggesting highly active subnetworks in the cortex with potentially
specialized sources and targets [45, 46]. We set x = 0.5 and the size of the selective sets to be
10% each. For the 2-patch-unbalanced distribution (Fig 6B), there are three times as many tar-
gets as sources, inspired by the fact that different layers have different numbers of neurons
[47]. For the 4-patch distribution (Fig 6C), there are two disjoint sets of sources and targets,
each putatively representing input-output activity from adjacent columns or layers. For the
4-patch Hubel-Wiesel distribution (Fig 6D), the second set of sources are shut-off and never
Fig 4. Simulation results for network optimization. (A) Efficiency (lower is better), (B) the number of unroutable pairs (disconnected source-target test
pairs), and (C) robustness (higher is better) using the 2-patch distribution. For the growing algorithm, there are no unroutable pairs due to the initial spanning
tree construction, which ensures connectivity between every pair to begin with.
doi:10.1371/journal.pcbi.1004347.g004
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 9 / 23
drawn from and their corresponding targets are recruited by the first set of sources, mimicking
monocular deprivation [16].
Overall, decreasing rates produced the most efficient and robust networks across all distri-
butions. which further supports the generality of our model and experimental observations.
Analysis of network motifs
To test if our model can replicate statistics of non-random circuits, we detected network motifs
within the final network generated using decreasing-rate pruning. We counted all possible
3-node motifs and compared these counts to those expected in a random network [48]. Inter-
estingly, when using the 2-patch distribution, where sources and targets are drawn uniformly
from the two sets, we found no over-represented motifs. However, when we considered the 2s-
patch distribution (where a subset of sources and targets are selectively more active than the
others, as one might expect in real cortical circuits [45, 46, 49, 50]), we found feed-forward
motifs to be statistically over-represented when compared to random networks (P < 0.01, Z-
score = 2.82). This motif has been widely observed in many biological and computational net-
works and is known for its role in signal propagation and noise control [48].
Fig 5. Similar advantages of decreasing pruning rates are observed when using a network-flow-based model of network activity. A) 2-patch input
distribution and B) 4-patch input distribution.
doi:10.1371/journal.pcbi.1004347.g005
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 10 / 23
Fig 6. Additional source-target distributions. Decreasing rates are more efficient and robust than all other rates and algorithms for all four distributions: A)
2s-patch. B) 2-patch-unbalanced. C) 4-patch. D) 4-patch Hubel & Wiesel distribution, where during development one input source is lost entirely.
doi:10.1371/journal.pcbi.1004347.g006
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 11 / 23
Theoretical basis of optimal pruning rates
Given a small, initial sampling of the training source-target pairs, it is relatively easy to deter-
mine many connections that will likely not be important. Decreasing rates eliminate these con-
nections quickly, and then provide longer time for the network to fine-tune itself and
accommodate indirect pathways while eliminating fewer connections. On the other hand,
increasing rates can gather more inform ation early, but then are forced to drastically alter net-
work topology towards the final pruning intervals, which can sever pathways and fragment the
network. Interestingly, if the network construction process were guided by a centralized coordi-
nator, then pruning only in the last interval would clearly be a superior strategy because the
longer the coordinator waits, the more data is available to determine which edges are most
important to inform the centralized design process. However, the distributed nature of the
optimization problem forces a different strategy. Indeed, we found more network fragmenta-
tion (unroutable pairs) between sources and targets using increasing rates versus decreasing
(Fig 4B).
To capture these intuitive notions more formally, we theoretically analyzed the effect of
pruning rates on network efficiency. Analysis was simplified in the following way: (1) we only
considered efficiency (routing distance) as the optimization target [51]; (2) we assumed the
2-patch routing distribution used for simulation (Fig 3A); and (3) we approximated the topol-
ogy of the final network using three-parameter Erdős-Rényi random graphs. In these graphs,
directed edges between sources S ! S or targets T ! T exist independently with probability p,
edges from S ! T exist with probability q, and edges T ! S existed with probability z (S1 Text,
S11A Fig; z = 0 in optimal sparse networks).
We derived a recurrence to predict the final p/q ratio given a pruning rate and analytically
related the final p/q ratio to efficiency, the expected path length between source-target pairs (S1
Text, S11B and S11C Fig). Decreasing rates led to networks with near-optimal p/q ratios, result-
ing in the best efficiency compared to other rates. Increasing rates yield larger values of q
(direct source-target edges) because these edges initially represent the shortest routing path for
source-target pairs observed during training when the network is very dense. However, these
exact pairs are unlikely to be seen again during testing, which leads to over-fitted networks.
From both simulations and theoretical analysis, we found that the regime where decreasing
rates are better than increasing rates lies mostly in sparse networks; i.e. where there are O(kn)
edges, where k is a small constant. For example, with n = 1000 nodes, we find k in the range of
26 to show the most significant differences between rates. This level of sparsity is in line with
many real-world geometric networks [52].
Real-world application to improve airline routing using pruning
algorithms
To demonstrate the utility of decreasing-rate pruning on real-world data, we used it to con-
struct airline routing networks using real traffic data denoting the frequency of passenger travel
between US cities. Here, nodes are cities and directed edges imply a direct flight from one city
to another (Fig 7A). Due to budgetary constraints, only a subset of routes can be offered based
on traffic demands from passengers. We collected data from the Department of Transportation
detailing how many passengers flew between the top 1000 source and target city pairs in the
United States (e.g. San Francisco to Los Angeles) during the 3rd quarter of 2013 [53]. These
frequencies were converted into a distribution (D) denoting the probability of travel between
two cities. For this data, a source can also be a target and vice-versa. There were 122 nodes (cit-
ies) in the graph. Training and evaluation was done as before.
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 12 / 23
Decreasing-rate pruning once again outperformed constant and increasing rates, enhancing
efficiency by 510% with similar robustness when using the same number of edges (Fig 7B and
7C). In other words, these networks can reduce travel time for passengersespecially when
travel to some cities is shut down by emergenciesand can reduce overall load of air traffic
control systems. While in practice airline routing networks can be designed in a centralized
and offline manner, we used this example to show in principle how our technique could work,
using real data.
Discussion
Motivated by new experimental data, we showed that decreasing pruning rates lead to more
efficient and robust routing networks compared to other pruning rate strategies and growth-
based methods, when learning is distribut ed and online. While pruning is initially resource-
intensive, early hyper-connectivity facilitates rapid convergence to the most important subset
of connections. Our experimental and theoretical results may appear counter-intuitive since
decreasing rates eliminate more connections early and thus cannot utilize information received
later, compared to increasing rates. However, similar to many large-scale engineered systems,
the brain is built distributedly, with many concurrent processes that do not have access to a sin-
gle global planner [54]. Increasing rates prune aggressively at the end, and such last-minute
drastic changes in topology leave the network fragmented. Decreasing rates provide the best of
both worlds in this regard. They retain extensively used connections and provide more time for
the network to fine-tune pathways by making only relatively minor topological modifications
in later pruning intervals. Moreover, decreasing rates require the least overall energy to imple-
ment because most edges are pruned early in development. This confers an additional practical
advantage to their usage. Our results applied to networks designed using both a shortest -path
model and a flow-based model.
Simultaneously enhancing both efficiency and robustness, a result achieved by decreasing
rates, is not a trivial task. A network in which each node is only connected to a single super-
hub can be used to route every source-target pair using at most 2 hops; however, if the primary
hub fails, the graph will be entirely disconnected, leading to a fragile network. On the other
hand, random networks will have many paths between two nodes, but these paths are not
Fig 7. Improving airline efficiency and robustness using pruning algorithms. (A) Actual data of travel frequency amongst 122 popular cities from the 3rd
quarter of 2013 was used to define a source-target distribution. (B-C) Efficiency (travel time in terms of number of hops) and robustness (number of
alternative routes with the same number of hops) comparison using different algorithms. Decreasing-rate pruning produced more efficient networks with
similar robustness.
doi:10.1371/journal.pcbi.1004347.g007
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 13 / 23
efficient for specific source-target distributions. The fact that decreasing rates outperformed
the other rates for both measures attests to its overall power. Given the importance of dynamic
construction of distributed networks, for example in wireless computing [55], decreasing-rate
pruning may be a viable alternative to current network design methods. These results may be
further improved by optimizing the actual rate of decrease in which we prune edges. Further, a
more rigorous analysis of the regimes where decreasing rates outperform other rates, including
their affect on network robustness, is left for future work.
Prior work studying synapse elimination have primarily focused on the molecular mecha-
nisms controlling this process, including the genes, proteins, and signaling pathways involved
[1, 2], and the role of microglia [4]. Quantitative measurements of synaptic density over devel-
opment have been made in several species, including human (frontal cortex [6], prefrontal cor-
tex [56], visual cortex [57], striate cortex [58]) and mouse (DLGN [59, 60], neuro-muscular
junction [61], barrel cortex [9, 33]), amongst others [7, 8]. However, unlike our study that
focused on determining pruning rates, the primary goal of these studies was to demonstrate
that pruning exists in these areas and to identify the time-period over which it occurs. In some
of these studies, decreasing pruning rates can be inferred [56, 58], which further strengthens
our findings. However, given their focus as mentioned above, no attempt is made in these prior
studies to determine the statistical significance of the observed decreasing rate, and these rates
were not linked to network-level information processing (routing), which is our primary con-
tribution. Prior computational modeling of synaptic pruning has used Hopfield networks as an
optimization model [5]; while this work also does not analyze pruning rates, our results may
shed light on the robustness of memory recall and storage under such a model. Finally, Goyal
et al. [62] used expression levels of known synaptic markers to study synapse elimination in
human; such expression patterns can potentially also be used to model co-occurring rates of
synapse growth and energy consumption (e.g. ATP) during development. There may also be
additional pruning parameters important to extract and analyze, such as pruning differences
amongst different cell types, the addition of afferents from other brain areas at delayed time-
points, and the involvement of glia in synaptic pruning.
Our experimental analysis of pruning rates in the neocortex shows that rates are decreasing
over time. This finding has important biological implications for how networks mature during
development or reorganize during learning. Given similar levels of activity over the network
construction period, these results suggest that the threshold for activation of signaling path-
ways that initiate synaptic weakening or loss should increase over time. Previous experimental
data provides some support to this view, indicating that nascent connections are particularly
vulnerable to synaptic depression [63] or elimination [64]. Decreasing pruning rates are also
consistent with the developmental time-course of myelination, which shows sharp sigmoidal
growth soon after pruning begins [65, 66]. By pruning aggressively early, myelin is not unduly
wasted on axons that may ultimately be lost. Clinically, many neurological disorders show
abnormal pruning levels during critical development periodseither too many synapses (Frag-
ile X syndrome [6769 ]) or too few synapses (Rett syndrome [7072])and these phenotypes
may also affect network function. While our exper imental analysis allowed us to coarsely quan-
tify pruning rates, further challenges remain in longitudinal analysis of synaptic changes within
a single animal and automatic synapse detection from large volumes of tissue. Both advances
can enable temporally-finer analyses, which can be used to establish more precise pruning
rates. Further, any continuous pruning rate that eventua lly stabilizes will have a time bin over
which the rate decreases; our data showed that this decrease persists over multiple days, th ough
finer analyses may be warranted to uncover more precise elimination rates.
Our main goal in this paper was to explore whether a pruning process that mimics how neu-
ral networks are formed can be used to construct efficient and robust computational
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 14 / 23
communication networks. To this end, our model abstracted away many other information
processing goals of neural networks, including synchronization and transformation of input
signals. In addition, we do not model many properties of neural circuits, including connection
weights, coincident activation of multiple neurons, spike-timing dependent plasticity, and the
presence of inhibitory transmission. Our intention in this study was to highlight the potential
importance of pruning rates on global circuit function and to show how this unusual strategy
can be applied in various computing applications. Further study will be required to experimen-
tally perturb pruning rates in vivo to understand how they affect neural function and behavior.
Our approach of abstracting broad-scale, algorithmic principles from neural networks is likely
to provide further insights into the construction of engineered networks and further exemplifies
how bi-directional studies can benefit both biology and computer science [11, 7375].
Materials and Methods
Ethics statement
All experiments were carried out in accordance with NIH Guidelines for animal care and use,
and were approved by Carnegie Mellons institutional animal care and use committee (IACUC
protocol AS13-37).
Electron microscopy imaging and image processing
To experimentally quantify the rate of pruning, we focused on layer 4 of the somatosensory
cortex. We extracted, fixed, and sectioned 50μm-thick tissue from wildtype C57b l6 (Harlan)
mice at different ages. A mitochondrial stain (cytochrome oxidase) was used to visualize the
barrelfield, and the D1 barrel was extracted using a dissecting light microscope.
To enable unbiased and high-throughput classification of synapses, we leveraged a staining
technique that uses ethanolic phosphotungstic acid (EPTA) to pronounce electron opacity at
synaptic sites by targeting proteins in contact zones [38, 39]. This technique typically leaves
non-synaptic structures (e.g. plasma membranes, neurotubules, and vesicles) less stained,
though considerable variation can exist across samples due to differences in histological chem-
istry, microscope lighting, etc. Tissue was prepared for electron microscopy (EM) imaging
using the same proc edure previously described [26]. Both excitatory and inhibitory synapses
are stained by this technique [26, 38, 39].
We previously developed a machine learning method that uses support vector machines
(SVM) to detect synapses in EPTA-EM images using texture- and shape-based features [26].
The SVM model was trained on data collected in this paper from all 16 developmental time-
points. This compromised 3,708 positive examples (synapses) and 39,163 negative examples
(non-synapses) across all ages studied. Overall, the classifier was highly accurate and achieved
a precision of 90.4% with a recall of 50.0% under 10-fold cross-validation. To ensure that syn-
apse densities were comparable across samples (animals), especially those with variable stain-
ing quality, we manually classified synapses in roughly 20 images per sample, applied the
classifier (which was built on training data from all the other samples) to these images, and
then selected the classification threshold that resulted in 50% recall with 80+% precision (S1
Text, S1 Fig). Recall is defined as: TP / (TP + FN), i.e. the percentage of true synapses correctly
predicted by the classifier. Precision is defined as: TP / (TP + FP), i.e. the percentage of pre-
dicted synapses that are truly synapses. This means that within each sample, we detected
roughly half the synapses, and if the classifier identified a synapse, it was indeed a synapse at
least 80% of the time. If precision was < 80% at 50% recall, the sample was removed from the
analysis. Table 1 shows average precision and recall values for samples in each time-point.
Although we carefully provided our classifier example synapses with a wide variety of
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 15 / 23
structures, shapes, and sizes, there may still be some bias towards classifying certain types of
synapses over others. Full details of the imaging method and synapse classification pipeline,
including their novelty compared to analysis of conventional electron microscopy images, was
previously discussed [26].
A potential method to improve accuracy is to classify synapses in 3D volumes rather than
2D images. Due to challenges related to imaging, alignment, segmentation, and reconstruction
across serial sections, such 3D analysis is currently difficult to fully automate [76, 77], which
makes it difficult to reason statistically about fine-scale pruning rates. To help control for vari-
ability in synapse density in the tissue itself, four regions were sampled from within the barrel
(S2 Fig) and counts were averaged. While this approach of sampling multiple regions within
the same 2D plane may miss synapses, the same procedure was applied to each animal in each
time point, and hence the relative number of synapses per unit area can still be fairly compared
to infer a temporal pruning rate.
To perform the statistical analysis of the pruning rates, we binned the data into 12 bins: P14
only, P17 only, P19 only, P21 and P22, P23 and P24, P26 only, P28 only, P30 only, P32 and
P33, P34 and P36, P38 only, P40 only. By removing one sample or time-point at a time from
the dataset and re-computing the pruning rate using the remaining dataset (known as leave-
one-out cross-validation), we statistically determined whether a single sample or time-point
was responsible for the observed pruning rate.
A theoretical framework for distributed network design
We developed a computational model for designing and evaluating distributed routing net-
works. The problem is as follows:
Problem: Given a set V of n nodes and an online stream of source-target pairs s
i
; t
i
Þg
p
i¼1
,
where s
i
, t
i
2 V are drawn from some distribution D, return a graph G with at most B edges that
is efcient and robust with respect to D.
The source-target pairs are drawn from an a priori unknown distribution D. This distribution
captures some structure in activity (input-output signals) that the network needs to learn during
the training phase in which the network is constructed. For example, half the nodes can be
sources and the other half are targets (the 2-patch distribution; Fig 3A), though the identity of
which node belongs to which class is not known a-priori. The sources and targets are individual
nodes in the network. The source-target pairs are drawn online, which means they are provided
one at a time to the network and thus cannot be processed in bulk, mimicking real-time informa-
tion processing constraints in many types of networks. The pairs are drawn randomly and hence
the same pair may appear multiple times in the training or testing sets.
After p source-target pairs are seen, the goal is to output an unweighted, directed network G
with some fixed number of edges (defined as the budget B). This budget represents the total
allowable cost that the system can maintain (i.e. the number of physical or wireless
connections).
Measuring the quality of a network: Efficiency and robustness
The quality of the final network G is evaluated according to its efficiency and robustness when
processing an additional p pairs drawn from the same distribution D (the testing phase).
During testing, the network is fixed and no changes are made to its connectivity. The test and
train pairs may overl ap (both are drawn from the same distribution), though they are both
likely to also have non-overlapping pairs. This emulates the fact that activity patterns observed
during development mimic those expected later but are not exactly the same. Hence, the chal-
lenge is to design a network that generalizes the training data and does not over-fit.
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 16 / 23
Efficiency is defined as the average shortest-path routing distance over all test pairs [78]:
efficiencyðGÞ¼
1
p
P
u;v2D
dðu; vÞ; where p is the number of source-target pairs observed in the
test phase, and d(u, v) is the shortest-path distance between source u and target v in the nal
network. If there does not exist a path between a pair, then we set its penalty distance to a large
constant.
Robustness is a measure of how tolerant the network is to the deletion of nodes. We adopt a
standard measure for robustness based on vertex connectivity [79]: for each source-target pair,
we compute the number of alternative paths that have up to one additional hop compared to
the true shortest path. This is computed implicitly by removing nodes along the shortest path,
and then finding the length of the next shortest path, etc. This definition of robustness ensures
that if a primary route is attacked or damaged, an alternative route not only exists but is one
that is not much worse than the shortest path.
These definitions are broad, well-established from a graph-theoretic perspective, and appli-
cable to many computing scenarios, but they are not meant to capture all the requirements of
information processing in the brain.
Pruning-based algorithms for distributed network design
To test the impact of pruning and pruning rates we use the following algorithm which is partic-
ularly suitable for routing appli cations. The algorithm begins with a fully connected graph (a
clique) on n nodes. For each source-target pair, the source routes its message to the target via
the shortest path in the graph (computed using a distributed routing table [80, 81]). Initially,
all shortest paths will be direct source-to-target paths. Each edge keeps track of the number of
times it has been used to satisfy a request(i.e. if an edge u ! v lies on the shortest path from
source s
i
to target t
i
, then edge u ! v updates its usage value by 1). All edges initially have a
usage of 0.
The above method is appropriate for simulating computational networks. In contrast, neu-
rons likely cannot route signals via shortest paths in networks. We thus tested another simula-
tion model which is more biologically plausible, yet still abstract. Rather than routing, this
simulation uses a flow-based model that performs a breadth-first search from the source node
(counting all paths between the pairs). Such search does not require any global shortest path
computation. In this model, the usage of edges along every successful path that reaches the tar-
get is upweighted by 1. This model assumes there is feedback to the circuit that rewards every
edge active along a source-to-target response [82]. To further mimic synapse failure (signal
loss) widely present in neural circuits [83], we assumed a constant signal loss probability of
0.65. This means that with probability 0.65, an edge will fail and will not propagate the signal
onwards. Similar values of the signal loss probability led to similar results. This flow process
repeats for each source and target during training. Edges are pruned iteratively according to
different pruning rates (see below).
For the simulations, the pruning period is divided into 10 discrete intervals, each occurring
after 10% of the source-target pairs have been processed. After each interval i, some r
i
-percent-
age of edges are removed (where r
i
depends on the pruning rate, see below). In each interval
the pruned edges are those with the lowest-usage (ties are broken randomly).
Pruning rate strategies
We divided the pruning period into 10 discrete intervals, and after each interval, some r
i
per-
centage of existing connections were pruned. We considered four pruning rate strategies:
increasing, decreasing, constant, and ending (S3 Fig).
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 17 / 23
1. Constant rate: r
1
= r
2
= ...= r
10
. Elimination rates are kept constant (i.e. the same percent-
age of existing connections are removed in each interval).
2. Increasing rate: r
1
< r
2
< ...< r
10
. Elimination begins very slowly and becomes aggressive
later.
3. Decreasing rate: r
1
> r
2
> ...> r
10
. Elimination begins aggressively and then decelerates
over time.
4. Ending rate: r
1
= r
2
= ...= r
9
= 0 and r
10
¼
B
nðn1Þ
. Elimination only occurs in the final
interval and immediately reduces the network from a clique to exactly B edges.
See S1 Text for complete details on how these rates are applied. The Ending rate produced
highly overfit networks with only direct edges connecting a subset of source-target pairs seen
during training. This yielded the worst efficiency and robustness over all rates.
Additional network design algorithms: growing and no-learning
We also tested a growth-based algorithm for solving the network design problem that adds
connections over time starting from a backbone spanning tree (which are commonly used in
engineered systems [17]). See S1 Text for details.
The no-learning algorithm simply selects B random directed edges to form the final network
and ignores the training data.
Supporting Information
S1 Text. Supplementary methods and results.
(PDF)
S1 Fig. Controlling for image quality in EPTA-EM images. A) First, positive (synapses) and
negative (non-synapses) examples were manually labeled in 20 images in the new sample s .B)
Second, the classifier (trained on images from all other samples, excluding s) was applied to the
labeled data for s and the threshold τ that yielded a recall of 50% with precision > 80% was
selected. C) Third, the classifier was applied to all images in s using τ as the classifier threshold.
(TIFF)
S2 Fig. Electron microscopy imaging within a barrel. To control for variability in synapse
density in different areas in the barrel, 4 regions of the barrel were imaged. Tissue was placed
on a mesh copper grid. White circles depict electron beam residue after images were taken.
Approximately 240 images per animal (60 images x 4 regions) were taken covering a total of
6,000μm
2
of tissue per animal.
(TIFF)
S3 Fig. Four pruning rate strategies. Constant rates (red) prune an equal percentage of exist-
ing connections in each pruning interval . Decreasing rates (blue) prune aggressively early-on
and then slower later. Increasing rates (black) are the opposite of decreasing rates. Ending rates
only prune edges in the final iteration. A) Number of edges remaining after each pruning inter-
val. B) Percentage of edges pruned in each pruning interval. Here, n = 1000.
(TIFF)
S4 Fig. Synapse density in adult mice (P65).
(TIFF)
S5 Fig. Pru ning rate with 3D-count adjustment. Adjusted pruning rate per volume of tissue
plotted using A) the raw data (where each point corresponds to a single animal) and B) the
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 18 / 23
binned data (where each point averages over animals from a 2-day window).
(TIFF)
S6 Fig. Pru ning with multiple periods of synaptogenesis and pruning.
(TIFF)
S7 Fig. Comp aring pruning and growing for denser networks.
(TIFF)
S8 Fig. Comp aring the efficiency and robustness of two growing algorithm variants.
(TIFF)
S9 Fig. Comp aring efficiency and robustness of pruning algorithms that start with variable
initial connectivity. A) Initial density is 60% (i.e. each edge exists independently with proba-
bility 0.6. B) Initial density is 80%.
(TIFF)
S10 Fig. Cumulative energy consumed by each pruning algorithm. Energy consumption at
interval i is the cumulative number of edges present in the netwo rk in interval i and all prior
intervals. Here, n = 1000 and it is assumed that the network initially starts as a clique.
(TIFF)
S11 Fig. Theoretical results for network optimization. (A) Example edge-distribution using
decreasing pruning rates and the 2-patch distribution. (B) Prediction of final network p/q ratio
given a pruning rate. Bold bars indicate simulated ratios, and hashed bars indicate analytical
predictions. (C) Prediction of source-target efficiency given a p/q ratio.
(TIFF)
Acknowledgments
We thank Joanne Steinmiller for anim al care.
Author Contributions
Conceived and designed the experiments: SN ALB ZBJ. Performed the experiments: SN ALB.
Analyzed the data: SN. Contributed reagents/materials/analysis tools: SN ALB ZBJ. Wrote the
paper: SN ALB ZBJ.
References
1. Stoneham ET, Sanders EM, Sanyal M, Dumas TC (2010) Rules of engagement: factors that regulate
activity-dependent synaptic plasticity during neural network development. Biol Bull 219: 8199. PMID:
20972254
2. Tessier CR, Broadie K (2009) Activity-dependent modulation of neural circuit synaptic connectivity.
Front Mol Neurosci 2: 8. doi: 10.3389/neuro.02.008.2009 PMID: 19668708
3. Lichtman JW, Colman H (2000) Synapse elimination and indelible memory. Neuron 25: 269278. doi:
10.1016/S0896-6273(00)80893-4 PMID: 10719884
4. Paolicelli RC, Bolasco G, Pagani F, Maggi L, Scianni M, et al. (2011) Synaptic pruning by microglia is
necessary for normal brain development. Science 333: 14561458. doi: 10.1126/science.1202529
PMID: 21778362
5. Chechik G, Meilijson I, Ruppin E (1998) Synaptic pruning in development: a computational account.
Neural Comput 10: 17591777. doi: 10.1162/089976698300017124 PMID: 9744896
6. Huttenlocher PR (1979) Synaptic density in human frontal cortexdevelopmental changes and effects
of aging. Brain Res 163: 195205. doi: 10.1016/0006-8993(79)90349-4 PMID: 427544
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 19 / 23
7. Markus EJ, Petit TL (1987) Neocortical synaptogenesis, aging, and behavior: lifespan development in
the motor-sensory system of the rat. Exp Neurol 96: 262278. doi: 10.1016/0014-4886(87)90045-8
PMID: 3569455
8. Bourgeois JP, Rakic P (1993) Changes of synaptic density in the primary visual cortex of the macaque
monkey from fetal to adult stage. J Neurosci 13: 28012820. PMID: 8331373
9. White EL, Weinfeld L, Lev DL (1997) A survey of morphogenesis during the early postnatal period in
PMBSF barrels of mouse SmI cortex with emphasis on barrel D4. Somatosens Mot Res 14: 3455.
doi: 10.1080/08990229771204 PMID: 9241727
10. Cowan WM, Fawcett JW, OLeary DD, Stanfield BB (1984) Regressive events in neurogenesis. Sci-
ence 225: 12581265. doi: 10.1126/science.6474175 PMID: 6474175
11. Navlakha S, Kingsford C (2011) Network archaeology: uncovering ancient networks from present-day
interactions. PLoS Comput Biol 7: e1001119. doi: 10.1371/journal.pcbi.1001119 PMID: 21533211
12. Gautrais J, Thorpe S (1998) Rate coding versus temporal order coding: a theoretical approach. Biosys-
tems 48: 5765. doi: 10.1016/S0303-2647(98)00050-1 PMID: 9886632
13. Alstott J, Breakspear M, Hagmann P, Cammoun L, Sporns O (2009) Modeling the impact of lesions in
the human brain. PLoS Comput Biol 5: e1000408. doi: 10.1371/journal.pcbi.1000408 PMID: 19521503
14. Feldmeyer D, Radnikow G (2009) Developmental alterations in the functional properties of excitatory
neocortical synapses. J Physiol (Lond) 587: 18891896. doi: 10.1113/jphysiol.2009.169458
15. Albert R, Jeong H, Barabasi AL (2000) Error and attack tolerance of complex networks. Nature 406:
378382. doi: 10.1038/35019019 PMID: 10935628
16. LeVay S, Wiesel TN, Hubel DH (1980) The development of ocular dominance columns in normal and
visually deprived monkeys. J Comp Neurol 191: 151. doi: 10.1002/cne.901910102 PMID: 6772696
17. Lynch NA (1996) Distributed Algorithms. San Francisco, CA, USA: Morgan Kaufmann Publishers Inc.
18. Yuste R (2011) Dendritic spines and distributed circuits. Neuron 71: 772781. doi: 10.1016/j.neuron.
2011.07.024 PMID: 21903072
19. Laughlin SB (2001) Energy as a constraint on the coding and processing of sensory information. Curr
Opin Neurobiol 11: 475480. doi: 10.1016/S0959-4388(00)00237-3 PMID: 11502395
20. Laughlin SB, Sejnowski TJ (2003) Communication in neuronal networks. Science 301: 18701874.
doi: 10.1126/science.1089662
PMID: 14512617
21. Hasenstaub A, Otte S, Callaway E, Sejnowski TJ (2010) Metabolic cost as a unifying principle govern-
ing neuronal biophysics. Proc Natl Acad Sci USA 107: 1232912334. doi: 10.1073/pnas.0914886107
PMID: 20616090
22. Moore D, Shannon C, Brown J (2002) Code-Red: a case study on the spread and victims of an Internet
worm. In: SIGCOMM/USENIX Internet Measurement Workshop. Marseille, France, pp. 273284.
23. Albert R, Albert I, Nakarado GL (2004) Structural vulnerability of the North American power grid. Phys
Rev E Stat Nonlin Soft Matter Phys 69: 025103. doi: 10.1103/PhysRevE.69.025103 PMID: 14995510
24. Carle J, Simplot-Ryl D (2004) Energy-efficient area monitoring for sensor networks. Computer 37: 40
46. doi: 10.1109/MC.2004.1266294
25. Feldmeyer D (2012) Excitatory neuronal connectivity in the barrel cortex. Front Neuroanat 6: 24. doi:
10.3389/fnana.2012.00024 PMID: 22798946
26. Navlakha S, Suhan J, Barth AL, Bar-Joseph Z (2013) A high-throughput framework to detect synapses
in electron microscopy images. Bioinformatics 29: 917. doi: 10.1093/bioinformatics/btt222
27. Dayan P, Abbott LF (2005) Theoretical Neuroscience: Computational and Mathematical Modeling of
Neural Systems. The MIT Press.
28. Dean J, Corrado G, Monga R, Chen K, Devin M, et al. (2012) Large scale distributed deep networks. In:
Advances in Neural Information Processing Systems 25. pp. 12321240.
29. Barabási AL, Albert R (1999) Emergence of scaling in random networks. Science 286: 509512. doi:
10.1126/science.286.5439.509 PMID: 10521342
30. Watts DJ, Strogatz SH (1998) Collective dynamics of small-world networks. Nature 393: 440442.
doi: 10.1038/30918 PMID: 9623998
31. Vazquez A, Flammini A, Maritan A, Vespignani A (2003) Modeling of protein interaction networks. Com-
plexus 1: 3844. doi: 10.1159/000067642
32. Navlakha S, Bar-Joseph Z (2011) Algorithms in nature: the convergence of systems biology and
computational thinking. Nature Mol Syst Biol 7: 546. doi: 10.1038/msb.2011.78
33. De Felipe J, Marco P, Fairen A, Jones EG (1997) Inhibitory synaptogenesis in mouse somatosensory
cortex. Cereb Cortex 7: 619634. doi: 10.1093/cercor/7.7.619 PMID: 9373018
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 20 / 23
34. Crair MC, Malenka RC (1995) A critical period for long-term potentiation at thalamocortical synapses.
Nature 375: 325328. doi: 10.1038/375325a0 PMID: 7753197
35. Feldman DE, Nicoll RA, Malenka RC (1999) Synaptic plasticity at thalamocortical synapses in develop-
ing rat somatosensory cortex: LTP, LTD, and silent synapses. J Neurobiol 41: 92101. doi: 10.1002/
(SICI)1097-4695(199910)41:1%3C92::AID-NEU12%3E3.0.CO;2-U PMID: 10504196
36. Ashby MC, Isaac JT (2011) Maturation of a recurrent excitatory neocortical circuit by experience-
dependent unsilencing of newly formed dendritic spines. Neuron 70: 510521. doi: 10.1016/j.neuron.
2011.02.057 PMID: 21555076
37. Lefort S, Tomm C, Floyd Sarria JC, Petersen CC (2009) The excitatory neuronal network of the c2 bar-
rel column in mouse primary somatosensory cortex. Neuron 61: 301316. doi: 10.1016/j.neuron.2008.
12.020 PMID: 19186171
38. Bloom FE, Aghajanian GK (1966) Cytochemistry of synapses: selective staining for electron micros-
copy. Science 154: 15751577. doi: 10.1126/science.154.3756.1575 PMID: 5924927
39. Bloom FE, Aghajanian GK (1968) Fine structural and cytochemical analysis of the staining of synaptic
junctions with phosphotungstic acid. J Ultrastruct Res 22: 361375. doi: 10.1016/S0022-5320(68)
90027-0 PMID: 4173151
40. Seabrook TA, El-Danaf RN, Krahe TE, Fox MA, Guido W (2013) Retinal input regulates the timing of
corticogeniculate innervation. J Neurosci 33: 1008510097. doi: 10.1523/JNEUROSCI.5271-12.2013
PMID: 23761904
41. Annexstein FS, Berman KA, Jovanović MA (2006) Broadcasting in unstructured peer-to-peer overlay
networks. Theoretical computer science 355: 2536. doi: 10.1016/j.tcs.2005.12.013
42. Hebb DO (1949) The Organization of Behavior: A Neuropsychological Theory. New York: Wiley.
43. Turney SG, Lichtman JW (2012) Reversing the outcome of synapse elimination at developing neuro-
muscular junctions in vivo: evidence for synaptic competition and its mechanism. PLoS Biol 10:
e1001352. doi: 10.1371/journal.pbio.1001352 PMID: 22745601
44. Leskovec J, Backstrom L, Kumar R, Tomkins A (2008) Microscopic evolution of social networks. In:
Proc. 14th ACM SIGKDD Intl. Conf. on Knowledge Discovery and Data Mining (KDD). New York, NY,
USA: ACM, KDD08, pp. 462470.
45. Yassin L, Benedetti BL, Jouhanneau JS, Wen JA, Poulet JF, et al. (2010) An embedded subnetwork of
highly active neurons in the neocortex. Neuron 68: 10431050. doi: 10.1016/j.neuron.2010.11.029
PMID: 21172607
46. Ko H, Hofer SB, Pichler B, Buchanan KA, Sjostrom PJ, et al. (2011) Functional specificity of local syn-
aptic connections in neocortical networks. Nature 473: 8791. doi: 10.1038/nature09880 PMID:
21478872
47. Meyer HS, Wimmer VC, Oberlaender M, de Kock CP, Sakmann B, et al. (2010) Number and laminar
distribution of neurons in a thalamocortical projection column of rat vibrissal cortex. Cereb Cortex 20:
22772286. doi: 10.1093/cercor/bhq067 PMID: 20534784
48. Milo R, Shen-Orr S, Itzkovitz S, Kashtan N, Chklovskii D, et al. (2002) Network motifs: simple building
blocks of complex networks. Science 298: 824827. doi: 10.1126/science.298.5594.824 PMID:
12399590
49. Song S, Sjostrom PJ, Reigl M, Nelson S, Chklovskii DB (2005) Highly nonrandom features of synaptic
connectivity in local cortical circuits. PLoS Biol 3: e68. doi: 10.1371/journal.pbio.0030068 PMID:
15737062
50. Perin R, Berger TK, Markram H (2011) A synaptic organizing principle for cortical neuronal groups.
Proc Natl Acad Sci USA 108: 54195424. doi: 10.1073/pnas.1016051108 PMID: 21383177
51. Royer E, Toh CK (1999) A review of current routing protocols for ad hoc mobile wireless networks. Per-
sonal Communications, IEEE 6: 4655. doi: 10.1109/98.760423
52. Newman M (2010) Networks: An Introduction. New York, NY, USA: Oxford University Press, Inc.
53. DoT (2013). Department of transportation airfare reportthird quarter 2013. http://www.dot.gov/office-
policy/aviation-policy/table-1-domestic-airline-airfare-report-third-qquarter-2013. Accessed: 2014-05-
13.
54. Kleinberg JM (2000) Navigation in a small world. Nature 406: 845. doi: 10.1038/35022643 PMID:
10972276
55. Romer K, Mattern F (2004) The design space of wireless sensor networks. Wireless Communications,
IEEE 11: 5461. doi: 10.1109/MWC.2004.1368897
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 21 / 23
56. Petanjek Z, Judas M, Simic G, Rasin MR, Uylings HB, et al. (2011) Extraordinary neoteny of synaptic
spines in the human prefrontal cortex. Proc Natl Acad Sci USA 108: 1328113286. doi: 10.1073/pnas.
1105108108 PMID: 21788513
57. Huttenlocher PR, de Courten C, Garey LJ, Van der Loos H (1982) Synaptogenesis in human visual cor-
texevidence for synapse elimination during normal development. Neurosci Lett 33: 247252. doi: 10.
1016/0304-3940(82)90379-2 PMID: 7162689
58. Huttenlocher PR, de Courten C (1987) The development of synapses in striate cortex of man. Hum
Neurobiol 6: 19. PMID: 3583840
59. Bickford ME, Slusarczyk A, Dilger EK, Krahe TE, Kucuk C, et al. (2010) Synaptic development of the
mouse dorsal lateral geniculate nucleus. J Comp Neurol 518: 622635. doi: 10.1002/cne.22223 PMID:
20034053
60. Hong YK, Chen C (2011) Wiring and rewiring of the retinogeniculate synapse. Curr Opin Neurobiol 21:
228237. doi: 10.1016/j.conb.2011.02.007 PMID: 21558027
61. Barber MJ, Lichtman JW (1999) Activity-driven synapse elimination leads paradoxically to domination
by inactive neurons. J Neurosci 19: 99759985. PMID: 10559405
62. Goyal MS, Raichle ME (2013) Gene expression-based modeling of human cortical synaptic density.
Proc Natl Acad Sci USA 110: 65716576. doi: 10.1073/pnas.1303453110 PMID: 23576754
63. Wen JA, DeBlois MC, Barth AL (2013) Initiation, labile, and stabilization phases of experience-depen-
dent plasticity at neocortical synapses. J Neurosci 33: 84838493. doi: 10.1523/JNEUROSCI.3575-
12.2013 PMID: 23658185
64. Yang G, Pan F, Gan WB (2009) Stably maintained dendritic spines are associated with lifelong memo-
ries. Nature 462: 920924. doi: 10.1038/nature08577 PMID: 19946265
65. LaMantia AS, Rakic P (1990) Axon overproduction and elimination in the corpus callosum of the devel-
oping rhesus monkey. J Neurosci 10: 21562175. PMID: 2376772
66. III DCD, OMuircheartaigh J, Dirks H, Waskiewicz N, Lehman K, et al. (2014) Modeling healthy male
white matter and myelin development: 3 through 60 months of age. NeuroImage 84: 742752. doi: 10.
1016/j.neuroimage.2013.09.058 PMID: 24095814
67. Nimchinsky EA, Oberlander AM, Svoboda K (2001) Abnormal development of dendritic spines in
FMR1 knock-out mice. J Neurosci 21: 51395146. PMID: 11438589
68. Pfeiffer BE, Huber KM (2009) The state of synapses in fragile X syndrome. Neuroscientist 15: 549
567. doi: 10.1177/1073858409333075
PMID: 19325170
69. Patel AB, Loerwald KW, Huber KM, Gibson JR (2014) Postsynaptic FMRP promotes the pruning of
cell-to-cell connections among pyramidal neurons in the L5A neocortical network. J Neurosci 34:
34133418. doi: 10.1523/JNEUROSCI.2921-13.2014 PMID: 24573297
70. Glaze DG (2004) Rett syndrome: of girls and micelessons for regression in autism. Ment Retard Dev
Disabil Res Rev 10: 154158. doi: 10.1002/mrdd.20030 PMID: 15362175
71. Johnston MV, Blue ME, Naidu S (2005) Rett syndrome and neuronal development. J Child Neurol 20:
759763. doi: 10.1177/08830738050200082601 PMID: 16225832
72. Na ES, Monteggia LM (2011) The role of MeCP2 in CNS development and function. Horm Behav 59:
364368. doi: 10.1016/j.yhbeh.2010.05.014 PMID: 20515694
73. Abbott LF (2008) Theoretical neuroscience rising. Neuron 60: 489495. doi: 10.1016/j.neuron.2008.
10.019 PMID: 18995824
74. Tero A, Takagi S, Saigusa T, Ito K, Bebber DP, et al. (2010) Rules for biologically inspired adaptive net-
work design. Science 327: 439442. doi: 10.1126/science.1177894 PMID: 20093467
75. Gordon DM (2014) The ecology of collective behavior. PLoS Biol 12: e1001805. doi: 10.1371/journal.
pbio.1001805 PMID: 24618695
76. Jain V, Seung HS, Turaga SC (2010) Machines that learn to segment images: a crucial technology for
connectomics. Curr Opin Neurobiol 20: 653666. doi: 10.1016/j.conb.2010.07.004 PMID: 20801638
77. Morgan JL, Lichtman JW (2013) Why not connectomics? Nat Methods 10: 494500. doi: 10.1038/
nmeth.2480 PMID: 23722208
78. Goldberg AV, Harrelson C (2005) Computing the shortest path: A search meets graph theory. In: Pro-
ceedings of the Sixteenth Annual ACM-SIAM Symposium on Discrete Algorithms. Philadelphia, PA,
USA: Society for Industrial and Applied Mathematics, SODA05, pp. 156165.
79. White DR, Newman MEJ (2001) Fast approximation algorithms for finding node-independent paths in
networks. Working papers, Santa Fe Institute.
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 22 / 23
80. Gavoille C (2001) Routing in distributed networks: overview and open problems. SIGACT News 32:
3652. doi: 10.1145/568438.568451
81. Thorup M, Zwick U (2001) Compact routing schemes. In: Proc. 13th annual ACM Symp. on Parallel
Algorithms and Architectures (SPAA). New York, NY, USA: ACM, pp. 110.
82. Hu H, Real E, Takamiya K, Kang MG, Ledoux J, et al. (2007) Emotion enhances learning via norepi-
nephrine regulation of AMPA-receptor trafficking. Cell 131: 160173. doi: 10.1016/j.cell.2007.09.017
PMID: 17923095
83. Barth AL, Poulet JF (2012) Experimental evidence for sparse firing in the neocortex. Trends Neurosci
35: 345355. doi: 10.1016/j.tins.2012.03.008 PMID: 22579264
Pruning Optimizes Construction of Efficient and Robust Networks
PLOS Computational Biology | DOI:10.1371/journal.pcbi.1004347 July 28, 2015 23 / 23
The Erdős–Rényi (ER) random graph model (G(n,M) model) creates all possible graphs (networks) with n nodes and M edges, Each network engages different nodes with different combination of edges, such that all graphs are equally probably and meet the parameters exactly. It's significance in this paper is just to highlight that there was completely random deletion of edges from a fully connected (all edges engaged) initial state. You may ask, why don't we just link the source node to the target node directly for each pair? Then there'd be an effective "highway" for signals, made of frequently used edges. Well, insofar, there is no "most effective way" to send information. Which nodes are source-target pairs, and which edges will be needed, are unknown and not programmed beforehand, but completely random (apriori unknown). To link source-target nodes together with one large edge, we would have to know which are S-T pairs before iterations, and such paths would limit robustness. By having several edges, we could use some edges in a specific pathway for a specific S-T pair, to connect bits traveling to other target nodes. This dramatically saves resources; the concept of sharing is like giving a computer to every student in a school, versus opening a computer cluster that allows students open access to PCs, as needed. Scientists theorize that the success of decreasing rate pruning is because it can establish “hyper-connectivity”. We may feel that the algorithm cannot properly gauge the potential that an edge has before deleting it. We can test the edge’s potential contribution to robustness and efficiency by gathering more information, which an increasing rate pruning algorithm may allow. The study monitors synapse density over time in an area responsible for sensory details (touch) from a specific whisker. From the image, B.) is a visual of spines on a dendrite, which are like distinct ridges all over its surface, each receiving a signal from a particular axon. D.) further explains how pruning can increase efficiency and robustness of nerve networks, by selecting out parts that are no longer used. Depending on the strain, mice will typically reach sexual maturity at 6 weeks for females, 8 weeks for males. Data was collected in the developmental stages of the mice, as pruning rate and synapse density start to stabilize after intense pruning. The more edges existing, the more energy is consumed. An optimal network structure will also minimize the energy spent to theoretically support the edges. From Fig S10, increasing rates have the highest energy expenditure and greatest drastic increase in energy use during the first few pruning intervals, suggesting poor edge pruning decisions contrary to intuition. There’s no one format of the network, or distribution, that would represent the single most optimized system in terms of efficiency and robustness. In decreasing rate pruning trials, the system removes many edges at first, creating a smaller set of edges and nodes to test and manage. This promotes connectivity, and allows for optimization in that particular distribution. One may think that we might lose potentially useful nodes this way, by deleting them early on, before they get to be used frequently enough. Here, we are exploring why experimentally observed decreasing pruning rates are more effective in promoting efficiency and robustness, rather than the algorithm’s constant pruning rate that was assumed (a constant rate of removal of existing edges, each interval). We shift perspectives, from comparing growing, pruning and non-learning algorithms, to different rates of pruning, i.e. increasing, decreasing, and constant, on network efficiency and robustness. Keep in mind that efficiency and robustness are the measures with which we define network performance. Can evolution really help construct the brain in such a way that this complex network functions at its best? Or can machine-learning technologies find the real way to optimize the system? This study takes the questions a step further to optimizing a similarly constructed network in airline traffic. How can we construct airline plans to satisfy the demands of the most people, by offering the most alternative routes to and from the most popular cities, and decrease traveling distance? The most frequently flown cities were identified and placed as nodes on a map, with flights as edges. A similar resource limitation (edge limit) was set. It was observed that decreasing rate pruning algorithms outperformed other rates again. It was evaluated using the same measures of robustness and efficiency, to create a network of the more time efficient and flexible variety of flights. As q represents the likelihood of getting an S T edge, it may be intuitive to want to maximize these, for a lower p/q ratio. However, we should think back to the practicality of having alternative paths as robustness (against failure). An edge that connects a source node to another source node may not seem helpful because it doesn’t directly complete the objective of delivering packets of information from source to target, but can serve as a “stop” on the way to the target. Having a path segmented into multiple edges, (typically two) allows them to be used in routing for many different paths rather than simple between one source-target pair. Theoretically, this allows the system to remove edges that can make source to target travel time longer, or be a waste of resources, as we have a fixed limit of B edges that can exist simultaneously. 10 iterations were run. During each iteration for increasing rates, a percentage of edges were removed: Iteration 1 removes 10%, or r1= 0.10 (synapses remaining expressed by E0 × (1 − r1)) Iteration 2 removes 15%, or r2= 0.15 (remaining expressed by E0 × (1 − r1)(1 − r2))... Iteration 3, 20%, and so on, until Iteration 5 removes (rc) edges, such that E0(1 − rc) I = B. I= 10 iterations, B= number of final edges, and rc= r5. The rates of iterations 6 through 10 are calculated in relation to rate of iterations 4 through 1, respectively. For example, to find the rate of increasing pruning in iteration 10 (r10), we can evaluate E(1 − rc)(1 − rc) = E(1 − r1)(1 − r10), where r1= 0.1, and rc is found experimentally. For the decreasing rates, the sequence was reversed so it started with larger elimination rates. Previously used models, such as the Hopfield network, nodes are designed to receive data from different sources, compute something, and output a value. However, in this study, nodes simply passed "bits" along, much like sending an email from person to person. The EPTA staining method allows for improved visualization of synaptic active zones, by increasing contrast while not interfering with the results of other visualization protocols. It is cost-efficient, relatively easily prepared (PTA in powder form usually, mixed easily with ethanol) and effective in studies involving cerebral development of mice (including visualizing soft tissue in larger, embryos with micro-CT technology). Though slightly unrelated, this is a link to a video using images taken by the micro-CT and stitched together to present a detailed, 3-D overview of the mouse):https://commons.wikimedia.org/wiki/File%3AMicroCT-for-comparative-morphology-simple-staining-methods-allow-high-contrast-3D-imaging-of-1472-6793-9-11-S4.ogv The computer algorithm and network has many parallels with the natural synaptic network and its processes that govern neuroplasticity. Edges represent neuronal connections, since both nodes and neurons are components of a system that transfer data. The directional nature of synaptic messaging is represented in the artificial system with arrows from one node to another. The node from which the “signal” originates, and the destination node, are called the signal-target pair. Rate Comparisons: Increasing: Starts out slow, by eliminating a few edges, and gradually increases the rate of pruning. This gives the algorithm more time to evaluate the activity of the edge before making an edge pruning decision, intuitively, a more “informed” decision. Decreasing rate: Rapid elimination of edges will remove connections from the densely packed network, and tapers off. There is less time to observe and collect edge activity data, compared to increasing rates, but provides more time for network stabilization. Decreasing rates were able to optimize both efficiency and robustness, while one is often achieved at the cost of the other in most situations. Notes on how rates (r) for each iteration in each scenario were determines, are included in later annotations. To get an idea of the precision of the data, the team used a method of statistical analysis called leave-one-out-cross-validation, which is exactly what it sounds like. The data was separated into 12 groups (i.e. P14 only, P17 only, etc.) from specific postnatal days, and one time-point was removed. The pruning rate was recalculated without that point to determine if it significantly changes the observed pruning rate. Towards the end of the trial, at P33, the increase in pruning (by a relatively small amount of 7%) was attributed to the growth of nerve branches from peripheral regions, into the monitored layer 4, though the region was chosen as a good model for observation because it tends to remain undisturbed from activity from other cortical layers. Decreasing pruning rates are also found in the brain, consistent with simulations of neural pruning run with the corresponding algorithm. As time increases, it becomes increasingly more difficult to activate the pathways that control pruning mechanisms. Myelination is the growth of the outer, insulating layer of myelin, around and along the length of the nerve cell. By aggressively pruning at first, less myelin is wasted, as opposed to increasing rates which would prune rapidly after a period of myelination has occurred. ![Myelin diagram](https://psych-brain-trust.wikispaces.com/file/view/6.png/261102636/581x334/6.png "Myelination Support") SVMs are classifiers- they are machine learning models that are trained to assign new data into categories. In this case, subjects in the image are either synapses or not. The code is openly available at http://www.cs.cmu.edu/∼saketn/detect_synapses/. 20 images from each animal were evaluated by human analysis to ensure consistency. If one sample absorbs dye more effectively than others, the machine may be able to identify more synapses than it does in other images. In reality however, the synapse density is relatively consistent across the samples. This image can’t be properly evaluated and should be removed from the analysis. This classification task is a relatively simple decision for humans to make so the scientists determine the number of “true synapses” and “non-synapses” present in the image, which is compared with the classifier’s conclusions. The classifier tried to count synapses in the same image; its performance was evaluated as precision and recall (defined as “percentage of predicted synapses that are truly synapses”, and “percentage of true synapse correctly predicted by the classifier”, respectively). The classification threshold was set such that if 50% of synapses were found in the image, we are 80% confident that it is a true synapse. Images for which the classifier didn’t meet the threshold were basically not clear enough and could result in skewness if included in the data. We know that flipping a coin will either give us heads, or tails, and the chance of either is 50%, even before we flip the coin. The distribution is known. If we are unsure of the distribution, we can run iterations of the coin flip. Maybe the coin is biased such that it will turn up heads 75% of the time. This is what we discover from experimental data; the distribution was unknown. The distribution doesn’t mean the probability of each outcome, but is more like a set of influential traits at a particular time. We aren’t born with synapses already specially formed to receive stimuli from our environment. We don’t know if we’ll be born in a sunny place, a cold place, to a tribe of nomads or any environment that accentuates a particular stimulus, which in turn also biases stimulation of certain synapses. Therefore, the distribution of our synapses is unknown. Our brains are able to process stimuli and adapt accordingly to the effects of our environment on our bodies, especially during developmental periods. This is an “online” way of receiving and processing information that requires constant engagement and “updating” of internal systems. Though genes can give us thicker hair for harsh climates, or long limbs for frequent movement, these are changes that happen over thousands of years; the concept of a priori unknown distributions in our brains allows timely recognition and response to environmental variance. The body can determine what it needs and through selective gene expression, produce components such as ligands (signaling proteins) to alter certain pathways and regulate a function. For example, one of the molecular mechanisms that governs pruning is observed to be the NMDA receptors, which is also involved as a ligand-gated channel at the synpase, embedded in the nerve cell membrane. It requires two different molecules to attach, to cause a conformational change that allows the flow of calcium ions into the cell, causing a change in membrane potential (voltage) and allowing for LTP or LTD (long-term potentiation, or long-term depression, respectively). These are processes that can change synaptic strength, which affects communication between nerves. You can read more about the role of NMDA receptors in neural plasticity here: http://www.biolbull.org/content/219/2/81.long#ref-160 ![Synapse Chemistry]( https://o.quizlet.com/2q4iLfTL6exBkK9BOV6VPw_m.png "Synapse Chemistry") 2s-patch (even)- A specific group of source nodes is connected to a specific group of target nodes. This models how there are specialized, unique connections, as subnetworks in the cortex., rather than continuous and random connections throughout the brain. 2-patch unbalance- There are three times as many target nodes, as there are source nodes. Multiple edges can interact with a single node, and since the distribution is random, some nodes can also be completely uninvolved. This model simulates how there can be a single input of stimulus (such as an eye, as a source node), but when the signal is sent into the brain for processing, it interacts with many more components (contacts several target nodes, modeled similarly to divergent processing). 4 patch- There are independently functioning groups of source and target nodes. 4-patch Hubel-Wiesel- the second set of source is not functioning. This simulates when one source of data, of two, (i.e. one single eye) is compromised, in order to observe how neural systems change in response. The study tested whether different rates of pruning would still perform similarly with different source-target distributions that reflect different biological situations (signals may not always be able to move directly from a source point A to a target point B). Since performance results were consistent, the initial model with 2 patch distribution, with unidirectional paths and equal sampling probability, is an appropriate model for general synaptic network construction. One of the main goals of this paper is to identify characteristics in the natural pruning process that make the network more efficient (described as minimizing time and distance from point to point in the network), and robust ( described as "number of alternative routes between two nodes"), to apply to engineered networks, (such as the airline networks mentioned earlier) for performance optimization. Here is a diagram of nerve cells and how synaptic pruning allows for more effective distribution of activity (See figure D of the below image). ![Synaptic Pruning]( http://english.sibs.cas.cn/ns/es/201508/W020150807524629444734.jpg "Synaptic Pruning") ![Detail of the Synapse]( https://minakinwrites.files.wordpress.com/2012/01/48_05neuronstructure.jpg "Diagram of Nerve Cell") In Figure B: For the pruning algorithm, we start with a network that is densely packed with edges, to reflect the fully developed network of synapses we see in developmental stages of young mice, before the rapid synaptic pruning process or pruning algorithm, respectively, can be applied. This algorithm plays a crucial role in data mining, which allows a computer to identify patterns in large reservoirs of data and represent it differently to reveal new, useful data. In one common example, an algorithm that analyzes the register of sold items from a convenience store can identity that customers who purchased item A, also purchase item B. The association made between these two objects has distinguished a particular consumer behavior (buy A with B), and can help with identifying receptive customers to certain advertisements. Decision trees can be used to organize data into subsets based off of shared characteristics; it is the means to make the association “buy A with B”. In relation to this paper, the pruning algorithm is able to identify nodes and edges in a decision tree that are not frequently used, and remove them. Pruning helps reduce the risk of creating parameters for categorization that are too specific or affected by irrelevant data. How the pruning algorithm works: Pruning decisions were made locally, meaning that the edge between the source and target kept track of use frequency for itself, and would be pruned if it didn’t exceed the minimum uses were determined to be unimportant. This decision is made relative to the performance of other node pairs; the top 10% of pairs in a single iteration of the experiment are retained. Keep in mind that since we are comparing rates in pruning algorithms, the all experiments now start with the densely packed structure of Figure 3B. All nodes are connected, so we will have source to source and target to target edges existing. A no-learning algorithm wastes 25% of its edges by retaining these connections. It doesn't progress from the testing phase. This is like our "control", though the experiment doesn't focus on the control because it is more focused on providing statistical support that specific rates of pruning are advantageous, instead of pruning rates over constant and growing rates. Much of the study of neurobiology is heavily focused on the regulatory mechanisms that control synapse elimination: genetic on-off switches, and protein expression pathways are a few examples. These studies made observations on proof that synaptic pruning is a process in several organisms. Previous studies used Hopfield networks, that simplified the neural system into a network of artificial neurons like the image below. ![Hopfield Neuron]( http://home.agh.edu.pl/~vlsi/AI/hopf/hopfield_eng_pliki/image002.jpg "Diagram of Hopfield system") The neuron is updated after receiving input from each of its branches, to simulate the dendrites of a nerve cell. Each input branch has a corresponding weight; the sum of the products of each input and its weight is the output. The singular output end imitates the terminal branch of the axon, and is either +1 or -1 depending on whether the calculated output is positive or negative, respectively. The structure and components of our neural systems affect many functions, such as memory and learning. Studying neural plasticity, including the mechanisms governing synaptic density, such as synaptic sprouting and pruning, is essential to understanding the pathology of diseases and disorders such as Alzheimer's disease and schizophrenia. You can read more about this specific application here: http://www.ncbi.nlm.nih.gov/pubmed/7932285. You can also listen to this podcast, featuring a brief interview with Maggie Koerth-Baker and Joseph Stromberg, for a brief introduction, here: http://www.sciencefriday.com/segments/roots-of-schizophrenia-zebra-stripes-wind-chill/ In these studies, we are also introduced to the occurrence of over-pruning. Though it seems counterproductive to overproduce, scientists have observed that inactive synapses, or connections that are no longer used in later stages in life, are the ones being removed. While artificial systems are created with the intent to have optimal performance, human neural systems have to actively select and develop themselves to reach that level.