Download Mapping the Brain

Document related concepts

Synaptogenesis wikipedia , lookup

Neuroscience and intelligence wikipedia , lookup

Multielectrode array wikipedia , lookup

Cortical cooling wikipedia , lookup

Molecular neuroscience wikipedia , lookup

Subventricular zone wikipedia , lookup

Donald O. Hebb wikipedia , lookup

Neuromarketing wikipedia , lookup

Biochemistry of Alzheimer's disease wikipedia , lookup

Neuroesthetics wikipedia , lookup

Blood–brain barrier wikipedia , lookup

Single-unit recording wikipedia , lookup

Human multitasking wikipedia , lookup

Neural engineering wikipedia , lookup

Feature detection (nervous system) wikipedia , lookup

Selfish brain theory wikipedia , lookup

Artificial general intelligence wikipedia , lookup

Functional magnetic resonance imaging wikipedia , lookup

Clinical neurochemistry wikipedia , lookup

Brain wikipedia , lookup

Development of the nervous system wikipedia , lookup

Synaptic gating wikipedia , lookup

Neuroeconomics wikipedia , lookup

Activity-dependent plasticity wikipedia , lookup

Neurotechnology wikipedia , lookup

Neurolinguistics wikipedia , lookup

Human brain wikipedia , lookup

Neurogenomics wikipedia , lookup

Neural correlates of consciousness wikipedia , lookup

Aging brain wikipedia , lookup

Optogenetics wikipedia , lookup

Brain Rules wikipedia , lookup

Mind uploading wikipedia , lookup

Neuroplasticity wikipedia , lookup

Brain morphometry wikipedia , lookup

Cognitive neuroscience wikipedia , lookup

Neuropsychology wikipedia , lookup

Haemodynamic response wikipedia , lookup

Channelrhodopsin wikipedia , lookup

Neurophilosophy wikipedia , lookup

Neuroinformatics wikipedia , lookup

Holonomic brain theory wikipedia , lookup

Nervous system network models wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Metastability in the brain wikipedia , lookup

History of neuroimaging wikipedia , lookup

Neuroanatomy wikipedia , lookup

Connectome wikipedia , lookup

Transcript
Focus on Mapping the Brain
npg
© 2013 Nature America, Inc. All rights reserved.
W
Visualization of functional
connectivity in the human
cerebral cortex based on magnetic
resonance imaging data. Brain
image by Joachim Böttger and
Daniel Margulies (Max Planck
Institute for Human Cognitive and
Brain Sciences, Leipzig, Germany)
with compositing help by Tobias S.
Hoffmann. Cover composition by
Erin Dewalt.
e are entering a new era in neuroscience in which technological development will allow us to obtain full anatomical, high-resolution renderings of entire brain circuits and
to map the activity of ever larger cellular populations as an animal
performs specific behaviors. Assembling anatomical, molecular and
functional maps has the potential to greatly advance our understanding of how brains work.
In this Focus, experts outline the technologies needed to obtain
these maps and discuss what will be needed beyond them to understand brain function.
In a Historical Perspective, Cornelia Bargmann and Eve Marder
discuss what has been learned from invertebrate circuits whose
connectivity patterns are known and what will be needed beyond
anatomical maps to understand brain function in other organisms.
In a Commentary, Jeff Lichtman and Joshua Morgan express their
views about why obtaining detailed, high-resolution structural
maps should be an essential part of this endeavor. To understand
the deluge of data these maps will engender once generated, Olaf
Sporns argues in another Commentary that data representation and
modeling will be critical.
Other experts discuss the newest technologies available for
obtaining brain maps. Moritz Helmstaedter presents the state of
the art and current challenges of electron microscopy–based circuit reconstruction. In three papers, the potential of using light
to unveil the function and anatomy of brain circuits is presented.
Karl Deisseroth and Kwanghun Chung discuss their newly developed method named CLARITY for rendering mammalian brains
permeable to visible photons and molecules. Pavel Osten and Troy
Margrie review light-microscopy methods available for large-scale
anatomical tracing and discuss ways to integrate molecular identity,
activity recording and anatomical information. In a Resource, Josh
Sanes and colleagues present improved tools for mapping the mouse
brain using the Brainbow technology. Finally, Michael Milham, Stan
Colcombe and their colleagues review methods for functional and
anatomical analysis of human brains at the macroscale.
We are pleased to acknowledge the financial support of Carl
Zeiss Microscopy, Hamamatsu Corporation, LaVision BioTec,
TissueVision, Inc. and Chroma Technology Corp. Nature Methods
carries sole responsibility for all editorial content and peer review.
Contents
483 From the connectome
to brain function
C I Bargmann & E Marder
491 Making sense of brain
network data
O Sporns
494 Why not
connectomics?
J L Morgan &
J W Lichtman
501Cellular-resolution
connectomics:
challenges of dense
neural circuit
reconstruction
M Helmstaedter
508 CLARITY for mapping
the nervous system
K Chung & K Deisseroth
515 Mapping brain
circuitry with a light
microscope
P Osten & T W Margrie
524 Imaging human
connectomes at the
macroscale
R C Craddock, S Jbabdi,
C-G Yan, J T Vogelstein,
F X Castellanos,
A Di Martino, C Kelly,
K Heberlein, S Colcombe &
M P Milham
540 Improved tools for the
Brainbow toolbox
Dawen Cai, K B Cohen,
T Luo, J W Lichtman &
J R Sanes
Erika Pastrana
Editor, Nature Methods Daniel Evanko
Copy Editor Odelia Ghodsizadeh
Carol Evangelista, Ivy Robles
Focus Editor Erika Pastrana
Publisher Stephanie Diment
Managing Production Editor
Renee Lucas
Sponsorship David Bagshaw,
Yvette Smith, Reya Silao
Senior Copy Editor Irene Kaganman
Production Editors Brandy Cafarella,
Marketing Nazly De La Rosa
nature methods | VOL.10 NO.6 | JUNE 2013 | 481
focus on mapping the brain
historical perspective
From the connectome to brain function
npg
© 2013 Nature America, Inc. All rights reserved.
Cornelia I Bargmann1 & Eve Marder2,3
In this Historical Perspective, we ask what
information is needed beyond connectivity
diagrams to understand the function of nervous
systems. Informed by invertebrate circuits
whose connectivities are known, we highlight
the importance of neuronal dynamics and
neuromodulation, and the existence of parallel
circuits. The vertebrate retina has these features in
common with invertebrate circuits, suggesting that
they are general across animals. Comparisons across
these systems suggest approaches to study the
functional organization of large circuits based on
existing knowledge of small circuits.
An animal’s behavior arises from the coordinated activity of many interconnected neurons—“many” meaning
302 for Caenorhabditis elegans, 20,000 for a mollusc,
several hundred thousand for an insect or billions for
humans. Determining the connectivity of these neurons, via combined anatomical and electrophysiological methods, has always been a part of neuroscience. As
we were writing this, these ideas were being revisited
from the perspective of massively parallel methods for
dense reconstruction, or ‘connectomics’. One thread of
this analysis involves the detailed, high-density mapping of point-to-point connections between neurons
at synapses1–4. The specialized membrane structures
and synaptic vesicles of synapses can be visualized
with an electron microscope, and consequently dense
reconstructions of nervous-system connectomes rely
on electron microscopy of serial brain sections. In a
complementary approach, detailed electrophysiological analysis shows how synapses and circuits function
at high resolution, and is increasingly being applied to
large numbers of interconnected neurons.
The first approaches used to map complete circuits
came from studies of the smaller nervous systems
of invertebrates. In the 1960s and 1970s, systematic
electrophysiological recordings from neurons in
discrete ganglia enabled the identification of neuronal components of circuits that generate specific
behaviors5–7. In association with the recordings of
these individually recognizable, identified neurons, the
cells were filled with dye to visualize their structures
and projection patterns via light microscopy8–10. In
some cases, electron microscopy was used to observe
the anatomical synapses in these small circuits11–13.
But until the publication of the heroic electron micros­
copy reconstruction of the full nervous system of
C. elegans14 in the mid-1980s, it was unimaginable
that the electron microscope could be used to determine circuit connectivity rather than providing
ultrastructural detail to connectivity determined
either with physiological or light microscopy–based
anatomical methods.
Recent advances in electron microscopy and image
analysis have made it possible to scale up this ultrastructural approach: to serially section and reconstruct pieces
of both vertebrate and invertebrate nervous systems,
with the stated purpose of using detailed connectomes
to reveal how these circuits work4,15–18. Such largescale projects will provide new anatomical data that will
offer invaluable insights into the functional organization of the structures studied. An unbiased approach
to data acquisition always reveals surprises and new
insights. Moreover, because of the scope and size of
these projects, such efforts will generate unprecedented
amounts of data to be analyzed and understood.
Here we ask what additional information is needed
beyond connectivity diagrams to understand circuit
function, informed by the invertebrate circuits whose
connectivity is known. For the prototypical case, the
complete C. elegans nervous system, the anatomical
connectome was largely established over 25 years
ago14. In a variety of other invertebrate preparations,
connectivity was established using combinations of
electrophysiological recordings and neuronal tracing 30–
40 years ago, which enabled researchers to generate
a wiring map that incorporates activity information.
Despite their different starting points from anatomy and
electrophysiology, these two approaches have uncovered
1Howard Hughes Medical Institute, The Rockefeller University, New York, New York, USA. 2Volen Center, Brandeis University, Waltham,
Massachusetts, USA. 3Department of Biology, Brandeis University, Waltham, Massachusetts, USA. Correspondence should be addressed to
C.I.B. ([email protected]).
Received 27 February; accepted 5 April; published online 30 may 2013; doi:10.1038/nmeth.2451
nature methods | VOL.10 NO.6 | JUNE 2013 | 483
npg
© 2013 Nature America, Inc. All rights reserved.
historical perspectiveFOCUS ON MAPPING THE BRAIN
Figure 1 | Connectivity of two
well-studied invertebrate circuits.
(a) Connectivity diagram of the crab
STG based on electrophysio­logical
recordings. Red and blue background
shading indicates neurons that are
primarily part of the pyloric and
gastric circuits, respectively. Purple
shading indicates that some neurons
switch between firing in pyloric
and gastric time, and that there
is no fixed boundary between the
pyloric and gastric circuits. Yellow
highlights two neurons that are both
electrically coupled and reciprocally
inhibitory. Green highlights one
of many examples of neurons that
are coupled both monosynaptically
and polysynaptically. (b) The
connectome of C. elegans,
showing all 302 neurons and their
chemical synapses but not their
gap junctions. Each neuron has a
three-letter name, often followed
by a spatial designator. This topto-bottom arrangement (signal
flow view) is arranged to reflect
dominant information flow,
which goes from sensory neurons
(red) to interneurons (blue) to
motor neurons (green). Reprinted
from ref. 50.
a
AB
PD
LPG
LP
IC
LG
MG
Electrical synapse
GM
Chemical inhibitory synapses
PY
Int1
VD
DG
AM
b
ASJL
AIMR
AINL
ASIR
ALMR
AIML
ASIL
AWAR ASGL
AWALAINR
IL2R IL2L
IL2VL
IL2VR
IL2DL
IL2DR
RIH
ASEL
PHAR
ASJR
AWBL AWBR
ADFL
CEPDR
CEPVR
ASERASGR
AWCR
AWCL ADFR
BAGL
CEPDL
CEPVL
URBR
URXL
PLNR
PLNL
ADEL
ASKL
ALNL
PVQL
ADLR
ADLL
HSNL
similar principles and similar puzzles as to how circuit function
arises from the component neurons and their interactions.
What do functional and anatomical maps reveal?
We begin with the connectivity diagram of the stomatogastric ganglion (STG) of the crab, Cancer borealis (Fig. 1a) and a
graph of the connectome of C. elegans (Fig. 1b). In each case, the
number of neurons is small, ~27 neurons or 302 neurons, respectively, but the number of synaptic connections is much larger;
the neurons are extensively interconnected. The basic function of
each circuit is known: to generate rhythmic stomach movements
for the crab STG and to control locomotion behavior in response
to sensory inputs for C. elegans. The intellectual strength of the
STG system is the ability to relate neuronal connectivity to
neuronal activity patterns; the complementary strength of
C. elegans is the ability to relate neuronal connectivity to wholeanimal behavior.
The STG contains motor neurons and interneurons that generate two rhythmic motor patterns19. The pyloric rhythm is an
oscillating, triphasic motor pattern that is continuously active and
depends on a set of electrically coupled pacemaker neurons. The
gastric mill rhythm is episodically active and depends on descending modulatory inputs activated by sensory neurons for its generation19,20. Although these rhythms are easily studied separately,
a close look at the STG connectivity diagram reveals that the neurons that conventionally are thought to be part of the pyloric
circuit (neurons AB, PD, LP, PY, VD and IC) are highly interconnected with those conventionally thought part of the gastric mill
circuit (neurons DG, GM, LPG, MG, LG and Int1) (AM is part
of a third circuit that we will not discuss here). Indeed, many
484 | VOL.10 NO.6 | JUNE 2013 | nature methods
PHCR
ALML
FLPR
VC05AVM
SDQL
PDEL
ASHR
PVPR PLMR
ASHL
AVHR
BDUL
BDUR
AVHL
FLPL
PVNL
PVQR
ADER
PDER
ASKR
AVG
RIFR
SDQR
RIFL HSNR
ADAL
AFDR
RIR
AFDL
OLLL
AUALBAGR
URXRAUAR
URYVL ALNR
URYVR
URBL
PVT
AIYL
RMGR
RIS
AIAL
RMGL
ALA PVR
URAVR
ADAR
AIAR
AIZL
URADR
OLLR
AIZR
URAVL
IL1VR
RICL
IL1R
URADL
AIYR
URYDR
IL1L
SAAVL
OLQVL
IL1VL
RICR AIBL
OLQVR
URYDL
RIVR
SMBVR RMFL
RIAR
RIAL
OLQDR
SAAVR
IL1DL
RIVL
RIGL
RIGR AIBR
OLQDL
SAADR AVKR
SMBVL
SMBDR
IL1DR
AVKL
RMFR
RMHR
RMHL
SAADL
RIBL
AVEL
RMEL
AVER
RIBR
RIPL
SMBDL
RIPR
SIADR
RIMR
SIBDR
RMER
RIML
SIADL
RMEV
SIBVL
SIAVR
RMDL
RMED
SIAVLSIBDL SIBVR
RMDDR
SMDVL
RMDDL
RMDR
RMDVR SMDVR SMDDL
RMDVL
SMDDR
PHAL
PVM
AVFL
LUAR
PVDR
PHBR PVDL
PHBL VC04
PQR
PLMLPHCL
LUAL
PVWR
PVNR
AVFRAVJR
PVPL
AQR
DVA
PVWL PVCR
DVC
AVJL
VB01
AVDR
VA12
AVDL
PDA
VD11
PVCL DVB
VC02
AVL
AVBL
AVBR
VB08
VB11
AVAL
AVAR
VC03
DB01
VB09
RID
VA07
VB07 VB02
VB06
VA02
DB02
VC01
AS11
DA09
DB03
VA09 DB04
PDB VA08
VB04
VA01
VA04
AS01
AS02
VD10VA03
VD13
VA11
AS09
AS06
DA01
DA03
DA04 VA06
AS03
DB05
DB06
AS05
DA05
DA08 DA02
VA05 AS04
VB10
DB07
SABD
SABVR
SABVL
VD01
DA06
AS07
AS10
VD12
DA07
VA10
AS08
VD09
DD05
DD06
DD01
VD02
VD08 VD07
VD05
DD04DD02
DD03
VD03 VD04
VD06
STG neurons switch their activity between the two rhythms19,
and the separation of the STG’s connectivity into two discrete
circuits, although convenient for those who study the network,
does not really capture the highly interconnected reality of the
ganglion’s architecture.
Like all nervous systems, the circuit has many chemical synapses,
in which a presynaptic neuron releases a chemical neurotransmitter to activate receptors on the postsynaptic neuron. Chemical
synapses can be inhibitory or excitatory depending on the nature
of the receptor and associated ion channels; the chemical synapses
among STG neurons are inhibitory. Additional connections are
created by the widespread electrical synapses, mediated by direct
cytoplasmic communication through gap junctions, through
which current flows depending on the voltages of the coupled
neurons. In the STG circuit, there are many instances of neurons
that are connected by electrical synapses as well as by chemical
inhibitory synapses (Fig. 1a). There are also many instances of
neurons connected by reciprocal inhibition. These wiring motifs
contribute to circuit properties that are not easily predictable.
In addition, there are many ‘parallel pathways’ in which two neurons are connected via two or more synaptic routes, one direct
route and additional indirect routes (Fig. 1a). The complexity of
this connection map poses the essential question: are all synapses
important, or are some only important under certain conditions
(as appears to be the case)21? How do we understand the importance of synaptic connectivity patterns that seem to oppose each
other, such as the common motif of electrical coupling between
neurons that also inhibit each other?
The C. elegans wiring diagram was assembled in the nearcomplete absence of prior functional information. It allowed an
FOCUS ON MAPPING THE BRAIN
Anterior touch
ALM
9 classes
of neurons
Posterior touch
AVM
9 classes
of neurons
historical perspective
PLM
AVD
AVB
PVC
6 classes
of neurons
VA
VB
Gap junction
DA
DB
Chemical synapse
Sensory neuron
Interneuron
Motor neuron
Reversal Acceleration
npg
© 2013 Nature America, Inc. All rights reserved.
Figure 2 | C. elegans neurons essential for avoidance of light touch.
Inferred connections necessary for anterior and posterior touch avoidance
are in purple and orange, respectively; other synapses are in black. The
‘essential’ synapses here shown in orange and purple comprise less than
10% of the output synapses of the mechanosensory neurons. Image based
on ablation data from ref. 22.
immediate classification of neurons into large classes: sensory
neurons (with distinctive sensory dendrites and cilia), motor
neurons (with neuromuscular junctions) and interneurons
(a term that is used in C. elegans to describe any neuron that is
not evidently sensory or motor, encompassing projection neurons
and local neurons)14. In each group, neurons were subdivided into
unique types with similar morphologies and connections, collapsing the wiring diagram from 302 neurons to 119 neuronal types.
The flow of information through chemical synapses is predominantly from sensory to interneuron to motor neuron, with many
parallel pathways linking neurons both directly and indirectly
(as in the STG), as well as gap junctions that may form electrical
synapses (~10% of all synapses). Most neurons are separated from
each other by no more than two or three synaptic connections.
The C. elegans map was immediately used to define neurons
required for the touch-avoidance response, which is still the most
completely characterized of the animal’s behaviors22. Light touch
to the head elicits a reversal, and light touch to the tail elicits
a forward acceleration. The neurons required for the touchavoidance response were identified by killing cells with a laser
microbeam and assessing the behavioral repertoire of the worms.
Guided by the wiring diagram, this analysis revealed essential
mechanosensory neurons in the head and tail, key interneurons
required to propagate information, and motor neurons required
for forward and backward movement (Fig. 2). The success of
this approach inspired similar analyses of chemosensory behaviors, foraging, egg-laying, feeding and more. At this point, over
60% of C. elegans neuron types have defined functions in one or
more behaviors.
This notable success, however, hides a surprising failure. For
C. elegans, although we know what most of the neurons do, we
do not know what most of the connections do, we do not know
which chemical connections are excitatory or inhibitory, and we
cannot easily predict which connections will be important from
the wiring diagram. The problem is illustrated most simply by
the classical touch-avoidance circuit22 (Fig. 2). The PLM sensory
neurons in the tail are solely responsible for tail touch avoidance. PLM forms 31 synapses with 11 classes of neurons, but only
one of those targets is essential for the behavior—an interneuron
called PVC that is connected to PLM by just two gap junctions
and two chemical synapses. An even greater mismatch between
the number of synapses and their importance in behavior is seen
in the avoidance of head touch, where just two of 58 synapses
(again representing gap junctions) are the key link between the
sensory neurons (ALM and AVM) and the essential inter­neuron
(AVD). This general mismatch between the number of synapses and apparent functional importance has applied wherever
C. elegans circuits have been defined. As a result, early guesses
about how information might flow through the wiring diagram
were largely incorrect.
Clearly, the wiring diagram could generate hypotheses to test,
but solving a circuit by anatomical inspection alone was not successful. We believe that anatomical inspection fails because each
wiring diagram encodes many possible circuit outcomes.
Parallel and antagonistic pathways complicate circuits
Both of the wiring diagrams shown in Figure 1 are richly connected. In the STG, a large fraction of the synapses are electrical
synapses. In some cases, the electrical synapses connect multiple
copies of the same neuron, such as the two PD neurons in the
STG. Notably, many electrical synapses connect neurons with different functions. Almost invariably, the combination of electrical
and chemical synapses create ‘parallel pathways’, that is to say,
multiple pathways by which neuron 1 can influence neuron 2
(Fig. 1a). For example, in the STG, the PD neuron inhibits the IC
neuron through chemical synapses but also can influence the IC
neuron via the electrical synapse from LP to IC. Parallel pathways
such as those in the STG can be viewed as degenerate, as they
create multiple mechanisms by which the network output can be
switched between states23 (Fig. 3). A simulation study23 shows a
simplified five-cell network of oscillating neurons coupled with
electrical synapses and chemical inhibitory synapses. The f1 and
f2 neurons are connected reciprocally by chemical inhibitory
synapses, as are the s1 and s2 neurons. This type of wiring configuration, called a half-center oscillator, often but not universally
causes the neurons to be rhythmically active in alternation 24. In
this example, two different oscillating rhythms are generated, one
fast and one slow. The hub neuron at the center of the network
can be switched between firing in time with the fast f1 and f2
neurons to firing in time with the slow s1 and s2 neurons by three
entirely different circuit mechanisms: changing the strength of the
electrical synapses, changing the strength of the synapses between
f1 and s1 onto the hub neuron, and changing the strength of the
reciprocal inhibitory synapses linking f1 to f2 and s1 to s2 in the
half-center oscillators.
An example from the C. elegans connectome illustrates another
twist of circuit logic: divergent circuits that start at a common
point but result in different outcomes. In this example, gap junctions and chemical synapses from ADL sensory neurons generate
opposite behavioral responses to a C. elegans pheromone (Fig. 4a).
The chemical synapses drive avoidance of the pheromone, whereas
the gap junctions stimulate a pheromone-regulated aggregation
behavior25. Differing use of the chemical synapse subcircuit versus the gap junction subcircuit allows ADL to switch between
these two opposing behaviors in different contexts. ADL illustrates the point that is not possible to ‘read’ a connectome if it is
intrinsically ambiguous, encoding two different behaviors.
Parallel and divergent systems of synapses are widespread features of invertebrate and vertebrate networks alike, and can be
composed of sets of chemical synapses as well as sets of chemical
nature methods | VOL.10 NO.6 | JUNE 2013 | 485
historical perspectiveFOCUS ON MAPPING THE BRAIN
npg
© 2013 Nature America, Inc. All rights reserved.
Figure 3 | Similar changes in
a
b
c
d
circuit dynamics can arise from
f1
f1
f1
gsynB f1
three entirely different circuit
gsynA
g
mechanisms. (a) Circuit diagram
el
s2
f2
hn
s2
f2
hn
s2
f2
hn
s2
f2
hn
(top) shows the ‘control’ condition
gel
in which f1 and f2 are firing
gsynA
gsynB
s1
s1
s1
s1
in a fast rhythm (indicated by
red shading) and the remaining
f1
neurons are firing in a slow
f2
rhythm (shaded in blue). hn is the
hn
hub neuron. In these diagrams,
s2
electrical synapses are shown as
s1
resistor symbols and chemical
100 mV 1 s
inhibitory synapses with filled
circuits. Traces (bottom) show the
voltage waveforms of the five neurons. (b–d) Responses when the strength of the chemical synapses to the hub neuron (gsynA) was decreased (b), when
the strength of the electrical synapses (gel) was decreased (c) and when the strength of the chemical synapses between f1 and f2 and between s1 and s2
(gsynB) was decreased (d). Image modified from ref. 23.
and electrical synapses. To understand information flow, there
will be no substitute for recording activity. The methods for monitoring neuronal activity have improved dramatically in recent
years, with development of new multi-electrode recording techniques and a suite of genetically encoded indicators that can be
used to measure calcium, voltage and synaptic release at cellular
and subcellular levels. However, improved methods are needed
to detect electrical synapses, which can also be difficult to see
in electron micrographs. The regulation of electrical synapses
by voltage, neuromodulation, phosphorylation and small mole­
cules is understudied26,27. A chemical method for measuring gap
junctions, local activation of molecular fluorescent probes, is a
promising new direction that should spawn innovation28.
Neuromodulation reconfigures circuit properties
Superimposed on the fast chemical synapses and electrical synapses in the wiring diagram are the neuromodulators—biogenic
amines (serotonin, dopamine, norepinephrine and histamine)
and neuropeptides (dozens to hundreds, depending on species)29.
These molecules are often released together with a fast chemical
transmitter near a synapse, but they can diffuse over a greater distance. Modulators also can be released from neuroendocrine cells
that do not make defined synaptic contacts or can be delivered as
hormones through the circulation. As a result, the targets of neuromodulation are invisible to the electron microscope. Signaling primarily through G protein–regulated biochemical processes rather
than through ionotropic receptors, neuromodulators change neuronal functions over seconds to minutes, or even hours.
Many years of work on the effects of neuromodulators on the
STG have revealed that the functional connections that give rise
to a specific circuit output are specified, or in fact ‘configured’, by
the neuromodulatory environment29. Every synapse and every
neuron in the STG is subject to modulation; the connectivity diagram by itself only establishes potential circuit configurations,
whose availability and properties depend critically on which of
many neuromodulators are present at a given moment29. Under
some modulatory conditions, anatomically ‘present’ synaptic
connections may be functionally silent, only to be strengthened
under other modulatory conditions. Likewise, modulators can
qualitatively alter the neurons’ intrinsic properties, transforming
neurons from tonic spiking to those generating plateau potentials or bursts29. These effects of neuromodulators can activate
or silence an entire circuit, change its frequency and/or the phase
relationships of the motor patterns generated.
C. elegans has over 100 different neuropeptides as well as biogenic amine neuromodulators. The integration of neuromodulation into its fast circuits appears to selectively enhance the use of
particular connections at the expense of others. For example, a
‘hub-and-spoke’ circuit drives aggregation of C. elegans by coupling multiple sensory inputs through gap junctions with a common target neuron, RMG (Fig. 4b). Neuromodulation of RMG
by the neuropeptide receptor NPR-1 effectively silences this gapjunction circuit, while sparing other functions of the input sensory
neurons that are mediated through chemical synapses30.
Neuromodulators are prominent in all nervous systems, and
act as key mediators of motivational and emotional states such as
Figure 4 | Two views of a multifunctional
a
b
C9 pheromone
C. elegans circuit. (a) Ambiguous circuitry
Pheromone
Oxygen
of the ADL sensory neurons, which drive
ADL
ASK
URX
attraction
avoidance
AVA
NPR-1
avoidance of the ascaroside pheromone
C9 through chemical synapses onto multiple
AIA
Nociceptive
Pheromone
RMG
RMG
ADL
ASH
interneurons (right) but can also promote
AVD
avoidance
avoidance
aggregation (attraction toward pheromones)
AIB
through gap junctions with RMG (left). Image
IL2
AWB
modified from ref. 25. (b) Neuromodulation
Pheromone
Aggregation
avoidance
Aggregation
separates overlapping circuits. Multiple sensory
neurons form gap junctions with the RMG hub
Gap junction
Chemical synapse
Sensory neuron
Interneuron or motor neuron
neurons and promote aggregation through this
circuit, but each sensory neuron also has chemical synapses that can drive RMG-independent behaviors. The neuropeptide receptor NPR-1 inhibits
RMG to suppress aggregation. Image modified from ref. 30.
486 | VOL.10 NO.6 | JUNE 2013 | nature methods
FOCUS ON MAPPING THE BRAIN
npg
© 2013 Nature America, Inc. All rights reserved.
Neuronal dynamics shape the activity of circuits
The existence of parallel circuits and neuromodulation means
that connectivity alone does not provide adequate information
to predict the physiological output of circuits. Even without these
factors, the behavior of neurons over time is unpredictable from
anatomy because neuronal behavior is sensitive to intrinsic channels and electrical properties that vary within and between cell
types. Channels, synapses and biochemical processes interact
to generate explicitly time-delimited features, or dynamics, in
neurons and circuits.
The importance of neuronal dynamics in circuit function can
be seen most simply in a two-cell circuit (Fig. 5). Two isolated
neurons from the STG that are not normally synaptically coupled
were connected using the dynamic clamp, a computer-neuronal
interface that allows a user to manipulate biological neurons
with conductances that imitate ion channels and synaptic connections32. The neurons are connected reciprocally by dynamic
clamp-created inhibitory synapses so that the neurons rhythmically alternate their activity24. The dynamic clamp allows the
investigator to change the strength of the synapses as well as the
amount of one of the membrane currents, ­hyperpolarizationactivated inward current (Ih)—either of which dramatically
alters the period of the circuit oscillation (Fig. 5). Thus, a given
wiring diagram can produce widely different dynamics with
different sets of circuit parameters, and conversely, different
circuit mechanisms can give rise to similar oscillation dynamics.
Without knowing the strength and time course of the ­synaptic
connections as well as the numbers and kinds of membrane
currents in each of the neurons, it would not be possible to
simply go from the wiring diagram to the dynamics of even
two neurons. Synaptic connectivity alone does not sufficiently
constrain a system.
Understanding neuron-specific and circuit-specific dynamics
will be essential to understanding mammalian circuits as well as
invertebrate circuits. In some cases, unique dynamic properties
are characteristics of particular cell types—for example, different
classes of inhibitory cortical interneurons are distinguished as
much by their dynamics as by their connectivity33. In other cases,
neuronal dynamics are variable among similar cells or even within
one cell type. For example, pyramidal neurons in specific areas
of the cortex exhibit persistent activity associated with working
memory34, and neurons in brainstem modulatory systems switch
their properties between tonic and phasic firing modes ­depending
25
1
Period (s)
sleep, arousal, stress, mood and pain. To understand their release
and their effects on circuits, new methods are needed to monitor
neuromodulation in vivo. Electrophysiology remains the best tool
for characterizing the functional effects of neuromodulators but
is low-throughput. Biochemical methods can be used to reveal
the presence of neuromodulators in tissue or in bulk extracellular
fluid but are less effective for detecting them near a particular
synapse or release site. A new genetic method can be used to read
out the neuromodulatory state directly by monitoring receptor
activation but with a timeframe of hours, whereas endogenous
modulation can change within minutes or seconds31. Progress
is needed in all of these domains and beyond: there is a need to
move from individual neurons and modulators to physio­logically
relevant modulatory states, which are likely to include multiple
neuromodulators acting at many sites.
historical perspective
2
Synaptic
Ih
20
15
10
5
0
0 20 40 60 80 100
Conductance (nS)
Synaptic conductance
60 nS
50 nS
80 nS
2
22
22
1
11
11
10 mV
2s
lh conductance
80 nS
50 nS
30 nS
2
22
22
1
11
11
10 mV
2s
Figure 5 | Changing either intrinsic neuronal properties or synaptic
properties can alter network function. The dynamic clamp, a computerneuron interface (top) was used to vary either the strength of synaptic
connections between two neurons (synaptic) or the amount of an
intrinsic hyperpolarization-activated inward current (Ih), in one neuron
as graphed (right). Traces (bottom) show the action potentials generated
by the alternating, oscillating neuron pair as those properties were
varied. Image modified from ref. 24.
on behavioral states35. Finally, synaptic plasticity can occur on
rapid timescales to strengthen and weaken synapses based on use,
adding complexity to circuit-level dynamics36.
Analyzing neuronal dynamics often requires the circuit to be
simultaneously monitored and manipulated, as shown in the
example of the dynamic clamp. Emerging techniques of optogenetics and pharmacogenetics can be combined with recording
as well, but a limitation of all of these methods is that they act at
the level of neurons or groups of neurons. To understand functional connectivity, it will be useful to develop methods to silence
or activate specific channels and specific synaptic connections
between two specified neurons, without affecting all other functions of the same cells.
Vertebrate retina also has complex circuit properties
What lessons will emerge as connectomes are scaled up from
small-scale to large-scale circuits? Many features will be common to small and large circuits. Vertebrate circuits, like invertebrate circuits, have multiple cell types with nonuniform intrinsic
properties, extensive and massively parallel synaptic connectivity,
and neuromodulation. The balance of these components varies
between animals and brain regions (the STG has more electrical synapses than most vertebrate brain regions; C. elegans uses
mostly graded potentials instead of all-or-none action potentials),
but in reality, the diversity of circuits in the vertebrate brain is
at least as great as the difference between any one vertebrate
region and any invertebrate circuit. The essential distinction we
see in vertebrate brains is not a particular microcircuit property
but their repeating structure (for example, the many cortical
columns) and their enormous scale compared to the worm brain
and the STG.
nature methods | VOL.10 NO.6 | JUNE 2013 | 487
npg
© 2013 Nature America, Inc. All rights reserved.
historical perspectiveFOCUS ON MAPPING THE BRAIN
Examination of the vertebrate retina has begun to reveal the
relationships between performance of a large circuit and properties of a small circuit. The special power of studying the isolated retina is the ability to experimentally control visual input
while simultaneously recording output—the spikes from retinal
ganglion cells that project to the brain. The retina represents an
intermediate degree of complexity with features of both a small
circuit and a large circuit, and has been subject to the most complete anatomical and electrophysiological characterizations of
any vertebrate brain region. Current connectomics studies of
the retina, for example, include dense reconstruction of serialsection electron micrographs accompanied by analysis of the
neurotransmitter phenotype and activity patterns of the reconstructed neurons3,17. The combination of structure and function,
and a rich history of elegant experiments, make this the ideal
system for understanding neural computations in detail.
The retina contains millions of neurons that fall into five major
neuronal classes (photoreceptors, horizontal cells, bipolar cells,
amacrine cells and retinal ganglion cells), which are subdivided
into about 60 discrete cell types37. Each of the 60 cell types is
arrayed in a near-crystalline two-dimensional array, so that any
pixel viewed by the retina is covered by at least one neuron of each
cell type. Ultimately, information leaves the retina through the
20 classes of retinal ganglion cells, each of which is considered to
be a parallel but partially overlapping processing stream.
The first views of the retinal connectome show all of the properties that we highlight in small circuits: cellular complexity,
extensive interconnectivity, parallel circuits with chemical and
electrical synapses, and neuromodulation. The heterogeneity of
the 60 retinal cell types is substantial, echoing the heterogeneity
of individual neurons in the STG or C. elegans. Anatomically,
some dendritic arbors cover only a tiny area of the visual field, but
others arborize much more broadly. Their intrinsic physiological
properties are also extremely diverse, with some neurons that
spike (such as retinal ganglion cells), and many neurons that do
not spike (such as photoreceptors and bipolar cells)37. There are
even amacrine neurons that perform independent computations
in different parts of their complex arbors38.
Synaptic connections in the retina are extensive and diverse,
and electron microscopy reconstructions have revealed many
classes of synaptic connections that had not been observed in
physiological studies3,17. There is a great variety of excitatory
and inhibitory chemical synapses, and there are many electrical synapses, that all vary in their strength and their modification by experience. Both anatomical and physiological studies
demonstrate that the retina, like small circuits, consists of many
partly parallel circuits with overlapping elements. In particular, the retina operates over many orders of magnitude of light
intensity, and the properties of its circuits change with its visual
inputs. Within a few seconds in a new visual environment, retinal
ganglion cells shift their properties to encode relevant features
of light intensity, contrast and motion, drawing on different
features of the network39. Subsets of retinal ganglion cells change
their weighting of center and surround inputs in a switch-like
fashion as light levels change40. A brief period of visual stimulation can even reverse the apparent direction-selectivity of retinal
ganglion cells41.
Finally, neuromodulation has a role in retinal processing that
reshapes visual circuits. Dopamine is released from a subset of
488 | VOL.10 NO.6 | JUNE 2013 | nature methods
amacrine cells around dawn, under the control of acute light
stimuli and circadian rhythm42, and acts on cells throughout the
retina to switch them from properties appropriate to night vision
to day vision. The photoreceptors themselves are modulated, and
their coupling through gap junctions decreases to increase their
resolution but reduce their sensitivity. Downstream of the rod
photoreceptors, which dominate night vision, dopamine closes
the gap junctions between rod bipolar cells and AII amacrine
cells, effectively diminishing rod input to the retinal ganglion cell
output of the retina.
What differences are there between small and large circuits? The
sheer size of the retina shows a sharp transition compared to the
size of the STG and the worm brain, and the level of analysis moves
from single cells to cell classes. Understanding a single pixel is
not sufficient to understand the retina, and here the properties of
simple and complex circuits diverge. For example, long-range communication allows groups of retinal cells to perform computations
that a single cell cannot. Wide-field cells such as starburst amacrine
cells can make judgments about motion that no single-pixel neuron
could make but can then feed that information into narrow-field
single-pixel neurons to bias their properties. The scaling from fine
resolution to broad resolution and back again emerges from the
diversity of spatial scales across the structure of the retina.
Circuits interact to generate behavior
The entire nervous system is connected, but reductionist neuroscientists invariably focus on pieces of nervous systems. The
value of these simplified systems should not let us forget that
behavior emerges from the nervous system as a whole. At the
moment, obtaining the connectomes of even small parts of
the vertebrate nervous system is a heroic task. However, establishing the detailed pattern of connectivity for a small part of the
nervous system may not be sufficient to understand how that
piece functions in its full context. By parceling out small regions,
one invariably loses information about the long-range connections to and from that area.
The extent to which long-range connectivity clouds our understanding of connectomes will vary. For example, the vertebrate
retina is anatomically isolated, functionally coherent and lacks
recurrent feedback synapses from other brain areas that are
prominent in most other parts of the central nervous system. We
might imagine the retina as a two-dimensional circuit, whereas
most vertebrate circuits are three-dimensional; new principles
will certainly arise from connectomes that include recurrent
inputs. In the amygdala, for example, the intermixing of multiple cell types with different long-range inputs and outputs would
preclude a meaningful understanding based on local anatomy
alone43. Choosing well among brain regions, and combining connectomes with molecular and functional information about the
same cells, as is being done in the vertebrate retina3,17, will lead
to the most informative results.
How can we ‘solve’ the brain?
As we look to ways that other neural systems may be characterized with similar power to the three described here, we can
draw certain lessons. One is that precise circuit mapping and specific neuron identification have had great importance for unifying structural and functional data from different laboratories.
Extending this idea, other systems may not have individually
npg
© 2013 Nature America, Inc. All rights reserved.
FOCUS ON MAPPING THE BRAIN
named cells, but all nervous systems have cell types distinguished
by anatomy, connectivity and molecular profile that can serve as
the basis of a common vocabulary. Improvements in molecular
staining methods will only increase the power of a connectome
anchored in cellular identity.
We can also see that connectomic endeavors will need to be
supplemented by experiments that monitor, manipulate and
model circuit activity. Monitoring and manipulation of circuit
function have been considered above. To complement and inform
these experimental approaches, the third step is to develop models that describe how a system’s output results from the interactions of its components. There is a tension between the desire to
study abstract models that are amenable to precise mathematical
analysis and the desire to study models with sufficient biological realism to represent the system’s underlying structures and
functions. In small circuits, it is now possible to construct
models and families of models that can be quite instructive44.
In C. elegans, a few testable models emerged directly from analyses of anatomy. One was the concept of a motif, a set of connection
patterns between three or four neurons that are over-represented
in the wiring diagram compared to the statistical expectation
based on individual connections45. Perhaps these motifs perform a canonical computation, or a few canonical computations,
so that solving a few of them effectively solves a larger piece
of the diagram.
But how should we approach building models of large networks
without generating models that are as difficult to understand as
the biological systems that motivated and inspired them? The
beauty of the connectome is its precision and specificity, but it
is hard to imagine useful network models that implement all of
the details of cell-to-cell connectivity obtained with the electron
microscope, when building such models would require enormous
numbers of assumptions about other circuit parameters, and these
parameters are likely to change in different modulatory states. So
we face a conundrum: the new anatomical data will be instructive,
but it is not yet obvious what kinds of models will best reveal the
implications of these data for how circuits actually work.
We are in the midst of a fascinating international debate about
whether it is the right time to embark on a ‘big science’ project to
monitor and model large brain regions. There are those who argue
that we are now at the point at which investments in large-scale
projects will considerably advance the field in ways not possible
by a distributed small-lab approach46–48. Big science works best
when the goals of a project are well-defined and when the outcomes can be easily recognized. Both were true about the human
genome project, but neither is true, yet, about large-scale attempts
to understand the brain. Moreover, this is well-recognized, and all
of the proponents of large-scale initiatives are acutely aware of the
necessity to develop new technology48 and of the extraordinary
complexity of biological systems49. That said, the largest challenge
we face in future attempts to understand the dynamics of large
circuits is not in collecting the data: what is most needed are new
methods that allow our human brains to understand what we find.
Humans are notoriously bad at understanding multiple nonlinear
processes, although we excel at pattern recognition. Somehow, we
have to turn the enormous data sets that are already starting to be
generated into a form we can analyze and think about. Otherwise,
we will be doomed to creating a machine that will understand the
human brain better than we can!
historical perspective
Acknowledgments
We thank M. Meister for sharing his knowledge of the retina. C.I.B. is funded by
the Howard Hughes Medical Institute. Research in the Marder laboratory relevant
to this piece is funded by the US National Institutes of Health (NS17813, NS
81013 and MH46742).
COMPETING FINANCIAL INTERESTS
The authors declare no competing financial interests.
Reprints and permissions information is available online at http://www.nature.
com/reprints/index.html.
1. Bock, D.D. et al. Network anatomy and in vivo physiology of visual
cortical neurons. Nature 471, 177–182 (2011).
2. Briggman, K.L. & Bock, D.D. Volume electron microscopy for neuronal
circuit reconstruction. Curr. Opin. Neurobiol. 22, 154–161 (2012).
3. Briggman, K.L., Helmstaedter, M. & Denk, W. Wiring specificity in the
direction-selectivity circuit of the retina. Nature 471, 183–188 (2011).
4. Kleinfeld, D. et al. Large-scale automated histology in the pursuit of
connectomes. J. Neurosci. 31, 16125–16138 (2011).
5. Burrows, M. Monosynaptic connexions between wing stretch receptors and
flight motoneurones of the locust. J. Exp. Biol. 62, 189–219 (1975).
6. Fentress, J.C. Simpler Networks and Behavior (Sinauer Associates, 1976).
7. Getting, P.A., Lennard, P.R. & Hume, R.I. Central pattern generator
mediating swimming in Tritonia. I. Identification and synaptic
interactions. J. Neurophysiol. 44, 151–164 (1980).
8. Stretton, A.O. & Kravitz, E.A. Neuronal geometry: determination with a
technique of intracellular dye injection. Science 162, 132–134 (1968).
9. Remler, M., Selverston, A. & Kennedy, D. Lateral giant fibers of cray fish:
location of somata by dye injection. Science 162, 281–283 (1968).
10. Goodman, C.S. Isogenic grasshoppers: genetic variability in the morphology
of identified neurons. J. Comp. Neurol. 182, 681–705 (1978).
11. Maynard, E.A. Electron microscopy of the stomatogastric ganglion in the
lobster, Homarus americanus. Tissue Cell 3, 137–160 (1971).
12. King, D.G. Organization of crustacean neuropil. I. Patterns of synaptic
connections in lobster stomatogastric ganglion. J. Neurocytol. 5, 207–237
(1976).
13. King, D.G. Organization of crustacean neuropil. II. Distribution of synaptic
contacts on identified motor neurons in lobster stomatogastric ganglion.
J. Neurocytol. 5, 239–266 (1976).
14. White, J.G., Southgate, E., Thomson, J.N. & Brenner, S. The structure of
the nervous system of the nematode Caenorhabditis elegans. Phil. Trans.
R. Soc. Lond. B 314, 1–340 (1986).
15. Seung, H.S. Reading the book of memory: sparse sampling versus dense
mapping of connectomes. Neuron 62, 17–29 (2009).
16. Meinertzhagen, I.A. & Lee, C.H. The genetic analysis of functional
connectomics in Drosophila. Adv. Genet. 80, 99–151 (2012).
17. Anderson, J.R. et al. Exploring the retinal connectome. Mol. Vis. 17,
355–379 (2011).
18. Lu, J., Tapia, J.C., White, O.L. & Lichtman, J.W. The interscutularis muscle
connectome. PLoS Biol. 7, e32 (2009).
19. Marder, E. & Bucher, D. Understanding circuit dynamics using the
stomatogastric nervous system of lobsters and crabs. Annu. Rev.
Physiol. 69, 291–316 (2007).
20. Nusbaum, M.P. & Beenhakker, M.P. A small-systems approach to motor
pattern generation. Nature 417, 343–350 (2002).
21. Thirumalai, V., Prinz, A.A., Johnson, C.D. & Marder, E. Red pigment
concentrating hormone strongly enhances the strength of the feedback
to the pyloric rhythm oscillator but has little effect on pyloric rhythm
period. J. Neurophysiol. 95, 1762–1770 (2006).
22. Chalfie, M. et al. The neural circuit for touch sensitivity in Caenorhabditis
elegans. J. Neurosci. 5, 956–964 (1985).
23. Gutierrez, G.J., O’Leary, T. & Marder, E. Multiple mechanisms switch an
electrically coupled, synaptically inhibited neuron between competing
rhythmic oscillators. Neuron 77, 845–858 (2013).
24. Sharp, A.A., Skinner, F.K. & Marder, E. Mechanisms of oscillation in
dynamic clamp constructed two-cell half-center circuits. J. Neurophysiol. 76,
867–883 (1996).
25. Jang, H. et al. Neuromodulatory state and sex specify alternative
behaviors through antagonistic synaptic pathways in C. elegans.
Neuron 75, 585–592 (2012).
26. Pereda, A.E. et al. Gap junction-mediated electrical transmission: regulatory
mechanisms and plasticity. Biochim. Biophys. Acta 1828, 134–146 (2013).
27. Neyton, J. & Trautmann, A. Physiological modulation of gap junction
permeability. J. Exp. Biol. 124, 93–114 (1986).
nature methods | VOL.10 NO.6 | JUNE 2013 | 489
28. Guo, Y.M. et al. Imaging dynamic cell-cell junctional coupling in vivo
using Trojan-LAMP. Nat. Methods 5, 835–841 (2008).
29. Marder, E. Neuromodulation of neuronal circuits: back to the future.
Neuron 76, 1–11 (2012).
30. Macosko, E.Z. et al. A hub-and-spoke circuit drives pheromone attraction
and social behaviour in C. elegans. Nature 458, 1171–1175 (2009).
31. Inagaki, H.K. et al. Visualizing neuromodulation in vivo: TANGO-mapping
of dopamine signaling reveals appetite control of sugar sensing. Cell 148,
583–595 (2012).
32. Sharp, A.A., O’Neil, M.B., Abbott, L.F. & Marder, E. Dynamic clamp: computergenerated conductances in real neurons. J. Neurophysiol. 69, 992–995 (1993).
33. Kawaguchi, Y. & Kondo, S. Parvalbumin, somatostatin and cholecystokinin
as chemical markers for specific GABAergic interneuron types in the rat
frontal cortex. J. Neurocytol. 31, 277–287 (2002).
34. Egorov, A.V., Hamam, B.N., Fransen, E., Hasselmo, M.E. & Alonso, A.A.
Graded persistent activity in entorhinal cortex neurons. Nature 420,
173–178 (2002).
35. Aston-Jones, G. & Cohen, J.D. An integrative theory of locus coeruleusnorepinephrine function: adaptive gain and optimal performance. Annu.
Rev. Neurosci. 28, 403–450 (2005).
36. Zucker, R.S. & Regehr, W.G. Short-term synaptic plasticity. Annu. Rev.
Physiol. 64, 355–405 (2002).
37. Masland, R.H. The neuronal organization of the retina. Neuron 76,
266–280 (2012).
38. Grimes, W.N., Zhang, J., Graydon, C.W., Kachar, B. & Diamond, J.S. Retinal
parallel processors: more than 100 independent microcircuits operate
within a single interneuron. Neuron 65, 873–885 (2010).
npg
© 2013 Nature America, Inc. All rights reserved.
historical perspectiveFOCUS ON MAPPING THE BRAIN
490 | VOL.10 NO.6 | JUNE 2013 | nature methods
39. Hosoya, T., Baccus, S.A. & Meister, M. Dynamic predictive coding by the
retina. Nature 436, 71–77 (2005).
40. Farrow, K. et al. Ambient illumination toggles a neuronal circuit
switch in the retina and visual perception at cone threshold. Neuron 78,
325–338 (2013).
41. Rivlin-Etzion, M., Wei, W. & Feller, M.B. Visual stimulation reverses
the directional preference of direction-selective retinal ganglion cells.
Neuron 76, 518–525 (2012).
42. Witkovsky, P. Dopamine and retinal function. Doc. Ophthalmol. 108, 17–40
(2004).
43. Ciocchi, S. et al. Encoding of conditioned fear in central amygdala
inhibitory circuits. Nature 468, 277–282 (2010).
44. Marder, E. & Taylor, A.L. Multiple models to capture the variability in
biological neurons and networks. Nat. Neurosci. 14, 133–138 (2011).
45. Milo, R. et al. Network motifs: simple building blocks of complex
networks. Science 298, 824–827 (2002).
46. Markram, H. The blue brain project. Nat. Rev. Neurosci. 7, 153–160
(2006).
47. Koch, C. & Reid, R.C. Neuroscience: Observatories of the mind.
Nature 483, 397–398 (2012).
48. Alivisatos, A.P. et al. The brain activity map project and the challenge
of functional connectomics. Neuron 74, 970–974 (2012).
49. Koch, C. Systems biology. Modular biological complexity. Science 337,
531–532 (2012).
50. Varshney, L.R., Chen, B.L., Paniagua, E., Hall, D.H. & Chklovskii, D.B.
Structural properties of the Caenorhabditis elegans neuronal network.
PLoS Comput. Biol. 7, e1001066 (2011).
focus on mapping the brain
commentary
Making sense of brain network data
Olaf Sporns
npg
© 2013 Nature America, Inc. All rights reserved.
New methods for mapping synaptic connections and recording neural signals generate rich and complex data
on the structure and dynamics of brain networks. Making sense of these data will require a concerted effort
directed at data analysis and reduction as well as computational modeling.
New empirical methodologies deliver an
ever increasing amount of data on patterns
of synaptic connectivity. For any given nervous system, the complete map of its neural
components and their synaptic interconnections corresponds to the connectome1. The
connectome delivers a description of structural brain connectivity, in popular parlance
often referred to as a ‘wiring diagram’. This
wiring diagram may describe brain connectivity at different scales, capturing synaptic
connections among single neurons or axonal
projections among brain regions. An important distinction must be made between such
connectome data and the highly variable and
dynamic patterns of neural activity and interactions in the connectome, also called ‘functional connectivity’.
Dynamic patterns of functional connectivity are distinct from connectome maps as
they are not based on synaptic connections
but on statistical measures derived from neuronal recordings, ranging from simple correlation to sophisticated inferences of causal
dependence. Despite these important differences between structural and functional
connectivity, both can be represented as networks comprising a set of discrete nodes and
edges (Fig. 1), corresponding to neural elements and their pairwise associations.
Creating accurate maps of brain networks
presents many methodological challenges.
For example, the construction of connectome
data from observations of micrometer-scale
anatomical patterns among individual neurons2 or from imaging of millimeter-scale
projections among neural populations or
regions 3 poses challenges regarding the
Olaf Sporns is in the Department of Psychological
and Brain Sciences, Indiana University, Bloomington,
Indiana, USA.
e-mail: [email protected]
precision, sensitivity and reproducibility
of network maps, partly because of various
methodological biases. Consequently, much
attention in the area of brain networks has
been devoted to addressing potential problems and pitfalls in data acquisition, regarding the reconstruction and inference of
connectional anatomy as well as the measurement of dynamic neural interactions.
Clearly, progress critically depends on methodological developments that increase the
completeness and accuracy of empirically
measured network maps.
But the scientific challenges with regard to
the study of brain networks do not end once
data are acquired. The sheer volume and
complexity of these data requires sophisticated new approaches to enable efficient
data integration and sharing via public neuroinformatics resources4,5. And another key
question looms on the horizon: how can we
use brain network data to gain new insight
into fundamental neurobiological structures
and processes? In this Commentary, I argue
that the shift toward an explicitly networkoriented approach to studying brain function
requires the development of new strategies
for the analysis and modeling of brain network data.
Shift toward networks
The rise of genomics at the end of the 20th
century ushered in a period of profound and
still ongoing change in the biomedical sciences. New technological developments such
as the capability to sequence whole genomes
and to create comprehensive inventories of
biomolecules and their chemical interactions have not only transformed the empirical study of biological systems, but also radically altered our conceptual understanding
of complex biological processes. Complex
Nodes
Hubs
Modules
Edges
Figure 1 | Simple network. Schematic illustrates
nodes and edges of the network, and its
community structure, with network modules and
hubs highlighted.
biological functions—once thought to
depend on the expression of single genes or
on the action of single molecules—are now
more commonly thought to depend on large
ensembles of biological components interacting in complex signaling, metabolic and
gene-regulatory networks.
The genomic revolution has given rise to
a new discipline, systems biology, dedicated
to unraveling the structure and dynamics of molecular and cellular networks that
underpin all aspects of biological function6.
Accomplishing this task requires tackling ‘big
data’ with new computational tools and theoretical models.
With the arrival of new experimental
methods for the comprehensive mapping of
neural structures and for large-scale recording of neural activity both in single cells and
in populations, neuroscience appears poised
to embark on its own ‘omics’ revolution7.
Connectomics has emerged as a major focus
of a new ‘systems biology of the brain’, centering on the study of anatomical networks
across multiple scales as well as the complex
nature methods | VOL.10 NO.6 | JUNE 2013 | 491
npg
© 2013 Nature America, Inc. All rights reserved.
focus on mapping the brain
dynamics these networks generate when
becoming active. In parallel to recent developments driven by modern genomics and
systems biology, the aim of connectomics is
to disclose the architecture of structural brain
networks and to explain the mechanisms by
which these structural networks shape and
constrain brain function.
New methods for mapping the layout and
operation of brain networks deliver huge
volumes of ‘big data’ that are shared in public
repositories and neuroinformatics databases.
These data pose fundamental challenges
for analysis and modeling, and will require
extensive data integration across multiple
domains of brain and behavior. Network
theory has become vitally important in this
endeavor. By relying on networks, connectomics follows a path similar to the one
already taken in systems biology, harnessing
some of the same opportunities as well as
encountering some of the same challenges.
Researchers in both fields continue to grapple
with methodological problems concerning
the accurate and complete mapping of system
interactions, the efficient representation and
sharing of big data sets, and the application of
large-scale computational models.
Analysis of brain network data
Brain networks are built from empirical
data that are extraordinarily rich and complex. There is ongoing debate concerning
the proper definition of network edges and
nodes. Structural brain network data usually
comprise multiple measures of connectional
anatomy, the neurobiological interpretation
of which is not straightforward, as is the
case for inferences of connection weight or
strength from synaptic morphology8, or from
axonal microstructure or myelination status9.
Functional brain network data derived from
time-series recordings of neuronal activity
can be expressed in many ways, from simply recording linear statistical dependencies (for example, Pearson correlations) to
sophisticated inferences on dynamic causal
interactions10. Which of these structural
and dynamic network measures best represents the underlying neurobiological reality
remains unclear11,12. Progress in this area is
important because the initial definition of
nodes and edges can have profound consequences for how brain networks are configured and interpreted.
Of particular importance for revealing the
organization of complex networks are methods for detection of network communities
or modules13 (Fig. 1). Generally speaking,
492 | VOL.10 NO.6 | JUNE 2013 | nature methods
modules correspond to sets of network elements that are densely linked with each other
and less so with other elements in the network (that is, other modules).
Module detection is an appealing strategy
for analyses of brain network data sets for a
host of reasons: (i) many studies have shown
that brain network modules correspond to
prominent functional subdivisions, for example, so-called resting-state networks reliably
encountered in spontaneous and task-evoked
brain dynamics14; (ii) the concept of modularity has validity and can be productively
applied across both structural and functional
networks, thus enabling comparisons across
the two domains; (iii) the topography and
relations among modules provide important
means for characterizing individual differences or developmental patterns in network
organization; (iv) modular decomposition
offers a low-dimensional description of
complex network data, which is useful for
purposes of data reduction and compression;
(v) modules help to define the roles of nodes
and edges in the global network topology, for
example, as connectors or bridges that crosslink different communities15; and (vi) modularity is thought to promote robust network
function16. Modularity may also become an
important tool for joining brain network data
across multiple scales. For example, modularity should allow the identification of structurally and functionally coherent populations
of neurons in cellular-scale data, which may
then be aggregated into larger-scale network
communities at regional and systems levels.
Establishing links between the connectome on the one side and the dynamic patterns of neural activity and the interactions
on the other requires relating data across
the two different domains of structural and
functional networks. To reiterate the difference between structural and functional
network data, edges in structural networks
refer to aspects of the physical infrastructure
of brain connectivity, that is, synaptic connections jointly comprising the ‘wiring diagram’, whereas edges in functional networks
reflect aspects of statistical dependencies
among neuronal time series, corresponding
to simple correlation or covariance or more
sophisticated measures of nonlinear coupling
or causal dependence.
These differences in edge definition entail
clear differences in the way structural and
functional network data should be analyzed
and interpreted. For example, communication processes governing the exchange
of information along connections do not
commentary
occur in functional networks; instead functional networks are a manifestation of these
processes as they unfold in the structural
connectome. Therefore, studies aiming to
identify important network nodes (so-called
‘hubs’) or to probe the robustness of a network after lesion (involving deletion of nodes
and/or edges) are more appropriately carried
out on structural networks than on functional networks. Whereas the identification
of hubs in structural networks is relatively
straightforward, it remains somewhat ambiguous in the context of functional networks17.
Brain networks are dynamic and change
across time, on both long time scales (for
example, plasticity and remodeling during
development and learning) as well as short
time scales (for example, fast remodeling of
dendritic spines in structural connectivity or
rapid edge dynamics in functional connectivity, the latter in part resulting from neuromodulation18). Hence, capturing characteristic features of how networks change through
time will be as important as characterizing
the structure of nodes and edges at each
instant in time. In this area, there is currently
a dearth of data-driven approaches, in part
because network studies in other disciplines
(social networks, protein networks and others) are only just beginning to explicitly take
network dynamics into account. The problem is particularly pressing in the analysis of
functional networks where the topology and
strengths of dynamic interactions are continually and rapidly modulated by endogenous
changes in state and exogenous changes in
input and task. Studies of the human brain
with electrophysiological or imaging methods consistently show that functional brain
networks rapidly reconfigure with changes
in the momentary demands of the environment19 and exhibit nonstationarity during
‘rest’20. Development of methods to characterize these network dynamics is becoming
critical in the area of data representation and
analysis.
Recently, far-ranging proposals for
cataloging all functional activity maps a
given brain can generate (referred to as
the brain’s ‘functional connectome’) have
been made 21 , culminating in the Brain
Research through Advancing Innovative
Neurotechnologies (BRAIN) initiative
announced in April 2013 by President
Barack Obama. Brain networks are central to achieving the goals of this ambitious
plan. Dynamic recordings of functional
activity maps can naturally and efficiently
be represented in the form of functional
npg
© 2013 Nature America, Inc. All rights reserved.
commentary
networks. Such networks, especially if
recorded at single-cell and millisecond resolution, will be highly dynamic and variable, thus posing difficult computational
and analytic challenges.
To address these challenges, the BRAIN
initiative must include intense efforts directed not only at data analysis and modeling but
also toward extending fundamental network
science and theory to the brain. Such efforts
have proven productive in other areas of neuroscience. Catalogs of functional activation
maps recorded with noninvasive neuroimaging methods in the human brain are available
as public repositories that can be data-mined
and analyzed online22,23. These databases
have enabled important insights regarding
the link between functional activations, networks of coactivation patterns and functional
connectivity. And these functional network
maps have been linked to various domains
of behavior and cognition.
Growing role of computational models
One of the key rationales behind studies of
protein networks is that patterns of proteinprotein interactions can be predictive of a
protein’s functional roles as well as its involvement in disease states24. The notion that
structure can predict function also informs
one of the chief goals of connectomics, which
is to furnish network models that can bridge
brain structure and function1,7. A large body
of evidence suggests that neuronal structure, and especially connectional anatomy,
is crucial for endowing neural elements with
specific functional properties. This structurefunction relationship holds across scales and
species, and it can even predict individual
variations in brain responses and behavior25.
The structure-function relationship opens
a crucial role for computational models that
build on structural data to predict brain
dynamics and associated functional attributes. Typically, such models consist of neuronal units whose dynamics are described by
differential equations based on biophysics
(for example, ionic membrane conductance).
These neuronal units are then interconnected
via a structural network of synaptic links that
allows units to communicate and influence
each other’s dynamic state. Neural communication may involve noisy signal transmission and conduction delays, and neural
activity is typically tracked by measuring
activation level, firing rate or membrane
potential across time. Hence, such models
deliver time-series data that can be analyzed
using methods very similar to those used for
focus on mapping the brain
the analysis of empirically measured brain
dynamics. Connectome data on structural
connectivity can supply the coupling matrix
for such computational models, and initial
results suggest that these models can partly
reproduce the spatial pattern and temporal
dynamics of empirically observed functional
networks26. First applications of such computational network models have already
provided considerable theoretical insight into
the network basis of functional connectivity
in the resting brain27.
As such computational models become
more realistic and sophisticated, they may
also become important tools for explaining
empirical observations and predicting new
findings. A highly ambitious project to build
a working model of a human brain incorporating cellular and subcellular detail has
commenced28, and structural connectome
data coming from both micro and macro
scales will likely be an important ingredient.
A parallel effort to construct a ‘virtual brain’
builds on whole-brain connectome data sets
to generate complex neural dynamics whose
spatiotemporal signatures can be compared
against empirical brain recordings29, with the
explicit aim to create a computational tool
that can be of clinical benefit. Simulations of
brain activity in human patients, informed by
mapping their individual connectome, may
be just over the horizon.
The confluence of connectome mapping
and computational modeling once again
parallels ongoing developments in genomics
and systems biology, where computational
models have become an integral component of many research projects. The idea
that computer models can help understand
the workings of a cell has a long history, and
the availability of comprehensive data on cell
structure, genes and gene products has now
put the construction of a working model of a
cell within reach30.
As is the case with the virtual cell, building
a virtual brain will open a wealth of new possibilities for testing and manipulating brain
networks at all levels and probing for architectural features that are necessary for efficient and flexible network function. As the
emerging field of connectomics continues to
mature, computational models will become
indispensable tools for guiding empirical
studies of brain networks.
Conclusion
The promise of concerted efforts to map the
brain’s structural and functional connections
is that a renewed focus on the brain as a com-
plex networked system can deliver fundamentally new insights and knowledge. This
promise can only be fulfilled with the help of
new approaches to data analysis and reduction, driven by network theory and computational modeling. As increasing amounts of
brain network data become available, there
will be a growing need for neuroscientists to
work more closely with colleagues in other
disciplines who are tackling complex technological, social and biological systems.
The adoption of a network perspective, the
development of new ways to analyze and
describe brain connectivity, and the deployment of sophisticated integrative models are
all important steps toward making sense of
brain network data.
ACKNOWLEDGMENTS
Supported by the J.S. McDonnell Foundation.
COMPETING FINANCIAL INTERESTS
The author declares no competing financial interests.
1. Sporns, O., Tononi, G. & Kötter, R. PLoS Comput.
Biol. 1, 245–251 (2005).
2. Denk, W., Briggman, K.L. & Helmstadter, M. Nat.
Rev. Neurosci. 13, 351–358 (2012).
3. Van Essen, D.C. & Ugurbil, K. Neuroimage 62,
1299–1310 (2012).
4. Akil, H., Martone, M.E. & van Essen, D.C. Science
331, 708–712 (2011).
5. Marcus, D.S. et al. Front. Neuroinform. 5, 4 (2011).
6. Hood, L. et al. Science 306, 640–643 (2004).
7. Sporns, O. Discovering the Human Connectome.
(MIT Press, 2012).
8. Bourne, J.N. & Harris, K.M. Curr. Opin. Neurobiol.
22, 372–382 (2012).
9. Jbabdi, S. & Johansen-Berg, H. Brain Connect. 1,
169–183 (2011).
10. Friston, K.J. Brain Connect. 1, 13–36 (2011).
11. Smith, S.M. Neuroimage 62, 1257–1266 (2012).
12. Wig, G.S., Schlaggar, B.L. & Petersen, S.E. Ann. NY
Acad. Sci. 1224, 126–146 (2011).
13. Fortunato, S. Phys. Rep. 486, 75–174 (2010).
14. Power, J.D. et al. Neuron 72, 665–678 (2011).
15. Guimerà, R. & Amaral, L.A.N. Nature 433, 895–
900 (2005).
16. Kashtan, N. & Alon, U. Proc. Natl. Acad. Sci. USA
102, 13773–13778 (2005).
17. Zuo, X.N. et al. Cereb. Cortex 22, 1862–1875
(2012).
18. Bargmann, C.I. Bioessays 34, 458–465 (2012).
19. Bassett, D.S. et al. Proc. Natl. Acad. Sci. USA 108,
7641–7646 (2011).
20. Jones, D.T. et al. PLoS ONE 7, e39731 (2012).
21. Alivisatos, A.P. et al. Neuron 74, 970–974 (2012).
22. Fox, P.T. et al. Hum. Brain Mapp. 25, 185–198
(2005).
23. Yarkoni, T. et al. Nat. Methods 8, 665–670 (2011).
24. Vazquez, A. et al. Nat. Biotechnol. 21, 697–700
(2003).
25. Saygin, Z.M. et al. Nat. Neurosci. 15, 321–327
(2012).
26. Adachi, Y. et al. Cereb. Cortex 22, 1586–1592
(2012).
27. Deco, G., Jirsa, V.K. & McIntosh, A.R. Nat. Rev.
Neurosci. 12, 43–56 (2011).
28. Markram, H. Sci. Am. 306, 50–55 (2012).
29. Jirsa, V.K. et al. Arch. Ital. Biol. 148, 189–205
(2010).
30. Karr, J.R. et al. Cell 150, 389–401 (2012).
nature methods | VOL.10 NO.6 | JUNE 2013 | 493
commentary
FOCUS ON MAPPING THE BRAIN
Why not connectomics?
Joshua L Morgan & Jeff W Lichtman
npg
© 2013 Nature America, Inc. All rights reserved.
Opinions diverge on whether mapping the synaptic connectivity of the brain is a good idea. Here we argue
that albeit their limitations, such maps will reveal essential characteristics of neural circuits that would
otherwise be inaccessible.
Neuroscience is in its heyday: large initiatives in Europe1 and the United States2 are
getting under way. The annual neuroscience
meetings, with their tens of thousands of
attendees, feel like small cities. The number of neuroscience research papers published more than doubled over the last two
decades, overtaking those from fields such
as biochemistry, molecular biology and cell
biology3. Given the number of momentous
advances we hear about, surely must we not
be on the threshold of knowing how the
healthy brain works and how to fix it when
it does not?
Alas, the brain remains a tough nut to
crack. No other organ system is associated with as long a list of incurable diseases.
Worse still, for many common nervous system illnesses there is not only no cure but no
clear idea of what is wrong. Few psychiatric
illnesses, learning disorders or even severe
pain syndromes such as migraine have
pathognomonic signs: no blood tests, radiographic or electrophysiological findings, or
even brain biopsies enable diagnoses. This
predicament is unlike that for other organ
systems, where disease is nearly always
associated with tissue pathological and/or
biochemical signs. A patient may come for
help because of pain in their belly, but the
physician discovers the cellular or molecular cause of the pain before a treatment regimen begins. Why is this so different for the
brain?
Those of us with an interest in a deeper
understanding of the structure of the brain,
Joshua L. Morgan and Jeff W. Lichtman are in the
Department of Molecular and Cellular Biology and
The Center for Brain Science, Harvard University,
Cambridge, Massachusetts, USA.
e-mail: [email protected] or
[email protected]
494 | VOL.10 NO.6 | JUNE 2013 | nature methods
not surprisingly, see the problem in structural terms. The nervous system is a physical tissue that is quite unlike other organ
systems4. For one thing, it contains far more
types of cells. The retina alone has more
than 50 different kinds of cells5, whereas the
liver has closer to five. Second, neurons, by
virtue of their complicated geometry, connect (via synapses) to many more cellular
partners than the limited associations of
immediately adjacent cells in other organs.
Third, the neuronal contacts give rise to
directional circuits that have no analogs in
other tissues. Fourth, the fine structure of
these neural circuits is quite diverse and differs from the marked structural redundancy
in other organs where the same multicellular motif (for example, the renal nephron)
is iterated many times. Fifth, perhaps the
most intriguing difference between neural
tissue and other organs is that the cellular
structure of neural tissue is a product of
both genetic instruction and experience.
Thus, the structure of each of our nervous
systems is personalized by our own set of
experiences.
The special structural features of the
nervous system are likely the reason why it
is more difficult to understand than other
organs are. We contend, however, that a more
complete rendering of neural circuit synaptic
connectivity (that is, connectomics6) would
go a long way to solving this problem. With
this understanding, diseases that manifest as
abnormalities of behavior, thought or learning, or as pain might become as clearly linked
to underlying pathological structure, as is
the case for diseases in other organ systems.
Knowing what is wrong is a good first step to
finding a solution.
For some proponents of connectomics,
the potential clinical payoffs provide suf-
ficient justification for such studies. But
not everyone agrees. First, connectomics
can require an industrialized effort that is
akin to initiatives that allowed genomics
to flourish. In light of present severe budget limitations, connectomics might be
ill-advised. Moreover, some have argued
that pursuing connectomics would be a
waste of money, even if it were free. They
have argued that anatomical maps fundamentally do not reveal how the brain
works. Below we address some of these
arguments.
Top ten arguments against connectomics
Number ten: circuit structure is different
from circuit function. One argument put
forward is that the nervous system’s macroscopic functions (that is, behaviors) are
derived directly from the functional (that
is, electrical) properties of the neurons
rather than the anatomical connections
between neurons. Hence it is the relationship between the firing patterns of action
potentials of neurons and function that is
the key to bridge the gap between the cells
of the nervous system and behavior. The
recent proposal of an initiative to get a ‘brain
activity map’ reflects this view, as do many
initiatives for Ca2+ imaging from ever larger
numbers of nerve cells.
Although no one doubts that neural
connections underlie signaling between
nerve cells, the focus on neuronal function
(action potentials) is based in part on the
belief that the structural wiring details per
se are insufficient to derive the firing patterns. Part of the problem is that the spiking properties of a neuron come not only
from the electrical signals they receive from
their presynaptic partners but also from
the distribution and characteristics of their
FOCUS ON MAPPING THE BRAIN npg
© 2013 Nature America, Inc. All rights reserved.
intrinsic voltage-gated channels. Although
in principle connectomics can reveal all the
presynaptic partners, distributions of ion
channels would not be easily revealed by a
connectional map. Furthermore, even if one
knew the molecular identities and precise
location of all the voltage-gated channels in
every neuron (determined by very sophisticated immunological labeling, for example),
this information would still not be sufficient
to provide the activity patterns underlying behavior. The reason is that behavior
is often driven by the particular trains of
action potentials set up by sensory inputs.
This sensory experience is extrinsic to the
nervous system and hence inaccessible from
just looking at a connectional map.
Response. Brains can encode experiences
and learned skills in a form that persists for
decades or longer. The physical instantiation of such stable traces of activity is not
known, but it seems likely to us that they
are embodied in the same way intrinsic
behaviors (such as reflexes) are: that is, in
the specific pattern of connections between
nerve cells. In this view, experience alters
connections between nerve cells to record
a memory for later recall. Both the sensory
experience that lays down a memory and its
later recall are indeed trains of action potentials, but in-between, and persisting for long
periods, is a stable physical structural entity
that holds that memory. In this sense, a map
of all the things the brain has put to memory
is found in the structure—the connectional
map. An ‘activity map’ of the brain that only
shows trains of action potentials would certainly be an incomplete map, as most behaviors and memories will not be visible in any
finite recording session. Decoding the way
experience via electrical activity becomes
stably embedded in physical neuronal networks is the unmet challenge that connectomics attempts to solve.
In our view, more challenging perhaps
will be ferreting out the relationships
between neural circuits and behaviors that
are intrinsic and unlearned. In most animals, the behavioral repertoire is mostly
genetically encoded. Such inherited circuits
may have arbitrary features that came about
accidentally at some instant in an animal’s
evolution. We therefore think that genetically encoded behaviors could be instantiated in quite diverse ways compared to a
more limited number of rules governing
experience-dependent formation of circuits.
From this perspective, the structural under-
pinnings of a behavior of Caenorhabditis
elegans may actually be more difficult to
decode than the circuit underlying a learned
behavior in a mammal.
Number nine: signals without synapses
and synapses without signals. There is
abundant evidence that neurotransmitter released at a synapse is not always
restricted to the ‘intended’ postsynaptic
target. Spillover of glutamate from excitatory synapses has been shown to affect
nearby postsynaptic sites7. It is also well
known that neuronal activity can affect
the behavior of nearby glial cells, which
can then convey this activity to other glial
cells through junctions between glia. In
all brains, there are also many signals that
pass between brain cells and other organ
systems via hormones. Steroids originating in the adrenal cortex have effects on
brain function as do those from sex organs.
Growth hormone, thyroid hormone, insulin, leptin and many others also affect brain
function. Furthermore, some forms of neurotransmitter release do not rely on classical synapses8. Lastly, neurons often release
peptides that act over large areas, using volume transmission as opposed to restricted
transmission at adjacent synapses9. All of
the aforementioned examples show that
brain function is profoundly affected by
chemical cues in ways that would be difficult or impossible to infer from anatomical
wiring diagrams.
Conversely, it is generally accepted that a
substantial fraction of excitatory synapses
can be structurally present but functionally
silent10. These synapses can be switched on
via a Hebbian learning step. Obviously in
the absence of knowing which synapses are
silent, a connectome provides a distorted
picture of the functionally useful connections in the brain.
Response. It is true that that a map of synaptic connectivity is not identical to a map
of the signaling pathways of the brain. In
terms of signals without synapses, there is
no denying the importance of the brain’s
chemical milieu on behavior. The ability of
pharmacological agents to rapidly induce
sleep, tranquility, excitement, hallucinations
and so on means that the behavioral state
can be dramatically altered probably without any modification to the connectome.
Therefore, we should be careful not to
confuse the goals of connectomics with
the aims of maps of the activity patterns
commentary
in a nervous system. The connectome certainly would not be a map of the behavioral
state at the moment the brain is preserved.
We think the connectome would be much
more than that, as it could provide the
entire behavioral capability of the brain. But
extracting such information may require
detailing an unprecedented amount of anatomical data. This anatomically stringent
view of connectomics contrasts with interesting recent ideas about determining the
connections between brain cells without
information about anatomy or physiology11.
A connectional matrix minus the anatomy
information, of course, can reveal neither
the sites of spillover nor the proximities necessary for the spread of peptides, whereas
an anatomical connectome could reveal
less direct paths for neurotransmitter and
neuropeptide signaling. However, the hormonal milieu via blood flow or cerebrospinal fluid would still remain invisible unless
the effects of diffusible factors on behavior
had a structural correlate12.
There are also anatomical approaches to
deal with the problem of synapses without
signals. Some of the earliest evidence for
a class of silent synapses came from serial
electron microscopy reconstructions13, and
we imagine that such information might
become available for glutaminergic synapses as well. If silent synapses are structurally
different from transmitting ones by virtue
of the neurotransmitter receptor, then in
principle that difference can be revealed.
Immunolabeling for AMPA-type glutamate
receptors should solve this problem.
Number eight: ‘junk’ synapses. Given the
trillions of synapses in a human brain, most
individual synapses are functionally negligible. The brain is probably organized such
that the precise number and identity of synapses are unimportant, and what really matters is the general likelihood of connectivity.
Thus, some fraction of the synaptic connections could rightly be called ‘junk’: they have
no functional role but have so little cost that
such scraps persist. If junk synapses are common, then going to the trouble of itemizing
all the synaptic connections is both a waste
of time and a distraction from more important functional questions.
Response. Our view is that it is premature
to assign an influence to individual synapses in wiring diagrams when so few complete
wiring diagrams have been described. The
lesson from genomics is that many noncodnature methods | VOL.10 NO.6 | JUNE 2013 | 495
commentary
npg
© 2013 Nature America, Inc. All rights reserved.
ing sequences that initially were thought to
be unimportant and sometimes called junk
turned out to have functional importance.
Moreover, we are not convinced that the
nervous system tolerates large numbers of
synapses that are not serving a useful function. In fact, the synapses observed in the
adult may reflect the small percentage of
synapses that survive an extended period of
synapse elimination during development14.
Number seven: same structure, many
functions. Connectomics might be tractable if each behavior was partitioned to a different circuit element (that is, one circuit,
one behavior). However, good evidence
shows that multiple sensations or behaviors
use the same neurons in different ways15.
Neurons can rapidly switch their functional
roles in response to chemical signals such
as peptides, other neuromodulators or
activity levels. Because of this flexibility, it
is really not possible to assign a neuron to
a function without knowing the behavioral
state—something that is invisible in connectomics.
Response. Without an understanding of
the physiological responses of neurons to
various chemical messengers such as peptides, we agree it will be difficult to define
the roles of neurons that switch their function depending on the chemical milieu. The
extent to which this kind of switching is a
general feature of nervous systems is not yet
known.
But whatever the case, it is important to
emphasize one limitation of connectomics:
it is not a replacement for insights gained
by physiological or pharmacological studies. Rather, connectomics may associate
specific physiological phenomena with
specific neural-circuit motifs so that the
next time that motif is observed in the
same tissue, it will signify a physiological
process without the requirement of repeating the physiological analysis each time.
Because in many nervous systems the same
neuron types are duplicated many thousands or millions of times, it seems likely
that the same motifs will be repeated multiple times. Thus, a little physiology may
go a long way.
Number six: same function, many structures What if there is great variability in the
circuits that give rise to a single behavioral
output? Without a stereotyped pattern of
neural circuits that underlie a particu496 | VOL.10 NO.6 | JUNE 2013 | nature methods
FOCUS ON MAPPING THE BRAIN
a
33% connected
b
Figure 1 | Potential results from two approaches
to studying circuit connectivity. (a) Probing
pairs of neurons in multiple subjects determines
the probability that neighboring neurons are
connected. (b) Connectome of the same tissue
reveals network motifs.
lar function, it will be impossible to relate
structure to function.
Response. In mammals, at least, variability seems to be the rule. Even the pattern
of nerve-muscle connections in the same
muscle is quite different from one instantiation to the next16. However, variability in
connections does not necessarily mean that
the result is incomprehensible. In muscle,
for example, there is a skewed distribution
of the size of motor units that is the same
in each muscle even though the exact location and branching pattern of each axon is
unique. Thus, each instantiation appears to
be a variation on the same common theme,
just as every chess match is different, but
they all obey the same rules. One of the reasons we believe connectomics is necessary,
is that it is the only way to derive network
principles despite variability in the connectivity of individual neurons.
Indeed, deriving network principles
without connectomics may be nearly
impossible in some cases. Without connectomics, neural circuit diagrams can only be
assembled by identifying connected pairs
of cells in many different subjects. This
approach, although accurate as far as it
goes, cannot reveal wiring rules as simple
as ‘cell A only connects to cell C when cell
B connects to cell C’. Such conditional rules
have been revealed from fortuitous circuits that lend themselves to easy analysis.
For example, triad synapses in the lateral
geniculate nucleus allow the connectivity
of cell A, B and C to all be seen in a single
electron microscopy section as all the synapses are adjacent17. If more complicated
patterns of conditional connectivity exist
(as well they might18,19), then there is no
way to find these without reconstructing
the entire circuit and including all the cellular elements (Fig. 1).
Number five: statistical synapses should
suffice. Statistical mechanics was a great
advance in physics: when many similar
events occur at the same time (such as the
collisions of molecules in a gas), the behavior of the system can be predicted with
accuracy without ascertaining the behavior
of each and every particle’s trajectory. Given
the trillions of synapses in a brain, is there
really any alternative except to take a statistical approach to synaptic connectivity?
Members of the European Union’s initiative on Future and Emerging Technologies
have decided, for example, to fund the large
‘Human Brain Project’. In this case, the
connectional associations will be determined by characterizing “...the morphologies of different types of neuron present in
different regions of the human brain.
Combined with modeling, the results
would enable the project to predict a large
proportion of the short-range connectivity between neurons, without measuring
the connectivity experimentally”1. That is,
they will create the wiring diagram without
having to go to the trouble of obtaining an
anatomic connectome.
Response. Because of the heterogeneous
nature of brain tissue, statistical approaches to studying neural-circuit wiring are, of
course, much more difficult than those for
a homogeneous gas. For a statistical model
to capture how brains actually generate
behavior, ideally it would need to include
all the neuronal connectivity motifs.
Unfortunately, the number of ways the
dozens of different cell types are interconnected, and especially how these connections are contingent on the full cohort of
synaptic partners of each cell, is presently
unknown. But the advocates of statistical
approaches to studying connectivity are
not unduly concerned because they believe
that models that combine a limited amount
of connectivity data with the right set of
FOCUS ON MAPPING THE BRAIN npg
© 2013 Nature America, Inc. All rights reserved.
learning rules will produce fully functional
neural circuits.
We, however, think the only way to know
whether the results of this strategy are actually consistent with a biological brain, is to
compare the predicted wiring to an actual
connectome. This is why we submit that
‘measuring the connectivity experimentally’ is a good idea: it provides all the circuit
motifs; with analysis, it may also provide the
physical instantiation of the learning rules
(see below) and is a direct path to building a
working model of the brain.
Number four: the mind is no match for
the complexity of the brain. Might it be
the case that ‘connectomisists’ have bit off
more than they can chew? Might it be that
the brain’s structural complexity far outstrips the complexity of the thoughts that
emanate from even the best and the brightest brains? Although some may hope that
there are organizing principles and regularities that will permit substantial compression of the connectional information
into a simpler framework, is there really
any biological reason such regularities
are inevitable? Could the most succinct
description that embodies all that a brain
does be the brain itself? If so, then there
is little point in mapping the connections
in great detail because in the best case,
one would end up with a description that
would be as complicated and intractable as
the brain itself.
Response. The proximate goal of connectomics is to generate detailed renderings of
the connections between neurons in large
volumes of the brain. The purposes that
such connectional maps could be put to are
numerous. Comparisons between healthy
and diseased brains might point to the physical underpinnings of psychiatric diseases.
Comparisons between young and old brains
might provide an understanding of what
kinds of network changes are associated
with development and aging. Comparisons
between human and other primate brains
may provide insight into what underlies
intelligence. So even if ‘understanding’ the
brain is not likely to be an early triumph of
connectomics, there are potentially many
fundamental things that this kind of data
will reveal. To put it perhaps a bit too bluntly, we think that understanding big data,
may be overrated as a goal. Connectomes
can immediately provide insights even if
understanding lags.
Number three: merely descriptive neuroanatomy, just more expensive. The only
hope for curing diseases and uncovering the ways brains work is to get insights
into underlying mechanisms. Mechanistic
insights come from experiments that test
hypotheses by manipulating variables.
There is of course a time and a place for
descriptive studies in a field, and neuroscience owes Cajal a great debt for his landmark description of the cellular underpinnings of neural tissue, but the time for mere
description is long past. Indeed, connectomics sounds modern but is really just a
throwback: it is neuroanatomy gussied up
with a fancy new word and ultra-expensive
machines. Connectomics harkens to a time
when description was all we could do. We
can do more now.
Response. The Hubble telescope, archaeology, the human genome and much of
structural biology are also merely descriptive, but few would argue that they have
not provided fundamental insights. The
hope of descriptive approaches is that they
provide specific data that lead us to general
hypotheses (inductive reasoning). They can
reveal associations and frameworks that
were not readily apparent before. In a variety of fields, big-data initiatives are being
used to challenge the sacrosanct ‘hypothesis
first’ world view of scientific investigation.
Number two: not much was learned from
the connectomes we have. The 10-year
effort to generate the connectome of
C. elegans20 is widely cited, but it has had
less utility than had been imagined originally. No one can claim that the relationship
between structure and function has been
settled in this very small nervous system. In
some respects, connectomes distract from
more mechanistic analyses because they
reveal many more synaptic interconnections
between neurons than would seem necessary. Given the struggle to make sense of the
mind of a worm with about 300 neurons, it
is very unlikely that we will get anywhere
trying to fathom a mammalian nervous
system that could be roughly a billion times
bigger.
Speaking of big brains, there is already a
big Human Connectome project21, so why
bother with maps of mice and flies?
Response. At the very least the worm connectome has been helpful in constraining
circuit analysis by providing for each neu-
commentary
ron the list of upstream and downstream
neurons. This is no doubt beneficial, having been the starting point for understanding the neural pathways for touch-induced
movements22, for example. But the worm
connectome may not be as useful as initially imagined because of its surprising complexity. The high level of interconnectivity
among the 300 neurons—a revelation in
itself—does not lend itself to easy analysis.
The interconnectedness of C. elegans neurons potentially provides for a much more
adaptable and diverse behavioral repertoire
than one with a simpler wiring diagram but
at the expense of easy comprehension by
humans.
A particular technical limitation of the
worm’s connectome is that its inhibitory
and excitatory connections cannot be differentiated in electron microscopy images
(unlike the situation for mammals), reducing the overall power of the wiring diagram
to reveal new concepts. In addition, there
are particular challenges in interpreting a
map of connectivity in a highly differentiated nervous system where each neuron has
many subcellular compartments, where all
the signaling is due to local potentials, and
evolution has specialized nearly every cell
into a unique type. This specification contrasts to the millions of cells of a single class
found in many parts of the (perhaps less
evolved) vertebrate nervous system. Thus,
the worm’s nervous system may use a fundamentally different strategy than the one
predominating in much larger nervous systems in which cell types reflect populations
of neurons whose connectivity is organized
by experience.
As to the Human Connectome project,
this is an important and ambitious multiinstitution study to gain information about
the organization of the brains of many individuals including many twins. One of its
goals is to map the pathways that project
between various brain regions (‘projectomics’). However, none of its many goals
relate to describing the synaptic connectivity of the brain.
Number one. “If you want to Google on
connectome and look [it] up, you can see
some gorgeous pictures that are being made
of the wiring diagram of the human brain,
showing you how the wires move from front
to back and up and down, side to side. But
again, it’s a static picture. It’d be like, you
know, taking your laptop and prying the
top off and staring at the parts inside, you’d
nature methods | VOL.10 NO.6 | JUNE 2013 | 497
commentary
npg
© 2013 Nature America, Inc. All rights reserved.
be able to say, yeah, this is connected to
that, but you wouldn’t know how it worked”
(Francis Collins, NPR Science Friday;
5 April 2013).
Response. Maybe picking an argument in
this case is not so wise, but prying the lid
off a computer to see how everything is
connected does not seem like such a bad
idea, at least as a first step. Without knowing the parts list and what is connected to
what, how could one ever really know how
it works?
In our view, the ideal brain-imaging
technology would provide both a complete
map of synaptic connectivity as well as a
complete map of the activity of all neurons
and synapses in real time during normal
behaviors. Even better, would be to do this
in a human being who can report on their
thoughts while behaving. Unfortunately, we
are a long way from such technologies; so
what do we do in the meantime?
We think the computer analogy gives us
a lead. The computer can be turned off and
on without losing much data because the
instructions that make it work are embedded in its ‘static’ physicality. A deep understanding of how information is stably stored
in the structure of hard discs, the input and
output wires of each chip, the physical structures that explain the working of those chips
and so on would be enormously helpful in
making sense of a computer. Might the same
be said of nervous systems?
We think that the static connectivity of
the brain has embedded within it much
of the ‘instruction book’ that guides the
billions of impulses through networks of
neurons to ultimately generate outward
behavior. We see a connection between
the relationship between the connectome
and the brain’s functional properties to not
only computers but also to the way the static
human genome encodes much of how an
organism works.
Structural defects in the genome that
give rise to a range of diseases are perhaps
a good analogy for defects in wiring that
give rise to diseases of brain function.
In each case, tracing the causal linkages
between the static structure (be it genotype
or connectotype) and the ultimate phenotype (be it cancer or schizophrenia) is difficult. But this difficulty has not deterred
cancer biologists from seeking the ultimate causes (the physical genes) that affect
the likelihood of cancer, and we think
should not deter neuroscientists from
498 | VOL.10 NO.6 | JUNE 2013 | nature methods
FOCUS ON MAPPING THE BRAIN
a
b
Figure 2 | Two types of wiring diagrams. (a) Schematic of mammalian retina drawn by
Santiago Ramón y Cajal in 1901. Image courtesy of J. De Carlos. (b) Dense connectivity of the
Twitter network of David Rodrigues as rendered by D. Rodrigues (http://www.flickr.com/photos/
dr/2048034334/).
seeking the physical wiring abnormalities
that underlie brain disorders, even if this
requires high resolution connectomics.
In both situations, the structural data is
overwhelming but nonetheless holds fundamental truths.
To quote Francis Collins referring to
the human genome project: “When you
have for the first time in front of you this
3.1-billion-letter instruction book that
conveys all kinds of information and all
kinds of mystery about humankind, you
can’t survey that going through page after
page without a sense of awe. I can’t help
but look at those pages and have a vague
sense that this is giving me a glimpse of
God’s mind”23. We likewise feel that the
static maps of the brain may engender awe
and it might not be so unrealistic to hope
that in staring into such a map we might
get a glimpse of the human mind.
Why we still want to do connectomics
The large (but we are sure still incomplete)
list of arguments above might seem discouraging enough to put a damper on all but the
most irrationally enthusiastic proponents of
connectomics. In addition to the reasons we
provided in our rebuttals, we remain committed to pursuing connectomics because
of the absence of one fundamental kind of
information about nervous systems that we
do not think is attainable in any other way.
What Cajal did not do for neuroscience.
Thanks to both his genius and the extraordinary power of the sparse labeling of the
Golgi stain, Cajal developed a view about
neurons that has had a dominating effect on
the way in which we imagine neural circuits
to work. Schematic textbook diagrams of
retinal circuits, cortical circuits, cerebellar
circuits, among many others, provide stu-
dents with a sense of which cell types connect with one another. These diagrams also
serve as a model of how information flows
through a circuit of cells because many of
these diagrams include arrows that harken
back to the arrows found in the original
Cajal diagrams (Fig. 2a).
The sense one might take from these
kinds of diagrams is that this kind of structural class-wise connectivity information
explains the way a neural circuit works.
One may quibble over the extent to which
the details of these connectional maps have
been worked out, but the implication is that
knowing this kind of neuronal-class connectivity is sufficient to get some understanding about how neurons process information.
In certain cases, this view is probably
correct. In an animal in which the majority of neurons are each genetically unique
and each has a stereotyped connectivity,
connections of individual neurons are the
same thing as a class-wise wiring diagram
(because to a first approximation each cell
is its own class). Thus, one can imagine a
detailed map describing exactly which
cells are connected, and that map would be
exactly the same as what actually exists in
the animal.
But in other cases these diagrams are
quite different from the actual connectional array. They lack both quantitative
information and ignore the different wiring patterns of the cells of the same class.
Thus, a retinal schematic diagram shows
what classes of cells are interconnected
but does not show critical details such as
how many different amacrine cells converge on each retinal ganglion cell or which
particular amacrine cells are making those
connections. A cortical schematic diagram
does not show whether two interconnected
npg
© 2013 Nature America, Inc. All rights reserved.
FOCUS ON MAPPING THE BRAIN pyramidal cells have a common third input.
A cerebellar schematic diagram does not
show whether the cohort of parallel fibers
connected to a given Purkinje cell predicts
which parallel fibers will connect with adjacent Purkinje cells.
This kind of information is, however,
being addressed by new generation of
connectomic circuit maps that reveal connectivity between dozens or hundreds
of cells in a single piece of tissue17,24–26.
These new efforts and several earlier
attempts 20,27, are beginning to shift the
paradigm from class-wise connectivity questions to questions about the particular pattern of connections in a cell
class. Variability in the connections of
cells of the same class, is the essential
feature that is lacking in classical wiring
diagrams but that exists in actual neural
circuits. Because the vertebrate brain is
mostly composed of many copies of each
of many cell types (unique neurons such
as the Mauthner cell in teleosts are quite
exceptional), this aspect of neural circuit
organization cannot be ignored. In sum,
Cajal’s use of sparse labeling did not provide a means for coming to grips with the
way redundant populations are used in circuits. Connectomics, however, does.
Engramics? In 1904, Richard Semon, a
German evolutionary biologist, coined
the term ‘engram’ as the physical memory trace that is somehow embedded in
an organism after an experience. He was
motivated to think about the physicality
of memory in an attempt to formulate a
means of inheritance of acquired characteristics 28. Although this Lamarckian
notion is refuted for gene-based inheritance, there is little doubt that humans do
acquire information (by learning) during
their lives that affects their behaviors and
then pass this information on to their children (among others) to alter their behaviors in nongenetic ways. A longstanding
goal of neuroscience is to determine the
physical basis of an engram.
An interesting feature of the highly
redundant populations of neurons in
mammalian nervous systems is that they
undergo dramatic changes in their connectivity in early postnatal life. Because these
alterations are activity-dependent and
transform diffusely connected networks
into many distinct subnetworks, this process could be the physical underpinnings
of memory14.
Regardless of the mechanism, we
wonder whether eventually a subfield of
neuroscience will exist that is devoted to
understanding the encoding and decoding of experience-based changes in neural networks. Engramics would certainly
require a hefty amount of connectomics
along with sophisticated analysis and ultimately simulations to test the idea that a
particular neural network encodes a particular memory.
What is ahead for connectomics?
Assuming that connectomics has a future,
several mostly technical obstacles will need
attention.
Statistical and analytic tools. In many
ways, maps of complete connectomes might
look less like the canonical wiring diagrams
of Cajal and more like modern renderings of
social networks (Fig. 2b). How to compare
and analyze these connectional graphs of
the brain is a new challenge that will require
new mathematical techniques for analyzing
graphs, new statistical tests to compare one
circuit with another and finally a cadre of
neuroscientists who will find mining big
data appealing. Ultimately, however, these
tools will only be useful if connectomes can
be produced easily.
Faster data collection. There are two reasons to collect connectomics data faster.
First, the production of connectomes must
eventually be easy enough so that multiple
experimental conditions can be compared
and findings can be replicated. Second,
the variety of questions that can be asked
increases as volumes enlarge and multiple
samples can be processed. If a complete
map of every synapse in a human brain is
wanted, a serious obstacle is the fact that
at current speeds it would take 10 million
years to complete.
How fast can we go? The best technology currently available for producing connectomics data sets is high-throughput
electron microscopy. Large-scale electron
microscopy data sets are now being produced that are about 300 cubic micrometers. For many circuits, this size is large
enough to encompass multiple complete axonal and dendritic arbors. There
are a number of ways that this technology could be modified to produce a
10–50 fold increase in data-acquisition
rates in the next few years. However, the
commentary
current bottleneck for obtaining connectomics data is not image acquisition but
image segmentation29. Many connectomics
data sets rely on a human painting voxels on
a computer screen.
One potential speedup is crowdsourcing
(http://eyewire.org/). Alternatively learning algorithms can use human segmentations to guide computer efforts30. At the
moment the results still require substantial
human editing. All of these directions are
being pursued, with competitions between
groups stimulating new approaches to this
daunting problem (http://brainiac.mit.edu/
SNEMI3D/). Ironically, much of this effort
in image analysis is to make machines do
what human brains do easily—something
that we might better understand once connectomics analysis of the visual system is
complete.
Proteome meets the connectome. The
structural mapping of connections will
only be useful if the cell type of each of
the neurons involved in a circuit is known.
Serial reconstruction of a neuron’s shape
and location will in some cases be sufficient for the type of neuron to be identified. But it is likely that there is heterogeneity among neurons that look alike.
Fortunately the past few decades have seen
an explosion in knowledge of the molecular classes of nerve cells31. What remains
to be done is finding ways to insert that
kind of information into mapping studies. Correlative fluorescence immunostaining with electron microscopy is one
approach32. Other techniques are certainly
on the horizon.
Conclusion
Despite the many arguments against
undertaking a connectomics analysis, we
think it must be done for three reasons.
First, the conditional patterns of synaptic
connectivity generated during development and by experience are inaccessible
to techniques that sample from only a
few cells at a time. Second, neuroscientists cannot claim to understand brains as
long as the network level of brain organization is uncharted; without this detailed
information, neuronal physiology is
connected to systems physiology by a
black box. Third, there is a high likelihood that such exploration will reveal
unexpected properties by shining a
light (or electrons) on this most mysterious tissue.
nature methods | VOL.10 NO.6 | JUNE 2013 | 499
commentary
ACKNOWLEDGMENTS
Our work was supported by a Conte Center grant (US
National Institute of Mental Health), the US National
Institutes of Health, the Gatsby Charitable Trust
and Center for Brain Science Harvard University. We
thank D. Rodrigues for use of his rendering of his
Twitter network and J. De Carlos for Cajal’s drawing
of the retina.
COMPETING FINANCIAL INTERESTS
The authors declare no competing financial interests.
npg
© 2013 Nature America, Inc. All rights reserved.
1. Markram, H. et al. A Report to the European
Commission. <http://www.humanbrainproject.
eu/files/HBP_flagship.pdf> (April 2012).
2. Collins, F. & Prabhakar, A. The White
House Blog <http://www.whitehouse.gov/
blog/2013/04/02/brain-initiative-challengesresearchers-unlock-mysteries-human-mind>
(2 April 2013).
3. Pautasso, M. Sustainability 4, 3234–3247
(2012).
4. Lichtman, J.W. & Denk, W. Science 334, 618–
623 (2011).
5. Masland, R.H. Curr. Opin. Neurobiol. 11, 431–
436 (2001).
500 | VOL.10 NO.6 | JUNE 2013 | nature methods
FOCUS ON MAPPING THE BRAIN
6. Lichtman, J.W. & Sanes, J.R. Curr. Opin.
Neurobiol. 18, 346–353 (2008).
7. Diamond, J.S. Nat. Neurosci. 5, 291–292
(2002).
8. Oláh, S. et al. Nature 461, 1278–1281 (2009).
9. Bargmann, C.I. Bioessays 34, 458–465 (2012).
10. Kerchner, G. & Nicoll, R. Nat. Rev. Neurosci. 9,
813–825 (2008).
11. Zador, A.M. et al. PLoS Biol. 10, e1001411
(2012).
12. Butcher, A.J. et al. J. Biol. Chem. 286, 11506–
11518 (2011).
13. Jahromi, S.S. & Atwood, H.L. J. Cell Biol. 63,
599–613 (1974).
14. Lichtman, J.W. & Colman, H. Neuron 25, 269–
278 (2000).
15. Weimann, J.M. & Marder, E. Curr. Biol. 4, 896–
902 (1994).
16. Lu, J., Tapia, J.C., White, O.L. & Lichtman, J.W.
PLoS Biol. 7, 13 (2009).
17. Sherman, S.M. & Guillery, R.W. J. Neurophysiol.
76, 1367–1395 (1996).
18. Seung, H.S. Neuron 62, 17–29 (2009).
19. Song, S., Sjöström, P.J., Reigl, M., Nelson, S. &
Chklovskii, D.B. PLoS Biol. 3, e68 (2005).
20. White, J.G., Southgate, E., Thomson, J.N. &
Brenner, S. Phil. Trans. R. Soc. Lond. B Biol. Sci.
314, 1–340 (1986).
21. Sporns, O., Tononi, G. & Kötter, R. PLoS Comput.
Biol. 1, e42 (2005).
22. Chalfie, M., Sulston, J.E., Thomson, J.N. &
White, G. J. Neurosci. 5, 956–964 (1985).
23. Swinford, S. The Sunday Times of London
(11 June 2006).
24. Anderson, J.R. et al. Mol. Vis. 17, 355–379
(2011).
25. Bock, D.D. et al. Nature 471, 177–182 (2011).
26. Briggman, K.L., Helmstaedter, M. & Denk, W.
Nature 471, 183–188 (2011).
27. Freed, M.A. & Sterling, P. J. Neurosci. 8, 2303–
2320 (1988).
28. Semon, R. W. & Simon, L. The Mneme. (G. Allen
& Unwin Ltd., 1921).
29. Helmstaedter, M. Nat. Methods 10, 501–507
(2013).
30. Kaynig, V. & Vazquez-Reina, A. IEEE Trans. Med.
Imaging 1, 1–7 (2012).
31. Lein, E.S. et al. Nature 445, 168–176 (2007).
32. Micheva, K.D., Busse, B., Weiler, N.C., O’Rourke, N.
& Smith, S.J. Neuron 68, 639–653 (2010).
focus on mapping the brain
perspective
Cellular-resolution connectomics: challenges
of dense neural circuit reconstruction
npg
© 2013 Nature America, Inc. All rights reserved.
Moritz Helmstaedter
Neuronal networks are high-dimensional graphs that
are packed into three-dimensional nervous tissue at
extremely high density. Comprehensively mapping
these networks is therefore a major challenge.
Although recent developments in volume electron
microscopy imaging have made data acquisition
feasible for circuits comprising a few hundreds to a
few thousands of neurons, data analysis is massively
lagging behind. The aim of this perspective is to
summarize and quantify the challenges for data
analysis in cellular-resolution connectomics and
describe current solutions involving online crowdsourcing and machine-learning approaches.
Brains are unique not in the number of cells they
comprise (about 85 billion neurons in the case of the
human brain1,2) but in the extent of direct and specific communication between their cells via synaptic
connections (each neuron has on the order of 1,000
synaptically coupled partner neurons). Mapping the
resulting complex connectivity graph is the goal of
connectomics. At the coarse level, inter-areal projections are tracked either noninvasively using variants
of diffusion imaging in humans3 (see Review4 in this
issue) or invasively using tracer injections combined
with high-throughput imaging in mice (Mouse Brain
Architecture Project5 (http://brainarchitecture.org/)
and Allen Brain Connectivity Atlas (http://connectivity.
brain-map.org/) among other initiatives; see Review6
in this issue). However, the mapping coverage achieved
using these approaches is still not sufficient to resolve
the complete set of neuronal networks contained in
the tissue. One voxel of magnetic resonance imaging
data with typical one-cubic-millimeter resolution, for
example, contains millions of neuronal cell bodies and
several kilometers of neuronal wires.
At the other end of the resolution spectrum, the
aim of connectomics is to image and analyze neuronal
circuits densely in sufficiently large volumes but at
single-cell and single-neurite resolution. Neuroscience
is very data-poor in this respect, contrary to the
assumptions made by contemporary simulation initiatives (The Human Brain Project7; http://www.human
brainproject.eu/), and it is not known whether there are
cases in which the measurement of cellular-resolution
connectivity graphs can in fact be replaced by simplified low-order statistical assumptions about neuronal
wiring. To date, only a few neuronal circuits have been
analyzed comprehensively—with the connectivity map
of the entire nervous system of Caenorhabditis elegans,
comprising 302 neurons8,9, being the most notable and
largest for decades.
Neuronal-circuit reconstruction is so difficult
because of the small size of neuronal processes and
the extremely high packing density of the neuropil
(Fig. 1a; here the term neuropil is used to describe
nervous tissue containing densely packed axons and
dendrites, even if interspersed with cell bodies, glia
cells and blood vessels). For more than a century, the
analysis of neuronal circuitry was therefore focused
on using very sparse labeling techniques that stained
only every 10,000th to 100,000th neuron in a given
volume (Fig. 1a).
Only recent advances in high-throughput electron
microscopy have made possible the imaging of substantial volumes of neuronal tissue at high resolution,
allowing the reconstruction of larger circuits at a much
faster pace. The main challenge, however, is the analysis of imaging data, which is currently the limiting step
by several orders of magnitude. Overcoming this gap
in data analysis is therefore the main methodological
focus of cellular-resolution connectomics.
Resolution requirements
The mapping of dense circuits requires the reconstruction of a large fraction of the neuronal wires in a
given volume of nerve tissue (Fig. 1a). In most cases,
Structure of Neocortical Circuits Group, Max Planck Institute of Neurobiology, Munich-Martinsried, Germany. Correspondence should be
addressed to M.H. ([email protected]).
Received 11 February; accepted 15 April; published online 30 may 2013; doi:10.1038/nmeth.2476
nature methods | VOL.10 NO.6 | june 2013 | 501
perspective
focus on mapping the brain
neuropil is largely isotropic: neuronal processes can locally turn
in any direction, even in cases where there is a preferred direction
such as the light axis in the retina or the radial developmental axis
in the neocortex. This means that the lowest-resolution dimension
of the imaging techniques used should account for the smallest
neurite diameter in the chosen tissue volume. Examples of the
smallest neurites in several parts of the nervous system (Fig. 1b)
are dendritic spine necks (40–50 nanometer (nm) diameter, for
example, in mouse hippocampus10 and neocortex), the thinnest
parts of mouse cortical axons (~50 nm; K.M. Boergens and M.H.;
unpublished data), amacrine cell dendrites in the mouse retina
(~50 nm (ref. 11; K.L. Briggman, personal communication) and
dendrites in the fly mushroom body (~30 nm (ref. 12)). The minimal required imaging resolution is then roughly half the smallest
neurite diameter, if one requires each neurite diameter to be represented by at least two voxels in the three-dimensional (3D) image
data set. It may be possible in some cases to relieve this resolution
requirement slightly based on prior knowledge about the shape of
neurites and the continuity of intracellular organelles. Thus, the
minimal required imaging resolution is ~20–30 nm for most circuits but can be as small as 10–15 nm in certain model systems.
a
A
B
C
D
E
F
10
–5
b Mushroom body
dendrites
30
10
–4
–3
40
–2
10
10
Fraction of labeled neurons
Dendritic spine
neck
10
Amacrine cell
dendrite
50
60
Smallest neurite diameter (nm)
–1
1
Pyramidal
cell axon
70
80
Volume requirements: minimal circuit dimension
When mapping neuronal circuits (Fig. 2a), it is important to
detect synaptic contacts between neurons, but it is in many cases
even more important to be able to exclude synaptic connectivity
between neurons to determine the structure of a wiring diagram
(measuring the ‘zeros’ of the connectivity matrix). To exclude the
existence of a synapse between two neurons, one has to image at
least one of these two neurons in entirety. More precisely, in cases
where neurons have an input part (for example, pyramidal cell
dendrites, which are assumed to be only postsynaptic for chemical synapses) and an output part (for example, the axon of a local
interneuron), at least one of the two neuronal arbors have to be
fully contained in the imaged volume.
This notion implies that for each circuit to be mapped, one
can define a minimal circuit volume that fulfills the criterion of
sufficient completeness for a sufficient number of the relevant
neurons (Fig. 2b–e). Then, the imaging technique should be capable of imaging volumes such that the smallest imaged dimension
is at least as large as the minimal required circuit dimension.
~1 mm
~300 µm
~500 µm
502 | VOL.10 NO.6 | june 2013 | nature methods
~300 µm
~80 µm
a Connectome
b Mouse retina c Mouse S1 cortex e Olfactory bulb
Figure 2 | Minimal circuit dimensions.
inner plexiform layer
layer 4
glomerulus
(a) A ‘connectome’, the connectivity matrix
Postsynaptic
~40 µm
‘1’
between a set of presynaptic and postsynatptic
neurons, with 15% positive entries (‘1’, blue
dots) representing the existence of a synaptic
connection between two neurons and 85%
negative entries (‘0’, white) representing the
absence of a synaptic connection. Detection
of a synaptic connection is a local decision
Mouse cortex layer 2/3
d
‘0’
(on the order of a few micrometers; top right).
To exclude synaptic connectivity between
two neurons, at least one of them has to be
imaged in its entirety, yielding the notion
of a minimal circuit volume (dashed box)
>1–2 mm
and the minimal circuit dimension (the smallest
of the volumes’ dimensions). (b–e) Approximate
minimal circuit dimensions for several example
circuits, based on the requirement to measure
the existence and the absence of synaptic connections and the spatial extent of the relevant neurons in the circuit. Not drawn to scale. Dashed boxes
indicate the respective minimal circuit volumes.
Presynaptic
npg
© 2013 Nature America, Inc. All rights reserved.
Figure 1 | Density of neuronal circuits and minimal resolution
requirements. (a) Fraction of labeled neurons using light microscopy (left)
and electron microscopy (right). In light microcopy images single neurons
can be detected only because of extremely sparse staining. Images show
(left to right): two synaptically coupled neurons from layers 4 and 2/3
in rat somatosensory cortex (modified from ref. 48), sketch of cortical
neurons (from ref. 49, image from Wikimedia), population of fluorescently
labeled neurons in neocortex that were presynaptic to one postsynaptic
pyramidal neuron (modified from ref. 50), reconstruction of mouse retina
inner plexiform layer (modified from ref. 36) and depiction of C. elegans
neurons (from wormbase.org and openworm.org, courtesy of C. Grove and
S. Larson). (b) Minimal resolution requirements for cellular connectomics.
The thinnest structures in various neuronal circuits are indicated, from left
to right: mushroom body dendrite (modified from ref. 12) dendrite of layer
4 neuron in mouse neocortex (image by B. Cowgill), starburst amacrine
cell dendrite, modified from ref. 11, GABAergic axon in mouse neocortex
(image by K. Boergens and N. Marahori). Not drawn to scale. Red arrows
indicate examples of the respective neurites.
focus on mapping the brain
a
perspective
e
Manual
ultrathin serial sectioning
TEM(CA)
Human cortex
b
Automated
serial sectioning,
tape collection
c
SEM
d
Minimal circuit dimension
1 cm
M.ret.
w.f.
Z.f. larv.
w.b.
M.o.b.
glom.
D.m.
m.b.
1 mm
100 µm
Mouse cortex
S1 layer 2/3
*
ssTEM(CA)
C. elegans w.b.
SB
EM
M. retina
s.f.
10 µm
D.m.
medulla
FIB-SEM
npg
© 2013 Nature America, Inc. All rights reserved.
Diamond knife,
serial block-face
SEM
FIB,
serial block-face
SEM
0
20
ATUMssSEM
40
60
Minimal required resolution (nm)
80
Figure 3 | Volume electron microscopy techniques for cellular connectomics and their spatial resolution and scope. (a–d) Sketches of the four most widely
used methods for dense-circuit reconstruction: conventional manual ultrathin sectioning of neuropil (a) followed by TEM or TEMCA imaging16 (a), ATUMSEM17 (b), SBEM18 (c) and FIB-SEM19 (d). In a,b, tissue is first sectioned and then (potentially later) transferred into the electron microscope for imaging.
In c,d, the tissue block is abraded while imaging inside the electron microscope. (e) Approximate minimal resolution and smallest spatial dimension typically
attainable with the imaging techniques in a–d, based on published results (gray shading); dashed lines indicate likely future extensions. Values also depend on
the quality of staining and neurons of interest in a circuit. Approximate minimal resolution and minimal circuit dimension required to image indicated circuits.
C. elegans w.b., C. elegans whole-brain reconstruction8,9,35; solid line indicates longest series from one worm and dashed line, the combined series length from
three worms9. D.m. m.b., Drosophila melanogaster mushroom body; minimal required resolution based on estimate of smallest dendrites (30 nm diameter; Fig. 1b);
D.m. medulla, D. melanogaster medulla, 1 cartridge (diameter of ~6 µm) with smallest processes less than 15 nm diameter. Human cortex, minimal circuit volume
containing entire L5 pyramidal neuron dendrites and their local axons. Mouse cortex S1 layer 2/3, minimal circuit volume (Fig. 2d). *, mouse cortex S1 layer 4
minimal circuit volume (Fig. 2c). M.o.b.glom., mouse olfactory bulb, 1 glomerulus, only intraglomerular circuitry (Fig. 2e). M. retina s.f., mouse retina, small
field (Fig. 2b). M.ret. w.f., mouse retina, wide field (including the largest amacrine and ganglion cells). Z.f. larv. w.b.: zebrafish larva whole brain.
For example, the minimal circuit volume for bipolar-toganglion cell connectivity in the mouse retina is dictated by the
size of the axon of bipolar cells, which is contained in a volume
of ~20 µm × 20 µm × 40 µm for most bipolar cells (20 µm in the
plane of the retina, and 40 µm along the light axis; Fig. 2b). If
one wants to map the circuits of roughly a dozen bipolar cells per
bipolar cell type, the resulting minimal circuit volume is ~80 µm ×
80 µm × 40 µm (Fig. 2b); thus the minimal circuit dimension is
40 µm. Other examples are the mouse olfactory bulb, where the
entire circuit within one glomerulus, excluding mitral cells, is
contained in a volume of ~300 µm × 300 µm × 300 µm (Fig. 2e).
If one aims to include mitral cells into the circuit analysis, the
minimal volume increases to ~1 mm × 500 µm × 500 µm (Fig. 2e;
A. Schaefer; personal communication). In mouse neocortex, the
minimal circuit volume is thought to be smallest in layer 4 of
primary somatosensory cortex (~300 µm × 300 µm × 300 µm,
Fig. 2c), it is ~1 mm × 1 mm × 500 µm in layer 2/3, ~3 mm ×
3 mm × 1 mm in layer 5 and at least as large in other cortical areas
where neurons have widespread axonal projections (Fig. 2d).
Together, the required minimal imaging resolution, as dictated by
the smallest neurite diameters, and the minimal circuit dimension,
as dictated by the types and spatial extent of the relevant neurons,
differ widely between different circuit model systems. To map these
circuits, it is therefore crucial to choose the appropriate volume
imaging method (Fig. 3).
Volume electron microscopy techniques
All current imaging techniques used for cellular-resolution connectomics rely on sequential two-dimensional (2D) imaging of
tissue combined with slicing, abrading or evaporating the tissue.
Then images are combined into a 3D image volume. Only in this
sense are today’s imaging methods volume imaging methods.
The key techniques are serial-section transmission electron
microscopy (ssTEM9,13–15) in some cases combined with fast
camera arrays (TEMCA16, Fig. 3a), automated serial-section
tape-collection scanning electron microscopy (ATUM-SEM17;
Fig. 3b), serial block-face SEM (SBEM11,18, Fig. 3c) and focused
ion beam SEM (FIB-SEM19, Fig. 3d); these microscopy techniques are reviewed in ref. 20. Briefly, these methods differ in
the sequence and quality of cutting and imaging, and provide different minimal imaging resolutions and dimensions. Essentially,
the resolution is currently lowest, and the acquirable dimension
is smallest, along the cutting axis.
In ssTEM (Fig. 3a), tissue is first cut into ultrathin sections,
which are then imaged using TEM or speed-enhanced TEMCA.
As sections are cut manually and are physically maintained for
the imaging step, these techniques have a minimal resolution of
~40 nm. They provide large imaging areas in the plane of imaging but yield typically small dimensions along the cutting axis
because manually cutting and maintaining more than a few thousand sections is hardly possible. ATUM-SEM (Fig. 3b) overcomes
the cutting-thickness limitation of ssTEM by automated slicing
and automated tape collection of the sections, yielding a minimal
resolution of up to 30 nm and providing long series of collected
sections. These are imaged using SEM as the collection tape is
electron-dense. The two block-face techniques (Fig. 3c,d) interlace
and invert the sequence of imaging and cutting: the entire tissue
block is transferred into the electron microscope chamber, and
nature methods | VOL.10 NO.6 | june 2013 | 503
perspective
focus on mapping the brain
a
b
d
Contouring
1
Difficult configuration
Synapse
2
....
1
c
6
Skeleton tracing
e
Missed branchpoint
3
Axon
npg
© 2013 Nature America, Inc. All rights reserved.
4
5
Dendrite
6
Figure 4 | Manual and automated reconstruction challenges in electron microscopy–based connectomics. (a) Volume reconstruction of a synaptic contact between
an excitatory axon and dendritic spine in mouse somatosensory cortex imaged using SBEM (top right; K.M. Boergens and M.H., unpublished data). Images show
electron microcopy slices at various depths; distance between slices 1 and 6 is 500 nm. (b,c) Manual data annotation is performed mostly either by contouring
of neurite cross-sections (b) or center-line reconstruction (skeleton tracing, c). Contouring is ~50 times faster but provides no direct volume reconstruction.
Modified from ref. 51. (d) Sketch of three possible neurite configurations for a configuration that automated algorithms typically cannot resolve but human
annotators can (panels show three possible configurations). Dashed lines, imaging plane in which analysis is most difficult. Note that one neurite runs parallel
to imaging plane (blue and green). Such a configuration is only resolvable if minimal resolution is greater than the smallest neurite diameter. Analysis may be
facilitated when including priors about neurite diameters (pink neurite). (e) Branchpoint that can be missed by human annotators (unforced attention-related
error25) corresponding to the dashed line in scheme on the left (from ref. 25; blue and white, skeletons traced independently by two expert annotators).
images are taken from the surface of the block using SEM. Then,
the top of the tissue block is abraded either using a diamondknife microtome installed into the electron microscope chamber
(SBEM, Fig. 3c) or using a focused ion beam attached to the electron microscope (FIB-SEM, Fig. 3d). Diamond knife–based cutting currently limits the abrasion thickness to 25 nm but allows
for long imaging sequences (smallest dimension is currently
300 µm (ref. 11), K.M. Boergens and M.H.; unpublished
data), whereas FIB-SEM is unique in providing a minimal
resolution of 4 nm, but so far limited in its smallest dimension
(~40 µm (ref. 19)).
The relationship between the available electron microscopy
imaging techniques and the resolution and size requirements of
several example circuits is summarized in Figure 3e. In a nutshell,
ssTEM is most amenable for circuits that can tolerate a minimal
required resolution of more than 40 nm and a minimal circuit
dimension of less than a few hundred micrometers, whereas FIBSEM is uniquely suited for circuits that require very high imaging
resolution, such as most circuits in the fly nervous system, but
can tolerate smaller volumes. ATUM-SEM and SBEM are currently best suited for dense-circuit reconstruction in systems with
~25–30 nm minimal resolution and several hundred micrometers minimal circuit dimension. The aims of ongoing method
developments are to increase the spatial scope along the cutting
dimension, increase imaging speed in the plane of imaging and
potentially decrease the cutting or abrasion thickness (Fig. 3e).
Circuit reconstruction
All high-resolution imaging techniques generate large volumes of
images, which totaled several hundred gigabytes per data set for
published projects11,16,21, are currently around a dozen terabytes
per data set and will likely be several petabytes per data set in
504 | VOL.10 NO.6 | june 2013 | nature methods
the foreseeable future. To reconstruct neuronal circuits from the
imaging data, synapses have to be identified, and the pre- and
postsynaptic neurites contributing to each synapse have to be followed back to their respective cell bodies to attribute the synapse
to the correct neuron. As soon as the imaging resolution and
staining quality are sufficient for detecting postsynaptic densities
and synaptic vesicles (which are ~40 nm in diameter and thus at
the same scale as most minimal neurite diameters, see above), the
detection of synapses becomes feasible. Synapse detection is more
difficult in systems where one presynaptic site can have multiple
postsynaptic partners (such as in the mammalian retina and the
fly optical system), presynaptic vesicle pools are small and/or
postsynaptic specializations are difficult to identify (for example,
symmetrical synapses of inhibitory axons in the neocortex).
In all cases, volume imaging substantially improves the
detection of synapses compared to the identification of synapses in two-dimensional images. For example, the 3D imaging of a synapse in mouse neocortex (Fig. 4a, at 25 nm slice
thickness using SBEM, K.M. Boergens and M.H., unpublished
data) substantially improves the reliability of synapse detection: judging the synapse from just one of the 2D images may
be questionable, but the sequence of several 2D images of
the same synaptic location relieves uncertainty of detection.
Synapse detection is additionally improved by high-resolution
3D imaging as obtained using FIB-SEM22 and can in these cases
be automated22,23. Automated synapse-detection methods for
SBEM and ssTEM data are likely to become available soon.
The main challenge for cellular connectomics, however, is not
the detection of synapses but the reconstruction of neuronal wires
(Fig. 4b–e). This immense difficulty has two main origins: first,
neurites vary greatly in diameter and in local entanglement, generating a substantial frequency of locations at which the path of
focus on mapping the brain
npg
© 2013 Nature America, Inc. All rights reserved.
a
Data acquisition
Reconstruction
Gaming
109
106
Completed
Analysis
gap
Completed
***
SBEM
ssTEM
* * * *
**
3
10
ssTEMCA
SBEM
Time (hours)
Figure 5 | Imaging and analysis times illustrating the analysis
gap in cellular connectomics. (a) Data acquisition times and
respective reconstruction durations for completed projects and
estimates for ongoing or planned projects. Blue and red shading
indicates experiment and reconstruction, respectively.
Data-acquisition projections assume ~5 megahertz imaging
speed and SBEM experiment (*); human cortex minimum
circuit volume–acquisition projection assumes an additional
~200-fold increase in imaging speed as potentially attainable
with multibeam electron microscopes (**). A recent C. elegans
data analysis had a tenfold increase in the speed of annotation26
(dashed line, ***). Gaming estimates are provided for successful
science games in total annotation hours and an online casual
game (Angry Birds) in annotation hours per year. FoldIt estimate
is based on the number of players; Galaxy Zoo estimate is based
on the reported number of classifications and an estimate of
on average 30 seconds per classification. (b,c) Screen shots
of ongoing online gaming initiatives in cellular connectomics:
Eyewire (b; image courtesy of H.S. Seung) and Brainflight (c).
perspective
) y
7) 9) 4) 6) 4 s /3 in it
3) 9) 4) 6)
us 4 /3 in nse
ar
43 om
6, f. . 1 3 x L ulu L2 ra rcu e
,3 f. 1 3
ul x L L2 ra
ye
er rte ex le b
de (ref. ron
s. (re (ref (ref. rte mer ex le b ci um ,24 (re (ref. (ref.
f
x
er
t
e
p
s
om co ort ho
(re tina ex se 1 co glo cort ho um vol 6,7 rse se se
g
l
t
r
g 1 c w
.
in a rds
ns re ort en S lb 1 a w nim
co
fs spa par en
ld o
lb S S1 va
fo Zo
Bi
ga e c d e u S rv i
r
(re ina x s a d y bu
an
y
le ous V1 ina ous ry b se la ex m
la
in axy
e
n
m
e
t
e
i
r
e
u
h
t
t
h
s
t
gr
u
t
t al
o
s
n Re or re
t
n
i
o
H
C. M use e re Macto Mo afis cor
r
c
e
A
F
c e
p G
d
f
o
lfa
br n
it
s
ol
M ous
V1 ous
O
ld
an
Ze uma
se
M
M
Fo
eg
l
H
ou
e
M
C.
neurites can only be revealed after intense inspection (Fig. 4d).
Second, and most notably, errors of neurite continuity have a
highly correlated effect such that any neurite break along the path
from a synapse to its soma creates an error in the assignment of
the synapse to a neuron. As these entangled paths can be as long
as several millimeters in many circuits (note that this is the neuronal path length between synapse and soma, not the Euclidean
distance), neurite reconstruction has to be extremely reliable to
provide tolerable error rates in the connectivity matrix.
In spite of these difficulties, human annotators can reconstruct
electron microscopy data by either contouring the neurites in
sequential image planes (Fig. 4b (refs. 21,24)) or by reconstructing only the neurite center line in three dimensions (Fig. 4c
(refs. 25,26). Most notably, human annotators can resolve difficult
locations with high accuracy (Fig. 4d). However, such annotation is very slow, and annotators make many unforced attentionrelated mistakes25 (Fig. 4e).
In an attempt to speed up reconstruction, a main focus in connectomic data analysis was to devise automated reconstruction
algorithms15,22,27–33. Unfortunately, however, these algorithms so
far do not provide sufficient reconstruction accuracy to replace
human annotators. Currently, automated reconstructions generate erroneous neurite breaks or erroneous neurite mergers every
few tens of micrometers of neuronal path length, which results
in the loss or misassignment of most of the synapses of a given
neuron. Automated algorithms can, however, substantially support data annotation by human annotators, as discussed below.
Why do automated algorithms fail at resolving what human
annotators can resolve? Figure 4d is an illustration of a local neurite configuration that could be interpreted in multiple ways by only
subtle local changes of the membranes separating the respective
neurites. Presently available automated algorithms are notoriously
bad at resolving such locations, especially when the concerned
neurites are very thin. It can be suspected that such neurite configurations can only be resolved when explicitly comparing the
different possible segmentation solutions (Fig. 4d), which is what
the human visual system may compute when inspecting such a
location in the data. Recent developments in algorithms aim to
account for such segmentation-based data analysis31,34.
Analysis of electron microscopy data is thus meeting a substantial challenge: humans can solve difficult locations but are slow
b
c
and make attention-based errors, whereas machines are efficient
at solving easy locations but fail when neurites become small and
their packing is dense.
The key approach to resolving the reconstruction problem in
electron microscopy–based connectomics has therefore been
to combine massive human data-annotation efforts with automated image analysis. In essence, human annotators provide the
long-range connectivity information and solve the most difficult
annotation problems, and automated algorithms provide the local
volume reconstruction at less difficult locations. This analysis
approach has been implemented in two variants: either automated
algorithms are first used to presegment the imaging data, followed by manual inspection and correction (iterated proofreading15,21,35, usually involving well-trained full-time annotators),
or manual annotation is parallelized, followed by automated
consensus computation, and by combination with local volume
segmentation (forward-only annotation, usually using lightly
trained part-time annotators such as undergraduate students25).
The analysis gap in cellular connectomics
What are the resulting experiment and analysis times for cellularresolution connectomics today? In Figure 5a, durations for the
accomplished circuit imaging and reconstruction experiments are
reported as well as estimates for ongoing and envisioned projects.
Data acquisition took around several hundred to a thousand
hours for the reconstruction of the C. elegans connectome9
(ssTEM), the direction-selectivity circuit analysis in the mouse
retina11 (SBEM), a local circuit analysis experiment in V1 mouse
cortex16 (ssTEMCA) and the dense local reconstruction of the
inner plexiform layer in mouse retina36 (SBEM). This time was
spread over several months for the ssTEM methods and compressed into 6–8 weeks in the SBEM experiments.
Imaging and acquisition speed has since substantially increased,
bringing much larger volumes within the reach of a several-month
continuous experiment: for example, one glomerulus in mouse
nature methods | VOL.10 NO.6 | june 2013 | 505
npg
© 2013 Nature America, Inc. All rights reserved.
perspective
olfactory bulb (~300 µm × 300 µm × 300 µm), the minimal circuit
volume in layer 4 of mouse S1 cortex (~300 µm × 300 µm ×
300 µm) or the entire brain of the larva-stage zebra fish
(~400 µm × 500 µm × 500 µm). Highly parallel SEM may promise
an additional speed-up of more than two orders of magnitude,
which would even enable the imaging of a relevant volume of
human cortex (~2 cm × 2 cm × 1 cm). Improvements in imaging and acquisition rates appear to have moved many circuits of
interest into the realm of experimental feasibility, which is ~1,000–
2,000 hours of imaging time (Fig. 5a; continuous experiments of
up to a year (~8,000 hours) may become feasible with improvements in automation).
However, if one considers the required reconstruction time,
cellular-resolution connectomics imposes daunting challenges. The
two recently published mouse connectomics studies made use of very
sparse circuit reconstructions: only around 30 neurons were reconstructed for the direction-selectivity circuit analysis11 and the V1
cortex study16, already amounting to ~1,000 hours of analysis time.
For dense-circuit reconstruction, the C. elegans reconstruction
took around 12 years of part-time work by one scientist9,26, also
amounting to around a few thousand hours of work. Including
the additional studies8,35, the C. elegans connectome may have
consumed around 10,000–20,000 hours spread across three
decades. The most recent mammalian dense-circuit reconstruction in mouse retina36 took ~30,000 hours, by hundreds of parttime undergraduate annotators.
Thus, all cellular-resolution connectomics studies to date have
involved thousands to tens-of-thousands of hours of reconstruction. Although imaging speed has substantially increased,
reconstruction speed is massively lagging behind. For any of
the proposed dense-circuit reconstructions (mouse neocortex,
­olfactory bulb, fish brain and human neocortex; Fig. 5a), analysistime estimates are at least one if not several orders of magnitude
larger than what has been accomplished to date (Fig. 5a): requirements for these envisioned projects are around several hundred
thousand hours of manual labor per project. These enormous
numbers constitute what can be called the analysis gap in cellular
connectomics: although imaging larger circuits is becoming more
and more feasible, reconstructing them is not.
Addressing the analysis problem
How is the field approaching this massive methodological challenge? There are two lines of solutions. First, improvements in
automated algorithms promise to increase the efficiency of reconstruction by human annotators. Second, massive online crowdsourcing initiatives are being built to ‘recruit’ the required work
hours over the internet.
The current accuracy of automated classifiers is at least three
orders of magnitude less than what is needed for reconstruction
of entire neurons. It is therefore unlikely to expect that automated
algorithms will soon take over image analysis for connectomics
completely. Improvements in algorithms have saturated at the
level of voxel classification (‘voxel classifiers’27,28,32,37–39), and
the current focus is on segmentation-based classifiers (‘supervoxel classifiers’31,34,40). Rather than waiting for such algorithms
to reach the required accuracy, automated methods have been
made useful in reducing the analysis challenge when they were
well integrated with human annotation15,25,41–44. For example,
506 | VOL.10 NO.6 | june 2013 | nature methods
focus on mapping the brain
the estimates of circuit reconstruction reported in Figure 5 are
based on manual-algorithmic integration that already yielded a
10–50-fold increase in reconstruction efficiency.
Thus, on one hand, the key prospect for the improvement of automated analysis remains in lowering error rates (that is, generating
longer correctly reconstructed neurite segments), but on the other
hand, the focus is on better integration with human annotation.
In an interesting twist, all automated methods rely heavily on training data. It is typically very expensive, however, to generate training data, and this in many cases limits algorithm improvements.
Therefore, the attempts to recruit annotators to perform large
amounts of human annotation, as described next, also may provide
a critical boost to improve automated reconstruction methods.
Online crowd-sourcing has been successful in a few scientific
fields: most notably the protein-folding game FoldIt45 and the galaxyclassification initiative Galaxy Zoo46. Although these approaches
proved very fruitful in their respective settings, the time requirements
in cellular connectomics are still large by comparison. Galaxy Zoo has
(over several years) probably recruited the work time required for
the reconstruction of one dense cortical circuit (Fig. 5a). The comparison to casual gaming, however, seems promising: even a single
successful game is reported to be played millions of hours every day,
amounting to as many hours a year as would be needed to reconstruct
a relevant piece of human cortex (~1 billion hours; Fig. 5a).
This setting of successful citizen-science games and the order of
magnitude of hours spent online playing games has fueled hopes
of recruiting the required analysis via the internet. As a first prerequisite, data annotation had to move online, overcoming the
challenges of streaming the required data to clients’ browsers.
The Open Connectome project using the collaborative annotation toolkit for massive amounts of image data (CATMAID) software47 focused on ssTEM and ssSEM data analysis, the Eyewire
project for SBEM data (http://www.eyewire.org/; Fig. 5b) and the
Brainflight project (http://www.brainflight.org/; Fig. 5c) are notable examples of online annotation, with Eyewire and Brainflight
using game-like features to enhance engagement.
It is an exciting endeavor today to explore whether citizen science can help to overcome the massive analysis gap we are facing in
cellular-resolution connectomics. Only if hundreds of thousands of
hours can be recruited will dense-circuit reconstruction be feasible.
It is also possible that improved automation will provide the crucial
gain in the efficiency of analysis, but most likely it is going to be a
combination of the two that will make dense-circuit reconstruction more and more realistic. Notably, it can be expected that once
such large data sets are being reconstructed, the large amounts of
training data needed to create more powerful analysis algorithms
will be available. Then, finally, can connectomics analysis become
a routine method with applications in many neuroscience laboratories, and cellular-resolution connectomics may further contribute
to unraveling the computations neuronal circuits perform.
Acknowledgments
I am grateful to the members of my laboratory for many fruitful discussions,
specifically to K.M. Boergens, Y. Buckley, F. Isensee, N. Marahori, A. Mohn and
H. Wissler for help with generating figures, and E. Dow for discussions concerning
game development. I thank M. Berning, K.M. Boergens, E. Dow and A. Schaefer
for helpful comments on the manuscript.
COMPETING FINANCIAL INTERESTS
The author declares no competing financial interests.
focus on mapping the brain
npg
© 2013 Nature America, Inc. All rights reserved.
Reprints and permissions information is available online at http://www.nature.
com/reprints/index.html.
1. Herculano-Houzel, S. The human brain in numbers: a linearly scaled-up
primate brain. Front. Human Neurosci. 3, 31 (2009).
2. Azevedo, F.A. et al. Equal numbers of neuronal and nonneuronal cells
make the human brain an isometrically scaled-up primate brain. J. Comp.
Neurol. 513, 532–541 (2009).
3. Wedeen, V.J. et al. The geometric structure of the brain fiber pathways.
Science 335, 1628–1634 (2012).
4. Craddock, R.C. Imaging human connectomes at the macroscale.
Nat. Methods 10, 524–539 (2013).
5. Bohland, J.W. et al. A proposal for a coordinated effort for the
determination of brainwide neuroanatomical connectivity in model
organisms at a mesoscopic scale. PLOS Comput. Biol. 5, e1000334
(2009).
6. Osten, P. & Margrie, T.W. Mapping brain circuitry with a light microscope.
Nat. Methods 10, 515–523 (2013).
7. Markram, H. The blue brain project. Nat. Rev. Neurosci. 7, 153–160 (2006).
8. Varshney, L.R., Chen, B.L., Paniagua, E., Hall, D.H. & Chklovskii, D.B.
Structural properties of the Caenorhabditis elegans neuronal network.
PLOS Comput. Biol. 7, e1001066 (2011).
9. White, J.G., Southgate, E., Thomson, J.N. & Brenner, S. The structure of
the nervous system of the nematode Caenorhabditis elegans. Philos. Trans.
R. Soc. Lond. B Biol. Sci. 314, 1–340 (1986).
10. Chicurel, M.E. & Harris, K.M. Three-dimensional analysis of the structure
and composition of CA3 branched dendritic spines and their synaptic
relationships with mossy fiber boutons in the rat hippocampus. J. Comp.
Neurol. 325, 169–182 (1992).
11. Briggman, K.L., Helmstaedter, M. & Denk, W. Wiring specificity in the
direction-selectivity circuit of the retina. Nature 471, 183–188 (2011).
12. Butcher, N.J., Friedrich, A.B., Lu, Z., Tanimoto, H. & Meinertzhagen, I.A.
Different classes of input and output neurons reveal new features in
microglomeruli of the adult Drosophila mushroom body calyx. J. Comp.
Neurol. 520, 2185–2201 (2012).
13. Fahrenbach, W.H. Anatomical circuitry of lateral inhibition in the eye
of the horseshoe crab, Limulus polyphemus. Proc. R. Soc. Lond. B
Biol. Sci. 225, 219–249 (1985).
14. Harris, K.M. et al. Uniform serial sectioning for transmission electron
microscopy. J. Neurosci. 26, 12101–12103 (2006).
15. Mishchenko, Y. et al. Ultrastructural analysis of hippocampal neuropil
from the connectomics perspective. Neuron 67, 1009–1020 (2010).
16. Bock, D.D. et al. Network anatomy and in vivo physiology of visual
cortical neurons. Nature 471, 177–182 (2011).
17. Hayworth, K.J., Kasthuri, N., Schalek, R. & Lichtman, J.W. Automating the
collection of ultrathin serial sections for large volume TEM reconstructions.
Microsc. Microanal. 12 (suppl. 2), 86–87 (2006).
18. Denk, W. & Horstmann, H. Serial block-face scanning electron microscopy
to reconstruct three-dimensional tissue nanostructure. PLoS Biol. 2, e329
(2004).
19. Knott, G., Marchman, H., Wall, D. & Lich, B. Serial section scanning
electron microscopy of adult brain tissue using focused ion beam milling.
J. Neurosci. 28, 2959–2964 (2008).
20. Briggman, K.L. & Bock, D.D. Volume electron microscopy for neuronal
circuit reconstruction. Curr. Opin. Neurobiol. 22, 154–161 (2012).
21. Cardona, A. et al. An integrated micro- and macroarchitectural analysis
of the Drosophila brain by computer-assisted serial section electron
microscopy. PLoS Biol. 8, e1000502 (2010).
22. Kreshuk, A. et al. Automated detection and segmentation of synaptic
contacts in nearly isotropic serial electron microscopy images. PLoS
ONE 6, e24899 (2011).
23. Becker, C., Ali, K., Knott, G. & Fua, P. Learning context cues for synapse
segmentation in EM volumes. Med. Image Comput. Comput. Assist. Interv.
15, 585–592 (2012).
24. Fiala, J.C. Reconstruct: a free editor for serial section microscopy.
J. Microsc. 218, 52–61 (2005).
25. Helmstaedter, M., Briggman, K.L. & Denk, W. High-accuracy neurite
reconstruction for high-throughput neuroanatomy. Nat. Neurosci. 14,
1081–1088 (2011).
Algorithm-based apparatus for combining the reconstructions of
many slightly trained annotators into a consensus reconstruction,
a prerequisite for large-scale crowd-sourcing.
26. Xu, M. et al. Computer assisted assembly of connectomes from electron
micrographs: application to Caenorhabditis elegans. PLoS ONE 8, e54050
(2013).
perspective
27. Jain, V. et al. Supervised learning of image restoration with convolutional
networks. IEEE 11th International Conference on Computer Vision 1–8
(2007).
28. Jain, V., Seung, H.S. & Turaga, S.C. Machines that learn to segment
images: a crucial technology for connectomics. Curr. Opin. Neurobiol. 20,
653–666 (2010).
29. Turaga, S.C. et al. Convolutional networks can learn to generate
affinity graphs for image segmentation. Neural Comput. 22, 511–538
(2010).
30. Chklovskii, D.B., Vitaladevuni, S. & Scheffer, L.K. Semi-automated
reconstruction of neural circuits using electron microscopy. Curr. Opin.
Neurobiol. 20, 667–675 (2010).
31. Andres, B. et al. 3D segmentation of SBFSEM images of neuropil by
a graphical model over supervoxel boundaries. Med. Image Anal. 16,
796–805 (2012).
32. Andres, B., Köthe, U., Helmstaedter, M., Denk, W. & Hamprecht, F.
Segmentation of SBFSEM volume data of neural tissue by hierarchical
classification. in Pattern Recognition, Proc. DAGM 2008, Lecture Notes in
Computer Science (ed., G. Rigoll) 5096, 142–152 (Springer, 2008).
33. Funke, J., Andres, B., Hamprecht, F.A., Cardona, A. & Cook, M. Efficient
automatic 3D-reconstruction of branching neurons from EM data. Proc.
IEEE Conference on Computer Vision and Pattern Recognition 1004–1011
(2012).
34. Jain, V. et al. Learning to agglomerate superpixel hierarchies. in Advances
in Neural Information Processing Systems (eds., J. Shawe-Taylor,
R.S. Zemel, P. Bartlett, F.C.N. Pereira and K.Q. Weinberger) 24, 648–656
(2011).
35. Jarrell, T.A. et al. The connectome of a decision-making neural network.
Science 337, 437–444 (2012).
36. Helmstaedter, M., Briggman, K.L., Turaga, S., Jain, V., Seung, H.S. &
Denk, W. Connectomic reconstruction of the inner plexiform layer in the
mouse retina. Nature (in the press).
37. Turaga, S.C. et al. Convolutional networks can learn to generate affinity
graphs for image segmentation. Neural Comput. 22, 511–538 (2010).
38. Mishchenko, Y. Automation of 3D reconstruction of neural tissue from
large volume of conventional serial section transmission electron
micrographs. J. Neurosci. Methods 176, 276–289 (2009).
39. Seyedhosseini, M. et al. Detection of neuron membranes in electron
microscopy images using multi-scale context and radon-like features.
Med. Image Comput. Comput. Assist. Interv. 14, 670–677 (2011).
40. Vazquez-Reina, A. et al. Segmentation fusion for connectomics. IEEE
International Conference on Computer Vision 177–184 (2011).
41. Sommer, C., Straehle, C., Koethe, U. & Hamprecht, F.A. ilastik: interactive
learning and segmentation toolkit. Proc. IEEE International Symposium on
Biomedical Imaging 8, 230–233 (2011).
42. Straehle, C.N., Kothe, U., Knott, G. & Hamprecht, F.A. Carving: scalable
interactive segmentation of neural volume electron microscopy images.
Med. Image Comput. Comput. Assist. Interv. 14, 653–660 (2011).
43. Jeong, W. et al. Ssecrett and NeuroTrace: interactive visualization and
analysis tools for large-scale neuroscience data sets. IEEE Comput.
Graph. Appl. 30, 58–70 (2010).
44. Roberts, M. et al. Neural process reconstruction from sparse user scribbles.
Med. Image Comput. Comput. Assist. Interv. 14, 621–628 (2011).
45. Cooper, S. et al. Predicting protein structures with a multiplayer online
game. Nature 466, 756–760 (2010).
Example of a computer game used for scientific discovery.
46. Land, K. et al. Galaxy Zoo: the large-scale spin statistics of spiral galaxies
in the Sloan Digital Sky Survey. Mon. Not. R. Astron. Soc. 388, 1686–1692
(2008).
47. Saalfeld, S., Cardona, A., Hartenstein, V. & Tomancak, P. CATMAID:
collaborative annotation toolkit for massive amounts of image data.
Bioinformatics 25, 1984–1986 (2009).
Online annotation platform for electron microscopy image data
presented in successive 2D image planes.
48. Feldmeyer, D., Lubke, J., Silver, R.A. & Sakmann, B. Synaptic connections
between layer 4 spiny neurone-layer 2/3 pyramidal cell pairs in juvenile
rat barrel cortex: physiology and anatomy of interlaminar signalling within
a cortical column. J. Physiol. (Lond.) 538, 803–822 (2002).
49. Ramón y Cajal, S. Textura del sistema nervioso del hombre y de los
vertebrados (Imprenta N. Moya, 1904).
50. Rancz, E.A. et al. Transfection via whole-cell recording in vivo: bridging
single-cell physiology, genetics and connectomics. Nat. Neurosci. 14,
527–532 (2011).
51. Helmstaedter, M. & Mitra, P.P. Computational methods and challenges for
large-scale circuit mapping. Curr. Opin. Neurobiol. 22, 162–169 (2012).
nature methods | VOL.10 NO.6 | june 2013 | 507
perspective
focus on mapping the brain
CLARITY for mapping the nervous system
npg
© 2013 Nature America, Inc. All rights reserved.
Kwanghun Chung1,2 & Karl Deisseroth1–4
With potential relevance for brain-mapping work,
hydrogel-based structures can now be built from
within biological tissue to allow subsequent
removal of lipids without mechanical disassembly
of the tissue. This process creates a tissue-hydrogel
hybrid that is physically stable, that preserves fine
structure, proteins and nucleic acids, and that is
permeable to both visible-spectrum photons and
exogenous macromolecules. Here we highlight
relevant challenges and opportunities of this
approach, especially with regard to integration with complementary methodologies for brainmapping studies.
Mammalian brains are staggeringly complex in terms
of both scale and diversity; many billions of neurons
are present, among them likely at least hundreds of
genetically distinct cell types, with each type of cell
represented by many distinct projection patterns.
CLARITY1 is a newly developed technology that can
be used to transform intact biological tissue into a
hybrid form in which tissue components are removed
and replaced with exogenous elements for increased
accessibility and functionality. CLARITY has the
potential to facilitate rapid extraction of anatomicalprojection information important for many fields of
neuroscience research2,3; such information can be
collected along with molecular-phenotype information at the resolution of single cells. Alone or in combination with other methods4,5, such an approach could
contribute to the study of function and dysfunction in
this complex system.
In general, obtaining system-wide detailed information from neural tissue is a formidable challenge (to
say nothing of subsequent data curation and analysis).
In the mammalian central nervous system, seamlessly
intertwined neural processes leave little extracellular
space, creating barriers to macromolecule diffusion
for in situ hybridization, antibody staining or other
forms of molecular phenotyping deeper than the first
few cellular layers of intact tissue6. The high density
of irregularly arranged lipid interfaces that characterizes this tissue likewise creates an effective scattering
barrier to photon penetration for optical interrogation of mammalian brains7, unlike the Caenorhabditis
elegans (worm) or larval Danio rerio (zebrafish)
nervous systems, which are more accessible owing
in part to smaller size and less myelination. Singlephoton microscopy can provide optical transmission of information from only about 50 micrometers
below the mammalian brain surface, and even welloptimized two-photon microscopy cannot be used to
image deeper than about 800 micrometers, far short
of enabling visualization of full projection patterns
and global arrangement of cell populations in the
intact brain8.
Over the past few decades, a great deal of technological innovation has been stimulated by these
challenges8–21. First, newer automated methods for
mechanical sectioning of tissue have overcome some
of the drawbacks of traditional sectioning methods
that were laborious, expensive and damaged the
tissue. Serial block-face mechanical 9–13 or optical­ablative14 methods, in combination with imaging readouts such as two-photon tomography14,15, electron
microscopy16 or array tomography17, have been used
to map ­macroscopic to nanoscopic brain structures
(see Review18 in this issue). In some cases, molecular labeling is built into these processes, and ongoing
work includes approaches for addressing additional
challenges such as generation of contrast in tissue
before sectioning19,20. Tools for automated analysis,
efficient reconstruction and error-free alignment also
continue to be developed21 as well as for registration
with activity information, and indeed detailed wiring
information linked to activity has been obtained from
well-defined volumes22,23.
Second, optical clearing methods have been
­developed that involve immersion of the specimen in
medium that matches the refractive index of the tissue,
thereby reducing light scattering and extending
1Department of Bioengineering, Stanford University, Stanford, California, USA. 2CNC Program, Stanford University, Stanford, California, USA.
3Department of Psychiatry and Behavioral Sciences, Stanford University, Stanford, California, USA. 4Howard Hughes Medical Institute, Stanford
University, Stanford, California, USA. Correspondence should be addressed to K.C. ([email protected]) and K.D. ([email protected]).
Received 18 March; accepted 22 April; published online 30 may 2013; doi:10.1038/nmeth.2481
508 | VOL.10 NO.6 | JUNE 2013 | nature methods
focus on mapping the brain
perspective
npg
© 2013 Nature America, Inc. All rights reserved.
Figure 1 | Imaging of nervous system projections
in the intact mouse brain with CLARITY. Adult
Thy1-EYFP H line mice (4 months old) were
perfused transcardially followed by hybridization
of biomolecule-bound monomers into a hydrogel
mesh and lipid removal as described1. The
clarified mouse brain was imaged from the
dorsal region to the ventral region (3.4 mm to
the midpoint) using a 10× water-immersion
objective. Image adapted from ref. 1.
the range of optical imaging24–26. For
example, benzyl alcohol−benzyl benzoate is an organic solvent that effectively
renders biological specimens transparent but reduces the stability of fluorescent protein signals24,25; in Scale, another
clearing method, an aqueous solution is
used to preserve fluorescence signals, but
the rate and the extent of clearing remain
limiting26. Notably all current tissue1 mm
clearing methods leave the densely packed
lipid bilayers intact and therefore still
face challenges with regard to penetration by visible-spectrum
light and molecules, making these methods largely incompatible
with whole-tissue molecular phenotyping.
The CLARITY approach1 helps address ongoing challenges
by enabling molecular and optical interrogation of large assembled biological systems, such as the entire adult mouse brain.
Light-microscopy (Fig. 1) and biochemical-phenotyping (Fig. 2)
techniques can be used to rapidly access the entire intact clarified mouse brain with fine structural resolution and molecular
detail (to the level of spines, synapses, proteins, single-amino-acid
neuro­transmitters and nucleic acids) while in the same preparation maintaining global structural information including brainwide macroscopic connectivity. To clarify tissue (Fig. 3), lipid
bilayers are replaced with a more rigid and porous hydrogel-based
a
3D rendering
b
DR
infrastructure (in a process conceptually akin to petrification or
fossilization, except that not only structure but also native biomolecules such as proteins and nucleic acids are preserved). This
outcome is achieved by first infusing small organic hydrogelmonomer molecules into the intact brain along with crosslinkers and thermally triggered polymerization initiators;
sub­sequent temperature elevation triggers formation, from within
the brain, of a hydrogel meshwork covalently linked to native proteins, small molecules and nucleic acids but not to lipids, which
lack the necessary reactive groups (Fig. 3a). Subsequent wholebrain electrophoresis in the presence of ionic detergents actively
removes the lipids (Fig. 3b), resulting in a transparent brainhydrogel hybrid that both preserves, and makes accessible, structural and molecular information for ­visualization and analysis.
DR
RR
z = 2,460 µm
SNR
SNR
c
VTA
VTA
PO
z = 1,005 µm
d
CPu
P
PO
CPu
V
D
A
z = 420 µm
Figure 2 | Intact mouse brain molecular
phenotyping and imaging with CLARITY.
(a) Three-dimensional (3D) visualization of
immunohistology data, showing tyrosine
hydroxylase (TH)-positive neurons and fibers in
the mouse brain. The intact clarified brain was
stained for 6 weeks as described1, with primary
antibody for 2 weeks, followed by a 1-week wash,
then stained with secondary antibody for 2 weeks
followed by a 1-week wash and imaged 2,500 µm
from ventral side using the 10× water-immersion
objective. D, V, A and P indicate dorsal, ventral,
anterior and posterior, respectively. (b–d) Optical
sections at different depths, corresponding
respectively to the upper, middle and lower
dashed box regions in a. Note that TH-positive
neurons are well-labeled and clearly visible even
at a depth of 2,460 µm in the intact brain.
CPu, caudate putamen; PO, preoptic nucleus;
VTA, ventral tegmental area; SNR, substantia
nigra; RR, retrorubral nucleus; DR, dorsal raphe.
Scale bars, 700 µm (a) and 100 µm (b–d).
Image adapted from ref. 1.
nature methods | VOL.10 NO.6 | JUNE 2013 | 509
perspective
npg
© 2013 Nature America, Inc. All rights reserved.
Figure 3 | CLARITY technology and instrumentation. (a) Tissue is crosslinked with formaldehyde in the presence of infused hydrogel monomers.
Thermally triggered polymerization then results in a hydrogel-tissue hybrid
which physically supports tissue structure and chemically incorporates
native biomolecules into the hydrogel mesh. (b) In electrophoretic tissue
clearing (ETC), an electric field is applied across the hybrid immersed in
an ionic detergent solution to actively transport ionic micelles into the
hybrid and extract membrane lipids out of the tissue, leaving structures
and cross-linked biomolecules in place and available for imaging and
molecular phenotyping1. The ETC setup consists of the custom ETC
chamber, a temperature-controlled buffer circulator (RE415, Lauda),
a buffer filter (McMaster) and a power supply (Bio-Rad). The sample
is electrophoresed by applying 20–60 volts to the electrodes. Buffer
solution is circulated through the chamber to maintain temperature
and the composition of the buffer solution constant throughout the
clearing process. The cut-through view (bottom right) shows placement
of the hydrogel-embedded tissue in the sample holder (Cell Strainer,
BD Biosciences) located in the middle of the chamber between the two
electrodes. The end of each electrode exposed outside the chamber is
connected to a power supply. Image adapted from ref. 1.
focus on mapping the brain
a
Step 1: hydrogel monomer infusion (days 1–3)
Proteins
DNA
ER
4 °C
Vesicle
510 | VOL.10 NO.6 | JUNE 2013 | nature methods
+
Formaldehyde Hydrogel
monomer
Plasma
membrane
Step 2: hydrogel-tissue hybridization (day 3)
37 °C
Hydrogel
b
Step 3: ETC (days 5–8)
Extracted lipids
in SDS micelle
SDS micelle
Here we address opportunities and challenges for CLARITY
methods in brain mapping, in combination with other emerging
genetic and imaging technologies.
Imaging methods for CLARITY
CLARITY is compatible with most fluorescence microscopy techniques. Conventional laser-scanning approaches, such as confocal
and multiphoton microscopies, are well-suited to image samples
prepared using CLARITY because the excitation and emission
wavelengths of light involved can penetrate deep into the transparent tissue. Objective working distance can be limiting in this
setting, although already single-photon imaging of the intact
clarified adult mouse brain with a 3.6-millimeter working distance objective has been achieved without noticeable degradation in resolution (Fig. 1)1. Even greater depth of imaging, for
larger brains, tissues or organisms, is possible with longer working distance and high-numerical-aperture objectives (such as the
5–8-millimeter working distance versions available from several
manufacturers), but it will be important for optics to be refined
and for objectives to be developed that are matched to the refractive index of clarified tissue to minimize aberration and maintain
both high resolution and long working distance.
Single-photon laser-scanning confocal imaging is in many
respects appropriate for CLARITY, given the broad range of
suitable fluorescent labels available for multicolor confocal
interrogation of clarified tissue. But confocal microscopy does
expose the entire depth of tissue to excitation light and therefore
induces substantial photobleaching of fluorescent molecules, a
particularly acute problem in the setting of slow, high-­resolution
whole-brain imaging. Nonlinear imaging techniques such as
two-photon microscopy address this out-of-focus photobleaching issue, but both single and two-photon imaging methods suffer
from low image acquisition rate. With these scanning methods,
imaging a mouse brain may take days at single-cell resolution
and months at single-neurite resolution—a key practical issue
when dealing with the larger samples that are characteristic of
CLARITY approaches.
For imaging large samples at high resolution and at high dataacquisition rates, we suggest that selective-plane illumination
microscopy (SPIM) may be a method of choice. In this technique,
+
Temperature-controlled
buffer circulator
Top view: cut through
Electrode
ETC
chamber
Target tissue
Buffer
filter
Inlet port
Electrode
Electrophoresis
power supply
Electrode
connector
the sample is illuminated with a thin sheet of light from the side
and fluorescence emission is collected along an axis perpendicular to the plane of illumination by wide-field acquisition26–31.
This unique illumination and imaging modality substantially
reduces both photobleaching and imaging time (for imaging
whole mouse brain, closer to hours at single-cell resolution, and
days at single-neurite resolution). CLARITY will naturally work
well with SPIM; for example, the high and uniform transparency
of clarified tissue will minimize light-plane broadening that causes
degradation in resolution toward the center of the tissue in SPIM
and therefore enable uniform high-resolution imaging of large
samples. Other rapidly developing imaging technologies, such
as structured illumination and Bessel-beam illumination31, will
enhance resolution and may potentially allow imaging in large
and intact tissue specimens even beyond the diffraction limit. The
CLARITY-SPIM combination of high resolution, independence
from mechanical sectioning and reconstruction, and fast dataacquisition rate may become particularly useful for distinguishing
and tracing neural projections with high throughput and accuracy
in the course of generating maps of brain connectivity.
When indicated, after global maps of intact systems are generated, individual portions of tissue such as the downstream target
of an imaged projection may be extracted for analyses such as
electron microscopy that require sectioning. CLARITY is compatible with subsequent electron microscopy and preserves some
ultrastructural features such as postsynaptic densities1, although
because of the absence of lipid, conventional electron microscopy
staining does not currently provide enough contrast to identify all
npg
© 2013 Nature America, Inc. All rights reserved.
focus on mapping the brain
relevant ultrastructures and boundaries, pointing to the need for
optimization of the electron microscopy preparation and ­staining
process. Alternatively, when combined with array tomography
and/or super-resolution imaging in neural ­ applications10,32,
clarified tissue may provide for nanoscopic molecular characterization of synaptic connections defined by paired presynaptic and postsynaptic proteins, so that the synaptic wiring logic
of incoming projections to a brain region can be established in
detail. Of course, no single imaging modality that may be used
with CLARITY provides all potentially important information on
functional connectivity. For example, conventional light microscopy may not readily provide definitive identification of synaptic
contacts, whereas electron microscopy analysis does not typically
reveal rich molecular information on detected synapses relevant
to functional properties, nor would an absence of synaptic junctions between cells identified by any method demonstrate absence
of direct cellular communication, which can occur via nonlocal and volume-transmission influences, as in the case of neuro­
modulators or extrasynaptic neurotransmitter action. Consistent
with possible integrative strategies, clarified tissue appears compatible with diverse imaging readout modalities that complement
each other; moreover, the multiround molecular phenotyping
capacity of CLARITY (see below) may be suitable for providing enriched detail on the composition and relationships of sub­
cellular structures such as synapses33.
Integrating circuit maps with molecular information
A unique feature of CLARITY is its potential for intact-tissue
molecular phenotyping studies. The hydrogel-tissue hybridization preserves endogenous biomolecules ranging from
­neurotransmitters to proteins and nucleic acids1; soluble proteins and cell-membrane proteins alike are secured by chemical
tethering to the hydrogel mesh. Moreover, removal of the lipid
membrane makes this retained molecular information accessible
via passive diffusion of macromolecular probes into the tissue,
and the enhanced structural integrity of clarified tissue allows
multiple rounds of antibody staining, elution or destaining and
restaining that are not typically feasible with conventionally
fixed tissue1,34. Immunostaining of 1-millimeter-thick clarified
tissue blocks takes days, and immunostaining of entire intact
adult mouse brains is possible on practical timescales of several
weeks (Fig. 2). This process could be accelerated by increasing
probe diffusion rate with electrophoresis or other methods. The
fact that CLARITY supports multiple (at least three) rounds of
molecular phenotyping may be of value to define cell types and
to link form and function in brain-mapping studies, allowing
the integration of rich local and global morphological details
(for example, type of synapse, shape of cell body, and information about dendritic arborization and axonal projections) with
genetic fingerprints.
In addition, the combination of CLARITY with genetic methods for identification of synapses or for labeling specific circuits
is also of interest. First, GFP reconstitution across synaptic
partners (GRASP) has emerged as a light microscopy–based
­synapse-­detection technique35. In this method, two nonfluorescent split-GFP fragments can be virally expressed in the synaptic
membrane of two separate neuronal populations; these two fragments reconstitute fluorescent GFP only across a synaptic cleft
so that the location of synapses between the two ­ populations
perspective
can be visualized35,36. This approach was validated with electron microscopy and was used to map synaptic connectivity in
the nematode35 and the fruit fly37; most recently, applicability
to mammalian systems was demonstrated: mammalian GRASP
(mGRASP) revealed labeling of both excitatory and inhibitory
synapses38. Second, in recent years many mouse lines and viral
targeting techniques have been developed for intersectional
labeling of spatially, genetically, synaptically and functionally
defined neuronal circuit elements. For instance, an engineered
rabies virus encoding EGFP can be injected into a particular
brain region to label neurons that are presynaptic to the injection
site39–41. More recently, activity-dependent expression of channelrhodopsin (ChR2) conjugated with EYFP was used to label and
drive a subset of synaptically recruited neurons associated with
fear memory42, and a technique has been developed and validated
that allows permanent marking (by filling a cell with fluorescent
protein labels) of neurons across the mouse brain that were active
during a relatively restricted time window (on a time scale of
hours)43. These versatile targeting approaches, in combination
with CLARITY, could provide a high-throughput approach for
globally mapping synaptically connected and synaptically activated populations across the brain that could complement the use
of electron microscopy in some settings.
Although new genetic strategies such as these are quite power­
ful, for studying and mapping the human brain, more conventional antibody-based (nongenetic) molecular phenotyping
alone is particularly important. Lipophilic dyes (for example,
1,1′-dioctadecyl-3,3,3′3′-tetramethylindocarbocyanine perchlor­
ate) can be injected into postmortem human brain to effectively
label local projections through passive diffusion44; however,
in fixed tissue without active axonal transport, diffusion rates
(and therefore tracing distance) are substantially limited and
may not be ­ideally suited to mapping long-range connectivity.
With CLARITY, molecular markers have already been used to
identify individual structures and projections (cell bodies and
fibers) in human tissue on the millimeter length scale1; although
this approach has not yet been tested for tracking long-range
projections in the human brain, neurofilament protein staining
reliably highlights a major subset of axonal fibers, and because
continuity of labeled projections is preserved in the intact tissue,
fibers can be traced with diminished risk of alignment and/or
reconstruction error. Cell type–specific markers (for example,
parvalbumin and tyrosine hydroxylase) can also be used to label
both cell bodies and projections in human brain tissue1. CLARITY
may in this way help unlock a rich source of clinically relevant
information, providing a means to interrogate and make accessible brain-bank samples that are unique or rare, for the purpose of
understanding disease mechanisms as well as the native structure,
organization and complexity of the human brain.
Limitations, challenges and opportunities
CLARITY is currently in the early stages of development; innovation will be needed over the coming years. First, we have noted
that tissues can expand after electrophoretic tissue clearing and
return to the original size after refractive-index matching1;
although macroscopic observation indicates that the changes in
volume are largely isotropic and reversible, quantitative monitoring at microscopic and nanoscopic resolution will be required
to confirm that loss in structural connectivity or occurrence of
nature methods | VOL.10 NO.6 | JUNE 2013 | 511
npg
© 2013 Nature America, Inc. All rights reserved.
perspective
other tissue artifacts is minimal. Second, although adherence to
the described protocol1 will minimize tissue damage, the precise
extent to which CLARITY under best current practice secures
specific molecular information represented in proteins, nucleic
acids and small molecules must be further explored. Only ~8% of
total protein content is lost in CLARITY1, substantially lower than
with other methods over the same timescale (for example, ~24%
loss with conventional fixative (paraformaldehyde) and detergent (0.1% Triton X-100) histological solution compositions 1).
However, the nature of the remaining loss should be investigated,
and we note that the current chemistry of CLARITY (by design)
does not preserve lipids or other molecules lacking functional
groups required for chemical tethering to the hydrogel mesh, such
as phosphatidylinositol 4,5-bisphosphate, or PIP2. Additional
chemical-treatment strategies may need to be explored for
specialized experimental questions.
It also remains to be determined how accessible different classes
of biological information from clarified tissue may be, in clinical
or animal settings, for high-content quantitative data-extraction
pipelines. Mapping this landscape will be of value for developing
strategies to maximize information extraction after global (brainwide) projection mapping has been completed, and may be carried
out in the context of normal function, disease states and treatment
regimens (in mapping effects of brain-stimulation treatments or
in drug screening, for example). Another open question is how
long clarified tissue may be maintained or stored for such analysis, assessment, imaging or remodeling in subsequent rounds of
CLARITY processing.
CLARITY can be seen as a prototype for a general approach for
building new structures and installing new functions from within
biological systems—an approach that may find other instantiations as the technology is developed for ­introducing ­components
or monomers into biological tissue that are then triggered subsequently to form a polymer, gel, mesh, network or assembled
structure with desired physical parameters (for example, stiffness, transparency, pore size, conductivity and permeability) or
active properties (for interfacing, catalysis or functionalization).
Relevant exogenous components may include proteins, oligonucleotides, stains, chemicals or even small mechanical, electronic
or optical components. These introduced components could be
designed for either constitutive or inducible functionality—in the
latter case, such that they can be activated by external elements
including heat, mechanical force, redox changes, electromagnetic
triggers such as light, and other accelerators, thereby enabling
temporally precise initiation of the structural and functional tissue transformation. The capacity to trigger functionality could
help enable versions of this general approach (not involving lipid
removal) that may be compatible with ongoing vital functions
of the tissue.
On the brain-mapping front, as pioneering technologies such as
array tomography and serial block face scanning electron microscopy have already proven, turning large data sets into useful and
tractable deliverables still poses an immense challenge. Relevant
computational approaches for registration in three dimensions
will need to be developed, and particularly in the setting of
variable, quenching and/or bleaching fluorescent signals, automated image acquisition, segmentation and tracing algorithms
will be helpful. We note that molecular landmarks are useful for
navigating three-dimensional data, and here the intact-tissue
512 | VOL.10 NO.6 | JUNE 2013 | nature methods
focus on mapping the brain
­ olecular phenotyping capability of CLARITY becomes particm
ularly important. As in conventional histology, broadly staining
markers (which are dense yet distinct) can be useful as landmarks
to navigate and identify brain regions of interest45. In this regard,
DAPI staining is compatible with CLARITY1 and appears particularly useful as a Nissl-like registration method.
Registration of CLARITY data with other data is both a
­challenge and an opportunity. For example, in an animalsubject brain in which a particular set of neurons becomes known
(via imaging of activity 46–49 and/or via optogenetic control4) to
be involved in a specific behavioral function or dysfunction, to
then clarify the same preparation and obtain brain-wide wiring
and molecular information on those same cells (and their connection partners) would be of substantial value. This approach
could be used to address fundamental questions in both basic
and clinical or preclinical neuroscience but will require the
development of efficient workflows for the registration of the
different types of experimental data. CLARITY may also help
to conceptually link future high-resolution activity maps50 with
structural or functional macroscale maps (for example, http://
www.neuroscienceblueprint.nih.gov/connectome/), providing
an anatomical foundation that could help researchers decipher
the meaning of brain-activity patterns linked to health and
disease4,50. Beyond neuroscience, CLARITY is currently being
explored for the evaluation, diagnosis and prognosis of pathological states including cancer, infection, autoimmune disease
and other clinical conditions as well as for the study of normal
tissue, organ and organism function, including development
and relationships of cells and tissues. Resources that may help
enable the general user to establish the methodology are available online (http://clarityresourcecenter.org/).
Acknowledgments
We acknowledge all members of the Deisseroth laboratory for discussions and
support. This work was funded by a US National Institutes of Health Director’s
Transformative Research Award (TR01) to K.D. from the National Institute
of Mental Health, as well as by the National Science Foundation, the Simons
Foundation, the President and Provost of Stanford University, and the Howard
Hughes Medical Institute. K.D. is also funded by the National Institute on Drug
Abuse and the Defense Advanced Research Projects Agency Reorganization
and Plasticity to Accelerate Injury Recovery program, and the Wiegers, Snyder,
Reeves, Gatsby, and Yu Foundations. K.C. is supported by the Burroughs Wellcome
Fund Career Award at the Scientific Interface.
COMPETING FINANCIAL INTERESTS
The authors declare competing financial interests: details are available in the
online version of the paper.
Reprints and permissions information is available online at http://www.nature.
com/reprints/index.html.
1. Chung, K. et al. Structural and molecular interrogation of intact biological
systems. Nature advance online publication, doi:10.1038/nature12107
(10 April 2013).
2. Petersen, C.C.H. The functional organization of the barrel cortex. Neuron
56, 339–355 (2007).
3. Mombaerts, P. et al. Visualizing an olfactory sensory map. Cell 87,
675–686 (1996).
4. Deisseroth, K. Optogenetics and psychiatry: applications, challenges, and
opportunities. Biol. Psychiatry 71, 1030–1032 (2012).
5. Kasthuri, N. & Lichtman, J.W. The rise of the ‘projectome’. Nat. Methods 4,
307–308 (2007).
6. Nicholson, C. Diffusion in brain extracellular space. Brain 6, 1277–1340
(2008).
7. Cheong, W., Prahl, S. & Welch, A. A review of the optical properties of
biological tissues. IEEE J. Quantum Electronics 26, 2166–2185 (1990).
npg
© 2013 Nature America, Inc. All rights reserved.
focus on mapping the brain
8. Helmchen, F. & Denk, W. Deep tissue two-photon microscopy.
Nat. Methods 2, 932–940 (2005).
9. Rauschning, W. Surface cryoplanning: a technique for clinical anatomical
correlations. Uppsala J. Med. Sci. 91, 251–255 (1986).
10. Toga, A.W., Ambach, K., Quinn, B., Hutchin, M. & Burton, J.S.
Postmortem anatomy from cryosectioned whole human brain. J. Neurosci.
Methods 54, 239–252 (1994).
11. Ewald, A.J., McBride, H., Reddington, M., Fraser, S.E. & Kerschmann, R.
Surface imaging microscopy, an automated method for visualizing whole
embryo samples in three dimensions at high resolution. Dev. Dyn. 225,
369–375 (2002).
12. McCormick, B.H. et al. Construction of anatomically correct models of
mouse brain networks. Neurocomputing 58–60, 379–386 (2004).
13. Li, A. et al. Micro-optical sectioning tomography to obtain a highresolution atlas of the mouse brain. Science 330, 1404–1408 (2010).
14. Tsai, P. et al. All-optical histology using ultrashort laser pulses. Neuron
39, 27–41 (2003).
15. Ragan, T. et al. Serial two-photon tomography for automated ex vivo
mouse brain imaging. Nat. Methods 9, 255–258 (2012).
16. Denk, W. & Horstmann, H. Serial block-face scanning electron microscopy
to reconstruct three-dimensional tissue nanostructure. PLoS Biol. 2, e329
(2004).
17. Micheva, K.D. & Smith, S.J. Array tomography: a new tool for imaging the
molecular architecture and ultrastructure of neural circuits. Neuron 55,
25–36 (2007).
18. Osten, P. & Margrie, T.W. Mapping brain circuitry with a light microscope.
Nat. Methods 10, 515–523 (2013).
19. Livet, J. et al. Transgenic strategies for combinatorial expression
of fluorescent proteins in the nervous system. Nature 450, 56–62
(2007).
20. Tsai, P. et al. Correlations of neuronal and microvascular densities in
murine cortex revealed by direct counting and colocalization of cell nuclei
and microvessels. J. Neurosci. 18, 14553–14570 (2009).
21. Kleinfeld, D. et al. Large-scale automated histology in the pursuit of
connectomes. J. Neurosci. 31, 16125–16138 (2011).
22. Bock, D.D. et al. Network anatomy and in vivo physiology of visual
cortical neurons. Nature 471, 177–182 (2011).
23. Briggman, K.L., Helmstaedter, M. & Denk, W. Wiring specificity in the
direction-selectivity circuit of the retina. Nature 471, 183–188 (2011).
24. Dodt, H., Leischner, U. & Schierloh, A. Ultramicroscopy: three-dimensional
visualization of neuronal networks in the whole mouse brain. Nat. Methods
4, 331–336 (2007).
25. Ertürk, A., Mauch, C., Hellal, F. & Förstner, F. Three-dimensional imaging
of the unsectioned adult spinal cord to assess axon regeneration and glial
responses after injury. Nat. Med. 18, 166–171 (2012).
26. Hama, H. et al. Scale: a chemical approach for fluorescence imaging and
reconstruction of transparent mouse brain. Nat. Neurosci. 14, 1481–1488
(2011).
27. Keller, P.J. et al. Fast, high-contrast imaging of animal development with
scanned light sheet-based structured-illumination microscopy. Nat.
Methods 7, 637–642 (2010).
28. Truong, T.V., Supatto, W., Koos, D.S., Choi, J.M. & Fraser, S.E. Deep and
fast live imaging with two-photon scanned light-sheet microscopy.
Nat. Methods 8, 757–760 (2011).
perspective
29. Tomer, R., Khairy, K., Amat, F. & Keller, P.J. Quantitative high-speed
imaging of entire developing embryos with simultaneous multiview
light-sheet microscopy. Nat. Methods 9, 755–763 (2012).
30. Krzic, U., Gunther, S., Saunders, T.E., Streichan, S.J. & Hufnagel, L.
Multiview light-sheet microscope for rapid in toto imaging. Nat. Methods
9, 730–733 (2012).
31. Keller, P.J., Schmidt, A.D., Wittbrodt, J. & Stelzer, E.H.K. Reconstruction
of zebrafish early embryonic development by scanned light sheet
microscopy. Science 322, 1065–1069 (2008).
32. Planchon, T.A. et al. Rapid three-dimensional isotropic imaging of living
cells using Bessel beam plane illumination. Nat. Methods 8, 417–423
(2011).
33. Dani, A., Huang, B., Bergan, J., Dulac, C. & Zhuang, X. Superresolution
imaging of chemical synapses in the brain. Neuron 68, 843–856 (2010).
34. Micheva, K.D., Busse, B., Weiler, N.C., O’Rourke, N. & Smith, S.J.
Single-synapse analysis of a diverse synapse population: proteomic
imaging methods and markers. Neuron 68, 639–653 (2010).
35. Wählby, C. & Erlandsson, F. Sequential immunofluorescence staining and
image analysis for detection of large numbers of antigens in individual
cell nuclei. Cytometry 47, 32–41 (2002).
36. Feinberg, E.H. et al. GFP reconstitution across synaptic partners (GRASP)
defines cell contacts and synapses in living nervous systems. Neuron 57,
353–363 (2008).
37. Wickersham, I.R. & Feinberg, E.H. New technologies for imaging synaptic
partners. Curr. Opin. Neurobiol. 22, 121–127 (2012).
38. Gordon, M.D. & Scott, K. Motor control in a Drosophila taste circuit.
Neuron 61, 373–384 (2009).
39. Kim, J. et al. mGRASP enables mapping mammalian synaptic connectivity
with light microscopy. Nat. Methods 9, 96–102 (2012).
40. Wickersham, I. et al. Retrograde neuronal tracing with a deletion-mutant
rabies virus. Nat. Methods 4, 47–49 (2007).
41. Miyamichi, K. et al. Cortical representations of olfactory input by
trans-synaptic tracing. Nature 472, 191–196 (2011).
42. Liu, X. et al. Optogenetic stimulation of a hippocampal engram activates
fear memory recall. Nature 484, 381–385 (2012).
43. Guenthner, C.J., Miyamichi, K., Yang, H., Heller, H.C. & Luo, L. Permanent
genetic access to transiently active neurons using targeted recombination
in active populations (TRAP). Neuron (in the press).
44. Godement, P., Vanselow, J., Thanos, S. & Bonhoeffer, F. A study in
developing visual systems with a new method of staining neurones and
their processes in fixed tissue. Development 101, 697–713 (1987).
45. Helmstaedter, M. Cellular-resolution connectomics: challenges of dense
neural circuit reconstruction. Nat. Methods 10, 501–507 (2013).
46. Dombeck, D.A., Graziano, M.S. & Tank, D.W. Functional clustering of
neurons in motor cortex determined by cellular resolution imaging in
awake behaving mice. J. Neurosci. 29, 13751–13760 (2009).
47. Ahrens, M.B. et al. Brain-wide neuronal dynamics during motor adaptation
in zebrafish. Nature 485, 471–477 (2012).
48. Ziv, Y. et al. Long-term dynamics of CA1 hippocampal place codes.
Nat. Neurosci. 16, 264–266 (2013).
49. Ahrens, M.B., Orger, M.B., Robson, D.N., Li, J.M. & Keller, P.J. Whole-brain
functional imaging at cellular resolution using light-sheet microscopy.
Nat. Methods 10, 413–420 (2013).
50. Alivisatos, P. et al. The brain activity map. Science 339, 1284–1285 (2013).
nature methods | VOL.10 NO.6 | JUNE 2013 | 513
FOCUS ON MAPPING THE BRAINreview
Mapping brain circuitry with a light
microscope
npg
© 2013 Nature America, Inc. All rights reserved.
Pavel Osten1 & Troy W Margrie2,3
The beginning of the 21st century has seen a renaissance in light microscopy and
anatomical tract tracing that together are rapidly advancing our understanding of the
form and function of neuronal circuits. The introduction of instruments for automated
imaging of whole mouse brains, new cell type–specific and trans-synaptic tracers, and
computational methods for handling the whole-brain data sets has opened the door
to neuroanatomical studies at an unprecedented scale. We present an overview of the
present state and future opportunities in charting long-range and local connectivity in
the entire mouse brain and in linking brain circuits to function.
Since the pioneering work of Camillo Golgi and
Santiago Ramón y Cajal at the turn of the last
century1,2, advances in light microscopy (LM) and
neurotracing methods have been central to the progress
in our understanding of anatomical organization in
the mammalian brain. The Golgi silver-impregnation
method allowed the visualization of neuron morpho­
logy, providing the first evidence for cell type–based
and connectivity-based organization in the brain. The
introduction of efficient neuroanatomical tracers in
the second half of the last century greatly increased
the throughput and versatility of neuronal projection mapping, which led to the identification of many
anatomical pathways and circuits, and revealed the basic
principles of hierarchical and laminar connectivity
in sensory, motor and other brain systems3,4.
The beginning of this century has seen a methodsdriven renaissance in neuroanatomy, one that is distinguished by a focus on large-scale projects that yield
unprecedented amounts of anatomical data. Instead of
the traditional ‘cottage-industry’ approach to studying
one anatomical pathway at a time, the new projects aim
to generate complete data sets—so-called projectomes
and connectomes—that can be used by the scientific
community as resources for answering specific experimental questions. These efforts range in scale and resolution from the macroscopic to the microscopic: from
studies of the human brain by magnetic resonance
imaging to reconstructions of dense neural circuits in
small volumes of brain tissue by electron microscopy
(see Review5 and Perspective6 in this Focus).
Advancements in LM methods, the focus of this
Review, are being applied to the mapping of point-topoint connectivity between all anatomical regions in
the mouse brain by means of sparse reconstructions of
anterograde and retrograde tracers7. Taking advantage
of the automation of LM instruments, powerful dataprocessing pipelines, and combinations of traditional
and modern viral vector-based tracers, teams of scientists at Cold Spring Harbor Laboratory (CSHL), Allen
Institute for Brain Science (AIBS) and University of
California Los Angeles (UCLA) are racing to complete a connectivity map of the mouse brain—dubbed
the ‘mesoscopic connectome’—which will provide the
scientific community with online atlases for viewing
entire anatomical data sets7. These efforts demonstrate
the transformative nature of today’s LM-based neuroanatomy studies and the astonishing speed with which
large amounts of data can be disseminated online, and
have an immediate impact on research in neuroscience
laboratories around the world.
As the mouse mesoscopic connectomes are being
completed, it is clear that LM methods will continue
to impact the evolution of biological research and specifically neuroscience: new trans-synaptic viral tracers
are being engineered to circumvent the need to resolve
synapses, which has constrained the interpretation of
cell-to-cell connectivity in LM studies, and new assays
1Cold Spring Harbor Laboratory, Cold Spring Harbor, New York, USA. 2Department of Neuroscience, Physiology and Pharmacology,
University College London, London, UK. 3Division of Neurophysiology, The Medical Research Council National Institute for Medical
Research, London, UK. Correspondence should be addressed to P.O. ([email protected]) or T.W.M. ([email protected]).
Received 2 march; accepted 15 April; published online 30 may 2013; doi:10.1038/nmeth.2477
nature methods | VOL.10 NO.6 | JUNE 2013 | 515
review
b
Ti:sapphire
PMT
c
PMT
PMT
Camera
Piezo element
Microtome
y
Camera
515-nm laser
Detection objective
Illumination objective
Knife
Light sheet
y
x
y
x
x
z
z
npg
© 2013 Nature America, Inc. All rights reserved.
Motorized x-y-z stage
Motorized x-y-z stage
Motorized x-y-z stage
z
J. Kuhl
a
FOCUS ON MAPPING THE BRAIN
Figure 1 | Whole-brain LM methods. (a) In STP tomography, a two-photon microscope is used to image the mouse brain in a coronal plane in a mosaic grid
pattern, and a microtome sections off the imaged tissue. Piezo objective scanner can be used for z-stack imaging (image adapted from ref. 15). (b) In fMOST,
confocal line-scan is used to image the brain as 1-micrometer thin section cut by a diamond knife (image adapted from ref. 16). (c) In LSFM, the cleared brain
is illuminated from the side with a light sheet (blue) through an illumination objective (or cylinder lens 19) and imaged in a mosaic grid pattern from the top
(image adapted from ref. 20). In all instruments, the brain is moved under the objective on a motorized x-y-z stage; PMT, photomultiplier tube.
combining anatomical and functional measurements are being
applied to bridge the traditional structure-function divide in the
study of the mammalian brain. In this Review, we aim to provide
an overview of today’s state of the art in LM instrumentation and
to highlight the opportunities for progress as well as the challenges
that need to be overcome to transform neuronal-tracing studies
into a truly quantitative science that yields comprehensive descriptions of long-range and local projections and connectivity in whole
mouse brains. We also discuss present strategies for the integration
of anatomy and function in the study of mouse brain circuits.
Automated light microscopes for whole-brain imaging
The field of neuroanatomy has traditionally been associated with
labor-intensive procedures that greatly limit the throughput of
data collection. Recent efforts to automate LM instrumentation
have standardized and dramatically increased the throughput
of anatomical studies. The main challenge for these methods is
to maintain the rigorous quality of traditional neuroanatomical
studies, which results from detailed visual analysis, careful data
collection and expert data interpretation.
There are currently two approaches to the automation of LM
for imaging three-dimensional (3D) whole-brain data sets: one
based on the integration of block-face microscopy and tissue sectioning and the other based on light-sheet fluorescence micros­
copy (LSFM) of chemically cleared tissue. The first approach has
been developed for wide-field imaging, line-scan imaging, confocal microscopy and two-photon microscopy8–16. Common to
all these instruments is the motorized movement of the sample
under the microscope objective for top-view mosaic imaging, followed by mechanical removal of the imaged tissue before the next
cycle of interleaved imaging and sectioning steps (Fig. 1a,b). As
the objective is always near the tissue surface, it is possible to use
high-numerical-aperture lenses to achieve submicrometer resolution close to the diffraction limits of LM.
516 | VOL.10 NO.6 | JUNE 2013 | nature methods
Three instruments have been designed that combine twophoton microscopy17 with subsequent tissue sectioning by
ultrashort laser pulses in all-optical histology10, by a milling
machine in two-photon tissue cytometry12 or by a vibrating blade
microtome in serial two-photon (STP) tomography15 (Fig. 1a).
Whereas in both all-optical histology and two-photon tissue
cytometry the sectioning obliterates the imaged tissue, the integration of vibratome-based sectioning in STP tomography allows the
collection of the cut tissue for subsequent analysis by, for example,
immunohistochemistry (see below). In addition, the tissue preparation by simple formaldehyde fixation and agar embedding in
STP tomography has minimal detrimental effects on fluorescence
and brain morphology. This makes STP tomography applicable
to a broad range of neuroanatomical projects that use genetically
encoded fluorescent protein–based tracers, which are sensitive to
conditions used for fixation, dehydration and tissue clearing. This
method is also versatile in terms of the mode and resolution of
data collection. For example, imaging the mouse brain as a data
set of 280 serial coronal sections, evenly spaced at 50 micrometers
and at x-y resolution of 1 micrometer, takes about ~21 hours and
generates a brain atlas–like data set of ~70 gigabytes. A complete
visualization can be achieved by switching to 3D scanning of
z-volume stacks between the mechanical sectioning steps, which
allows the entire mouse brain to be imaged, for instance, at
1-micrometer x-y resolution and 2.5-micrometer z resolution in
~8 days, generating ~1.5 terabytes of data15. The instrument is
commercially available from TissueVision Inc. The Allen Brain
Institute is using this methodology for its Mouse Connectivity
project (see below).
Two instruments have been designed to combine bright-field linescan imaging and ultramicrotome sectioning of resin-embedded
tissue in methods named knife-edge scanning microscopy (known
as KESM)13 and micro-optical sectioning tomography (MOST)14
(Fig. 1b). The latter was used to image Golgi-stained mouse brain
npg
© 2013 Nature America, Inc. All rights reserved.
FOCUS ON MAPPING THE BRAINreview
at 0.33 × 0.33 × 1.0 micrometer x-y-z resolution, generating >8
terabytes of data in ~10 days13,14. The MOST instrument design
was also recently built for fluorescence imaging (fMOST) by
confocal laser scanning microscopy, with the throughput of one
mouse brain at 1.0-micrometer voxel resolution in ~19 days16.
Knife-edge scanning microscopy imaging is now also available
as a commercial service from 3Scan.
The second approach for automated whole-brain imaging
is based on LSFM (also known as selective-plane illumination
microscopy18 and ultramicroscopy19; Fig. 1c). This approach
allows fast imaging of chemically cleared ‘transparent’ mouse
brains without the need for mechanical sectioning19,20 but,
at least until now, with some trade-offs for anatomical tracing applications. The chemical clearing procedures reduced
the signal of fluorescent proteins, but this problem appears to
be solved by a new hydrogel-based tissue transformation and
clearing method termed CLARITY21 (see Perspective about this
methodology in this Focus22). The spatial resolution of LSFM
for the mouse brain also has been limited by the requirement for
large field-of-view objectives with low power and low numerical aperture that were used to visualize the whole brain19,23.
However, new objectives with long working distance and high
numerical aperture, such as 8-millimeter working distance and
0.9 numerical aperture objective from Olympus, promise to enable LSFM of the whole mouse brain at submicrometer resolution.
If necessary, LSFM can also be combined with one of several
forms of structured illumination to reduce out-of-focus background fluorescence and improve contrast24–26. Taken together,
these modifications are likely to enhance the applicability of
LSFM to anterograde tracing of thin axons at high resolution
in the whole mouse brain, as done by STP tomography in the
AIBS Mouse Connectivity project (see below) and by fMOST
in a recent report16. In addition, LSFM is well-suited for retrograde tracing in the mouse brain, which relies on detection
of retrogradely fluorescence-labeled neuronal soma that are
typically >10 micrometers in diameter. Such application was
recently demonstrated for mapping retrograde connectivity of
granule cells of the mouse olfactory bulb20 using rabies viruses
that achieve high levels of fluorescent protein labeling27,28.
Mesoscopic connectivity-mapping projects
The labeling of neurons and subsequent neuroanatomical
tract tracing by LM methods has been used for over a century
to interrogate the anatomical substrate of the transmission of
information in the brain. Throughout those years, the credo of
neuroanatomy, ‘the gain in brain is mainly in the stain’, signified
that progress was made mainly through the development of new
anatomical tracers. Yet despite the decades of neuroanatomical
research, the laborious nature of tissue-processing and datavisualization has kept the progress in our knowledge of brain
circuitry at a disappointingly slow pace7. Today, neuroanatomy
stands to greatly benefit from the application of high-throughput
automated LM instruments and powerful informatics tools for the
analysis of mouse brain data29,30. The high-resolution capacity
LM methods afford, and the fact that an entire brain data
set can be captured, makes these systems well-suited for the
systematic charting of the spatial profile and the connectivity of
populations of neurons and even individual cells projecting over
long distances.
The pioneering effort in the field of anatomical projects applied
at the scale of whole animal brains was the Allen Mouse Brain
Atlas of Gene Expression, which cataloged in situ hybridization
maps for more than 20,000 genes in an online 3D digital mouse
brain atlas29,31,32. The proposal by a consortium of scientists led
by Partha Mitra to generate similar LM-based atlases of ‘brainwide neuroanatomical connectivity’ in several animal models7 has
in short time spurred three independent projects, each promising to trace all efferent and afferent anatomical pathways in the
mouse brain. The aim of the Mouse Brain Architecture Project
(http://brainarchitecture.org/) from the Mitra team at CSHL
is to image >1,000 brains; the Allen Mouse Brain Connectivity
Atlas project (http://connectivity.brain-map.org/) led by Hongkui
Zeng at AIBS has a goal of imaging >2,000 brains; and the Mouse
Connectome Project (http://www.mouseconnectome.org/) led
by Hong-Wei Dong at UCLA has a goal of imaging 500 brains,
with each brain injected with four tracers. Whereas the CSHL
and UCLA projects rely on automated wide-field fluorescence
microscopy (Hamamatsu Nanozoomer 2.0 and Olympus VS110,
respectively) to image manually sectioned brains, the Mouse
Connectivity project at the Allen Institute is being done entirely
by STP tomography15. The main strength of these efforts is in the
broad range of tracers used. Given that each tracer has its own
advantages and problems33, the information derived from all three
projects will ensure generalizable interpretation of the projection
results throughout the brain. The CSHL group uses a combination
of traditional anterograde and retrograde tracers, fluorophoreconjugated dextran amine34 and cholera toxin B (CTB) subunit35,
respectively, which are complemented by a combination of viral
vector–based tracers, GFP-expressing adeno-associated virus
(AAV)36 for anterograde tracing (Fig. 2a) and modified rabies
virus27 for retrograde tracing. Although the virus-based methods
are less well tested, they offer advantages in terms of the brightness of labeling and the possibility of cell type–specific targeting
using Cre recombinase–dependent viral vectors37 and transgenic lines expressing Cre recombinase from cell type–specific
promoters38–40. The AIBS team uses solely anterograde tracing
by AAV-GFP viruses that label axonal arborizations with GFP 41
(Fig. 2b), in many cases taking advantage of transgenic ‘driver’
mouse lines expressing Cre recombinase from cell type–specific
promoters to achieve anterograde tracing of specific neuronal cell
types. Finally, the team at UCLA is using a strategy of two injections per brain, each with a mix of anterograde and retrograde
tracers42: CTB together with Phaseolus vulgaris leucoagglutinin43
and FluoroGold44 together with biotinylated dextran amine42,45.
This approach has an added advantage of enabling direct visualization of the convergence of inputs and outputs from across
different areas in one brain42,46,47.
The unprecedented amounts of data being collected in these
projects means that the considerable person-hours historically
spent performing microscopy have largely shifted toward data
analysis. The first step of such data analysis comprises the compilation of the serial section images for viewing as whole-brain data
sets at resolutions beyond the minimum geometric volume of the
neuronal structures of interest: soma for retrograde tracing and
axons for anterograde tracing. All three projects offer a convenient way to browse the data sets online, including high-resolution
magnified views that in most cases are sufficient to visually determine labeled soma and axons. All three projects use the Allen
nature methods | VOL.10 NO.6 | JUNE 2013 | 517
review
a
npg
© 2013 Nature America, Inc. All rights reserved.
b
Figure 2 | Primary motor cortex projection maps. (a) Mouse Brain
Architecture data of AAV-GFP injected into the supragranular layers
and AAV–red fluorescent protein injected in the infragranular layers
(F. Mechler and P. Mitra; unpublished data). Front (left) and lateral (right)
views of the volume-rendered brain (top); and coronal section image from
the area marked by the dashed line (center) with magnification of the
lower boxed region showing axonal fibers in the cerebral peduncle (left)
and magnification of the upper boxed region showing projections to the
midbrain reticular nucleus (right). Scale bars, 1,000 µm (top) and 20 µm
(bottom). (b) Mouse Connectivity data of a similar AAV-GFP injection
show the primary motor cortex projectome reconstructed in the Allen
Brain Explorer48 (H. Zeng; unpublished data). Inset, magnified view and
coronal section overview of projections in the ventral posteromedial (VPM)
nucleus of the thalamus.
Mouse Brain Atlas for the registration of the coronal sections,
which will help in the cross-validation of results obtained from
the different tracers. The Allen Mouse Brain Connectivity Atlas
website also offers the option to view the data after projection
segmentation, which selectively highlights labeled axons, as well
as in 3D in the Brain Explorer registered to the Allen Mouse Brain
Atlas48 (Fig. 2b).
The second step of data analysis requires the development
of informatics methods for quantitation of the data sets, which
will facilitate the interpretation of the data available online. The
Allen Mouse Brain Connectivity Atlas online tools allow the user
to search the projections between injected regions and display
the labeled pathways as tracks in three dimensions in the Brain
Explorer. The CSHL and UCLA connectomes can currently be
viewed online as serial section data sets. The data from the Cre
518 | VOL.10 NO.6 | JUNE 2013 | nature methods
FOCUS ON MAPPING THE BRAIN
recombinase driver mouse lines in the AIBS project provide a
unique feature of cell-type specificity for the interpretation of
the anterograde projections. The main strength of the CSHL
and UCLA efforts lies in the multiplicity of the anatomical tracers used. The use of multiple retrograde tracers in ­ particular
will yield useful information, as retrogradely labeled soma
(>10 micrometers in diameter) are easier to quantify than thin
(<1 micrometer) axon fibers. These experiments will also provide an important comparison between the traditional CTB and
FluoroGold tracers and the rabies virus tracer that is also being
used in trans-synaptic labeling (see below) but is less well studied
and may show some variation in transport affinity at different
types of synapses. In summary, the LM-based mesoscopic mapping projects are set to transform the study of the circuit wiring
of the mouse brain by providing online access to whole-brain data
sets from several thousand injections of anterograde and retrograde tracers. The informatics tools being developed to search the
databases will greatly aid in parsing the large amounts of data and
in accessing specific brain samples for detailed scholarly analyses
by the neuroscience community.
Mapping connectivity using trans-synaptic tracers
In contrast to electron microscopy methods, which provide a
readout of neuronal connectivity with synapse resolution over
small volumes of tissue, the whole-brain LM methods permit
the assessment of projection-based connectivity between brain
regions and in some cases between specific cell types in those
regions but without the option of visualizing the underlying synaptic contacts. Trans-synaptic viruses that cross either multiple
or single synapses can help to circumvent the requirement to confirm connectivity at resolution achieved by electron microscopy
because such connectivity may be inferred from the known direction and mechanism of spread of the trans-synaptic tracer. Transsynaptic tracers based on rabies virus, pseudorabies virus and
herpes simplex virus, which repeatedly cross synaptic connections
in a retrograde or anterograde direction, are powerful tools for
studying multistep pathways upstream and downstream from the
starter cell population49–51. Furthermore, modified trans-synaptic
rabies viruses have been developed that are restricted in their
spread to a single synaptic jump and thus can be used to identify monosynaptic connections onto and downstream of specific
neuronal populations and even individual cells27,52–58 (Fig. 3).
Rabies virus spreads from the initially infected cells in a transsynaptic retrograde manner49,59. Rabies virus infection does not
occur via spurious spread or uptake by fibers of passage and,
because it cannot cross via electrical synapses, it is an effective
tool for unidirectional anatomical tracing60. In a modified rabies
virus system, the infection can also be cell type–targeted by encapsulating glycoprotein-deficient rabies virus with an avian virus
envelope protein (referred to as ‘SAD-∆G-EnvA’). This restricts
infection to only those cells that express an avian receptor protein TVA that is natively found in birds but not in mammals61,62.
Thus, the delivery of vectors driving the expression of both TVA
and rabies virus glycoprotein (RV-G) into a single cell28,54,56
(see below) or a specific population of cells55,63, ensures that only
the targeted cell or cells will (i) be susceptible to initial infection
and (ii) provide the replication-incompetent virus with RV-G
required for trans-synaptic infection64. In this system, the virus can
spread from the primarily infected cell or cells to the ­presynaptic
FOCUS ON MAPPING THE BRAINreview
npg
© 2013 Nature America, Inc. All rights reserved.
Figure 3 | Mapping the function and
a
b
c
Action potential response 2 Hz
connectivity of single cells in the mouse brain
Synaptic response 24 V × s
in vivo. (a) Experimental setup for combined
0
315
45
10
mV
single-cell physiology and trans-synaptic
100 ms
connectivity mapping. Patch pipettes have
270
90
TVA RV-G
Rabies virus
solutions containing DNA vectors used to drive
135
225
–78 mV
the expression of the TVA and RV-G proteins.
Synaptic tuning
Local synaptic
Intrinsic properties
(b) Patch pipettes are used to perform a
connectivity
20 mV
V2ML
whole-cell recording of the intrinsic and
V1
200 ms
330°
sensory-evoked synaptic properties of a
V2L
300°
single layer-5 neuron in primary visual cortex.
240°
Synaptic responses are averages of five traces.
210°
DLG
(c) Then the encapsulated modified rabies virus
150°
DLG
120°
is injected into the brain in close proximity
60°
to the recorded neuron. After a period of up
30°
to 12 days to ensure retrograde spread of the
Retrograde labeling of local
Synaptic properties
and long-range inputs
modified rabies from the recorded neuron, the
brain is removed and imaged for identification
of the local and long-range presynaptic inputs underlying the tuning of the recorded neuron to the direction of visual motion (polar plot). Fluorescence
image on the right shows injection site with the starter cell (yellow) in the middle. Scheme (bottom left) and imaged data (bottom right) of long-range
retrograde tracing. V1, primary visual cortex; V2L, secondary visual cortex (lateral); V2ML secondary visual cortex (medio-lateral). Inset shows long-range
inputs from the dorsal lateral geniculate nucleus (DLG). Scale bars, 300 µm (top) and 50 µm (bottom), respectively. Images modified from ref. 54.
input cells, which become labeled by expression of a fluorescent
protein. However, as the presynaptic cells do not express RV-G28,
the virus cannot spread further. This approach thus allows the discovery of the identity and location of the upstream input network
relative to a defined population of neurons57,58.
Brain-region specificity and cell-type specificity for mapping connectivity by the modified rabies virus system can be
achieved by using a Cre recombinase–dependent helper virus
driving expression of TVA and RV-G and transgenic driver
mouse lines that express Cre recombinase in specific cell types
or cortical layers38,39,63. This strategy is particularly useful for
brain regions comprising many different cell types that could
not be otherwise selectively targeted. Moreover, the engineering
of other neurotropic trans-synaptic viruses is adding new tools
for anatomical tracing, including Cre recombinase–dependent
anterograde tracers based on a modified H129 strain of herpes
simplex virus65 and vesicular stomatitis virus66, and retrograde
tracers based on a modified pseudorabies virus (H. Oyibo and
A. Zador, personal communication). The use of retrograde and
anterograde trans-synaptic viruses, in combination with wholebrain LM methods, thus promises to afford unprecedented
access to the upstream and downstream connectivity of specific
cell types in the mouse brain.
Present challenges and opportunities for whole-brain LM
As highlighted above, LM instruments for whole-brain imaging are
expected to make a substantial contribution in large-scale projects
that focus on anatomical connectivity at the level of the whole
mouse brain. It has also become clear that the use of these instruments will have an impact in many experimental applications in
different neuroscience laboratories. It is therefore imperative that
there exist broadly applicable image-processing, warping and analytical tools that will facilitate data sharing and across-­laboratory
collaboration and validation in future neuroscience studies
focusing on, for example, mapping whole-brain anatomical
changes during development and in response to experience.
One practical problem arising from the choice to scan entire
mouse brains at high resolution relates to the handling of large
data sets (up to several terabytes per brain), which necessitates
automated analytical pipelining. STP tomography is currently
the most broadly used method among the whole-brain LM
approaches, and there are freely available informatics tools for
compiling STP tomography image stacks and viewing them as
three-dimensional data, including algorithms that automate
seamless stitching15. Another key challenge for charting the distribution of the labeled elements in the whole mouse brain is
the process of accurate registration of the individual brain data
sets onto an anatomical reference atlas. To this end, scientists at
AIBS have generated the open-source segmented Allen Mouse
Brain Atlas for the adult C57BL/6 mouse29,31,32,48, which is also
available for registration of data sets generated by STP tomography (Figs. 2b and 4b). In addition, the so-called Waxholm space
for standardized digital atlasing67 allows comparisons of registered mouse brain data using multiple brain atlases, including the
Allen Mouse Brain Atlas, the digital Paxinos and Franklin Mouse
Brain Atlas68, and several magnetic resonance imaging reference
mouse brains. The continuing development of the Waxholm space
and other online data-analysis platforms30,69,70 will be essential
for standardized comparisons of mouse brain data collected in
different laboratories using different instruments.
The completion of the three mesoscopic connectome projects
in the next several years will yield a comprehensive map of pointto-point connectivity between anatomical regions in the mouse
brain7. Determining the cell-type identity of the neurons sending and receiving the connections in the brain regions will be
essential for interpreting the function of the brain-wide neural
circuits. Immunohistochemical analyses of labeled circuits have
proven invaluable for ascertaining the identity of specific classes
of neurons71–73 and synaptic connections52,74. The combination
of immunohistochemical analysis by array tomography75,76 and
anatomical tracing by the whole-brain LM instruments promises
to be particularly powerful, as it will bring together two largely
automated methodologies with complementary focus on synaptic
and mesoscopic connectivity, respectively. STP tomography outputs sectioned tissue (typically 50-micrometer-thick sections15),
which can be further resectioned, processed and reimaged by array
nature methods | VOL.10 NO.6 | JUNE 2013 | 519
review
npg
© 2013 Nature America, Inc. All rights reserved.
Figure 4 | Imaging induction of c-fos as a
means to map whole-brain activation.
(a) A 3D visualization of 367,378 c-fos–GFP
cells detected in 280 coronal sections of an STP
tomography data set of a mouse brain after the
mouse was allowed to explore a novel object
for 90 seconds. (b) Examples of anatomical
segmentation of the brain volume with the
Allen Mouse Brain Reference Atlas labels48
modified for the 280-section STP tomography
data sets: hippocampus (blue), prelimbic cortex
(aqua), infralimbic cortex (orange) and piriform
cortex (green). (c) Visualization of c-fos–GFP
cells in the hippocampus (38,170 cells),
prelimbic cortex (3,305 cells), infralimbic
cortex (3,827 cells) and piriform cortex
(10,910 cells) (P.O., Y. Kim and K. Umadevi
Venkataraju; unpublished data).
FOCUS ON MAPPING THE BRAIN
a
tomography for integrating cell type–specific information into the
whole-brain data sets. Industrial-level automation of slice capture
and immunostaining can be developed to minimize manual handling and enhance the integration of immunohistochemistry and
STP tomography. In addition, sectioning and immunostaining
can also be applied to LSFM-imaged mouse brains20.
A related, cell type–focused application of whole-brain LM
imaging will be to quantitatively map the distribution (the cell
counts) of different neuronal cell types in all anatomical regions
in the mouse brain. Several such cell count–based anatomical
studies have been done previously at smaller scales, revealing,
for example, cell densities with respect to cortical vasculature77
or the density of neuronal cell types per layer in a single cortical
column78–80. Using the whole-brain LM methods, a comprehensive anatomical atlas of different GABAergic inhibitory interneurons81 can now be generated by imaging cell type–specific Cre
recombinase–mediated knock-in mouse lines38,39 crossed with
Cre recombinase–dependent reporter mice expressing nuclear
GFP. These and similar data sets for other neuronal cell types
will complement the mesoscopic brain region connectivity data
and help the interpretation of the immunohistochemistry data by
providing a reference for total numbers of specific cell types per
anatomical brain region.
Integrating brain anatomy and function
The anterograde, retrograde and trans-synaptic tracing approaches
described above will yield the structural scaffold of anatomical
projections and connections throughout the mouse brain.
However, such data will not be sufficient to identify how
specific brain regions connect to form functional circuits driving
different behaviors. Bridging whole-brain structure and function is the next frontier in systems neuroscience, and the
develop­ment of new technologies and methods will be crucial in
achieving progress.
The structure-function relationship of single neurons can be
examined by in vivo intracellular delivery of the DNA vectors
required for targeting and driving trans-synaptic virus expression
via patch pipettes in loose cell-attached mode for electroporation56
or via whole-cell recording54 (Fig. 3). Used in combination with
two-photon microscopy, this single-cell delivery technique may
also be targeted at fluorescently labeled neurons of specific cell
types56,82,83. The whole-cell method is particularly ­informative, as
520 | VOL.10 NO.6 | JUNE 2013 | nature methods
b
c
its intracellular nature permits recording the intrinsic ­biophysical
profile of the target cell, which, in turn, may reflect its functional connectivity status in the local network 84. In ­ addition,
by recording sensory-evoked inputs, it is possible to compare
single-cell synaptic receptive fields and anatomical local and
long-range connectivity traced by LM methods54. This combinatorial approach, involving single-cell electrophysiology and
genetic manipulation designed for connection mapping, makes
it possible to test long-standing theories regarding the extent to
which emergent features of sensory cortical function manifest via
specific wiring motifs85.
As has recently been achieved for serial electron microscopy–
based reconstruction86,87, it will also be valuable to functionally
characterize larger local neuronal populations for registration
against LM-based connectivity data. In this sense, genetically
encoded calcium indicators, which permit physiological characterization of neuronal activity in specific cell types88–90, along with
viral vectors for trans-synaptic labeling and LM-based tracing,
will have critical complementary roles. Large-volume in vivo twophoton imaging of neuronal activity before ex vivo whole-brain
imaging will establish the extent to which connectivity patterns
relate to function91 at the level of single cells, and local and longrange circuits. Interpolation of such experiments will rely on the
ability to cross-register in vivo functional imaging with complete
ex vivo LM connectivity data. Preliminary experiments, which
already hint at the spatial spread of monosynaptic connectivity
of individual principal cortical cells, suggest that combination
of functional imaging and traditional anatomical-circuit reconstruction may only be feasible at the local network level where
connection probability is the highest92–94. Given the broad, sparse
expanse of connectivity in most brain regions and especially in
cortical areas, high-throughput whole-brain LM methods will
be imperative for complete anatomical-circuit reconstruction of
functionally characterized local networks.
The amalgamation of whole-brain LM and physiological methods for single neurons and small networks offers a powerful means
to study the mouse brain. An exciting application of this approach
will be to trace the synaptic circuits of neurons functionally characterized in head-fixed behaving animals engaged in tasks related
to spatial navigation, sensorimotor integration and other complex brain functions95–97. This research will lead to the generation of whole-brain structure-function hypotheses for specific
npg
© 2013 Nature America, Inc. All rights reserved.
FOCUS ON MAPPING THE BRAINreview
behaviors, which can then be tested for causality by optogenetic
methods ­targeted to the identified cell types and brain regions98.
Furthermore, LM, physiological and optogenetic methods can be
applied to interrogate entire brain systems in large-scale projects,
as is currently being done for the mouse visual cortex in an effort
led by C. Koch and R.C. Reid at AIBS99.
Finally, a discussion in the neuroscience community has been
initiated regarding the feasibility of mapping activity at cellular
resolution in whole brains and linking the identified activity patterns to brain anatomy100. Today, such experiments are possible
in small, transparent organisms, as was demonstrated by twophoton microscopy and LSFM-based imaging of brain activity in
larval zebrafish expressing the calcium indicator GCaMP89,101,102.
Understandably, LM-based approaches will not be useful for
in vivo whole-brain imaging in larger, nontransparent animals,
and the invention of new disruptive technologies will likely be
needed to achieve the goal of mapping real-time brain activity
at cellular resolution in, for example, the mouse. In contrast, LM
methods can be used to map patterns of whole-brain activation
indirectly, by post-hoc visualization of activity-induced expression of immediate early genes, such as mouse Fos (c-fos), Arc or
Homer1a (ref. 103). Transgenic fluorescent immediate early gene
reporter mice, such as c-fos–GFP or Arc-GFP mice104–106, can be
trained in a specific behavior, their brains subsequently can be
imaged ex vivo, and the exact distribution of GFP-positive neurons
can be mapped and analyzed by computational methods (Fig. 4).
In this approach, a statistical analysis of the counts of GFP-labeled
neurons can be used to identify brain regions and cell types activated during behaviors but without providing information on the
temporal sequence of brain region activation or the firing patterns
of the activated cells. However, the development of more sensitive,
for instance, fluorescent RNA–based methods, may allow calibration of the cellular signal with respect to the temporal window
and the pattern of activity related to the induction of immediate early genes. Such calibration would considerably enhance
the power of LM-based whole-brain mapping of the induction
of immediate early genes, which, in combination with the connectomics data, could then be used to begin to build cellularresolution models of function-based whole-brain circuits.
Conclusions
The advances in automated LM methods, anatomical tracers,
physio­logical methods and informatics tools have begun to transform our understanding of the circuit wiring in the mouse brain.
The focus on the mouse as an animal model is, of course, not accidental. In addition to the generation of cell type–specific driver
mouse lines38–40 that allow the study of specific neuronal populations in the normal brain, mouse genetics are used in hundreds of
laboratories to model gene mutations linked to heritable human
disorders, including complex cognitive disorders such as autism
and schizophrenia. Without a doubt, understanding the relationships between brain structure and function in the genetic mouse
models will be crucial to understanding the underlying brain circuit mechanisms of these disorders. The toolbox of LM methods
described here, and the continuing development of new methods,
promise to transform the study of brain circuits in animal models
and to decipher the structure-function relationships essential to
the understanding of complex brain functions and their deficits in
human brain disorders.
Acknowledgments
We thank P. Mitra, H. Zeng and Christian Niedworok for comments on the
manuscript and J. Kuhl for the graphics. P.O. is supported by the US National
Institute of Mental Health grant 1R01MH096946-01, McKnight Foundation,
Technological Innovations in Neuroscience Award and Simons Foundation for
Autism Research grants 204719 and 253447. T.W.M. is supported as a Wellcome
Trust Investigator and by the Medical Research Council MC U1175975156.
COMPETING FINANCIAL INTERESTS
The authors declare competing financial interests: details are available in the
online version of the paper.
Reprints and permissions information is available online at http://www.nature.
com/reprints/index.html.
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
Golgi, C. Sulla Struttura Della Sostanza Grigia del Cervello. Gazz. Med. Ital.
(Lombardia) 33, 244–246 (1873).
Ramón y Cajal, S. Textura del Sistema Nervioso del Hombre y de los
Vertebrados. Vol. 2 (Moya, 1904).
Felleman, D.J. & Van Essen, D.C. Distributed hierarchical processing in
the primate cerebral cortex. Cereb. Cortex 1, 1–47 (1991).
Rockland, K.S. & Pandya, D.N. Laminar origins and terminations of
cortical connections of the occipital lobe in the rhesus monkey. Brain
Res. 179, 3–20 (1979).
Craddock, R.C. et al. Imaging human connectomes at the macroscale.
Nat. Methods 10, 524–539 (2013).
Helmstaedter, M. Cellular-resolution connectomics: challenges of dense
neural circuit reconstruction. Nat. Methods 10, 501–507 (2013).
Bohland, J.W. et al. A proposal for a coordinated effort for the
determination of brainwide neuroanatomical connectivity in model
organisms at a mesoscopic scale. PLoS Comput. Biol. 5, e1000334 (2009).
This paper describes the rationale for mapping connectivity in the
whole mouse brain at the mesoscale level by LM.
Odgaard, A., Andersen, K., Melsen, F. & Gundersen, H.J. A direct method
for fast three-dimensional serial reconstruction. J. Microsc. 159, 335–342
(1990).
Ewald, A.J., McBride, H., Reddington, M., Fraser, S.E. & Kerschmann, R.
Surface imaging microscopy, an automated method for visualizing whole
embryo samples in three dimensions at high resolution. Dev. Dyn. 225,
369–375 (2002).
Tsai, P.S. et al. All-optical histology using ultrashort laser pulses.
Neuron 39, 27–41 (2003).
This study pioneered the approach of serial imaging by two-photon
microscopy and tissue sectioning for ex vivo collection of
neuroanatomical data.
Sands, G.B. et al. Automated imaging of extended tissue volumes using
confocal microscopy. Microsc. Res. Tech. 67, 227–239 (2005).
Ragan, T. et al. High-resolution whole organ imaging using two-photon
tissue cytometry. J. Biomed. Opt. 12, 014015 (2007).
Mayerich, D., Abbott, L. & McCormick, B. Knife-edge scanning microscopy
for imaging and reconstruction of three-dimensional anatomical structures
of the mouse brain. J. Microsc. 231, 134–143 (2008).
Li, A. et al. Micro-optical sectioning tomography to obtain a highresolution atlas of the mouse brain. Science 330, 1404–1408 (2010).
Ragan, T. et al. Serial two-photon tomography for automated ex vivo
mouse brain imaging. Nat. Methods 9, 255–258 (2012).
This study introduces the method of STP tomography and
demonstrates its use for anterograde and retrograde tracing in the
mouse brain.
Gong, H. et al. Continuously tracing brain-wide long-distance axonal
projections in mice at a one-micron voxel resolution. Neuroimage 74, 87–98,
(2013).
This study demonstrates the first long-range tracing of individual
axons in the mouse brain by fMOST.
Denk, W., Strickler, J.H. & Webb, W.W. Two-photon laser scanning
fluorescence microscopy. Science 248, 73–76 (1990).
Huisken, J., Swoger, J., Del Bene, F., Wittbrodt, J. & Stelzer, E.H. Optical
sectioning deep inside live embryos by selective plane illumination
microscopy. Science 305, 1007–1009 (2004).
Dodt, H.U. et al. Ultramicroscopy: three-dimensional visualization of
neuronal networks in the whole mouse brain. Nat. Methods 4, 331–336
(2007).
This study is the first to use LSFM for imaging the entire mouse
brain.
nature methods | VOL.10 NO.6 | JUNE 2013 | 521
npg
© 2013 Nature America, Inc. All rights reserved.
review
20. Niedworok, C.J. et al. Charting monosynaptic connectivity maps by twocolor light-sheet fluorescence microscopy. Cell Rep. 2, 1375–1386 (2012).
21. Chung, K. et al. Structural and molecular interrogation of intact
biological systems. Nature doi:10.1038/nature12107 (2013).
22. Chung, K. & Deisseroth, K. CLARITY for mapping the nervous system.
Nat. Methods 10, 508–513 (2013).
23. Leischner, U., Zieglgansberger, W. & Dodt, H.U. Resolution of
ultramicroscopy and field of view analysis. PLoS ONE 4, e5785 (2009).
24. Mertz, J. & Kim, J. Scanning light-sheet microscopy in the whole mouse
brain with HiLo background rejection. J. Biomed. Opt. 15, 016027
(2010).
25. Kalchmair, S., Jahrling, N., Becker, K. & Dodt, H.U. Image contrast
enhancement in confocal ultramicroscopy. Opt. Lett. 35, 79–81 (2010).
26. Keller, P.J. et al. Fast, high-contrast imaging of animal development
with scanned light sheet-based structured-illumination microscopy.
Nat. Methods 7, 637–642 (2010).
27. Wickersham, I.R., Finke, S., Conzelmann, K.K. & Callaway, E.M. Retrograde
neuronal tracing with a deletion-mutant rabies virus. Nat. Methods 4,
47–49 (2007).
28. Wickersham, I.R. et al. Monosynaptic restriction of transsynaptic tracing
from single, genetically targeted neurons. Neuron 53, 639–647 (2007).
This study describes a genetically modified rabies virus designed to
specifically label direct presynaptic input onto a given cell
population.
29. Ng, L. et al. An anatomic gene expression atlas of the adult mouse
brain. Nat. Neurosci. 12, 356–362 (2009).
30. Jones, E.G., Stone, J.M. & Karten, H.J. High-resolution digital brain
atlases: a Hubble telescope for the brain. Ann. NY Acad. Sci. 1225
(suppl. 1), E147–E159 (2011).
31. Lein, E.S. et al. Genome-wide atlas of gene expression in the adult
mouse brain. Nature 445, 168–176 (2007).
This study pioneered large-scale LM-based mouse brain anatomy and
introduced the Allen Mouse Brain Atlas and online data portal.
32. Dong, H.W. The Allen Reference Atlas: A Digital Color Brain Atlas of the
C57Bl/6J Male Mouse (John Wiley & Sons Inc., 2008).
33. Lanciego, J.L. & Wouterlood, F.G. A half century of experimental
neuroanatomical tracing. J. Chem. Neuroanat. 42, 157–183 (2011).
34. Glover, J.C., Petursdottir, G. & Jansen, J.K. Fluorescent dextran-amines
used as axonal tracers in the nervous system of the chicken embryo.
J. Neurosci. Methods 18, 243–254 (1986).
35. Llewellyn-Smith, I.J., Martin, C.L., Arnolda, L.F. & Minson, J.B. Tracertoxins: cholera toxin B-saporin as a model. J. Neurosci. Methods 103,
83–90 (2000).
36. Grinevich, V., Brecht, M. & Osten, P. Monosynaptic pathway from rat
vibrissa motor cortex to facial motor neurons revealed by lentivirus-based
axonal tracing. J. Neurosci. 25, 8250–8258 (2005).
37. Atasoy, D., Aponte, Y., Su, H.H. & Sternson, S.M. A FLEX switch targets
Channelrhodopsin-2 to multiple cell types for imaging and long-range
circuit mapping. J. Neurosci. 28, 7025–7030 (2008).
38. Taniguchi, H. et al. A resource of Cre driver lines for genetic targeting of
GABAergic neurons in cerebral cortex. Neuron 71, 995–1013 (2011).
39. Madisen, L. et al. A robust and high-throughput Cre reporting and
characterization system for the whole mouse brain. Nat. Neurosci. 13,
133–140 (2010).
40. Madisen, L. et al. A toolbox of Cre-dependent optogenetic transgenic
mice for light-induced activation and silencing. Nat. Neurosci. 15,
793–802 (2012).
41. Harris, J.A., Wook Oh, S. & Zeng, H. Adeno-associated viral vectors for
anterograde axonal tracing with fluorescent proteins in nontransgenic and
cre driver mice. Curr. Protoc. Neurosci. 1.20, 21–18 (2012).
42. Thompson, R.H. & Swanson, L.W. Hypothesis-driven structural
connectivity analysis supports network over hierarchical model of brain
architecture. Proc. Natl. Acad. Sci. USA 107, 15235–15239 (2010).
43. Gerfen, C.R. & Sawchenko, P.E. An anterograde neuroanatomical tracing
method that shows the detailed morphology of neurons, their axons and
terminals: immunohistochemical localization of an axonally transported
plant lectin, Phaseolus vulgaris leucoagglutinin (PHA-L). Brain Res. 290,
219–238 (1984).
44. Naumann, T., Hartig, W. & Frotscher, M. Retrograde tracing with FluoroGold: different methods of tracer detection at the ultrastructural level
and neurodegenerative changes of back-filled neurons in long-term
studies. J. Neurosci. Methods 103, 11–21 (2000).
45. Reiner, A. et al. Pathway tracing using biotinylated dextran amines.
J. Neurosci. Methods 103, 23–37 (2000).
522 | VOL.10 NO.6 | JUNE 2013 | nature methods
FOCUS ON MAPPING THE BRAIN
46. Hintiryan, H. et al. Comprehensive connectivity of the mouse main
olfactory bulb: analysis and online digital atlas. Front. Neuroanat. 6, 30
(2012).
47. Conte, W.L., Kamishina, H. & Reep, R.L. Multiple neuroanatomical tracttracing using fluorescent Alexa Fluor conjugates of cholera toxin subunit
B in rats. Nat. Protoc. 4, 1157–1166 (2009).
48. Sunkin, S.M. et al. Allen Brain Atlas: an integrated spatio-temporal
portal for exploring the central nervous system. Nucleic Acids Res. 41,
D996–D1008 (2013).
49. Ugolini, G. Advances in viral transneuronal tracing. J. Neurosci. Methods
194, 2–20 (2010).
50. Ekstrand, M.I., Enquist, L.W. & Pomeranz, L.E. The alpha-herpesviruses:
molecular pathfinders in nervous system circuits. Trends Mol. Med. 14,
134–140 (2008).
51. Song, C.K., Enquist, L.W. & Bartness, T.J. New developments in
tracing neural circuits with herpesviruses. Virus Res. 111, 235–249
(2005).
52. Callaway, E.M. Transneuronal circuit tracing with neurotropic viruses.
Curr. Opin. Neurobiol. 18, 617–623 (2008).
53. Wickersham, I.R. & Feinberg, E.H. New technologies for imaging synaptic
partners. Curr. Opin. Neurobiol. 22, 121–127 (2012).
54. Rancz, E.A. et al. Transfection via whole-cell recording in vivo: bridging
single-cell physiology, genetics and connectomics. Nat. Neurosci. 14,
527–532 (2011).
This study was the first to combine intracellular neuronal recording
with DNA delivery. The authors use this method to map the synaptic
function of a single cell in vivo and then target rabies-based
retrograde labeling of the cells’ synaptic input.
55. Miyamichi, K. et al. Cortical representations of olfactory input by transsynaptic tracing. Nature 472, 191–196 (2011).
56. Marshel, J.H., Mori, T., Nielsen, K.J. & Callaway, E.M. Targeting single
neuronal networks for gene expression and cell labeling in vivo. Neuron
67, 562–574 (2010).
This paper describes an electroporation method for single-cell
delivery of DNA for targeted infection of modified rabies virus.
57. Takatoh, J. et al. New modules are added to vibrissal premotor circuitry
with the emergence of exploratory whisking. Neuron 77, 346–360 (2013).
58. Arenkiel, B.R. et al. Activity-induced remodeling of olfactory bulb
microcircuits revealed by monosynaptic tracing. PLoS ONE 6, e29423
(2011).
59. Finke, S. & Conzelmann, K.K. Replication strategies of rabies virus.
Virus Res. 111, 120–131 (2005).
60. Ugolini, G. Specificity of rabies virus as a transneuronal tracer of motor
networks: transfer from hypoglossal motoneurons to connected secondorder and higher order central nervous system cell groups. J. Comp.
Neurol. 356, 457–480 (1995).
61. Federspiel, M.J., Bates, P., Young, J.A., Varmus, H.E. & Hughes, S.H.
A system for tissue-specific gene targeting: transgenic mice susceptible
to subgroup A avian leukosis virus-based retroviral vectors. Proc. Natl.
Acad. Sci. USA 91, 11241–11245 (1994).
62. Young, J.A., Bates, P. & Varmus, H.E. Isolation of a chicken gene that
confers susceptibility to infection by subgroup A avian leukosis and
sarcoma viruses. J. Virol. 67, 1811–1816 (1993).
63. Wall, N.R., Wickersham, I.R., Cetin, A., De La Parra, M. & Callaway, E.M.
Monosynaptic circuit tracing in vivo through Cre-dependent targeting and
complementation of modified rabies virus. Proc. Natl. Acad. Sci. USA 107,
21848–21853 (2010).
64. Etessami, R. et al. Spread and pathogenic characteristics of a G-deficient
rabies virus recombinant: an in vitro and in vivo study. J. Gen. Virol. 81,
2147–2153 (2000).
65. Lo, L. & Anderson, D.J.A. Cre-dependent, anterograde transsynaptic viral
tracer for mapping output pathways of genetically marked neurons.
Neuron 72, 938–950 (2011).
66. Beier, K.T. et al. Anterograde or retrograde transsynaptic labeling of CNS
neurons with vesicular stomatitis virus vectors. Proc. Natl. Acad. Sci. USA
108, 15414–15419 (2011).
67. Hawrylycz, M. et al. Digital atlasing and standardization in the mouse
brain. PLoS Comput. Biol. 7, e1001065 (2011).
68. Paxinos, G. & Franklin, K.B. The Mouse Brain in Stereotaxic Coordinates
(Gulf Professional Publishing, 2004).
69. Swanson, L.W. & Bota, M. Foundational model of structural connectivity
in the nervous system with a schema for wiring diagrams, connectome,
and basic plan architecture. Proc. Natl. Acad. Sci. USA 107, 20610–20617
(2010).
npg
© 2013 Nature America, Inc. All rights reserved.
FOCUS ON MAPPING THE BRAINreview
70. Moene, I.A., Subramaniam, S., Darin, D., Leergaard, T.B. & Bjaalie, J.G.
Toward a workbench for rodent brain image data: systems architecture
and design. Neuroinformatics 5, 35–58 (2007).
71. Klausberger, T. & Somogyi, P. Neuronal diversity and temporal dynamics:
the unity of hippocampal circuit operations. Science 321, 53–57
(2008).
72. O’Rourke, N.A., Weiler, N.C., Micheva, K.D. & Smith, S.J. Deep molecular
diversity of mammalian synapses: why it matters and how to measure it.
Nat. Rev. Neurosci. 13, 365–379 (2012).
73. DeFelipe, J. et al. New insights into the classification and nomenclature of
cortical GABAergic interneurons. Nat. Rev. Neurosci. 14, 202–216 (2013).
74. Emes, R.D. & Grant, S.G. Evolution of synapse complexity and diversity.
Annu. Rev. Neurosci. 35, 111–131 (2012).
75. Micheva, K.D., Busse, B., Weiler, N.C., O’Rourke, N. & Smith, S.J. Singlesynapse analysis of a diverse synapse population: proteomic imaging
methods and markers. Neuron 68, 639–653 (2010).
76. Micheva, K.D. & Smith, S.J. Array tomography: a new tool for imaging
the molecular architecture and ultrastructure of neural circuits. Neuron
55, 25–36 (2007).
77. Tsai, P.S. et al. Correlations of neuronal and microvascular densities in
murine cortex revealed by direct counting and colocalization of nuclei
and vessels. J. Neurosci. 29, 14553–14570 (2009).
78. Oberlaender, M. et al. Cell type-specific three-dimensional structure of
thalamocortical circuits in a column of rat vibrissal cortex. Cereb. Cortex
22, 2375–2391 (2012).
79. Meyer, H.S. et al. Inhibitory interneurons in a cortical column form hot
zones of inhibition in layers 2 and 5A. Proc. Natl. Acad. Sci. USA 108,
16807–16812 (2011).
80. Meyer, H.S. et al. Number and laminar distribution of neurons in a
thalamocortical projection column of rat vibrissal cortex. Cereb. Cortex
20, 2277–2286 (2010).
81. Ascoli, G.A. et al. Petilla terminology: nomenclature of features of
GABAergic interneurons of the cerebral cortex. Nat. Rev. Neurosci. 9,
557–568 (2008).
82. Margrie, T.W. et al. Targeted whole-cell recordings in the mammalian
brain in vivo. Neuron 39, 911–918 (2003).
83. Kitamura, K., Judkewitz, B., Kano, M., Denk, W. & Hausser, M. Targeted
patch-clamp recordings and single-cell electroporation of unlabeled
neurons in vivo. Nat. Methods 5, 61–67 (2008).
84. Angelo, K. et al. A biophysical signature of network affiliation and
sensory processing in mitral cells. Nature 488, 375–378 (2012).
85. Reid, R.C. From functional architecture to functional connectomics.
Neuron 75, 209–217 (2012).
86. Briggman, K.L., Helmstaedter, M. & Denk, W. Wiring specificity in the
direction-selectivity circuit of the retina. Nature 471, 183–188 (2011).
87. Bock, D.D. et al. Network anatomy and in vivo physiology of visual
cortical neurons. Nature 471, 177–182 (2011).
88. Wallace, D.J. et al. Single-spike detection in vitro and in vivo with a
genetic Ca2+ sensor. Nat. Methods 5, 797–804 (2008).
89. Akerboom, J. et al. Optimization of a GCaMP calcium indicator for neural
activity imaging. J. Neurosci. 32, 13819–13840 (2012).
90. Mank, M. et al. A genetically encoded calcium indicator for chronic
in vivo two-photon imaging. Nat. Methods 5, 805–811 (2008).
91. Ko, H. et al. Functional specificity of local synaptic connections in
neocortical networks. Nature 473, 87–91 (2011).
92. Holmgren, C., Harkany, T., Svennenfors, B. & Zilberter, Y. Pyramidal cell
communication within local networks in layer 2/3 of rat neocortex.
J. Physiol. 551, 139–153 (2003).
93. Thomson, A.M., West, D.C., Wang, Y. & Bannister, A.P. Synaptic
connections and small circuits involving excitatory and inhibitory neurons
in layers 2–5 of adult rat and cat neocortex: triple intracellular
recordings and biocytin labelling in vitro. Cereb. Cortex 12, 936–953
(2002).
94. Song, S., Sjostrom, P.J., Reigl, M., Nelson, S. & Chklovskii, D.B. Highly
nonrandom features of synaptic connectivity in local cortical circuits.
PLoS Biol. 3, e68 (2005).
95. Harvey, C.D., Collman, F., Dombeck, D.A. & Tank, D.W. Intracellular
dynamics of hippocampal place cells during virtual navigation. Nature
461, 941–946 (2009).
This study introduced the method of physiological recording in headrestrained mice on a spherical treadmill while the mice performed
spatial tasks in a virtual environment.
96. Harvey, C.D., Coen, P. & Tank, D.W. Choice-specific sequences in parietal
cortex during a virtual-navigation decision task. Nature 484, 62–68 (2012).
97. Huber, D. et al. Multiple dynamic representations in the motor cortex
during sensorimotor learning. Nature 484, 473–478 (2012).
98. Fenno, L., Yizhar, O. & Deisseroth, K. The development and application of
optogenetics. Annu. Rev. Neurosci. 34, 389–412 (2011).
99. Koch, C. & Reid, R.C. Neuroscience: observatories of the mind. Nature
483, 397–398 (2012).
100. Alivisatos, A.P. et al. Neuroscience. The brain activity map. Science 339,
1284–1285 (2013).
101. Ahrens, M.B., Orger, M.B., Robson, D.N., Li, J.M. & Keller, P.J.
Whole-brain functional imaging at cellular resolution using light-sheet
microscopy. Nat. Methods 10, 413–420 (2013).
102. Ahrens, M.B. et al. Brain-wide neuronal dynamics during motor
adaptation in zebrafish. Nature 485, 471–477 (2012).
103. Herrera, D.G. & Robertson, H.A. Activation of c-fos in the brain. Prog.
Neurobiol. 50, 83–107 (1996).
104. Barth, A.L., Gerkin, R.C. & Dean, K.L. Alteration of neuronal firing
properties after in vivo experience in a FosGFP transgenic mouse.
J. Neurosci. 24, 6466–6475 (2004).
105. Grinevich, V. et al. Fluorescent Arc/Arg3.1 indicator mice: a versatile tool
to study brain activity changes in vitro and in vivo. J. Neurosci. Methods
184, 25–36 (2009).
106. Reijmers, L.G., Perkins, B.L., Matsuo, N. & Mayford, M. Localization of a
stable neural correlate of associative memory. Science 317, 1230–1233
(2007).
nature methods | VOL.10 NO.6 | JUNE 2013 | 523
review
Focus on mapping the brain
Imaging human connectomes at
the macroscale
© 2013 Nature America, Inc. All rights reserved.
R Cameron Craddock1,2, Saad Jbabdi3,9, Chao-Gan Yan1,2,4,9, Joshua T Vogelstein1,5–7,
F Xavier Castellanos2,4, Adriana Di Martino4, Clare Kelly4, Keith Heberlein8,
Stan Colcombe2 & Michael P Milham1,2
At macroscopic scales, the human connectome comprises anatomically distinct brain
areas, the structural pathways connecting them and their functional interactions.
Annotation of phenotypic associations with variation in the connectome and
cataloging of neurophenotypes promise to transform our understanding of the human
brain. In this Review, we provide a survey of magnetic resonance imaging–based
measurements of functional and structural connectivity. We highlight emerging areas
of development and inquiry and emphasize the importance of integrating structural
and functional perspectives on brain architecture.
First introduced in 2005 (ref. 1), the term ‘connectome’
embodies the advances of over a century of neuroscientific innovation and reflects an agenda for a new era.
Initially defined as a complete map of neural connections in the brain, the connectome is a multiscale construct that can be examined at varying resolutions. At
the extremes are the microscale, which encompasses
individual neurons and their synaptic connections, and
the macroscale, which encompasses cortical tissues
(commonly a cubic centimeter or larger). The intermediate resolution is the mesoscale, which, in humans,
encompasses vertical columns of 80–120 neurons
(commonly referred to as micro- or mini-columns)1,2.
Although all scales of resolution are intimately associated, each provides unique perspectives on the connectome. At the macroscale, and particularly in studies of
the human brain, conceptualizations of the connectome
have grown to also include information about function3. In this Review, we will use the term connectome
to refer to brain areas, their anatomical connections
and their functional interactions.
At present, methodologies for analysis at macro­
scale resolution are best positioned for mapping and
annotating human connectomes with cognitive and
behavioral associations. The higher-order, albeit lowerresolution, representations captured at the macroscale
most directly relate to regulatory, cognitive and affective processes. Interpretation of macroscale-resolution
findings is most amenable to guidance from lesion and
brain-imaging studies. Comprehensive mapping and
annotation of the connectome is most feasible at the
lower-resolution macroscale, owing to lower computational and analytical demands. Moreover, noninvasive
tools for in vivo imaging the human connectome are
only available for analyses at the macroscale; in vivo
microscale-resolution studies are currently limited to
model organisms and neurosurgical patients.
Among the modalities used for macroconnectomics,
magnetic resonance imaging (MRI) is dominant,
partly because of widespread availability, safety and
spatial resolution. Diffusion-weighted MRI (dMRI)
and functional MRI (fMRI) are widely used for inferring structural and functional connectivity, respectively4. dMRI provides cubic-millimeter-resolution
portrayals of white-matter tracts and insights into
organizing principles that guide their orientation and
1Center for the Developing Brain, Child Mind Institute, New York, New York, USA. 2Nathan Kline Institute for Psychiatric Research,
Orangeburg, New York, USA. 3The Oxford Centre for Functional MRI of the Brain, University of Oxford, Oxford, UK. 4The Phyllis Green
and Randolph Cowen Institute for Pediatric Neuroscience, Child Study Center, New York University Langone Medical Center, New York,
New York, USA. 5Department of Statistical Science, Duke University, Durham, North Carolina, USA. 6Institute for Brain Sciences,
Duke University, Durham, North Carolina, USA. 7Institute for Data Intensive Engineering and Sciences, John Hopkins University, Baltimore,
Maryland, USA. 8Siemens Medical Solutions USA, Charlestown, Massachusetts, USA. 9These authors contributed equally to this work.
Correspondence should be addressed to S.C. ([email protected]) or M.P.M. ([email protected]).
RECEIVED 22 FEBRUARY; ACCEPTED 22 APRIL; published online 30 may 2013; doi:10.1038/nmeth.2482
524 | VOL.10 NO.6 | JUNE 2013 | nature methods
Defining nodes
Defining the nodes of a macroscale connectome is a complex
task as we lack agreement on how best to define the constituent
brain units. Depending on the scope of the investigation, the
specific brain subunits represented by nodes can range from
the small patches of cortex contained in individual MRI voxels to larger brain areas (for example, dorsolateral prefrontal
cortex). Early parcellation efforts used postmortem architectonic measurements (for example, cell morphology) of single
individuals. Although these resulting atlases are central to neuroscience, no functional or structural connectivity–based information was used to construct them, thus limiting their capacity
to accurately represent connectomes. For example, despite being
represented as a single area in anatomical atlases 6, the anterior
cingulate region contains subregions that are each characterized
by dramatically different functional7 and structural8 connectivity patterns. Although the large-scale human brain patterns
captured using different strategies of parceling data may bear
a gross similarity to one another, the specific details conveyed
vary substantially (Fig. 1).
Ideally, both brain-function and structural-connectivity
information should be used to delineate brain areas. Metaanalytic approaches can be used to define nodes on the basis of
task-based fMRI (T-fMRI) studies9. Alternatively, data-driven
clustering techniques can be used to subdivide the brain into
areas based on homogeneity of functional time series10,11, or
functional or structural connectivity profiles8,11,12. Blinded
source-separation techniques can also be used to define network nodes using spatial independence13. These methods
commonly involve pooling of information across individuals,
and most enforce specific properties on resulting brain areas11.
One drawback is the need to prespecify the number of areas
to be generated, which can be estimated based on homogeneity, accuracy, reproducibility or stability of the brain areas 10,11.
EZ
HO
TT
CC200
trajectories; fMRI reveals a universal functional architecture, with
variations among individuals meaningfully related to phenotypic
variables5 (for example, behavioral and psychiatric).
In this Review, we focus on the mapping, characterization and
analysis of macroscale connectomes. We structured our presentation in terms of a mathematical perspective that treats the connectome as a graph of interactions among brain areas. Nodes in
the graph are abstract representations of brain areas, and edges
represent pairwise relationships between nodes. We first review
approaches and challenges to subdividing the brain into discrete
subunits represented by nodes (here referred to as ‘parcellation’
efforts) and then review the imaging and analytic methodologies
used to map and quantify patterns of structural and functional
connectivity that are represented by edges in the connectome.
CC400
© 2013 Nature America, Inc. All rights reserved.
Figure 1 | Different parcellations of the human brain. Atlases of brain
areas generated using anatomical (top four rows) and functional (bottom
two rows) parcellation schemes show a lateral view (right) and top views
of the human brain. AAL (automated anatomical labeling)109 and Harvard
Oxford (HO)110 are derived from anatomical landmarks (sulci and gyral).
The EZ (Eickhoff–Zilles)111 and TT (Talariach Daemon)112 atlases are
derived from postmortem cyto- and myelo-architectonic segmentations.
The CC200 and CC400 atlases are derived from 200- and 400-unit
functional parcellations11.
review
AAL
Focus on mapping the brain
Optimal comparison of connectomes is likely to require
parcellation strategies that incorporate information across individuals and modalities, potentially at the cost of quality of fit for
single subjects and modalities.
Estimating structural connectivity
Structural connectivity encompasses the collection of axonal
and dendritic connections among neurons 1. Despite definitional simplicity, structural connectivity is difficult to measure
with noninvasive, in vivo imaging approaches. Before dMRI, our
knowledge was primarily derived from lesions and blunt dissection in humans or invasive tracing in nonhumans. These methodologies remain the gold standard for establishing connectivity,
but the noninvasive nature of dMRI makes it the de facto standard
for studies of human structural connectivity.
Acquisition. The basic principle underlying the inference of
structural connectivity from dMRI data is that water diffusion
in white matter is hindered and occurs primarily along the path
of axons. In contrast, water diffusion in gray matter and cerebral
spinal fluid occurs (almost) equally in all directions. By following
the motion of water, it is possible to map the orientation(s) of fibers passing through each voxel of white matter. In dMRI, a series
of images are acquired, each sensitive to diffusion along a specific
direction. The number of images for unique directions acquired
varies from six to hundreds. When combined, these images contain the information necessary to estimate the orientation(s) of
fibers passing through each voxel; this information is used to
reconstruct large-scale tracts of white matter (tractography). We
describe dMRI acquisition in more detail in Box 1 and Table 1.
nature methods | VOL.10 NO.6 | JUNE 2013 | 525
review
Focus on mapping the brain
Box 1 Optimizing fMRI and dMRI image acquisition
Echo-planar imaging (EPI) is the most common method for
acquiring functional magnetic resonance imaging (fMRI) and
diffusion MRI (dMRI) data. An EPI volume is a set of slices,
each acquired after a single excitation. EPI volumes are collected in sequence, representing time points in fMRI data
and diffusion directions in dMRI data. Gradient echo EPI in
fMRI only involves excitation and readout. However, in dMRI an additional 180° rephasing (spin echo) radiofrequency pulse
and bipolar diffusion–encoding gradients are applied.
© 2013 Nature America, Inc. All rights reserved.
SNR ∝ I0 × ∆ x × ∆ y × ∆ z × N acq × N x × N y × N z ×
1
Pf
(1)
Signal-to-noise ratio (SNR), spatial distortions and artifacts, and
image contrast determine image quality. Image SNR can be represented in terms of parameters such as available signal (I0), voxel
dimensions (∆x, ∆y, ∆z), number of image acquisitions (Nacq),
number of samples in each dimension (Nx, Ny, Nz), time between
samples (∆t) and parallel imaging factor (Pf) (equation (1))117. These parameters are described in Table 1. Equation (1)
demonstrates how different parameter settings affect the SNR;
for example, acquiring and averaging two volumes (Nacq = 2)
improves SNR by 2 . SNR for modern imaging technology is
greater than required for most imaging applications and can be
traded-off to optimize other imaging aspects (for example, to
reduce spatial distortions). Beyond spatial image considerations,
parameter optimizations must be evaluated in terms of impact
on the temporal signal in fMRI or diffusion signal in dMRI.
General considerations for EPI
Spatial distortion. The EPI readout scheme leads to extended
readout times and asymmetric bandwidth between directions
that result in prolonged spin-spin (T2) relaxation, which increases in-plane smoothing. Bandwidth corresponds to the number of frequencies allocated to a voxel; any shift in
frequency resulting from magnetic field imperfections spatially shifts the signal in proportion to the inverse of the
bandwidth118. The bandwidth in the phase-encoding direction
is a fraction of the bandwidth in the readout direction, making
it particularly susceptible to spatial distortions. Increasing
bandwidth can minimize spatial distortions, though costing
SNR. Parallel imaging techniques increase effective bandwidth
and image acquisition rate119. These techniques require calibration scans for proper reconstruction, and any misalignment (for example, head motion) between these scans
and image data produce reconstruction errors119. Alternatively,
spatial distortions can be mathematically corrected using maps of the spatial variation in the magnetic field15.
Image resolution. Image resolution can be increased by
reducing the field of view (FOV) or increasing the number of
samples along a dimension. Decreasing the FOV produces a
proportional decrease in the SNR; the FOV must remain larger
than the head to avoid excessive ghosting. Increasing the
number of samples has less impact on SNR (reduced by square
root of the increase117). However, increasing the number of 526 | VOL.10 NO.6 | JUNE 2013 | nature methods
samples in the phase encoding direction substantially increases readout time, which increases the amount of T2 decay experienced during slice acquisition and leads to greater in-plane
smoothness118; parallel imaging mitigates these effects119.
Considerations for fMRI
BOLD contrast. Spin phase (T2*) contrast is the most sensitive to the BOLD effect and is optimized using an echo
time equal to the T2* of gray matter (echo time of ~30 milliseconds at 3 tesla). BOLD can also be measured with spin-echo
sequences, which are sensitive to T2 rather than T2* and
hence have better spatial specificity, although less sensitivity
than gradient echo techniques120.
Minimizing dropout. Tilting and positioning imaging slices to
prevent slices from intersecting air-tissue interfaces can minimize signal dropout resulting from differences in magnetic susceptibility121. Slice positioning should focus on areas most vital
to the experiment. Reducing echo time, which decreases BOLD
contrast, and increasing resolution, which decreases SNR, should
be avoided. Specialty sequences (for example, spiral-in/out122
and z-shim techniques123) result in the acquisition of multiple
echoes with different dephasing characteristics to reduce susceptibility effects. Alternatively, spin-echo EPI avoids susceptibility
artifacts but is much less sensitive to BOLD contrast120.
Motion sensitivity. Beyond image misalignment, head motion
induces fMRI signal fluctuations owing to partial voluming and
spin-history effects39. Lower flip angles124, longer repetition
time125 and interleaved slices minimize spin-history effects by
allowing the signal in a slice to fully relax before acquisition
of the neighboring slice. Spiral sequences are less sensitive to
motion during slice acquisition126. The 3D imaging techniques
minimize the impact of within-volume motion, but betweenvolume motions remain problematic unless a long repetition
time is used; long readouts increase the interaction between
magnetic field inhomogeneity and motion127. Parallel imaging
can improve acquisition time, resolution and bandwidth but
requires calibration scans119. Misalignments between the acquired data and calibration scans induce reconstruction errors and motion-correlated imaging artifacts119.
Temporal resolution. Number of slices acquired determines
temporal resolution. Echo time fundamentally limits slice- acquisition time. Increasing bandwidth, partial Fourier imaging and parallel imaging can result in marginal speed gains118.
Considerations for dMRI
Diffusion contrast. Minimizing echo time, and therefore
T2 decay, is essential to dMRI. T2 reduces overall magnetic
resonance signal and image SNR, and increases in-plane image
blurring. Increasing bandwidth, partial Fourier imaging, parallel imaging and using stronger diffusion gradients (b values; high b values require longer diffusion gradient pulses) can
minimize echo time118. Higher-strength gradient systems
enable high b values (desirable for fiber tracking) in shorter
continued
Focus on mapping the brain
review
Box 1 Optimizing fMRI and dMRI image acquisition (continued)
time, though require a hardware upgrade for most systems.
Echo shifting also reduces echo time, by relaxing the spin echo
condition and shifting the echo train earlier in time118.
Scan time. Echo time and number of slices determine dMRI
volume repetition time. The number of diffusion directions, b values and averages acquired determine total scan time.
© 2013 Nature America, Inc. All rights reserved.
Emerging acquisition sequences
Multiband and multi-echo imaging are emerging acquisition
strategies. In multiband imaging, multiple slices are acquired
simultaneously, substantially improving temporal resolution,
Preprocessing. After acquisition, dMRI data must be preconditioned before directional information can be extracted and tractography can be performed. Little debate exists regarding dMRI
data preprocessing (see ref. 14 for an exception), though this may
reflect the difficulties of preprocessing and its complex impact
rather than consensus in the field.
Correcting image distortions. In vivo dMRI data are plagued by
spatial distortions, which are a central focus of preprocessing. In
particular, magnetic field inhomogeneities are a major contributor. Areas where materials that differ with respect to magnetic susceptibility (that is, extent of magnetization achieved in the MRI)
interface with one another (for example, air-tissue interfaces) are
particularly prone to such inhomogeneities. These local variations
in the magnetic field generate spatial distortions, which can be
reduced through parallel imaging techniques or mathematically
corrected using estimates of the field variations15. Another cause
of spatial distortion is the interaction between the static magnetic
field and the currents induced by rapid switching of gradients with
the magnetic field, known as eddy currents. These artifacts can be
or increases in spatial resolution per unit time128, though
at the cost of SNR and spatial smoothness. This improvement enables higher fMRI sampling rates and more diffusion
directions per b-value acquisitions in dMRI129, substantially improving estimation of macroscale connectomes130.
In multi-echo fMRI, two echoes of a slice are acquired at
each acquisition (one with short echo time and one with an
echo time optimized for BOLD imaging131). Systematic noise
(heartbeat, respiration, head motion and scanner instability)
are measured in the first echo (with the short echo time),
and removed from the second, enabling signal denoising with
practically no SNR cost.
reduced using bipolar gradients16. Image coregistration, or spatial
alignment of brain scans, is commonly used in preprocessing to
correct for eddy-current distortions and subject head motion17.
However, this method is ineffective for images acquired using
very strong diffusion gradients. Model-based approaches that
explicitly account for the effects of eddy currents during image
acquisition are emerging as the preferred option18.
Overlooked issues. The above preprocessing steps, combined with
visual inspection, constitute standard preprocessing. A few important details are often overlooked (see ref. 14 for examples). First,
modern scanners are often equipped with antennas that acquire
data in parallel through multiple channels that are subsequently
combined. Such parallel imaging techniques can increase noise
levels if the data are submitted to standard reconstruction19. This
can artifactually lower diffusivity estimates along the axons, which
increases the tendency of tractography approaches to overfit the
data and generate false positive connections. An alternative reconstruction method has recently been proposed to address this important issue19. Second, any correction of distortion must account
Table 1 | Magnetic resonance imaging parameters
Parameter
Echo time
Repetition time
Bandwidth (1/∆t)
Definition
Time between slice excitation and acquiring the center
of k space
The time between acquisitions of adjacent fMRI volumes
(sampling period)
The range of frequencies mapped to a voxel
Flip angle (α)
The amount of rotation applied to proton spins by the
excitation pulse
Spatial resolution
(∆x, ∆y, ∆z, Nx, Ny)
The volume of tissue sampled in a given voxel;
determined by field of view and the number of points
sampled in a slice
Parallel imaging (Pf)
Methods such as GRAPPA and SENSE can decrease
repetition time and reduce spatial distortions by sampling
k-space lines in parallel; decreases in repetition time are
rate-limited by echo time required for BOLD contrast
Number of acquisitions that are acquired and
subsequently averaged
Nacq
Impact
Determines the impact of spin-spin relaxation (T2) and spin dephasing
(T2*) on image
Impacts the signal available for imaging (I0), impacted by the number
of slices; longer repetition time durations reduce motion sensitivity
Lower bandwidth settings can increase artifacts owing to inadequate
shimming or susceptibility and distortions in the phase-encoding
direction (for example, ‘scalloping’)
Impacts I0. Flip angles larger or smaller that the Ernst angle for a
given repetition time will reduce I0. Low flip angles may reduce motion
sensitivity and in-flow effects and improve spin-lattice relaxation (T1)
contrast of images.
Signal-to-noise ratio (SNR) is substantially impacted by voxel volume;
higher spatial resolution have lower SNR, for example, a 2-mm isovoxel
has only ~30% of the SNR of a 3-mm isovoxel, holding all other
factors constant
Can allow faster image acquisition but at a reduction of SNR by
1/parallelization factor, holding all other factors constant; will
increase temporal noise resulting from head motion, respiration
or pulsatile effects
Improves SNR; does not make sense for fMRI but is commonly used
for dMRI
nature methods | VOL.10 NO.6 | JUNE 2013 | 527
review
Figure 2 | Diffusion imaging of structural
connectivity maps for a human brain. (a) dMRIbased map of principal tensor orientations
of the human brain viewed from the front.
Blue, superior-inferior; green, anteriorposterior; and red, medial-lateral. (b) DTI
fiber orientation estimates (red lines) from a
region of corpus callosum superimposed on a
fractional anisotropy image. Sample streamline
is in yellow. Image is magnified 4× relative
to that in a. (c) Pyramidal tract streamlines
based on deterministic (left) and probabilistic
(right) approaches (results superimposed on
fractional anisotropy image). Color bar indicates
confidence about the presence of the tract.
Images courtesy of S. Sotiropoulos (http://
etheses.nottingham.ac.uk/1164/).
Focus on mapping the brain
a
b
2 cm
c
© 2013 Nature America, Inc. All rights reserved.
High
for voxel-wise compression or expansion
during image coregistration. This is particularly important for dMRI, where distortions vary in direction and magnitude
between different gradient orientations, but
is largely ignored by current software packages. Finally, image coregistrations contain
a rotation component that is applied to each volume and must
therefore be applied to the concurrent gradient orientation20.
Estimating fiber orientation. Before delineating tracts and bundles via tractography, fiber orientation(s) must be inferred for
white-matter voxels individually. Sensitivity of the dMRI signal
to water-diffusion properties (such as rate and direction) enables the estimation of fiber orientation at each voxel. As each
voxel contains thousands of axonal fibers, not just one, the goal
of dMRI analysis is to infer a probability function for each voxel,
which captures the different fiber orientations present and their
relative proportions21. Estimation of this function— referred to
as the fiber orientation density function (fODF)—at each voxel
is the first step in estimating structural connectivity.
The diffusion tensor is a simplistic but viable model for the diffusion profile that provides a simple approximation to the fODF22.
Diffusion tensor imaging (DTI) uses a 3 × 3 matrix to provide an
abstract ellipsoid representation of the water-diffusion profile for
a given voxel. Mathematical decomposition of this matrix yields
information regarding the directions (x, y and z) of maximum
and minimum water motion (eigenvectors) as well as the amount
of diffusion that occurs along each direction (eigenvalues). The
direction of maximal diffusion, referred to as the principal diffusion direction, is taken as the best estimate of fiber orientation
within a voxel. Formally speaking, in the case of DTI, the fODF
is approximated using a delta function or peak-aligned with the
principal diffusion direction.
The diffusion tensor provides a good estimate of fiber orientations when axons are homogeneously aligned within a voxel.
However, this is not always the case. Fibers are known to disperse (fan), cross, merge and kiss (temporarily run adjacent to
one another)—all of which can happen within a single voxel and
lead to heterogeneity not accounted for by a simple delta function. More complex approximations of fODF can account for such
heterogeneity within a voxel, though they require greater angular
coverage23,24 (that is, more directions) and models that either
528 | VOL.10 NO.6 | JUNE 2013 | nature methods
Low
explicitly or implicitly account for interactions between fiber orientation and the diffusion signal (see ref. 21 for example methods).
Complex fODF models better estimate fiber trajectories,
particularly when several white-matter tracts intersect and allow
recovery of nondominant pathways invisible to DTI25.
Estimating edges. After estimation of voxel-wise fiber orientations, tractography approaches are used to establish structural
connectivity between connectome nodes. Three-dimensional
(3D) trajectories, referred to as ‘streamlines’, are used to trace
putative white-matter paths. Local fiber orientation information
guides the construction of streamlines along the fODF, allowing
us to trace major white-matter bundles26. Results are typically
visualized as 3D renderings of thin curves grouped into bundles
(Fig. 2), reminiscent of postmortem dissection photographs. The
individual streamlines do not represent actual axons; they depict
estimates of the average trajectories of axon bundles, given our
assumption that diffusion is least hindered along axons.
The specific process by which streamlines are developed varies depending on the complexity of the fODF approximations
available. With diffusion-tensor modeling, the principal diffusion
direction at each voxel guides the formation of the streamline;
specifically, it provides a candidate for the tangent to the streamline at each voxel. For more complex fODF models, streamlining
follows the same principle, though with multiple peak orientations available at each voxel rather than a single principal diffusion direction. This allows streamlines with differing orientations
to pass through the same voxel, which is crucial when heterogeneous fibers are present. Whereas traditional tractography
approaches are deterministic, probabilistic approaches account
for uncertainty in estimates of local fiber orientations, allowing
for estimation of probabilities for any given streamline.
Using streamlining methodologies, it is theoretically possible
to measure all connections between gray-matter areas. We can
estimate both the trajectories and the end points of anatomical pathways. In practice, however, inference of point-to-point
© 2013 Nature America, Inc. All rights reserved.
Focus on mapping the brain
connectivity using streamlining is imprecise and error-prone27;
improvements in both data quality and modeling are needed to
yield more accurate structural connectomes.
Ideally, we should not only be able to infer the existence or
absence of connections between nodes, but estimate connection
(edge) strengths as well. Anatomical connections are made up of
axons; features of these axons, such as density, size, length and
myelination, have important consequences on the propagation
of action potentials, and hence information transfer. Measures of
microstructural features based on more complex dMRI experiments are emerging28, and may become an important component
of connectomics. Related measures, such as fractional anisotropy
and diffusivity, serve as common proxies for these microstructural
complexities, but are extremely sensitive to confounding factors
such as partial volume (for example, voxels containing a mixture of white matter and gray matter) and axonal dispersion27,29.
Accordingly, these measures should be used with caution to
quantify connection strength. Other anatomical factors such as
dendrite densities, spine densities, number of synapses at axon
terminals and synaptic efficacy are much harder to determine
noninvasively, though potentially are more relevant.
Unfortunately, tractography does not result in quantification
of any of the above properties. Often, probabilistic tractography,
which provides an estimate in the uncertainty of streamline trajectories, is used to quantify connection strength. Strong connections are expected to have a more discernible trace in the diffusion
data and therefore lower uncertainty in their trajectories. This
approximation, however, can easily break. For instance, locally
nondominant pathways (for example, those that cross larger
bundles), have greater uncertainty. Uncertainty is also affected
by nonrelevant factors such as signal-to-noise ratio and partial
volume effects. Another issue specific to streamlining is that
uncertainty in the streamline’s path increases with the length of
tract. Because streamlining operates by propagating uncertainty
spatially, connection probabilities inevitably decrease with distance. As a result, tractography-based structural connections are
difficult to quantify, threshold, compare between groups and use
for other types of statistical analyses27,29.
Interpretation and considerations. Pitfalls of tractography
can be divided into two categories: accuracy (correctness) and
precision (reproducibility)14. Accuracy refers to our ability to
infer axonal organization from measurements of water diffusion.
In the ideal case, there is no instrument-based or physiological noise, yet we can still make erroneous inferences regarding
microstructure because of inaccurate modeling. A white-matter
voxel contains hundreds of thousands of axons, which do not
necessarily align30. The fODF models, which account for multiple directions in a voxel (crossing fibers) are replacing tensor
models for tracto­graphy. However, the subvoxel organization
of axons can be more complex than a simple crossing and may
not always be easily recovered from the diffusion profile. For
instance, a collection of axons that bend in a voxel will create a
diffusion pattern that may not be easily distinguishable from that
of fiber dispersion. One can easily imagine even more complex
situations where all these configurations (for example, bending,
dispersion and crossing) happen in the same voxel. Diffusion data
from a single voxel cannot be used to unambiguously resolve these
complexities. Future approaches may benefit from semiglobal
review
models that aggregate diffusion data across multiple adjacent
voxels to infer subvoxel features.
With regard to precision, measurement noise (for example,
instrument-based or physiological) and inadequate waterdiffusion modeling can compromise tensor and fODF estimation, inducing spurious variations in the generated streamlines. Additionally, use of a fixed step size in the generation of
streamlines (despite local variations in anatomy) and the discrete
nature of voxels (when tracts are continuous) increase measurement error. Probabilistic tractography algorithms try to quantify
these errors by estimating the uncertainty in the entire process.
Uncertainty in voxelwise fiber orientation can be quantified31 and
propagated into uncertainties regarding the location of streamlines. This process turns 3D point estimates of streamline trajectories into spatial histograms of their locations (Fig. 2c).
Beyond concerns regarding accuracy and precision, identification of fibers in their entirety requires knowledge of where
tracts terminate throughout cortex. This remains a challenge for
tractography algorithms, which are very good at estimating the
location of bundles in deep white matter but not good (yet) at
identifying where they project into gray matter27. These difficulties result from a lack of detail in white-matter architecture modeled through diffusion and biases in cortical projections.
Estimating functional connectivity
Although the concept of structural connectivity is relatively intuitive given the presence of physical connections between brain
areas, its functional counterpart can be more challenging to
define. The macroconnectomics field has adopted a neurophysio­
logical perspective of functional connectivity, defining it as the
synchronization of neurophysiological events between spatially
remote brain areas32. First quantified in early electroencephalography and multiunit recording studies, functional connectivity analyses were adopted for positron emission tomography
and fMRI in 1993 (ref. 32). Although functional connectivity
can be measured noninvasively using a variety of neuroimaging
modalities (for example, positron emission tomography, fMRI
and magnetoencephalography) and different indices related to
physiological function (for example, blood oxygenation level–
dependent (BOLD), cerebral blood flow and glucose metabolism
analyses), BOLD-based fMRI is the most widely used technique
for inferring functional connectivity.
Studies of functional connectivity may be dichotomized on the
basis of the presence or absence of a task (that is, T-fMRI versus
task-free or ‘resting state’ fMRI (R-fMRI)). Task-based approaches
focus on the detection of synchronous responses to extrinsic
stimulation or tasks, referred to here as evoked functional connectivity (eFC) or coactivation33. Evoked functional connectivity
can be quantified across the entire period of task performance
or in response to specific types of events. Approaches using
R-fMRI focus on the detection of synchronized spontaneous
activity occurring in the absence of experimenter-controlled tasks
or stimuli, referred to as intrinsic functional connectivity (iFC)34.
Although iFC and eFC patterns can be notably similar, especially
when eFC is assessed using meta-analytic techniques35 or broad
comparisons (for example, task versus rest), these analyses probe
different aspects of the functional architecture33. eFC patterns
obtained using one task will not necessarily generalize to another,
and aspects of iFC obtained during one state may not necessarily
nature methods | VOL.10 NO.6 | JUNE 2013 | 529
review
Focus on mapping the brain
Box 2 iFC and Conscious States
© 2013 Nature America, Inc. All rights reserved.
The impact of cognitive97, physiological132 and pathological
states133 on functional connectivity has been recently demonstrated in the literature. Consciousness has received particular attention, with many studies examining physiological states
(for example, sleep), induced states (for example, anesthesia
and hypnosis) and pathological states (for example, coma,
vegetative syndrome and minimally conscious state133). iFC patterns detected in these states are grossly similar to
those observed during wakefulness, though direct comparison
reveals state-related iFC changes. For example, the default and
lateral networks each exhibit decreased functional connectivity among network components (that is, within-network
connectivity) during sleep, suggesting decreased integration
generalize to another (for example, wakefulness and sleep; see
Box 2 for a discussion of states other than wakeful rest).
Acquisition. BOLD is the predominant fMRI technique used
in studies of functional connectivity (see ref. 36 for alternative
cerebral blood flow–based technique). BOLD is measured using
ultrafast imaging sequences that are sensitive to relative concentrations of deoxyhemoglobin, which is paramagnetic (that is,
dephases the magnetic resonance signal), in contrast to oxyhemoglobin, which is diamagnetic37; the resulting measurement is
an indirect measure of neural activation. Data sets for functionalconnectivity analysis using fMRI are typically obtained in 5–30
minutes, as a participant either performs an experimental task
(T-fMRI) or rests quietly in the scanner (typically while awake;
R-fMRI). We discuss determinants of BOLD imaging acquisition
quality in Box 1.
Preprocessing. The aim of preprocessing is to remove confounding variation from data and facilitate comparison across subjects.
Structured nuisance signals and anatomical variation can obscure
functional connectivity measurements if left unaccounted for.
Despite considerable effort, we lack consensus regarding the optimal
set of preprocessing steps, their ordering and their implementation.
Most preprocessing steps originated with task-activation
approaches. However, their use and implications are greater for
functional-connectivity approaches because of the greater risk of
spurious findings given that the independent and dependent variables can be contaminated by the same noise signals (for example,
motion and respiration). Comprehensive comparison of preprocessing strategies and their implications for eFC and iFC analyses
remain elusive, in part owing to a lack of objective benchmarks.
The preprocessing steps described below are post-hoc corrections.
However, optimization of acquisition strategies to minimize the
impact of noise sources is preferred (Box 1).
Slice timing correction. The slices of an fMRI volume are
acquired at different times, creating effective shifts in time series
obtained at different slices. Although some question its necessity38, correction by temporal interpolation is recommended to
avoid the potential for deleterious impact of these lags on signal
denoising and time-series extraction from brain areas.
530 | VOL.10 NO.6 | JUNE 2013 | nature methods
of information. Complementary findings of decreased negative
iFC between these networks and others may suggest decreased
segregation as well134. These results suggest that although
sleep-based studies may be useful in populations not amenable to examination in wakeful states (for example, toddlers),
comparison of findings across states may be problematic. Of note, changes in thalamocortical connectivity are reported
during anesthesia, non–rapid eye movement (non-REM) sleep and vegetative states135, but the specific role of thalamo­
cortical circuitry in consciousness remains underexplored.
These studies also suggest potential clinical applications of
iFC, such as improving recognition of consciousness after
recovery from coma.
Motion correction. Head motion results in a misalignment of brain
areas between volumes typically accounted for using 3D imageregistration techniques. Additionally, head motion induces artifactual fMRI signal fluctuations resulting from changes in slice tissue
composition (partial voluming) and residual magnetization from
prior slice excitations (spin-history effects)39. These motion artifacts
are typically modeled and removed in a regression framework, containing predictors calculated from motion parameters estimated during coregistration40. Although effective, modeling-based approaches
do not completely remove motion-related fluctuations in the fMRI
signal41–43. To address this issue, the ‘scrubbing’ of offending volumes
via removal41 or spike regression44 has been proposed. Excluding
time points alters the temporal structure of the data, thereby compromising analyses that rely on this structure (for example, temporal
dynamics, spectral analysis and estimation of temporal autocorrelation). Regardless of the motion-correction scheme used, it is necessary to account for motion in group-level analyses42,45.
Physiological noise correction. Cardiac pulsation and respiration can induce fMRI signal fluctuations, which had led to early
criticisms attributing iFC to these physiological signals rather
than neural signals46. The cardiac cycle generates pulsatile motion
throughout the brain47. Respiratory movement of the chest and
abdomen induce changes in the magnetic field, producing intensity fluctuations in fMRI images47. Additionally, changes in cardiac
rhythm as well as rate and depth of breathing create longer-term
effects. Respiration and pulse can be recorded to model and subsequently remove their impact47. Although this is accepted as
ideal, it is not commonly performed. Instead, signals present in
white matter and cerebrospinal fluid are taken as surrogates for
respiration and cardiac effects, and regressed from the fMRI time
series. Incorporating spatial variation in the noise captured by the
white-matter signal provides superior denoising (for example,
anatomy-based correlation corrections (ANATICOR)48). Blinded
source-separation techniques provide another means of physio­
logical correction (for example, Corsica49).
Global signal regression. The mean time series across the whole
brain is commonly regressed from the data. In this model, the global
signal is considered a nonspecific measure of noise, whose removal
improves the specificity of iFC50, decreases motion effects44, and
Focus on mapping the brain
© 2013 Nature America, Inc. All rights reserved.
removes intersession and intersite effects. However, awareness
that global signal regression centers the correlation distribution
at zero, and thus introduces negative connections51 and can alter
inter-individual differences52, has made its use controversial. In
this regard, it is important to note that functional correlation coefficients obtained after global signal regression are relative values,
not absolute. Additionally, electrophysiological demonstrations of
globally synchronous neural signals in gray matter53 call into question the interpretation of the global signal as simply noise.
Temporal filtering. Bandpass filtering is usually performed to
remove frequencies below 0.001 hertz and greater than 0.08 hertz
from the fMRI time series. This frequency range targets removal of
low-frequency scanner drift and frequencies above those traditionally associated with functional connectivity34,54. However, complete removal of physiological noise is unlikely because of artifacts
induced by the low-temporal-resolution of fMRI (for example,
aliasing)46. Concerns about temporal filtering include reductions
in the degrees of freedom for the time series and recent demonstrations of functional connectivity at frequencies greater than
0.1 hertz for several brain areas, suggesting that low-pass filtering
is removing valuable signal55. Thus, despite historical precedent,
inclusion of low-pass filtering merits additional consideration.
Spatial normalization and smoothing. Another aspect of preprocessing is conditioning the data for comparison across subjects. Spatial normalization addresses morphological variation
across individuals by transforming subject data to a common
stereotactic space; population- and study-specific templates are
increasingly used to optimize correspondence. Spatial smoothing
additionally improves the correspondence of brain areas across
individuals and increases the signal-to-noise ratio56.
Estimating edges. Several mathematical modeling techniques can
be used to define functional relationships, differing primarily in
the stringency with which they define functional connectivity.
Functional connectivity simply implies a statistical dependency
between activities observed in brain areas and is an umbrella term
for a wide range of dependency measures, each providing a different perspective32. For example, mutual information measures
statistical dependency from the joint-probability distribution
function and is sensitive to linear and nonlinear relationships,
whereas Pearson’s correlation is primarily sensitive to linear relationships57. Effective connectivity, in contrast, requires a mathe­
matically precise (directional) description of the interactions
between brain areas58. This leads to a plurality of graphs that can
be derived from functional connectivity, with each graph characterizing functional interactions from a different perspective.
Estimating iFC from R-fMRI data typically begins with extraction of the mean time series across voxels in each brain node (that
is, parcellation-defined brain area). Intrinsic functional connectivity is commonly estimated from bivariate tests for statistical
dependency (for example, Pearson’s correlation, mutual information and spectral coherence) between every possible pairing of
time series59. Although these approaches perform well in simple
simulations59, the limited number of observations results in noisy
estimates of statistical relationships, which can be reduced using
regularization (shrinkage) methods60. A limitation of bivariate
approaches is that they do not account for information from
review
multiple brain areas simultaneously. Hence they cannot be used
to distinguish direct from indirect interactions (mediated by
common relationships with other areas). Partial correlation
(related to the inverse covariance matrix) results in estimates of
the conditional linear dependency between two brain areas, after
accounting for interactions with every other area61. Although
this approach is preferred to bivariate approaches, the number of
brain areas commonly exceeds the degrees of freedom, preventing
unique specification of partial correlations. In these cases, regularization techniques (for example, graphical lasso and elastic net)
can be used to find a solution59. Additionally, information can be
pooled across individuals to optimize estimation parameters62.
Note that some, but not all, of these approaches enforce symmetry; in other words, one can obtain dependency in one direction
but not the other (Box 3 ).
Many approaches exist for estimating eFC. Several authors have
borrowed approaches used to examine iFC; such approaches are
based on the assumption that the time series spans the entire
task63 or concatenated blocks of specific task conditions64.
Psychophysiological interaction39 analyses directly model interactions between patterns of functional connectivity and the experimental stimulus design, potentially offering greater specificity of
findings. Others have measured eFC from ‘coactivation’ using fitted regression coefficients65 or binarized (by applying a threshold
to regression coefficients) time series66 generated from a first-level
task analysis. Regression coefficient series are then compared using
correlation or partial correlation33,65. Binarized time series can
be compared from the joint distribution of the two values using
measures akin to mutual information66. Finally, meta-analytical
approaches provide a means of measuring eFC across studies and
often tasks, enabling detection of patterns of coactivation across
statistical maps generated from data in the literature35,67.
A variety of data-driven techniques are also used for identifying
iFC and eFC patterns. Examples include self-organizing maps68,
principal component analysis, normalized cut clustering 69 and
independent component analysis70. These methods are more
appropriate for identifying nodes of connectome graphs than
edges, although exceptions exist59.
Once functional connectivity is estimated, some applications
require it to be thresholded or binarized (that is, to determine
whether a connection is present or not). Threshold selection is
not straightforward but can be accomplished by applying a test
of statistical significance to each edge. When using parametric
statistics, care must be taken to adjust the degrees of freedom
for temporal autocorrelation. Alternatively, this can be addressed
using nonparametric tests of significance such as wavestrapping71
or circular-block bootstrap 72. Sparse covariance estimation
methods can also be used.
Interpretations and considerations. A common pursuit of
T-fMRI and R-fMRI studies is to fractionate the connectome into a
set of spatially and functionally distinct networks that can each be
annotated in terms of the specific functionality domain they subserve (for example, cognitive, affective or visceral). It is impressive
that these T-fMRI and R-fMRI studies have converged on similar
definitions of 8–20 spatially and functionally distinct networks,
though studies have suggested the actual number of networks is
substantially greater13. The concordance of T-fMRI and R-fMRI
findings suggests the brain’s intrinsic functional architecture
nature methods | VOL.10 NO.6 | JUNE 2013 | 531
review
Focus on mapping the brain
Box 3 Directionality in functional connectomics
© 2013 Nature America, Inc. All rights reserved.
Commonly referred to as effective connectivity, the mapping
of directional relationships is essential for characterizing
information flow in functional-connectivity studies136. Identification of neural drivers can facilitate our understanding of
control systems as well as ectopic foci leading to pathological
conditions (for example, epilepsy137). Although invasive tracing and stimulation techniques are powerful tools for mapping
directional relationships in nonhuman populations138, they are
not generally applicable in humans. Here we provide an overview of noninvasive approaches to establishing directionality.
Statistical techniques
Before reviewing statistical effective connectivity approaches,
we note that despite the nomenclature, their findings should
not be interpreted as indicating causality but rather the directionality of information flow. Structural equation modeling (SEM) and dynamic causal modeling (DCM) approaches
evaluate the fit of hypothesized models of directional inter­
actions among nodes with measured fMRI data. SEM is a
covariance-based approach that represents each node as an
exogenous variable (predicting the activity of another node),
endogenous variable (activity is predicted by another node)
or both139; effective connectivity is modeled at the hemodynamic (BOLD response) level. In contrast, DCM is a generative
approach, modeling effective connectivity at the neuronal level
based on a given biophysical model and uses a forward model
to produce the downstream fMRI activity. Models can be evaluated individually; however, common practice is to compare
models representing competing hypotheses regarding causality
to identify the ‘best-fit’ model. Concerns exist regarding the
feasibility of successfully identifying a ‘best fit’ model from a large population of putative models140. SEM and DCM approaches are limited in the number of nodes and interactions they can model efficiently. Primarily intended for confirmatory analysis, exploratory implementations of SEM141 and
DCM136 are emerging for use with R-fMRI.
Granger causality analysis (GCA) is a model-free, data-driven
effective connectivity approach142 that investigates whether
past values of the time series for one node could improve the
prediction of the current value in another143. In contrast to SEM
and DCM, GCA can be used to assess many nodes simultaneously. Application of GCA to R-fMRI144 is particularly controversial
because of (i) limitations imposed by the sluggish and variable
provides a framework for moment-to-moment responses to the
external world. As summarized in ref. 67, it appears that “the full
repertoire of functional networks utilized by the brain in action
is continuously and dynamically ‘active’ even when at ‘rest’.”
When considering the visualization of functional connectivity (Fig. 3), an important question is: ‘what are we missing?’.
Whereas functional connectivity is often represented with static
graphs, neurophysiological models have long asserted the transient nature of many functional interactions. Specifically, distributed neural assemblies appear to change their patterns of
interaction with one another from one cognitive act or state to
another. Consistent with this notion, eFC studies have noted
532 | VOL.10 NO.6 | JUNE 2013 | nature methods
hemodynamic response function (HRF), (ii) slow sampling rate
(for example, repetition time ≥ 2 s)59 and (iii) assumptions that
HRF characteristics are constant across nodes59. Attempts to
rehabilitate GCA for R-fMRI include HRF deconvolution before
GCA137, accounting for regional variation in the HRF using
breath-hold scans and faster sampling rates145. Nonetheless,
using GCA with R-fMRI remains problematic.
Lesion studies
Lesion studies directly perturb the system and therefore can
be compelling in establishing causality when carried out in
a controlled manner. From this perspective, changes in the
activity of a node after disruption of inputs from another are
taken to infer causal influences. In the ideal paradigm, scans
are obtained before and after occurrence of a given lesion. For
example146, R-fMRI scans from the brain of a 6-year-old child
before and after callosotomy revealed disruptions specific to
interhemispheric connectivity. Lesion studies are, however,
commonly limited to scans after lesion, taken after natural occurrence of lesions (for example, stroke, congenital abnor­
mality, neoplasm and seizure focus). Although useful for
testing predictions or generating hypotheses, findings of such
studies cannot be considered definitive.
Brain stimulation
The gold standard for establishing directional influences is the
demonstration that direct stimulation of node A impacts node B but not the converse. In this regard, intraoperative studies of
corticocortically evoked potentials—which involve the tracking of electrical signals from stimulation sites to other locations—are powerful tools147, though with limited applicability
(that is, in neurosurgical patients). Noninvasive stimulation
techniques with a broader range of applications are slowly
evolving. Transcranial magnetic stimulation (TMS) can be used to produce temporary and reversible neural excitation or
suppression in targeted cortical areas. Concurrent TMS-fMRI
studies have revealed causal relationships in motor circuits148
and the visual system149. Transcranial direct-current stimulation and transcranial alternating-current stimulation approaches deliver weak currents to targeted brain areas, and are emerging as less expensive and technically demanding
alternatives to TMS, though delivering focal stimulation is challenging150.
substantial task-dependency in their findings, even when looking at the same regions33,39. Perhaps most exciting, recent iFC
studies have observed dynamic changes in iFC patterns over a
5-minute scan73. These findings suggest that commonly used
metrics of iFC are incomplete, only capturing the ‘mean’ connectivity over time. If true, the implications would be multifold:
(i) findings of hypo- or hyper-connectivity in population studies would need to be reassessed, as they may reflect a different
distribution of time spent in the various iFC configurations
between populations, and (ii) the detection of changing eFC
patterns over the course of task performance may prove to be a
means of explaining observed behavioral variability. More and
Figure 3 | Visualizing the connectome.
(a) Classical anatomical-tracing–style depiction
of iFC for posteromedial cortex subdivisions
(image reproduced from ref. 113). (b) Matrix
representation of functional brain connections
predictive of diagnostic status in depression.
dmThalamus, dorsomedial thalamus; vMF10,
ventral medial prefrontal cortex (Brodmann
area (BA) 10); sgCg24|25, subgenual cingulate
(BA 24 and 25); aINS, anterior insula; rACC24,
right anterior cingulate cortex (BA 24); dlPFC9,
dorsolateral prefrontal cortex (BA 9); OFC11,
orbitofrontal cortex (BA 11); dMF10, dorsal
medial prefrontal cortex (BA 10); MCC24,
midcingulate cortex (BA 24); vPCC, ventral
posterior cingulate cortex; scACC25, subcallosal
cingulate cortex (BA 25) (image reproduced
from ref. 90). (c) Flatmap-based representation
of the CoCoMac atlas of the macaque
connectome (image reproduced from ref. 114).
(d) Connectogram depicts brain areas (nodes)
as columns in the circular band, differing
connectivity metrics in separate layers and
connections with lines; lobes are differentiated
by color, and left or right halves corresponds
to hemispheres (reproduced with permission
from ref. 115).
review
a
b
dmThalamus
vMF10
sgCg24|25
Hypothalamus
Nucleus accumbens
aINS
rACC24
dlPFC9
OFC11
dMF10
MCC24
vPCC
Amygdala
Hippocampus
scACC25
c
6.30
5.91
5.51
5.12
4.73
4.33
3.94
3.54
3.15
2.76
2.36
1.97
1.58
1.18
0.79
0.39
0.00
–0.39
–0.79
–1.18
–1.58
–1.97
–2.36
–2.76
–3.15
–3.54
–3.94
–4.33
–4.72
–5.12
–5.51
–5.91
–6.30
Connection strength (z)
© 2013 Nature America, Inc. All rights reserved.
Focus on mapping the brain
d
CCR
CCP
PCM
CCS
CCA
S1
PCS
M1
PMCM
PCIP
PFCM
MACD
FCD
MCD
PCI
V1
PFCD
FCPCL
FEF
FCC
A2
S2
MCV
FCV
V2
more studies are highlighting the potential value of examining transition zones
between functional areas in the brain5.
Examination of temporal dynamics may
inform such efforts, by mapping changes
in the boundary over time and providing
greater clarity for findings.
In addition to naturally occurring variations in iFC over time,
studies have suggested that iFC can be systematically impacted
by cognitive demands before R-fMRI data acquisition. By comparing iFC during R-fMRI scans collected before and after
task performance74, studies show that iFC strength within and
between networks is altered in a task-dependent manner. For
example, R-fMRI–based functional connectivity between the
inferior frontal gyrus and visual areas varied depending on the
category of stimuli viewed before the R-fMRI scan75. Across
participants, the extent to which iFC was modulated correlated
with subsequent memory for the stimuli. In addition to exhibiting plasticity related to tasks performed close in time to the
measurements, iFC is modulated by direct-stimulation protocols
including median nerve stimulation 76, heat pain77, transcranial
magnetic stimulation78 and transcranial direct current stimulation79. This suggests that R-fMRI may have utility in the identification of targets for stimulation protocols as well as assessment
of their efficacy (for example, in the context of the treatment
of depression80).
This iFC-based evidence of experience-induced plasticity provides strong support for the hypothesis that iFC reflects a history of coactivation among areas. However, this also suggests a
corollary: correlated intrinsic activity has a role in learning and
memory consolidation81. The demonstration of brain-behavior
correlations between task-related modulations of iFC and subsequent behavior (for example, recall) supports this hypothesis. If
short-term iFC alterations reflect experience-induced plasticity,
MACV
TCV
A1
IP
TCC
FCOR8
TCS
IA
TCI
PHC
CPO
then enduring changes would be expected after extended practice
or training. Several studies suggest this is the case82. Studies of
long-term training–induced plasticity have the potential to inform
our understanding of mechanisms involved in remediation-based
recovery of function or even to index the efficacy of treatment
interventions. For example, in a preliminary retrospective study,
differences in iFC were observed between children with dyslexia
who were remediated by reading interventions versus children
who received no treatment83.
Statistical analysis of the connectome
Once connectome graphs are estimated, the next goal is to annotate them in terms of their relevance to higher-order cognitive processes, neuropsychiatric diagnoses or other phenotypic
variables5. These associations are most often inferred by performing a categorical or dimensional statistical analysis that
compares connectivity across a population of individuals or
within an individual across time or treatments 84. Many of the
same statistical approaches are appropriate for the analysis of
connectome graphs regardless of whether they were constructed
with functional or structural data sets. However, the interpretation of the results must always account for the idiosyncratic
differences between structural and functional connectivity. For
example, functional connections are typically weighted and can
be positive or negative. Structural connectivity graphs tend to
be unweighted and are strictly nonnegative27. Additionally,
structural connections can be thought of as pathways along
nature methods | VOL.10 NO.6 | JUNE 2013 | 533
review
© 2013 Nature America, Inc. All rights reserved.
which information can flow, but functional connections
cannot be interpreted in the same manner85.
A bag of edges. The simplest approach to compare graphs is to
treat them as a bag, or collection, of edges and perform statistical
analyses at each edge one at a time, without taking into account
interactions or relationships between them (that is, edgewise
statistics)84. Such univariate approaches (for example, t-tests,
F-tests or regression) allow researchers to identify easily interpretable relationships between categorical or dimensional variables and edge weights. However, this approach results in the
need to perform many statistical tests, which require correction
for multiple comparisons to adequately control for the number
of false positives. Standard correction techniques such as false
discovery rate86 that do not model the dependencies between
edges may result in overly liberal or conservative corrections87.
Alternate correction techniques such as the network-based statistic88 or group Benjamini-Hochberg89 leverage information about
the group structure of connectome graphs to increase statistical
power, while maintaining control of false positives.
Alternatively, multivariate regression and classification techniques evaluate the relationship between the entire connectome
graphs and their associated phenotypic variables with a single
statistical test90,91. Although powerful for the analysis of connectome-phenotype relationships, they obscure information about
the involvement of individual edges. Extracting this information,
if desired, requires a return to edge-specific tests, and the need for
multiple-comparison correction90. Although these multivariate
techniques tend to be applied to bag-of-edges representations,
which ignore graph structure, they can also be performed using
graph-distance measures that preserve topological information
when comparing graphs92.
Node and graph-level statistics: invariants. Graphical representations of connectomes contain a wealth of information about
brain architecture beyond the presence and strength of bivariate
connections, which can be described using a variety of node-level
and graph-level statistics. These measures are called ‘invariants’
in graph-theory parlance or ‘topological measures’ in networktheory parlance because they are not unique to particular representations of the graph. The most commonly used node invariants
are centrality measures that indicate a node’s relative influence in
a graph. Several different centrality metrics are available that mea­
sure a node’s importance on the basis of the number and strength
of direct connections (degree centrality93), the importance of
neighboring nodes (eigenvector94 or Page Rank95 centrality) and
their role in connecting other pairs of nodes (betweenness85).
The various measures provide different perspectives on a node’s
role in the graph and, when combined, can lead to a more holistic
understanding of connectome-phenotype relationships95.
Similarly, a range of graph-level invariants is used for studying
structural and functional connectivity. In particular, graphs are
commonly assessed in terms of their local and global efficiency.
Local efficiency assesses the extent to which neighbors are densely
interconnected, whereas global efficiency captures the number of
connections that must be traversed to connect any two nodes93. The
relationships of these two measures to what would be obtained from
random graphs with similar properties can be combined to assess
the ‘small-worldness’ of a graph93. Small-world graphs balance
534 | VOL.10 NO.6 | JUNE 2013 | nature methods
Focus on mapping the brain
i­ ntegration and segregation to obtain fast and cost-efficient propagation of information through the graph as well as robustness to
single-node failures96. The cost-efficiency of a graph can be inferred
from the difference between global efficiency and the number of
edges in the graph93. An additional invariant is modularity, which
quantifies the extent to which a graph can be segregated into densely
intraconnected but sparsely interconnected modules and allows
direct comparison of module membership between graphs85.
Each of the previously described node and graph invariants can be statistically evaluated to identify relationships with
categorical and dimensional phenotypes. Although invariants
can increase statistical power by decreasing the number of multiple comparisons, the resulting relationships can be more difficult to interpret. When comparing invariants between graphs,
it is important to consider the impact of potential differences in
graph properties (for example, number of edges) that can systematically differ between individuals or groups and confound
interpretation of findings93. Additionally, as the distribution
properties of most invariants are poorly characterized, nonparametric statistical tests are preferred93.
Predictive modeling
Finally, researchers frequently aim to identify connectivity
patterns predictive of a phenotypic variable (for example, diagnosis90, age91 or brain state97). Predictive modeling can be used
to directly assess the ability of a connectivity pattern to predict the phenotype of an individual, in contrast to inferential
statistics, which evaluate improbability of a set of relationships
arising by chance98. Predictive modeling is typically supervised,
with the training set consisting of connectivity graphs and their
associated phenotypes99. One can assess the predictive accuracy via cross-validation99 or other model-selection techniques.
Predictive modeling has primarily focused on invariants and
bag-of-edges–style90,100 approaches.
Translational connectomics
MRI-based approaches to connectomics research are rapidly
transforming neuroscience in animal models as well, by removing barriers to longitudinal examinations associated with invasive techniques (for example, animal killing and injection of toxic
chemicals). The recent Mouse Biomedical Informatics Research
Network initiative (http://www.loni.ucla.edu/BIRN/Projects/
Mouse/) provides an initial demonstration of the potential to
complement cross-sectional atlases of the developing brain generated using histology approaches with longitudinal atlases obtained
using dMRI. Simultaneously, R-fMRI is emerging as a powerful
tool for comparative functional neuroanatomy studies. Initial
work has demonstrated impressive correspondence between the
iFC observed in humans and macaques for homologous functional networks supporting an array of functions, including those
that are putatively ‘human’ (for example, language, self-referential
processes and cognition)101. Evidence of homologies with patterns of iFC in lower mammals, such as rats, underscores this
translational potential102. Armed with increasingly powerful
imaging-based tools, macroscale connectomics studies in animal models are poised to provide a mechanistic understanding of
brain function through the combination of noninvasive imaging
with direct structural, pharmacological, molecular and genetic
manipulations that are impossible in humans.
Focus on mapping the brain
review
Table 2 | Initiatives promising to accelerate macroscale connectomics research
© 2013 Nature America, Inc. All rights reserved.
Initiative (weblink when available; location)
Goals
Brain Genomics Superstruct (United States)
Aims to collect a large-scale imaging data set to explore brain–behavior relationships and their genetic
influences. The initiative has collected R-fMRI, dMRI and saliva samples from over 3,000 adults, along
with comprehensive phenotyping data (cognition, personality and lifestyle), and the resultant repository
containing 1,500 completed, quality-pass data sets is expected to be publicly available in 2013.
Brainnetome (http://www.brainnetome.org/;
China)
Attempts to characterize brain networks with multimodal neuroimaging techniques, from the
microscale (microtechnique, ultramicrotomy, staining and visualization techniques) to the macroscale
(electroencephalography, fMRI and dMRI). R-fMRI and diffusion-imaging data sets, along with
behavioral and blood data from more than 1,000 patients with schizophrenia, 300 patients with
Alzheimer’s disease and mild cognitive impairment, 120 patients who had a stroke, 50 patients with
glioma and 2,000 healthy controls collected from 11 hospitals and imaging centers.
Consortium of Neuroimagers for the Noninvasive
Exploration of Brain Connectivity and Tracts
(http://www.brain-connect.eu/; European Union)
Consortium focused on studying the brain’s microstructure, tracts and connectivity using
dMRI. Target deliverables include optimized acquisition protocols, analytic tools and a
connectivity atlas.
Developing Human Connectome Project
(European Union)
Initiative to comprehensively map and model the human connectome for 1,000 babies, including
in utero and in vivo imaging (20–44 weeks after conception).
US National Institutes of Health Human
Connectome Project: Washington University in
Saint Louis–University of Minnesota consortium
(http://humanconnectome.org/; United States)
State-of-the-art multimodal imaging initiative (R-fMRI, T-fMRI, dMRI and magnetoencephalography)
that makes use of a twin design (1,200 healthy adults, including twin pairs and their siblings from
300 families) to provide insights into relationships between brain connectivity, behavior and genetics.
The project uses multiband imaging sequences for R-fMRI (high spatial and temporal resolution) and
dMRI (high spatial resolution), which it has refined and is currently distributing to interested centers.
All data and tools developed through the initiative will be openly shared.
US National Institutes of Health Human
Connectome Project: MGH-Harvard-UCLA
consortium (http://humanconnectomeproject.org/;
United States)
Initiative focusing on unraveling the full connectivity map using the first ‘Connectome Scanner’,
which is designed to carry out diffusion using ultrahigh gradient strength (4–8 times the strength
of conventional systems). Efforts to optimize dMRI technology will focus on increasing the spatial
resolution, quality and speed of acquisition.
1000 Functional Connectomes Project (FCP)
(http://fcon_1000.projects.nitrc.org/; global)
Grass-roots data-sharing initiative that brought together over 1,200 previously collected R-fMRI
data sets from 33 independent sites around the world and released them openly to the scientific
community via the Neuroimaging Informatics Tool Resources Clearinghouse (http://www.nitrc.org/).
International Neuroimaging Data-sharing
Initiative (INDI) (global)
Second FCP initiative that was founded in an attempt to (i) expand the scope of open data sharing in
the functional neuroimaging community to include phenotypic data beyond age and sex (major INDI
data releases: ADHD-200 Consortium (http://fcon_1000.projects.nitrc.org/indi/adhd200/), the Autism
Brain Imaging Data Exchange (ABIDE; http://con_1000.projects.nitrc.org/indi/abide/)) and (ii) provide
a model for prospective, prepublication data sharing (major release: the Nathan Kline Institute-Rockland
Sample http://fcon_1000.projects.nitrc.org/indi/enhanced/). More than 5,000 R-fMRI data sets are
available through the FCP and INDI efforts combined as well as a growing number of dMRI data sets.
UK Biobank Imaging (http://www.ukbiobank.
ac.uk/; UK)
Building on an existing long-term prospective epidemiological study that has collected genetics,
blood samples and lifestyle information from a cohort of 500,000 subjects, the UK Biobank Imaging
Extension aims to resample 100,000 of the cohort using multimodal neuroimaging (including but not
limited to R-fMRI and dMRI), as well as cardiac MRI and rich phenotyping.
Large-scale initiatives from around the world that are promising to accelerate the pace of macroscale connectomics research through either the advancement of macroconnectomics research
through the generation and sharing of large-scale imaging data sets with phenotyping or innovation of data-acquisition and/or analysis techniques (see ref. 116 for additional information
regarding these initiatives and others).
Despite the rich promise of translational connectomics, methodological issues must also be addressed. For example, iFC can
be examined in awake rats that have been habituated to restraint
in the loud MRI environment103. However, most studies are
conducted under anesthesia—in particular, using the general
anesthetic isoflurane104, which can confound findings owing to
its effects on neural excitability. The sedating alpha-2 adrenergic agonist medetomidine may be preferable as it avoids such
confounds105. Dose-response studies of anesthesia are few104
and are essential. Initial translational studies in monkeys, rats
and mice have relied on preprocessing and analytical approaches
identical to those developed in humans103. Although their success is encouraging, differences in physiological (for example,
cardiac activity and respiration) and imaging parameters must be
explored to arrive at optimal strategies. Finally, we note that the
many questions raised regarding the interpretation of dMRI and
R-fMRI techniques in humans also apply to animal studies.
Toward neurophenotypes and clinical applications
An overarching goal of the connectomics era is the derivation
of ‘neurophenotypes’106, a concept that remains poorly specified
despite increasing enthusiasm from investigators. An individual’s macroscale connectome and its subgraphs contribute to the
specification of that individual’s neurophenotype. A central goal
of connectomics is to catalog neurophenotypes and relate them
to phenotypic profiles4. This can be accomplished through datadriven approaches focused on the detection of commonalities and
distinctions in connectomes or by differentiating populations of
neurophenotypes based on their phenotypic profiles. The breadth
of phenotyping can vary depending on the application, though
it typically consists of some combination of cognitive, affective,
behavioral, neurological or psychiatric variables. When cataloging
neurophenotypes based upon macroscale connectomes, the specificity of findings will depend on their nature, granularity of
node definitions and quality of neuroimaging data used. Similarly,
nature methods | VOL.10 NO.6 | JUNE 2013 | 535
review
Focus on mapping the brain
Table 3 | Connectomics studies in clinical populations
© 2013 Nature America, Inc. All rights reserved.
Neuropsychiatric diagnoses
Schizophrenia
Alzheimer’s disease
Depression
Epilepsy and seizures
Mild cognitive impairment
Other neurological disorders
Substance dependence
Attention deficit hyperactivity disorder
Autism spectrum disorders
Multiple sclerosis
Parkinson’s disease
Traumatic brain injury
Sleep disorders
Stroke
Anxiety disorders
Coma and vegetative state
Obsessive compulsive disorder
Amyotrophic lateral sclerosis
Bipolar disorder
R-fMRI count
dMRI count
100
89
81
62
58
45
40
32
26
22
19
18
17
15
15
15
12
10
10
283
201
101
179
115
50
94
37
86
211
73
259
7
182
16
9
27
84
61
Neuropsychiatric disorders most commonly studied using macroscale functional and structural
connectivity. Determined by the number (count) of R-fMRI and dMRI papers dedicated to the
neuropsychiatric diagnosis. Calculated from the Child Mind Institute Librarian (Resting State
and DTI libraries; (http://www.mendeley.com/profiles/cmi-librarian/).
field through the introduction of spurious, irreproducible findings
associated with suboptimal methodologies. Conversely, increased
attention to the acquisition of high-quality data, combined with
optimized preprocessing and analytic methodologies can serve to
accelerate the pace at which connectomes can be meaningfully
annotated and their variations cataloged.
Acknowledgments
This work was supported by grants from US National Institute of Mental Health
(BRAINS R01MH094639 to M.P.M. and K23MH087770 to A.D.M.), the Stavros
Niarchos Foundation (M.P.M.), the Brain and Behavior Research Foundation
(R.C.C.) and the Leon Levy Foundation (C.K. and A.D.M.). J.T.V. receives funding
from the London Institute for Mathematical Sciences HDTRA1-11-1-0048 and
US National Institutes of Health R01ES017436. Additional support was provided
by a gift from Joseph P. Healey to the Child Mind Institute (M.P.M.). We thank
D. Lurie for his assistance in the preparation of the manuscript and references
as well as Z. Shehzad, Z. Yang and S. Urchs for their helpful comments.
We acknowledge our colleagues who allowed us to reproduce their figures.
COMPETING FINANCIAL INTERESTS
The authors declare competing financial interests: details are available in the
online version of the paper.
Reprints and permissions information is available online at http://www.nature.
com/reprints/index.html.
1.
when sorting neurophenotypes on the basis of phenotypic profiles, specificity will be determined by the precision and comprehensiveness (that is, number and breadth of independent features)
of the phenotyping available to statistical analysis. Future work
will need to find a balance between categorical and dimensional
perspectives of neurophenotypes.
Beyond the derivation of a fundamental understanding of brain
architecture and its implications for behavior and cognition, a
major reason for the excitement surrounding connectomics is
the promise of clinical utility because of the ability to obtain individual-relevant reliable brain indices (see Table 2 for initiatives
that are accelerating the pace of macroconnectomics research).
Recent years have witnessed an explosion in the number of neurological and psychiatric disorders studied with dMRI and R-fMRI
(Table 3). Hopes of attaining clinically useful diagnostic tools are
increasingly espoused in the literature. However, leaders in the
field have recently suggested that the attainment of tools capable
of stratifying individuals based upon disease risk, prognosis and
treatment response may prove to be a more fruitful goal than
focusing on diagnosis107. Regardless, a key requirement remains:
attaining large-scale data sets representative of the human population. In this regard, the macroconnectomics community has
supported several large-scale data-sharing initiatives dedicated
to rapidly aggregating the necessary data108.
2.
Conclusion
The connectomics era is the culmination of more than a century of conceptual and methodological innovation. MRI-based
approaches to mapping and annotating the connectome at the
macroscale are transforming basic, translational and clinical
neuroscience research by overcoming barriers to progress faced
by more traditional invasive methodologies. In this Review we
broadly surveyed the many challenges that remain in the acquisition, preprocessing and analysis of brain-imaging data. Failure to
consider the many complexities could jeopardize this burgeoning
15.
536 | VOL.10 NO.6 | JUNE 2013 | nature methods
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
16.
17.
18.
Sporns, O., Tononi, G. & Kötter, R. The human connectome: A structural
description of the human brain. PLoS Comput. Biol. 1, e42 (2005).
Varela, F., Lachaux, J.P., Rodriguez, E. & Martinerie, J. The brainweb:
phase synchronization and large-scale integration. Nat. Rev. Neurosci. 2,
229–239 (2001).
Biswal, B.B. et al. Toward discovery science of human brain function.
Proc. Natl. Acad. Sci. USA 107, 4734–4739 (2010).
Behrens, T.E.J. & Sporns, O. Human connectomics. Curr. Opin. Neurobiol.
22, 144–153 (2012).
Kelly, C., Biswal, B.B., Craddock, R.C., Castellanos, X.F. & Milham, M.P.
Characterizing variation in the functional connectome: promise and
pitfalls. Trends Cogn. Sci. 16, 181–188 (2012).
Talairach, J. & Tournoux, P. Co-planar Stereotaxic Atlas of the Human Brain
(Thieme Classics, 1988).
Margulies, D.S. et al. Mapping the functional connectivity of anterior
cingulate cortex. Neuroimage 37, 579–588 (2007).
Beckmann, M., Johansen-Berg, H. & Rushworth, M.F.S. Connectivity-based
parcellation of human cingulate cortex and its relation to functional
specialization. J. Neurosci. 29, 1175–1190 (2009).
Dosenbach, N.U. et al. A core system for the implementation of task
sets. Neuron 50, 799–812 (2006).
Bellec, P. et al. Identification of large-scale networks in the brain using
fMRI. Neuroimage 29, 1231–1243 (2006).
Craddock, R.C., James, G.A., Holtzheimer, P.E., Hu, X.P. & Mayberg, H.S.
A whole brain fMRI atlas generated via spatially constrained spectral
clustering. Hum. Brain Mapp. 33, 1914–1928 (2012).
Cohen, A.L. et al. Defining functional areas in individual human brains
using resting functional connectivity MRI. Neuroimage 41, 45–57
(2008).
Kiviniemi, V. et al. Functional segmentation of the brain cortex using
high model order group PICA. Hum. Brain Mapp. 30, 3865–3886
(2009).
Jones, D.K. Challenges and limitations of quantifying brain connectivity
in vivo with diffusion MRI. Imaging 2, 341–355 (2010).
Andersson, J.L.R., Skare, S. & Ashburner, J. How to correct susceptibility
distortions in spin-echo echo-planar images: application to diffusion
tensor imaging. Neuroimage 20, 870–888 (2003).
Alexander, A.L., Tsuruda, J.S. & Parker, D.L. Elimination of eddy current
artifacts in diffusion-weighted echo-planar images: the use of bipolar
gradients. Magn. Reson. Med. 38, 1016–1021 (1997).
Haselgrove, J.C. & Moore, J.R. Correction for distortion of echo-planar
images used to calculate the apparent diffusion coefficient. Magn. Reson.
Med. 36, 960–964 (1996).
Anderson, J. et al. A comprehensive Gaussian Process framework for
correcting distortions and movements in diffusion images. in Proceedings
of the International Society for Magnetic Resonance in Medicine (ISMRM)
20th Annual Meeting and Exhibition (Melbourne, 2012).
© 2013 Nature America, Inc. All rights reserved.
Focus on mapping the brain
19. Sotiropoulos, S.N. et al. Effects of image reconstruction on fibre
orientation mapping from multichannel diffusion MRI: reducing the noise
floor using SENSE. Magn. Reson. Med. advance online publication,
doi:10.1002/mrm.24623 (7 February 2013).
20. Leemans, A. & Jones, D.K. The B-matrix must be rotated when correcting
for subject motion in DTI data. Magn. Reson. Med. 61, 1336–1349 2009).
21. Seunarine, K. & Alexander, D. Multiple fibers: beyond the diffusion
tensor. in Diffusion MRI: From Quantitative Measurement to in vivo
Neuroanatomy (eds., Johansen-Berg, H. and Behrens, T.E.J.) 55–72
(Academic Press, 2009).
22. Basser, P.J., Mattiello, J. & LeBihan, D. Estimation of the effective
self-diffusion tensor from the NMR spin echo. J. Magn. Reson. B 103,
247–254 (1994).
23. Tuch, D.S. et al. High angular resolution diffusion imaging reveals
intravoxel white matter fiber heterogeneity. Magn. Reson. Med. 48,
577–582 (2002).
24. Wedeen, V.J., Hagmann, P., Tseng, W.Y., Reese, T.G. & Weisskoff, R.M.
Mapping complex tissue architecture with diffusion spectrum magnetic
resonance imaging. Magn. Reson. Med. 54, 1377–1386 (2005).
25. Behrens, T.E., Berg, H.J., Jbabdi, S., Rushworth, M.F. & Woolrich, M.W.
Probabilistic diffusion tractography with multiple fibre orientations: What
can we gain? Neuroimage 34, 144–155 (2007).
26. Catani, M., Howard, R.J., Pajevic, S. & Jones, D.K. Virtual in vivo
interactive dissection of white matter fasciculi in the human brain.
Neuroimage 17, 77–94 (2002).
27. Jbabdi, S. & Johansen-Berg, H. Tractography: where do we go from here?
Brain Connect. 1, 169–183 (2011).
28. Assaf, Y., Blumenfeld-Katzir, T., Yovel, Y. & Basser, P.J. AxCaliber: a
method for measuring axon diameter distribution from diffusion MRI.
Magn. Reson. Med. 59, 1347–1354 (2008).
29. Jones, D.K., Knosche, T.R. & Turner, R. White matter integrity, fiber
count, and other fallacies: The do’s and don’ts of diffusion MRI.
Neuroimage 73, 239–254 (2013).
30. Mikula, S., Binding, J. & Denk, W. Staining and embedding the whole
mouse brain for electron microscopy. Nat. Methods 9, 1198–1201 (2012).
31. Behrens, T. & Jbabdi, S. MR diffusion tractography. in Diffusion MRI:
From Quantitative Measurement to in vivo Neuroanatomy (eds.,
Johansen-Berg, H. and Behrens, T.E.J.) 333–352 (Academic Press, 2009).
32. Friston, K.J., Frith, C.D., Liddle, P.F. & Frackowiak, R.S. Functional
connectivity: the principal-component analysis of large (PET) data sets.
J. Cereb. Blood Flow Metab. 13, 5–14 (1993).
33. Mennes, M., Kelly, C., Colcombe, S., Castellanos, F.X. & Milham, M.P.
The extrinsic and intrinsic functional architectures of the human brain
are not equivalent. Cereb. Cortex 23, 223–229 (2013).
34. Biswal, B., Yetkin, F.Z., Haughton, V.M. & Hyde, J.S. Functional
connectivity in the motor cortex of resting human brain using echoplanar MRI. Magn. Reson. Med. 34, 537–541 (1995).
35. Toro, R., Fox, P.T. & Paus, T. Functional coactivation map of the human
brain. Cereb. Cortex 18, 2553–2559 (2008).
36. Chuang, K.-H. et al. Mapping resting-state functional connectivity using
perfusion MRI. Neuroimage 40, 1595–1605 (2008).
37. Ogawa, S. et al. Intrinsic signal changes accompanying sensory
stimulation: functional brain mapping with magnetic resonance imaging.
Proc. Natl. Acad. Sci. USA 89, 5951–5955 (1992).
38. Wu, C.W. et al. Empirical evaluations of slice-timing, smoothing, and
normalization effects in seed-based, resting-state functional magnetic
resonance imaging analyses. Brain Connect. 1, 401–410 (2011).
39. Friston, K.J. et al. Psychophysiological and modulatory interactions in
neuroimaging. Neuroimage 6, 218–229 (1997).
40. Friston, K.J., Williams, S., Howard, R., Frackowiak, R.S. & Turner, R.
Movement-related effects in fMRI time-series. Magn. Reson. Med. 35,
346–355 (1996).
41. Power, J.D., Barnes, K.A., Snyder, A.Z., Schlaggar, B.L. & Petersen, S.E.
Spurious but systematic correlations in functional connectivity MRI
networks arise from subject motion. Neuroimage 59, 2142–2154 (2012).
42. Van Dijk, K.R.A., Sabuncu, M.R. & Buckner, R.L. The influence of head
motion on intrinsic functional connectivity MRI. Neuroimage 59, 431–438
(2012).
43. Satterthwaite, T.D. et al. Impact of in-scanner head motion on multiple
measures of functional connectivity: Relevance for studies of
neurodevelopment in youth. Neuroimage 60, 623–632 (2012).
44. Satterthwaite, T.D. et al. An improved framework for confound regression
and filtering for control of motion artifact in the preprocessing of
resting-state functional connectivity data. Neuroimage 64, 240–256
(2013).
review
45. Yan, C.G. et al. A comprehensive assessment of regional variation in the
impact of head micromovements on functional connectomics. Neuroimage
76C, 183–201 (2013).
46. Lund, T.E. fcMRI–mapping functional connectivity or correlating cardiacinduced noise? Magn. Reson. Med. 46, 628–629 (2001).
47. Birn, R.M. The role of physiological noise in resting-state functional
connectivity. Neuroimage 62, 864–870 (2012).
48. Jo, H.J., Saad, Z.S., Simmons, W.K., Milbury, L.A. & Cox, R.W. Mapping
sources of correlation in resting state fMRI, with artifact detection and
removal. Neuroimage 52, 571–582 (2010).
49. Perlbarg, V. et al. CORSICA: correction of structured noise in fMRI by
automatic identification of ICA components. Magn. Reson. Imaging 25,
35–46 (2007).
50. Fox, M.D., Zhang, D., Snyder, A.Z. & Raichle, M.E. The global signal and
observed anticorrelated resting state brain networks. J. Neurophysiol.
101, 3270–3283 (2009).
51. Murphy, K., Birn, R.M., Handwerker, D.a., Jones, T.B. & Bandettini, P.a.
The impact of global signal regression on resting state correlations:
are anti-correlated networks introduced? Neuroimage 44, 893–905
(2009).
52. Saad, Z. et al. Trouble at rest: how correlation patterns and group
differences become distorted after global signal regression. Brain
Connect. 2, 25–32 (2012).
53. Schölvinck, M.L., Maier, A., Ye, F.Q., Duyn, J.H. & Leopold, D.A. Neural
basis of global resting-state fMRI activity. Proc. Natl. Acad. Sci. USA
107, 10238–10243 (2010).
54. Cordes, D. et al. Frequencies contributing to functional connectivity in
the cerebral cortex in “resting-state” data. AJNR Am. J. Neuroradiol. 22,
1326–1333 (2001).
55. Niazy, R.K., Xie, J., Miller, K., Beckmann, C.F. & Smith, S.M. Spectral
characteristics of resting state networks. Prog. Brain Res. 193, 259–276
(2011).
56. Worsley, K.J., Marrett, S., Neelin, P. & Evans, A.C. Searching scale space
for activation in PET images. Hum. Brain Mapp. 4, 74–90 (1996).
57. Marrelec, G. et al. Regions, systems, and the brain: hierarchical measures
of functional integration in fMRI. Med. Image Anal. 12, 484–496
(2008).
58. Friston, K.J. & Frith, C.D. Time dependent changes in effective
connectivity measured with PET. Hum. Brain Mapp. 1, 69–79 (1993).
59. Smith, S.M. et al. Network modelling methods for fMRI. Neuroimage 54,
875–891 (2011).
60. Varoquaux, G., Gramfort, A. & Poline, J.B. Advances in Neural Information
Processing Systems. (Vancouver, Canada, 2010).
61. Marrelec, G. et al. Partial correlation for functional brain interactivity
investigation in functional MRI. Neuroimage 32, 228–237 (2006).
62. Varoquaux, G. et al. A group model for stable multi-subject ICA on fMRI
datasets. Neuroimage 51, 288–299 (2010).
63. Lowe, M.J., Dzemidzic, M., Lurito, J.T., Mathews, V.P. & Phillips, M.D.
Correlations in low-frequency BOLD fluctuations reflect cortico-cortical
connections. Neuroimage 12, 582–587 (2000).
64. Sun, F.T., Miller, L.M. & D’Esposito, M. Measuring interregional functional
connectivity using coherence and partial coherence analyses of fMRI
data. Neuroimage 21, 647–658 (2004).
65. Rissman, J., Gazzaley, A. & D’Esposito, M. Measuring functional
connectivity during distinct stages of a cognitive task. Neuroimage 23,
752–763 (2004).
66. Patel, R.S., Bowman, F.D. & Rilling, J.K. A Bayesian approach to
determining connectivity of the human brain. Hum. Brain Mapp. 27,
267–276 (2006).
67. Smith, S.M. et al. Correspondence of the brain’s functional architecture
during activation and rest. Proc. Natl. Acad. Sci. USA 106, 13040–13045
(2009).
68. Peltier, S.J., Polk, T.A. & Noll, D.C. Detecting low-frequency functional
connectivity in fMRI using a self-organizing map (SOM) algorithm.
Hum. Brain Mapp. 20, 220–226 (2003).
69. van den Heuvel, M., Mandl, R. & Hulshoff Pol, H. Normalized cut group
clustering of resting-state fMRI data. PLoS ONE 3, e2001 (2008).
70. Beckmann, C.F. Modelling with independent components. Neuroimage 62,
891–901 (2012).
71. Breakspear, M., Brammer, M.J., Bullmore, E.T., Das, P. & Williams, L.M.
Spatiotemporal wavelet resampling for functional neuroimaging data.
Hum. Brain Mapp. 23, 1–25 (2004).
72. Bellec, P., Marrelec, G. & Benali, H. A bootstrap test to investigate
changes in brain connectivity for functional MRI. Stat. Sin. 18,
1253–1268 (2008).
nature methods | VOL.10 NO.6 | JUNE 2013 | 537
© 2013 Nature America, Inc. All rights reserved.
review
73. Smith, S.M. et al. Temporally-independent functional modes of
spontaneous brain activity. Proc. Natl. Acad. Sci. USA 109, 3131–3136
(2012).
74. Waites, A.B., Stanislavsky, A., Abbott, D.F. & Jackson, G.D. Effect of prior
cognitive state on resting state networks measured with functional
connectivity. Hum. Brain Mapp. 24, 59–68 (2005).
75. Stevens, W.D., Buckner, R.L. & Schacter, D.L. Correlated low-frequency
BOLD fluctuations in the resting human brain are modulated by recent
experience in category-preferential visual regions. Cereb. Cortex 20,
1997–2006 (2010).
76. Klingner, C.M., Hasler, C., Brodoehl, S., Axer, H. & Witte, O.W. Perceptual
plasticity is mediated by connectivity changes of the medial thalamic
nucleus. Hum. Brain Mapp. advance online publication, doi:10.1002/
hbm.22074 (25 March 2012).
77. Riedl, V. et al. Repeated pain induces adaptations of intrinsic brain
activity to reflect past and predict future pain. Neuroimage 57, 206–213
(2011).
78. Vercammen, A., Knegtering, H., Liemburg, E.J., den Boer, J.A. &
Aleman, A. Functional connectivity of the temporo-parietal region in
schizophrenia: effects of rTMS treatment of auditory hallucinations.
J. Psychiatr. Res. 44, 725–731 (2010).
79. Keeser, D. et al. Prefrontal transcranial direct current stimulation changes
connectivity of resting-state networks during fMRI. J. Neurosci. 31,
15284–15293 (2011).
80. Fox, M.D., Buckner, R.L., White, M.P., Greicius, M.D. & Pascual-Leone, A.
Efficacy of transcranial magnetic stimulation targets for depression is
related to intrinsic functional connectivity with the subgenual cingulate.
Biol. Psychiatry 72, 595–603 (2012).
81. Tambini, A., Ketz, N. & Davachi, L. Enhanced brain correlations during
rest are related to memory for recent experiences. Neuron 65, 280–290
(2010).
82. Lewis, C.M., Baldassarre, A., Committeri, G., Romani, G.L. & Corbetta, M.
Learning sculpts the spontaneous activity of the resting human brain.
Proc. Natl. Acad. Sci. USA 106, 17558–17563 (2009).
83. Koyama, M.S. et al. Cortical signatures of dyslexia and remediation: an
intrinsic functional connectivity approach. PLoS ONE 8, e55454 (2013).
84. Varoquaux, G. & Craddock, R.C. Learning and comparing functional
connectomes across subjects. Neuroimage advance online publication,
doi:10.1016/j.neuroimage.2013.04.007 (11 April 2013).
85. Rubinov, M. & Sporns, O. Complex network measures of brain
connectivity: uses and interpretations. Neuroimage 52, 1059–1069
(2010).
86. Genovese, C.R., Lazar, N.A. & Nichols, T. Thresholding of statistical maps
in functional neuroimaging using the false discovery rate. Neuroimage
15, 870–878 (2002).
87. Efron, B. When should hypothesis testing problems be combined?
Ann. Appl. Stat. 2, 197–223 (2008).
88. Zalesky, A., Fornito, A. & Bullmore, E.T. Network-based statistic:
identifying differences in brain networks. Neuroimage 53, 1197–1207
(2010).
89. Hu, J.X., Zhao, H. & Zhou, H.H. False Discovery Rate Control With
Groups. J. Am. Stat. Assoc. 105, 1215–1227 (2010).
90. Craddock, R.C., Holtzheimer, P.E., Hu, X.P. & Mayberg, H.S. Disease state
prediction from resting state functional connectivity. Magn. Reson. Med.
62, 1619–1628 (2009).
91. Dosenbach, N.U.F. et al. Prediction of individual brain maturity using
fMRI. Science 329, 1358–1361 (2010).
92. Bunke, H. A graph distance metric based on the maximal common
subgraph. Pattern Recognit. Lett. 19, 255–259 (1998).
93. Bullmore, E. & Sporns, O. Complex brain networks: graph theoretical
analysis of structural and functional systems. Nat. Rev. Neurosci. 10,
186–198 (2009).
94. Bullmore, E.T. & Bassett, D.S. Brain graphs: graphical models of the
human brain connectome. Annu. Rev. Clin. Psychol. 7, 113–140 (2011).
95. Zuo, X.N. et al. Network centrality in the human functional connectome.
Cereb. Cortex 22, 1862–1875 (2012).
96. Bassett, D.S., Meyer-Lindenberg, A., Achard, S., Duke, T. & Bullmore, E.
Adaptive reconfiguration of fractal small-world human brain functional
networks. Proc. Natl. Acad. Sci. USA 103, 19518–19523 (2006).
97. Richiardi, J., Eryilmaz, H., Schwartz, S., Vuilleumier, P. & Van De Ville, D.
Decoding brain states from fMRI connectivity graphs. Neuroimage 56,
616–626 (2011).
98. Hansen, L.K. Multivariate strategies in functional magnetic resonance
imaging. Brain Lang. 102, 186–191 (2007).
538 | VOL.10 NO.6 | JUNE 2013 | nature methods
Focus on mapping the brain
99. Hastie, T., Tibshirani, R. & Friedman, J. The Elements of Statistical
Learning (Springer, 2001).
100. Zhu, C.Z. et al. Discriminative analysis of brain function at resting-state
for attention-deficit/hyperactivity disorder. Med. Image Comput. Comput.
Assist. Interv. 8, 468–475 (2005).
101. Vincent, J.L. et al. Intrinsic functional architecture in the anaesthetized
monkey brain. Nature 447, 83–86 (2007).
102. Pawela, C.P. et al. Resting-state functional connectivity of the rat brain.
Magn. Reson. Med. 59, 1021–1029 (2008).
103. Becerra, L., Pendse, G., Chang, P.-C., Bishop, J. & Borsook, D. Robust
reproducible resting state networks in the awake rodent brain. PLoS ONE
6, e25701 (2011).
104. Wang, K. et al. Temporal scaling properties and spatial synchronization of
spontaneous blood oxygenation level-dependent (BOLD) signal
fluctuations in rat sensorimotor network at different levels of isoflurane
anesthesia. NMR Biomed. 24, 61–67 (2011).
105. Pawela, C.P. et al. A protocol for use of medetomidine anesthesia in rats
for extended studies using task-induced BOLD contrast and resting-state
functional connectivity. Neuroimage 46, 1137–1147 (2009).
106. Sörös, P. & Stanton, S.G. On variability and genes: inter-individual
differences in auditory brain function. Front. Hum. Neurosci. 6, 150
(2012).
107. Kapur, S., Phillips, A.G. & Insel, T.R. Why has it taken so long for
biological psychiatry to develop clinical tests and what to do about it?
Mol. Psychiatry 17, 1174–1179 (2012).
108. Mennes, M., Biswal, B.B., Castellanos, F.X. & Milham, M.P. Making data
sharing work: the FCP/INDI experience. Neuroimage advance online
publication, doi:10.1016/j.neuroimage.2012.10.064 (30 October 2012).
109. Tzourio-Mazoyer, N. et al. Automated anatomical labeling of activations
in SPM using a macroscopic anatomical parcellation of the MNI MRI
single-subject brain. Neuroimage 15, 273–289 (2002).
110. Desikan, R.S. et al. An automated labeling system for subdividing the
human cerebral cortex on MRI scans into gyral based regions of interest.
Neuroimage 31, 968–980 (2006).
111. Eickhoff, S.B. et al. A new SPM toolbox for combining probabilistic
cytoarchitectonic maps and functional imaging data. Neuroimage 25,
1325–1335 (2005).
112. Lancaster, J.L. et al. Automated Talairach atlas labels for functional brain
mapping. Hum. Brain Mapp. 10, 120–131 (2000).
113. Margulies, D.S. et al. Precuneus shares intrinsic functional architecture in
humans and monkeys. Proc. Natl. Acad. Sci. USA 106, 20069–20074
(2009).
114. Knock, S.A. et al. The effects of physiologically plausible connectivity
structure on local and global dynamics in large scale brain models.
J. Neurosci. Methods 183, 86–94 (2009).
115. Van Horn, J.D. et al. Mapping connectivity damage in the case of
Phineas Gage. PLoS ONE 7, e37454 (2012).
116. Jiang, T. Brainnetome: a new -ome to understand the brain and its
disorders. Neuroimage advance online publication, doi:10.1016/j.
neuroimage.2013.04.002 (6 April 2013).
117. Haacke, E.M., Bornw, R.W., Thompson, M.R. & Venkatesan, R. Magnetic
Resonance Imaging: Physical Principles and Sequence Design (Wiley-Liss,
1999).
118. Bernstein, M.A., King, K.F. & Zhou, X.J. Handbook of MRI Pulse Sequences
(Academic Press, 2004).
119. Blaimer, M. et al. SMASH, SENSE, PILS, GRAPPA: how to choose the
optimal method. Top. Magn. Reson. Imaging 15, 223–236 (2004).
120. Yacoub, E. et al. Spin-echo fMRI in humans using high spatial resolutions
and high magnetic fields. Magn. Reson. Med. 49, 655–664 (2003).
121. Weiskopf, N., Hutton, C., Josephs, O. & Deichmann, R. Optimal EPI
parameters for reduction of susceptibility-induced BOLD sensitivity losses:
a whole-brain analysis at 3 T and 1.5 T. Neuroimage 33, 493–504
(2006).
122. Glover, G.H. & Law, C.S. Spiral-in/out BOLD fMRI for increased SNR and
reduced susceptibility artifacts. Magn. Reson. Med. 46, 515–522 (2001).
123. Heberlein, K.A. & Hu, X. Simultaneous acquisition of gradient-echo and
asymmetric spin-echo for single-shot z-shim: Z-SAGA. Magn. Reson. Med.
51, 212–216 (2004).
124. Gonzalez-Castillo, J., Roopchansingh, V., Bandettini, P.A. & Bodurka, J.
Physiological noise effects on the flip angle selection in BOLD fMRI.
Neuroimage 54, 2764–2778 (2011).
125. Grootoonk, S. et al. Characterization and correction of interpolation
effects in the realignment of fMRI time series. Neuroimage 11, 49–57
(2000).
© 2013 Nature America, Inc. All rights reserved.
Focus on mapping the brain
126. Noll, D.C., Cohen, J.D., Meyer, C.H. & Schneider, W. Spiral K-space MR
imaging of cortical activation. J. Magn. Reson. Imaging 5, 49–56 (1995).
127. Rabrait, C. et al. High temporal resolution functional MRI using parallel
echo volumar imaging. J. Magn. Reson. Imaging 27, 744–753 (2008).
128. Moeller, S. et al. Multiband multislice GE-EPI at 7 tesla, with 16-fold
acceleration using partial parallel imaging with application to high
spatial and temporal whole-brain fMRI. Magn. Reson. Med. 63,
1144–1153 (2010).
129. Setsompop, K. et al. Improving diffusion MRI using simultaneous
multi-slice echo planar imaging. Neuroimage 63, 569–580 (2012).
130. Feinberg, D.A. et al. Multiplexed echo planar imaging for sub-second
whole brain FMRI and fast diffusion imaging. PLoS ONE 5, e15710
(2010).
131. Bright, M.G. & Murphy, K. Removing motion and physiological artifacts
from intrinsic BOLD fluctuations using short echo data. Neuroimage 64,
526–537 (2013).
132. Heine, L. et al. Resting state networks and consciousness: alterations of
multiple resting state network connectivity in physiological,
pharmacological, and pathological consciousness States. Front. Psychol.
3, 295 (2012).
133. Boly, M. et al. Brain connectivity in disorders of consciousness.
Brain Connect. 2, 1–10 (2012).
134. Horovitz, S.G. et al. Decoupling of the brain’s default mode network
during deep sleep. Proc. Natl. Acad. Sci. USA 106, 11376–11381 (2009).
135. Spoormaker, V.I. et al. Development of a large-scale functional brain
network during human non-rapid eye movement sleep. J. Neurosci. 30,
11379–11387 (2010).
136. Friston, K. Dynamic causal modeling and Granger causality Comments on:
the identification of interacting networks in the brain using fMRI: model
selection, causality and deconvolution. Neuroimage 58, 303–305 (2011).
137. David, O. et al. Identifying neural drivers with functional MRI: an
electrophysiological validation. PLoS Biol. 6, 2683–2697 (2008).
138. Scannell, J.W., Burns, G.A., Hilgetag, C.C., O’Neil, M.A. & Young, M.P.
The connectional organization of the cortico-thalamic system of the cat.
Cereb. Cortex 9, 277–299 (1999).
review
139. Mclntosh, A.R. & Gonzalez-Lima, F. Structural equation modeling and its
application to network analysis in functional brain imaging. Hum. Brain
Mapp. 2, 2–22 (1994).
140. Lohmann, G., Erfurth, K., Müller, K. & Turner, R. Critical comments on
dynamic causal modelling. Neuroimage 59, 2322–2329 (2012).
141. Zhuang, J., LaConte, S., Peltier, S., Zhang, K. & Hu, X. Connectivity
exploration with structural equation modeling: an fMRI study of bimanual
motor coordination. Neuroimage 25, 462–470 (2005).
142. Goebel, R. Investigating directed cortical interactions in time-resolved
fMRI data using vector autoregressive modeling and Granger causality
mapping. Magn. Reson. Imaging 21, 1251–1261 (2003).
143. Granger, C.W. Investigating causal relations by econometric models and
cross-spectral methods. Econometrica 37, 424–438 (1969).
144. Sridharan, D., Levitin, D.J. & Menon, V. A critical role for the right
fronto-insular cortex in switching between central-executive and
default-mode networks. Proc. Natl. Acad. Sci. USA 105, 12569–12574
(2008).
145. Deshpande, G. & Hu, X. Investigating effective brain connectivity from
fMRI data: past findings and current issues with reference to Granger
causality analysis. Brain Connect. 2, 235–245 (2012).
146. Johnston, J.M. et al. Loss of resting interhemispheric functional
connectivity after complete section of the corpus callosum. J. Neurosci.
28, 6453–6458 (2008).
147. Matsumoto, R. et al. Functional connectivity in the human language
system: a cortico-cortical evoked potential study. Brain 127, 2316–2330
(2004).
148. Bohning, D.E. et al. Echoplanar BOLD fMRI of brain activation induced by
concurrent transcranial magnetic stimulation. Invest. Radiol. 33, 336–340
(1998).
149. Ruff, C.C. et al. Distinct causal influences of parietal versus frontal areas
on human visual cortex: evidence from concurrent TMS-fMRI. Cereb. Cortex
18, 817–827 (2008).
150. Datta, A. et al. Gyri-precise head model of transcranial direct current
stimulation: improved spatial focality using a ring electrode versus
conventional rectangular pad. Brain Stimulat. 2, 201–207 (2009).
nature methods | VOL.10 NO.6 | JUNE 2013 | 539
RESOURCE
Improved tools for the Brainbow toolbox
npg
© 2013 Nature America, Inc. All rights reserved.
Dawen Cai1,2, Kimberly B Cohen1,2, Tuanlian Luo1,2, Jeff W Lichtman1,2 & Joshua R Sanes1,2
In the transgenic multicolor labeling strategy called ‘Brainbow’,
Cre-loxP recombination is used to create a stochastic choice
of expression among fluorescent proteins, resulting in the
indelible marking of mouse neurons with multiple distinct
colors. This method has been adapted to non-neuronal cells in
mice and to neurons in fish and flies, but its full potential has
yet to be realized in the mouse brain. Here we present several
lines of mice that overcome limitations of the initial lines,
and we report an adaptation of the method for use in adenoassociated viral vectors. We also provide technical advice
about how best to image Brainbow-expressing tissue.
The discovery that recombinant jellyfish GFP fluoresces when
expressed in heterologous cells1 has led to a vast array of powerful
methods for marking and manipulating cells, subcellular compartments and molecules. The discovery or design of numerous
spectral variants2–13 (fluorescent proteins collectively called XFPs;
ref. 14) expanded the scope of the ‘GFP revolution’ by enabling
discrimination of nearby cells or processes labeled with contrasting colors. At least for the nervous system, however, two or three
colors are far too few because each axon or dendrite approaches
hundreds or thousands of other processes in the crowded neuropil
of the brain.
Several years ago, we developed a transgenic strategy called
Brainbow15 that addresses this problem by marking neurons with
many different colors. In this method, three or four XFPs are
incorporated into a transgene, and the Cre-loxP recombination
system16 is used to make a stochastic ‘choice’ of a single XFP to be
expressed from the cassette. Because multiple cassettes are integrated at a single genomic site, and the choice within each cassette
is made independently, combinatorial expression can endow individual neurons with 1 of ~100 colors, providing nearby neurons
with distinct spectral identities.
If Cre recombinase is expressed transiently, descendants of
the initially marked cell inherit the color of their progenitor.
Accordingly, the Brainbow method has been adapted for use in
lineage analysis in non-neural tissues of mice17–21. In addition, it
has been adapted for analyses of neuronal connectivity, cell migration and lineage in fish22,23 and flies24,25. In contrast, the method
has been little used in the mouse nervous system26. We believe
that the main reasons for this are limitations of the initial set of
Brainbow transgenic mice. These include suboptimal fluorescence
intensity, failure to fill all axonal and dendritic processes, and
disproportionate expression of the ‘default’ (that is, nonrecombined) XFP in the transgene. We have now addressed several of
these limitations, and we present here a new set of Brainbow reagents. In addition, we provide guidelines for imaging Brainbowexpressing tissue.
RESULTS
Design of Brainbow 3.0 transgenes
As a first step in improving Brainbow methods, we sought XFPs
with minimal tendency to aggregate in vivo, high photostability
and maximal stability with respect to paraformaldehyde fixation. Because some XFPs that ranked highly in cultured cells
performed poorly in vivo, we generated transgenic lines from 15
XFPs (Supplementary Table 1 and refs. 2–13). Of the XFPs tested
in this way, seven were judged suitable: mTFP1, EGFP, EYFP,
mOrange2, TagRFPt, tdTomato and mKate2.
From these XFPs, we chose three according to the criteria
of minimal spectral overlap and minimal sequence homology.
These XFPs were mOrange2 from a coral (excitation peak
(Ex) = 549 nm, emission peak (Em) = 565 nm), EGFP from a
jellyfish (Ex = 488 nm, Em = 507 nm) and mKate2 from a sea
anemone (Ex = 588 nm, Em = 635 nm)2,9,11. Our reason for minimizing sequence homology was to ensure that the XFPs would be
antigenically distinct, in contrast to spectrally distinct but antigenically indistinguishable jellyfish variants (EBFP, ECFP, EGFP
and EYFP). Exploiting this property, we generated antibodies to
the XFPs in different host species (rabbit anti-mCherry, antimOrange2 and anti-tdTomato; chicken anti-EGFP, anti-EYFP and
anti-ECFP; and guinea pig anti-mKate2, anti-TagBFP and antiTagRFP; Supplementary Table 1). Tests in transfected cultured
cells confirmed a lack of cross-reactivity (data not shown).
Next we addressed the need to fill all parts of the cell evenly.
Unmodified XFPs labeled somata so strongly that nearby processes were difficult to resolve, whereas palmitoylated derivatives,
which targeted the XFPs to the plasma membrane, were selectively
transported to axons and labeled dendrites poorly15. We therefore generated farnesylated derivatives27, which were trafficked
to membranes of all neurites (see below).
On the basis of these results, we generated ‘Brainbow 3.0’ transgenic lines incorporating farnesylated derivatives of mOrange2,
EGFP and mKate2. We retained the Brainbow 1 format15, in which
1Center for Brain Science, Harvard University, Cambridge, Massachusetts, USA. 2Department of Molecular and Cellular Biology, Harvard University, Cambridge,
Massachusetts, USA. Correspondence should be addressed to J.R.S. ([email protected]).
Received 3 October 2012; accepted 30 March 2013; published online 5 May 2013; doi:10.1038/nmeth.2450
540 | VOL.10 NO.6 | JUNE 2013 | nature methods
npg
© 2013 Nature America, Inc. All rights reserved.
RESOURCE
Figure 1 | Brainbow 3 transgenic mice.
(a) Brainbow constructs and transgenic mice.
Brainbow 1.0 is described in ref. 15. Brainbow
3.0 incorporates farnesylated (‘f’), antigenically
distinct XFPs: mOrange2f (mO2f), EGFPf and
mKate2f (mK2f). Brainbow 3.1 incorporates a
nuclear-targeted nonfluorescent XFP (ØNFPnls)
in the first (default) position. Brainbow 3.2
incorporates a woodchuck hepatitis virus posttranscriptional regulatory element (W) into
Brainbow 3.1. P, loxP site; 2, lox2272; N, loxN;
pA, polyadenylation sequence. (b,c) Low- (b)
and high-power (c) views of muscles from a
Brainbow 3.0 (line D) Islet-Cre mouse, showing
terminal axons and neuromuscular junctions in
extraocular muscle. (d) Rotated image along
dashed bar in c showing five motor axons
labeled in distinct colors. The open circles show
that farnesylated XFPs mark plasma membranes
more than cytoplasm. (e) Cerebellum from a
Brainbow 3.1 (line 3) L7-Cre mouse. The ten
Purkinje cells in this field are labeled by at
least seven distinct colors (antibody amplified
and numbered i–vii). Because Cre is selectively
expressed by Purkinje cells in the L7-Cre line,
no other cell types are labeled. (f) Cerebellum
from a Brainbow 3.2 (line 7) CAGGS-CreER
mouse showing granule native fluorescence in
red, pink, yellow, green, cyan, blue and brown.
P, parallel fibers in molecular layer. Purkinje
cell bodies, which are unlabeled, are outlined.
Scale bars: 50 µm (b), 20 µm (c,e), 5 µm (d),
10 µm (f).
a
Brainbow 1.0
Thy1
P 2
RFP
pA P
YFP
pA 2
CFP
pA
mO2f
pA 2
EGFPf
pA N
mK2f
pA
Brainbow 3.0
Thy1
2
N
Brainbow 3.1
Thy1
P 2 N
∅NFPnls
pA P
mO2f
pA 2
∅NFPnls
pA P
mO2f
W pA 2
EGFPf
pA N
pA
mK2f
Brainbow 3.2
Thy1
P 2 N
b
c
e
EGFPf
W pA N
W pA
mK2f
d
f
vii
v
P
iv
iii
ii
i
incompatible wild-type and mutant loxP
sites are concatenated so that Cre recombinase yields a stochastic choice among XFPs
(Fig. 1a). We also retained two other features of the Brainbow 1
strategy. First, we used neuron-specific regulatory elements from
the Thy1 gene14 because it promotes high levels of transgene
expression in many, although not all, neuronal types; other promoter-enhancer sequences that we tested support considerably
lower levels of expression21. Second, we generated transgenic lines
by injection into oocytes because this method leads to integration
of multiple copies of the cassette and, thus, a broad spectrum of
outcomes15,28; by contrast, knock-in lines generated by homologous recombination contain one or two copies of the cassette
(as heterozygotes or homozygotes, respectively) and, consequently,
a smaller number of possible color combinations17–20.
Design of Brainbow 3.1 and 3.2 transgenes
In Brainbow 1, 2 and 3.0 (Fig. 1a) and ref. 15, one XFP is expressed
‘by default’ in Cre-negative cells. The presence of a default XFP has
both drawbacks and advantages. In cases of limited Cre expression, this XFP is expressed in a majority of cells, reducing spectral
diversity among recombined neurons. On the other hand, incorporation of a default XFP allows one to screen numerous lines in the
absence of a Cre reporter to assess the number and types of cells in
which XFPs could be expressed following recombination.
To eliminate the default XFP while retaining the ability to assess
expression in the absence of Cre, we adopted the following strategy. First, we incorporated three rather than two pairs of incompatible loxP sites, which allowed insertion of a ‘stop’ cassette29
vi
i
ii
iii
iv
v
vi
vii
into a first position. With this modification, the three XFPs were
expressed only in Cre-positive neurons, giving more spectral
diversity. Second, we inserted a fourth XFP, Phi-YFP (from the
hydrozoon Phialidium)4, which is antigenically distinct from the
other three, into the stop cassette. We mutated Phi-YFP to eliminate its endogenous fluorescence (PhiYFP Y65A), fused it to a
nuclear localization sequence and generated antibodies against
it in rat. In Brainbow 3.1 mice generated from this cassette
(Fig. 1a), one can screen sections with rat anti–Phi-YFP before
Cre recombination (Supplementary Fig. 1).
Finally, we inserted a sequence that stabilizes mRNAs, called
a woodchuck hepatitis virus post-transcriptional regulatory element (WPRE), into the 3′ untranslated sequences following each
XFP. The WPRE has been used in many cases to increase protein
levels produced by viral vectors and transgenes30,31. We call lines
incorporating the WPRE ‘Brainbow 3.2’ (Fig. 1a).
Characterization of Brainbow 3 mice
We generated 31 lines of transgenic mice from the Brainbow 3.0,
3.1 and 3.2 cassettes. Offspring were crossed with several Cre transgenic lines32–37, leading to multicolor spectral labeling of neuronal
populations in numerous regions including cerebral cortex, brainstem, cerebellum, spinal cord and retina (Fig. 1b–f, Supplementary
Table 2, Supplementary Fig. 2 and Supplementary Videos 1
and 2). The intensity of expression varied markedly among lines,
making quantitative comparison of dubious value, but the most
nature methods | VOL.10 NO.6 | JUNE 2013 | 541
RESOURCE
npg
a
c
d
e
f
h
g
i
strongly expressing lines were those that
incorporated the WPRE (Brainbow 3.2).
Analysis of these lines confirmed their
advantages over Brainbow 1 and 2 lines15.
First, the use of farnesylated XFPs led the
XFPs to concentrate at the plasma membrane (Fig. 1d). As a consequence, somata
were less intensely labeled in Brainbow 3
than in Brainbow 1 mice, so processes
could be visualized without saturation of
the somata (Fig. 2a,b and Supplementary
Fig. 3). Use of farnesylated XFPs also
improved labeling of fine processes and
dendritic spines (Fig. 2a,b). Second, the ability to immunostain
all three XFPs led to enhancement of the intrinsic fluorescence
without loss of color diversity (Fig. 2c,d and Supplementary
Fig. 4). Third, XFP fluorescence was visible only in Crepositive cells, thereby correcting the color imbalance caused
by default XFP expression in cells that were Cre negative or
cells that expressed Cre at low levels in Brainbow 1 and 2 lines
(Fig. 2e–i and Supplementary Fig. 5). Thus, Brainbow 3 lines
are likely to be more useful than Brainbow 1 and 2 lines for
multicolor labeling.
Brainbow with self-excising Cre recombinase
In Brainbow 1–3 lines, the cassette encodes XFPs separated by
loxP sites; Cre recombinase is supplied from a separate transgene. For analysis of connectivity in mouse mutants, breeding
mice with two transgenes (Brainbow and Cre) into an already
complex background is cumbersome. We therefore attempted to
combine constructs encoding XFPs and Cre in a single cassette.
In this transgene, called ‘Autobow’, we substituted a self-excising
Cre recombinase38 for the stop sequences in Brainbow 3.1
(Fig. 3a). The Thy1 regulatory elements lead to expression of Cre
in differentiated neurons; Cre then simultaneously activates an
XFP and excises itself.
We generated three founder mice using this construct.
Two were analyzed as adults, and both of these exhibited
542 | VOL.10 NO.6 | JUNE 2013 | nature methods
b
% of total neurons
© 2013 Nature America, Inc. All rights reserved.
Figure 2 | Improved visualization of neurons
in Brainbow 3 mice. (a,b) Cerebellum from
Brainbow 1.0 (a) and Brainbow 3.1 (line 3)
L7-Cre (b) mice. Insets show high-magnification
views of boxed regions. The farnesylated XFPs
clearly label the fine processes and dendritic
spines (b, inset arrows), which are missing in
the cytoplasmic labeling (a, inset). (c,d) Retina
from a Brainbow 3.0 (line D) Islet-Cre mouse
expressing EGFP (blue), mOrange2 (green)
and mKate2 (red). Intrinsic XFP fluorescence
(c) and a nearby section immunostained
with chicken anti-GFP, rabbit anti-mOrange2
and guinea pig anti-mKate2 (d) are shown.
(e–h) Immunostained cerebellum section from
Brainbow 3.2 (line 7) parvalbumin-Cre mouse.
Separate channels of region boxed in e are
shown in f–h. (i) Fraction of all labeled neurons
that express EGFP, mOrange2 (mO2) or mKate2
(mK2) (n = 2,037 neurons from 15 regions of
two brains). Scale bars: 40 µm (a,b), 20 µm
(c,d), 25 µm (e–g), 10 µm (insets).
50
n = 2,037
40
30
20
10
0
EGFP+ mO2+ mK2+
combinatorial expression of XFPs in multiple neurons (Fig. 3b
and Supplementary Fig. 2e). We were concerned that precocious
Cre activation in the germ line might lead to loss of the cassette.
We therefore established a line from the third founder and examined mice in the second and sixth generations. Color range was
limited in this line, perhaps because only a few copies had been
integrated into the genome, but the variety of colors and level of
expression were similar in both generations (Fig. 3c,d). Thus,
Autobow transgenes can be stably maintained.
Brainbow using Flp recombinase and FRT sites
A second recombination system, orthogonal to Cre-loxP, could
be used to independently control expression of distinct XFPs in,
for example, excitatory and inhibitory neurons. In this way, the
color of a neuron could denote cell identity, a feature lacking in
currently available Brainbow lines15,17–20. We therefore tested a
second recombination system, in which Flp recombinase acts on
Flp recombinase target (FRT) sites. The Flp-FRT system has been
used in conjunction with Cre-loxP in mice16 and in Brainbow-like
transgenes in Drosophila24,25.
We tested previously described mutant FRT sites39,40 to find
incompatible sets (Supplementary Fig. 6) and used these sets to
construct ‘Flpbow’ lines (Fig. 4a). In one cassette, we fused the
XFPs to an epitope tag41, which allowed for discrimination of cells
labeled by Cre- and Flp-driven cassettes (Supplementary Fig. 7).
RESOURCE
a
2
1
3
pA P
Cerulean
pA 2
PhiYFP
pA N
mKate2
pA
1
P
Cerulean
pA 2
PhiYFP
pA N
mKate2
pA
2
P 2
3
P 2 N
Thy1 P 2 N
Cre
PhiYFP
mKate2
pA N
mKate2
pA
pA
d
npg
© 2013 Nature America, Inc. All rights reserved.
Cerulean
PhiYFP
mKate2
b
c
Cre
When Flpbow mice were mated to Flp-expressing mice42,43, we
observed multicolor labeling (Fig. 4b,c). Although the few lines
tested to date exhibit narrow expression patterns, these results
demonstrate that Flp- and Cre-based Brainbow systems can be
used in combination.
Brainbow adeno-associated viral vectors
In parallel to developing Brainbow transgenic lines, we generated adeno-associated viral (AAV) vectors to provide spatial
and temporal control over expression and to make the method
applicable to other species. Because the Brainbow 3.1 cassette
described above is >6 kilobase pairs (kb), but the capacity of AAV
vectors is <5 kb, we re-engineered the cassette. On the basis of
results from initial tests (Supplementary Fig. 8), we devised a
scheme in which loxP sites with left or right element mutations44
were used for unidirectional Cre-dependent inversion (Fig. 5a,b).
Farnesylated XFPs were positioned in reverse orientation to prevent
Figure 3 | Autobow. (a) Schematic of the
Autobow construct and the steps that lead
to expression of the three XFP combinations
(outcomes 1–3). P, loxP site; 2, lox2272;
N, loxN; pA, polyadenylation sequence.
(b) Labeling of hippocampal neurons by Autobow
in founder 3. The 20 large neurons (diameter
<5 µm) in this section are labeled in 20 distinct
colors. Antibody-amplified Cerulean, PhiYFP and
mKate2 are in blue, green and red, respectively.
(c,d) Cortical neurons of an Autobow mouse
line in the second (c) and sixth (d) generations.
Antibody-amplified Cerulean, PhiYFP and mKate2
are as in b. Scale bars, 50 µm.
Cre-independent expression, and WPRE
elements were added to increase expression. In this design, recombination can
lead to three outcomes from two XFPs:
XFP1, XFP2 or neither. We generated
two AAVs with two XFPs each, such that
co-infection would lead to a minimum of
eight hues (3 × 3 – 1; Fig. 5c). Because AAV
can infect cells at high multiplicity, the
number of possible colors is 8. An additional feature is that excision of the nonexpressed XFP in a second step (Fig. 5a)
enhances and equalizes expression of the
remaining XFP (Fig. 5d,e).
We infected cortex, cerebellum and retina
of Cre transgenic mice with these vectors.
When examined 3–5 weeks later, neurons
were labeled in multiple colors (Fig. 5f–j
and Supplementary Video 3). Near injection sites, high levels of infectivity led to
coexpression of all XFPs in single cells so
that neurons appeared gray or white. The
variety of colors increased with distance
from these sites and then decreased again
in sparsely injected regions, presumably
because each labeled neuron received only
one virion (Supplementary Fig. 9).
Methods to optimize Brainbow imaging
Obtaining high-quality images from of tissues expressing
Brainbow transgenes is challenging. Because colors are derived
by mixing images of multiple fluorophores over a wide range of
concentrations, factors that differentially affect the labels degrade
the final image. In addition, it is usually necessary to image a
tissue volume rather than a single section, so methods for taking
image stacks must be optimized. Here we summarize guidelines
for imaging Brainbow tissue.
Sample preparation. To minimize background, section thickness
should be less than the working distance of the objective, generally <100 µm for high–numerical aperture (high NA) lenses. It is
also important to match the refractive indices of the immersion
medium and the sample because chromatic aberrations caused by
mismatches between these values lead to spatial offsets between
color channels (Supplementary Fig. 10). Commercial antifade
nature methods | VOL.10 NO.6 | JUNE 2013 | 543
RESOURCE
npg
© 2013 Nature America, Inc. All rights reserved.
Figure 4 | Flpbow. (a) Schematics of Flpbow
transgenes. In Flpbow 1, the loxP sites of
Brainbow 1.0 were replaced by the incompatible
FRT sites Frt5T2 and Frt545. In Flpbow 3,
the loxP sites of Brainbow 3.2 were replaced
by incompatible FRT sites F3, Frt5T2 and
Frt545. XFPs were fused to SUMO-Star (‘S-’).
f indicates farnesylation; pA, polyadenylation
sequence; W, woodchuck hepatitis virus
post-transcriptional regulatory element.
(b) Neurons in caudate putamen of a Flpbow 1
Wnt-Flp double transgenic. Neurons are labeled
by at least nine colors (red, orange, yellow,
yellow-green, green, cyan, blue, purple and
pink). Native fluorescence of ECFP, tdTomato
and antibody-amplified PhiYFP are in blue,
red and green, respectively. (c) Mossy fibers
in a Flpbow3 Wnt-Flp double transgenic.
Fluorescence of antibody-amplified EGFP,
mOrange2 and mKate2 are in blue, red and
green, respectively. Scale bars: 50 µm (b),
20 µm (c).
a
Flpbow 1
Thy1 5T2 545
tdTomato
PhiYFP
pA 545
ECFP
pA
Flpbow 3
Thy1
F3 5T2 545 ØNFPnls
b
mountants such as Vectashield (Vector Labs) or ProLong Gold
(Invitrogen) that have refractive indices of ~1.47 are optimal for
objectives that use glycerin as the immersion medium. Polyvinyl
alcohol mountants (such as Mowiol 4-88; Sigma-Aldrich)
provide a better match for oil-immersion (refractive index of
~1.52) objectives.
Confocal laser scanning microscopy. Epifluorescence microscopy can be used for imaging thin sections (<10 µm) or monolayer cultures, but confocal microscopes are preferable for
thick specimens because they decrease contamination by light
from outside the plane of focus. Newly developed two-photon
multi-XFP imaging techniques are also useful45–47. Apochromatic
or fluorite microscope objectives that are corrected for three
or more colors are strongly recommended. Most lens manufacturers specify preferred oils and coverslips. Using the wrong
oil or coverslip degrades the sharpness of focus and increases
chromatic aberration.
Fluorophores with overlap in the excitation or emission spectra should be imaged sequentially rather than simultaneously
to minimize fluorescence cross-talk and thereby optimize color
separation. Laser power should be set as low as possible for several reasons. First, all planes are bleached as each image plane
is scanned, so generation of stacks leads to gradual bleaching
and decreased signal through the stack. Second, because each
fluorophore bleaches at a different rate, colors may shift during
imaging. Third, linear signaling requires that fluorophores emit
photons at submaximal rates; at higher excitation intensities, only
the out-of-focus signal is increased48. Fourth, if one fluorophore
species is saturated but another is not, small changes in laser
power will affect the fluorophore intensities differently, leading
to color change. With high-NA objectives, a laser power of just
a few milliwatts is saturating; this is generally a small percentage
of the total power the laser can provide.
With laser power adjusted to a low level, the photomultipliertube voltages and digital gains must be set to relatively high values. In some confocal microscopes it is possible to compensate for
signal loss from deep layers through automatic adjustment of laser
544 | VOL.10 NO.6 | JUNE 2013 | nature methods
pA 5T2
pA F3
S-mO2f
W pA 5T2 S-EGFPf
W pA 545
S-mK2f
W pA
c
power, photomultiplier-tube voltage or digital gain as a function
of depth. Imaging parameters can be adjusted to obtain images
with similar signal ranges throughout the stack.
Image processing. Brainbow images must be postprocessed to
maximize color information, but care is needed to avoid introducing artifacts. Often one begins by reducing noise. Because
confocal laser scanning of multicolor stacks is generally done at
speeds of ~1 µs per pixel or less to save time, the small number
of photons collected for each pixel gives rise to sufficient shot
noise to cause perceptible local color differences. This problem
can be minimized by slower scanning or averaging of multiple
scans, but when this is infeasible, simple filtering and deconvolution methods are helpful (Fig. 6a–c). For example, median or
Gaussian filters with radii of 0.5–2 pixels reduce color noise, but
at the expense of resolution. Deconvolution algorithms (Online
Methods) are more challenging to use than simple filters but can
remove color noise without compromising spatial resolution
(Supplementary Fig. 11).
Subsequent processing steps can expand the detectable color
range and correct for color shifts (Fig. 6d,e). To obtain easily
perceived color differences, pixel intensity values for each channel
in each image are normalized to the same minimum and maximum intensity values for that color in the whole image stack.
This linearly stretches all channels and images to the full dynamic
range. Color shifts also arise because illumination strength is generally uneven across the imaging field and differs among lasers.
This effect can be attenuated by intensity or shading correction for
each channel46,49. The resulting composite RGB images provide
maximum color separation for viewing by eye (Fig. 6f).
DISCUSSION
The goal of the work reported here was to design, generate and
characterize improved reagents for multicolor Brainbow imaging of neurons in mice. First, we generated new transgenic lines
that overcome some limitations of the Brainbow 1 and 2 lines
that are currently available15. The improvements were the substitution of XFPs (especially red and orange fluorescent proteins)
AAV no. 2
5′ITR
EF1
W
pA
3′ITR
W
pA
3′ITR
1
1
2
mChef
mTFPf
mTFPf
mChef
3
4
mTFPf
mTFPf
npg
EF1
mChef
5′ITR
mTFPf
AAV no. 1
TagBFPf
a
Figure 5 | Brainbow AAV. (a) AAV Brainbow
constructs and recombination scheme.
Farnesylated TagBFP and EYFP or mCherry and
mTFP were placed in reverse orientation between
mutant loxP sites. EF1α, regulatory elements
from elongation factor 1α gene;
W, woodchuck hepatitis virus post-transcriptional
regulatory element; pA, polyadenylation
sequence; mChef, mCherryf; triangles, loxP
mutants and their recombination products
(dark and light green sectors indicate wild-type
and mutant portions of loxP sites, respectively);
1–6, recombination events. (b) Outcomes of
recombination events numbered in a. Unfilled
arrows, direction of intervening cDNA. An X
signifies that a fully mutant (light green) loxP
site cannot serve as substrate for Cre. (c) Eight
color outcomes resulting from pairs of Brainbow
AAVs following recombination as shown in a.
This is a minimum value because it does not
account for differences in relative intensity
of the four XFPs. (d,e) Test of color balance.
The mTFPf-mChef AAV vector was injected at
low titer into the cortex of a Thy1-Cre mouse,
and neurons of each color were counted. An
image from the cortex (d) and the fraction of
neurons expressing mTFPf (58%) and mChef
(42%; n = 1,523 neurons in four sections of
three mice; s.d. in red) are shown. (f–h) Cortex
of parvalbumin (PV)-Cre mice injected with the
two Brainbow AAVs. High-magnification views of
adjacent planes from the yellow (g) and white
(h) boxed regions in f are shown. (i) Retina from
AAV-injected mouse expressing Cre in retinal
ganglion cells. Arrows, dendrites; arrowheads,
axons. (j) Cerebellum from AAV-injected PV-Cre
mouse. Fan-shaped dendrites of Purkinje cells
(white arrows) and interneurons (yellow arrows)
are indicated. In f–j, antibody-amplified mTFP
and EYFP are in green, TagBFP is in blue and
mCherry is in red. Scale bars: 100 µm (d),
20 µm (f), 10 µm (g,h), 50 µm (i,j).
EYFPf
mChef
5
b
6
c
AAV no. 1 AAV no. 2
Output
6
f
1
2
3
4
5
6
d
e 70
% of total neurons
© 2013 Nature America, Inc. All rights reserved.
RESOURCE
g
60
h
50
40
30
20
10
0
+
mTFPf
mChef+
j
i
that are more photostable and less prone
to aggregation than those used initially;
use of XFPs with minimal sequence homology so they could be
immunostained separately; farnesylation of the XFPs for even
staining of somata and the finest processes; insertion of a stop
cassette to increase color variety by eliminating broad expression
a
b
c
f
e
d
of a default XFP; inclusion of a nonfluorescent marker in the
default position to facilitate screening of multiple lines; and
insertion of a WPRE to boost expression (Figs. 1 and 2). These
lines, which incorporate regulatory elements from the Thy1 gene,
enable marking of many but not all neuronal types. To date,
elements tested other than those from the Thy1 gene do not support the high expression levels needed to image Brainbow 1 and
2 material. The ability to immunostain provided by Brainbow 3
cassettes may allow weaker promoters to be used.
Second, we designed two additional transgenes and performed
initial tests to demonstrate that they can be used effectively in vivo.
Figure 6 | Processing a Brainbow image. (a) Original image from parvalbuminCre mouse cortex injected with mixed Brainbow AAVs. (b) Region boxed in a.
(c) Deconvolution, to decrease noise without sacrificing spatial resolution.
(d) Intensity normalization, to expand perceptible color range. (e) Colorshift correction (Supplementary Fig. 10). (f) Fully processed image.
Yellow arrows indicate the sequence of image processing. White arrowheads
indicate corresponding objects in the original and color shift–corrected
images. Scale bars, 10 µm (a,f), 3 µm (b–e).
nature methods | VOL.10 NO.6 | JUNE 2013 | 545
npg
© 2013 Nature America, Inc. All rights reserved.
RESOURCE
One construct, called Autobow, incorporates a self-excising Cre
recombinase. Autobow lacks the temporal and spatial control
afforded by the use of specific Cre lines or ligand-activated Cre
(CreER). However, because it does not require the generation of
double transgenics, it may be useful for rapid screening of neuronal morphology in mutant mice or mice submitted to various
experimental interventions (such as drug treatments). The other
novel transgene, Flpbow, replaces loxP sites with FRT sites so that
recombination can be controlled by Flp recombinase rather than
Cre recombinase. Flpbow 3 also incorporates an epitope tag so
that XFPs in Flpbow can be distinguished immunohistochemically from XFPs in Brainbow. By using Cre and Flp transgenic
lines with distinct, defined specificities, it should be possible to
map separate sets of neurons in a single animal.
Finally, we generated Brainbow AAV vectors. These, along
with recently described Brainbow herpes viral vectors50, may be
more useful than Brainbow transgenic mice in some situations.
Similar to Autobow, they avoid the need for double-transgenic
animals. Because the time of infection can be varied, these vectors
provide an alternative to CreER for temporal control. Moreover,
localized delivery of AAVs enables the tracing of connections
from known sites to multiple targets and discrimination of longdistance inputs from local connections.
The three most broadly useful Brainbow 3 lines (Brainbow 3.0
line D, Brainbow 3.1 line 3, Brainbow 3.1 line 18 and Brainbow 3.2
line 7) have been provided to Jackson Laboratories (http://www.
jax.org/; stock numbers 21225–21227) for distribution. The two
AAVs shown in Figure 5 can be obtained from the University
of Pennsylvania Vector Core (http://www.med.upenn.edu/gtp/
vectorcore/). Plasmids used to generate Brainbow 3.0, 3.1, and
3.2, Autobow and Flpbow 1.1 and 3.1 mice are available through
Addgene (http://www.addgene.org/).
Methods
Methods and any associated references are available in the online
version of the paper.
Accession codes. Addgene plasmids: 45176, 45177 (Brainbow
3.0), 45178 (Brainbow 3.1), 45179 (Brainbow 3.2), 45182, 45187
(Autobow), 45180 (Flpbow 1.1), 45181 (Flpbow 3.1).
Note: Supplementary information is available in the online version of the paper.
Acknowledgments
This work was supported by grants from the US National Institutes of Health
(5U24NS063931) and the Gatsby Charitable Foundation and by Collaborative
Innovation Award no. 43667 from the Howard Hughes Medical Institute. We
thank S. Haddad for assistance with mouse colony maintenance; X. Duan,
L. Bogart and J. Lefebvre for testing Brainbow mice and AAVs; R.W. Draft for
valuable discussions and advice; R.Y. Tsien (University of California, San Diego)
for mOrange2 and TagRFPt; and D.M. Chudakov (Institute of Bioorganic Chemistry
of the Russian Academy of Sciences) for TagBFP, PhiYFP, mKate2 and eqFP650.
AUTHOR CONTRIBUTIONS
D.C., K.B.C. and T.L. performed experiments. D.C., J.W.L. and J.R.S. designed
experiments, interpreted results and wrote the manuscript.
COMPETING FINANCIAL INTERESTS
The authors declare no competing financial interests.
Reprints and permissions information is available online at http://www.nature.
com/reprints/index.html.
546 | VOL.10 NO.6 | JUNE 2013 | nature methods
1. Tsien, R.Y. The green fluorescent protein. Annu. Rev. Biochem. 67,
509–544 (1998).
2. Heim, R. & Tsien, R.Y. Engineering green fluorescent protein for improved
brightness, longer wavelengths and fluorescence resonance energy transfer.
Curr. Biol. 6, 178–182 (1996).
3. Rizzo, M.A., Springer, G.H., Granada, B. & Piston, D.W. An improved cyan
fluorescent protein variant useful for FRET. Nat. Biotechnol. 22, 445–449
(2004).
4. Shagin, D.A. et al. GFP-like proteins as ubiquitous metazoan superfamily:
evolution of functional features and structural complexity. Mol. Biol.
Evol. 21, 841–850 (2004).
5. Ai, H.W., Henderson, J.N., Remington, S.J. & Campbell, R.E. Directed
evolution of a monomeric, bright and photostable version of Clavularia
cyan fluorescent protein: structural characterization and applications
in fluorescence imaging. Biochem. J. 400, 531–540 (2006).
6. Subach, O.M. et al. Conversion of red fluorescent protein into a bright
blue probe. Chem. Biol. 15, 1116–1124 (2008).
7. Ai, H.W., Shaner, N.C., Cheng, Z., Tsien, R.Y. & Campbell, R.E. Exploration
of new chromophore structures leads to the identification of improved
blue fluorescent proteins. Biochemistry 46, 5904–5910 (2007).
8. Karasawa, S., Araki, T., Nagai, T., Mizuno, H. & Miyawaki, A. Cyanemitting and orange-emitting fluorescent proteins as a donor/acceptor
pair for fluorescence resonance energy transfer. Biochem. J. 381, 307–312
(2004).
9. Shaner, N.C. et al. Improving the photostability of bright monomeric
orange and red fluorescent proteins. Nat. Methods 5, 545–551 (2008).
10. Shaner, N.C. et al. Improved monomeric red, orange and yellow
fluorescent proteins derived from Discosoma sp. red fluorescent protein.
Nat. Biotechnol. 22, 1567–1572 (2004).
11. Shcherbo, D. et al. Far-red fluorescent tags for protein imaging in living
tissues. Biochem. J. 418, 567–574 (2009).
12. Shcherbo, D. et al. Near-infrared fluorescent proteins. Nat. Methods 7,
827–829 (2010).
13. Shaner, N.C., Steinbach, P.A. & Tsien, R.Y. A guide to choosing fluorescent
proteins. Nat. Methods 2, 905–909 (2005).
14. Feng, G. et al. Imaging neuronal subsets in transgenic mice expressing
multiple spectral variants of GFP. Neuron 28, 41–51 (2000).
15. Livet, J. et al. Transgenic strategies for combinatorial expression of
fluorescent proteins in the nervous system. Nature 450, 56–62 (2007).
16. Branda, C.S. & Dymecki, S.M. Talking about a revolution: the impact of
site-specific recombinases on genetic analyses in mice. Dev. Cell 6, 7–28
(2004).
17. Snippert, H.J. et al. Intestinal crypt homeostasis results from neutral
competition between symmetrically dividing Lgr5 stem cells. Cell 143,
134–144 (2010).
18. Red-Horse, K., Ueno, H., Weissman, I.L. & Krasnow, M.A. Coronary
arteries form by developmental reprogramming of venous cells.
Nature 464, 549–553 (2010).
19. Rinkevich, Y., Lindau, P., Ueno, H., Longaker, M.T. & Weissman, I.L.
Germ-layer and lineage-restricted stem/progenitors regenerate the mouse
digit tip. Nature 476, 409–413 (2011).
20. Schepers, A.G. et al. Lineage tracing reveals Lgr5+ stem cell activity in
mouse intestinal adenomas. Science 337, 730–735 (2012).
21. Tabansky, I. et al. Developmental bias in cleavage-stage mouse
blastomeres. Curr. Biol. 23, 21–31 (2013).
22. Gupta, V. & Poss, K.D. Clonally dominant cardiomyocytes direct heart
morphogenesis. Nature 484, 479–484 (2012).
23. Pan, Y.A., Livet, J., Sanes, J.R., Lichtman, J.W. & Schier, A.F. Multicolor
Brainbow imaging in zebrafish. Cold Spring Harb. Protoc. 2011,
pdb.prot5546 (2011).
24. Hampel, S. et al. Drosophila Brainbow: a recombinase-based fluorescence
labeling technique to subdivide neural expression patterns. Nat.
Methods 8, 253–259 (2011).
25. Hadjieconomou, D. et al. Flybow: genetic multicolor cell labeling for
neural circuit analysis in Drosophila melanogaster. Nat. Methods 8,
260–266 (2011).
26. Lang, C., Guo, X., Kerschensteiner, M. & Bareyre, F.M. Single collateral
reconstructions reveal distinct phases of corticospinal remodeling after
spinal cord injury. PLoS ONE 7, e30461 (2012).
27. Badaloni, A. et al. Transgenic mice expressing a dual, CRE-inducible
reporter for the analysis of axon guidance and synaptogenesis. Genesis
45, 405–412 (2007).
28. Lichtman, J.W., Livet, J. & Sanes, J.R. A technicolour approach to the
connectome. Nat. Rev. Neurosci. 9, 417–422 (2008).
29. Lakso, M. et al. Targeted oncogene activation by site-specific
recombination in transgenic mice. Proc. Natl. Acad. Sci. USA 89,
6232–6236 (1992).
30. Paterna, J.C., Moccetti, T., Mura, A., Feldon, J. & Büeler, H. Influence
of promoter and WHV post-transcriptional regulatory element on AAVmediated transgene expression in the rat brain. Gene Ther. 7, 1304–1311
(2000).
31. Madisen, L. et al. A robust and high-throughput Cre reporting and
characterization system for the whole mouse brain. Nat. Neurosci. 13,
133–140 (2010).
32. Hippenmeyer, S. et al. A developmental switch in the response of
DRG neurons to ETS transcription factor signaling. PLoS Biol. 3, e159
(2005).
33. Srinivas, S. et al. Cre reporter strains produced by targeted insertion of
EYFP and ECFP into the ROSA26 locus. BMC Dev. Biol. 1, 4 (2001).
34. Guo, C., Yang, W. & Lobe, C.G. A Cre recombinase transgene with mosaic,
widespread tamoxifen-inducible action. Genesis 32, 8–18 (2002).
35. Zhang, X.M. et al. Highly restricted expression of Cre recombinase in
cerebellar Purkinje cells. Genesis 40, 45–51 (2004).
36. Rossi, J. et al. Melanocortin-4 receptors expressed by cholinergic neurons
regulate energy balance and glucose homeostasis. Cell Metab. 13, 195–204
(2011).
37. Campsall, K.D., Mazerolle, C.J., De Repentingy, Y., Kothary, R. &
Wallace, V.A. Characterization of transgene expression and Cre recombinase
activity in a panel of Thy-1 promoter-Cre transgenic mice. Dev. Dyn. 224,
135–143 (2002).
38. Bunting, M., Bernstein, K.E., Greer, J.M., Capecchi, M.R. & Thomas, K.R.
Targeting genes for self-excision in the germ line. Genes Dev. 13,
1524–1528 (1999).
39. McLeod, M., Craft, S. & Broach, J.R. Identification of the crossover site
during FLP-mediated recombination in the Saccharomyces cerevisiae
plasmid 2 µm circle. Mol. Cell Biol. 6, 3357–3367 (1986).
40. Schlake, T. & Bode, J. Use of mutated FLP recognition target (FRT) sites
for the exchange of expression cassettes at defined chromosomal loci.
Biochemistry 33, 12746–12751 (1994).
41. Peroutka, R.J., Elshourbagy, N., Piech, T. & Butt, T.R. Enhanced protein
expression in mammalian cells using engineered SUMO fusions: secreted
phospholipase A2. Protein Sci. 17, 1586–1595 (2008).
42. Farley, F.W., Soriano, P., Steffen, L.S. & Dymecki, S.M. Widespread
recombinase expression using FLPeR (flipper) mice. Genesis 28, 106–110
(2000).
43. Awatramani, R., Soriano, P., Rodriguez, C., Mai, J.J. &
Dymecki, S.M. Cryptic boundaries in roof plate and choroid plexus
identified by intersectional gene activation. Nat. Genet. 35, 70–75
(2003).
44. Araki, K., Okada, Y., Araki, M. & Yamamura, K. Comparative analysis of
right element mutant loxP sites on recombination efficiency in embryonic
stem cells. BMC Biotechnol. 10, 29 (2010).
45. Entenberg, D. et al. Setup and use of a two-laser multiphoton microscope
for multichannel intravital fluorescence imaging. Nat. Protoc. 6,
1500–1520 (2011).
46. Mahou, P. et al. Multicolor two-photon tissue imaging by wavelength
mixing. Nat. Methods 9, 815–818 (2012).
47. Wang, K. et al. Three-color femtosecond source for simultaneous
excitation of three fluorescent proteins in two-photon fluorescence
microscopy. Biomed. Opt. Express 3, 1972–1977 (2012).
48. Conchello, J.A. & Lichtman, J.W. Optical sectioning microscopy.
Nat. Methods 2, 920–931 (2005).
49. Ducros, M. et al. Efficient large core fiber-based detection for multichannel two-photon fluorescence microscopy and spectral unmixing.
J. Neurosci. Methods 198, 172–180 (2011).
50. Card, J.P. et al. A dual infection pseudorabies virus conditional reporter
approach to identify projections to collateralized neurons in complex
neural circuits. PLoS ONE 6, e21141 (2011).
npg
© 2013 Nature America, Inc. All rights reserved.
RESOURCE
nature methods | VOL.10 NO.6 | JUNE 2013 | 547
© 2013 Nature America, Inc. All rights reserved.
npg
ONLINE METHODS
Brainbow constructs. cDNA encoding the following fluorescent
proteins were used: EGFP1, EYFP2, ECFP2, mCerulean3, PhiYFP4
(Evrogen), mTFP5 (Allele Biotechnology), TagBFP6 (Evrogen),
EBFP2 (ref. 7) (Addgene), Kusabira-Orange8, TagRFPt9, mOrange2 (ref. 9), tdTomato10, mCherry10, mKate2 (ref. 11) (Evrogen)
and eqFP650 (ref. 12) (Evrogen). A HRAS farnesylation sequence51
was used to tether XFPs to the cell membrane. A nuclear localization signal (NLS, APKKKRKV) was added to the N-terminal
end of PhiYFP(Y65A). WPRE sequence was used to stabilize
mRNA and enhance nuclear mRNA export30,31. Polyadenylation
signals were from the SV40 T antigen for mouse transgenes and
from the human growth hormone gene for AAV. Brainbow constructs were assembled by standard cloning methods. A cloning
scaffold containing concatenated loxP mutant sequences and
unique restriction digestion sites was synthesized (DNA2.0, Inc.)
to facilitate cloning.
Brainbow modules were cloned into the pCMV- N1 mammalian expression vector (Clontech) for transient mammalian cell
expression. Brainbow mouse constructs were cloned into a unique
XhoI site in a genomic fragment of Thy1.2 containing neuronspecific regulatory elements14. Brainbow AAV constructs were
cloned into vectors provided by the University of Pennsylvania
Virus Core. Constructs were tested by expression in HEK293 cells
(ATCC) before generation of mice or AAV.
Mice. Transgenic mice were generated by pronuclear injection at the
Harvard Genome Modification Core. Mice were maintained on C57B6
or CD-1 backgrounds. Brainbow mice were crossed to mice that
expressed Cre or Flp recombinases, including PV-Cre32, Islet-Cre33,
CAGGS-CreER34, L7-Cre35, ChAT-Cre36, Thy1-Cre37, Wnt-Flp43 and
Actin-Flp42. Both male and female mice were used. All experiments
conformed to NIH guidelines and were carried out in accordance with
protocols approved by the Harvard University Standing Committee
on the Use of Animals in Research and Teaching.
AAV injection. Two Brainbow AAVs were mixed to equal titer
(7.5 × 1012 genome copies per ml) before injection. For retina
injection, adult mice were anesthetized with ketamine-xylazine by
intraperitoneal injection. A small hole was made in the temporal
eye by puncturing the sclera below the cornea with a 30 1/2–G
needle. With a Hamilton syringe with a 33-G blunt-ended needle,
0.5–1 µl of AAV virus was injected intravitreally. After injections,
animals were treated with Antisedan (Zoetis) and monitored for
full recovery. For cortex injection, adult mice were anesthetized
with isoflurane via continuous delivery through a nose cone and
fixed to a stereotaxic apparatus. Surgery took place under sterile
conditions with the animal lying on a heating pad. One microliter
of 1:5 saline-diluted AAV mix (1.5 × 1012 genome copies per ml)
was injected over 10 min. The head wound was sutured at the
end of the experiment. One injection of the nonsteroidal antiinflammatory agent meloxicam was given at the end of the
surgery, and mice were kept on a heating pad with accessible
moistened food pellets and/or HydroGel until they had fully
recovered. The mice were given another dose of meloxicam 1 d
later and examined 4–6 weeks after infection.
Antibodies. Expression vectors were constructed to produce His
tag fusions of XFPs in Escherichia coli. Proteins were produced
nature methods
in bacteria, purified using His Fusion Protein Purification Kits
(Thermo Scientific), concentrated to >3 mg/ml, and used as
immunogens to produce rat anti-mTFP, chicken anti-EGFP, rat
anti-PhiYFP, rabbit anti-mCherry and guinea pig anti-mKate2.
(Covance). Chicken anti-EGFP IgY was purified from chicken
egg yolks using Pierce Chicken IgY Purification Kit (Thermo
Scientific). Other sera were used without purification. Other
antibodies used were: rabbit anti-GFP (ab6556, Abcam), rabbit
anti-PhiYFP (AB604, Evrogen), chicken anti-SUMOstar (AB7002,
LifeSensors), DyLight 405–conjugated goat anti-rat (Jackson
ImmunoResearch) and Alexa fluorescent dye–conjugated goat
anti-rat 488, anti-chicken 488 and 594, anti-rabbit 514 and 546,
and anti–guinea pig 647 (Life Technologies).
Histology. Mice were anesthetized with sodium pentobarbital
before intracardiac perfusion with 2%–4% paraformaldehyde
in PBS. Brains were sectioned at 100 µm using a Leica vt1000s
vibratome. Muscle and retina were sectioned at 20 µm in a Leica
CM1850 cryostat or processed as whole mounts. For immunostaining, tissues were permeabilized by 0.5% Triton X-100 with
0.02% sodium azide in StartingBloc (Thermo Scientific) at room
temperature for 2 h and then incubated with combinations of
anti-XFPs (see above) for 24–48 h at 4 °C. After extensive washing
in PBST (0.01 M PBS with 0.1% Triton X-100), all secondary antibodies (1:500) were added for 12 h at 4 °C. Finally, sections and
tissues were mounted in Vectashield mounting medium (Vector
Labs) and stored at –20 °C until they were imaged.
Antibody combinations used in figures are as follows. In
Figures 1e–h; 2b,d–g; and 4c and Supplementary Figures 2a–d,
4a and 5, primary antibodies are chicken anti-GFP (1:2,000),
rabbit anti-mCherry (1:1,000, for mOrange2) and guinea pig
anti-mKate2 (1:500). Secondary antibodies are Alexa dye–
conjugated goat anti-chicken 488, anti-rabbit 546, and anti–guinea
pig 647. In Figure 3b–d and Supplementary Figure 2e, primary
antibodies are chicken anti-GFP (for ECFP), rabbit anti-PhiYFP
(1:1,000) and guinea pig anti-mKate2. Secondary antibodies are
Alexa dye–conjugated goat anti-chicken 488, anti-rabbit 546 and
anti-guinea pig 647. In Figure 4b, rabbit anti-PhiYFP and Alexa
dye–conjugated goat anti-rabbit 514 were used. In Figure 5d,
primary antibodies are rat anti-mTFP and rabbit anti-mCherry.
Secondary antibodies are Alexa dye–conjugated goat anti-rat 488
and anti-rabbit 546. In Figure 5f–j and Supplementary Figure 9,
primary antibodies are guinea pig anti-mKate2 (for TagBFP), rat
anti-mTFP (1:1,000), chicken anti-GFP (for EYFP) and rabbit
anti-mCherry. Secondary antibodies are Alexa dye–conjugated
goat anti-rat 488, anti-chicken 488, anti-rabbit5 46 and antiguinea pig 647. In Supplementary Figure 1, primary antibodies
are rat anti-PhiYFP (1:1,000) and rabbit anti-mCherry (1:1,000,
for mOrange2). Secondary antibodies are Alexa dye–conjugated
goat anti-rat 488 and anti-rabbit 546. In Supplementary Figure 7,
primary antibodies are rabbit anti-EGFP (1:1,000, for ECFP) and
chicken anti-SUMOstar (1:1,000). Secondary antibodies are Alexa
dye–conjugated goat anti-rabbi 514 and anti-chicken 594.
Imaging. Fixed brain and muscle samples were imaged using a
Zeiss LSM710 confocal microscope. Best separation of multiple
fluorophores was obtained by using a 405-nm photodiode laser
for TagBFP and DyLight 405, a 440-nm photodiode laser for
mTFP, a 488-nm Argon line for EGFP and Alexa 488, a 514-nm
doi:10.1038/nmeth.2450
collected at 545–600 nm in channel 2. In the antibody-amplified
samples, conjugated Alexa dyes normally produced much stronger
fluorescence signal than XFPs. The Zeiss microscope we used was
optimized for imaging the Alexa 488/546/647 combination. The
fixed dichroic mirror was DM488/561/633. Alexa 488 and Alexa
647 were excited by 488-nm and 633-nm lasers simultaneously,
and fluorescence was collected at 495–590 nm in channel 1 and
638–780 nm in channel 2, respectively. In the subsequent scan,
a 561-nm laser was used to excite mOrange2, and fluorescence
was collected at 566–626 nm in channel 2.
51. Hancock, J.F., Cadwallader, K., Paterson, H. & Marshall, C.J.A. CAAX or a
CAAL motif and a second signal are sufficient for plasma membrane
targeting of ras proteins. EMBO J. 10, 4033–4039 (1991).
npg
© 2013 Nature America, Inc. All rights reserved.
Argon line for EYFP and Alexa 514, a 561-nm photodiode for
mOrange2 and Alexa 546, a 594-nm photodiode for mCherry,
mKate2 and Alexa 594 or a 633-nm photodiode for Alexa 647.
Images were obtained with 16× (0.8 NA), and 63× (1.45 NA) oil
objectives. Confocal image stacks for all channels were acquired
sequentially, and maximally or 3D-view projected using ImageJ
(NIH). Intensity levels were uniformly adjusted in ImageJ.
Optimal imaging for Brainbow 3 tissue used a Zeiss LSM710
with fixed dichroic mirror combinations of DM455+514/594
to reduce lag time between the two sequential scans. EGFP and
mKate2 were excited by 458-nm and 594-nm lasers simultaneously, and fluorescence was collected at 465–580 nm in channel 1
and 605–780 nm in channel 2, respectively. In a subsequent scan, a
514-nm laser was used to excite mOrange2, and fluorescence was
doi:10.1038/nmeth.2450
nature methods
NEW TOOLS FOR THE BRAINBOW TOOLBOX
Dawen Cai, Kimberly B. Cohen, Tuanlian Luo, Jeff W. Lichtman and Joshua R. Sanes
Supplementary Figures and Tables
Contents:
•
•
Supplementary Figures 1 – 11
Supplementary Tables 1 – 2
Nature Methods: doi:10.1038/nmeth.2450
Supplementary Figure 1. Use of the non-fluorescent default protein to analyze expression potential of a
Brainbow3.1 mouse line.
Sagittal section of brain from Brainbow 3.1 (line 3), stained with rat anti-PhiYFP. (a) Low power image. C,
cerebellum; H, Hippocampus. (b, c) Higher magnification view of cerebellar (C) and hippocampal (H) regions
boxed in a. Bars are 1mm in a, 200μm in b, 20μm in c.
Nature Methods: doi:10.1038/nmeth.2450
Supplementary Figure 2. Neuronal labeling in Brainbow3;cre double transgenic and Autobow mice.
(a) Whole mount of retina from Islet1-Cre;Brainbow 3.1 (line 10) showing retinal ganglion cells.
(b,c) Cerebellum from Brainbow 3.1 (line 3); parvalbumin-cre mouse. Low power image in b shows Purkinje
(yellow arrows) and basket cells (cyan arrow). Axon terminals of basket cells arborize on initial segments of
Purkinje cell axons. Higher power view in c shows axons from multiple basket cells, labeled in distinct colors,
wrapping a Purkinje cell body and axonal initial segment (yellow arrow).
(d) Brainstem from PV-Cre;Brainbow 3.1 (line 4) showing axonal terminals of inhibitory neurons. In a-d, antibody
amplified EGFP, mOrange2 and mKate2 are in blue, green and red, respectively.
(e) Labeling of hippocampal neurons by Autobow (line 2). Antibody amplified Cerulean, PhiYFP and mKate2 are
in blue, green and red, respectively. Bars are 30μm in a, 20μm in b, 10μm in c and d, 100μm in e.
Nature Methods: doi:10.1038/nmeth.2450
Supplementary Figure 3. Farnesylated XFPs label neurons more evenly than cytoplasmic XFPs.
Retinal bipolar cells are labeled by (a) cytoplasmic GFP and (b) farnesylated membrane GFP. To image processes
of neurons labeled with cytoplasmic GFP, it is necessary to saturate the somata. Bars are 5µm.
Nature Methods: doi:10.1038/nmeth.2450
Supplementary Figure 4. Immunostaining enhances Brainbow signals without compromising color diversity.
Motor axons innervating extraocular muscles of a Brainbow 3.0 mouse (line D). a shows native XFP fluorescence.
b shows a nearby section immunostained with non-crossreactive chicken-anti-GFP, rabbit-anti-mOrange2 and
guinea pig-anti-mKate2. Bars are 50μm.
Nature Methods: doi:10.1038/nmeth.2450
Supplementary Figure 5. XFP expression balance in Brainbow 3.
Immunostained cerebellum section from Brainbow 3.1 (line 3);PV-cre mouse. Merged image is shown in a and
separate channels in b (EGFP), c (mOrange2) and d (mKate2). (e) Percent of Purkinje cells expressing each XFP
(208 neurons from 12 regions). Bars are 20μm.
Nature Methods: doi:10.1038/nmeth.2450
Supplementary Figure 6. Tests for incompatible mutant Frt sites.
Canonical Frt site and multiple Frt variants have been reported to allow Flp-dependent recombination between
the same variants but not between each other17, 18. To verify incompatibility, three constructs were generated
and tested:
(a) Construct testing incompatibility between Frt5T2 and Frt545 sites. When co-transfected with this construct
and FLP recombinase, HEK293 cells showed high level of Kusabira-Orange but no detectable mKate2. This
indicated that Frt5T2 and Frt545 are incompatible.
(b) Adding FrtF3 site and TagBFP to the construct in a allowed a three-way test of FrtF3, Frt5T2 and Frt545. Cotransfection of this construct with Flp recombinase showed TagBFP expression only, indicating that the FrtF3 site
is incompatible with both Frt5T2 and Frt545.
(c) Adding the canonical Frt site and TagBFP to the construct in a allowed a three-way test of Frt, Frt5T2 and
Frt545. Co-transfection of this construct with Flp recombinase showed high level of TagBFP expression as well as
detectable levels of Kusabira-Orange and mkate2. This indicates that canonical Frt site undergoes Flp dependent
excision with Frt5T2 and Frt545 sites.
Bars are 100μm.
Nature Methods: doi:10.1038/nmeth.2450
Supplementary Figure 7. SUMOstar as epitope tag in transgenic mouse.
Section from the hippocampus from a transgenic mouse that expressed SUMOstar-ECFP fusion protein from the
Thy1 promoter. The section was stained with rabbit anti-GFP and chicken anti-SUMOstar, plus appropriate
secondary antibodies.
(a) Native fluorescence of ECFP.
(b) Rabbit anti-GFP
(c) Chicken anti-SUMOstar.
Bar is 50μm.
Nature Methods: doi:10.1038/nmeth.2450
Supplementary Figure 8. Tests of Brainbow AAV constructs.
In order to accommodate the 5kB packaging size of AAV yet achieve high expression level, several promoter
elements, polyadenylation sequences and stop mechanisms were tested in constructs based on Brainbow1.
(a) This series tested the performance of a 1kB hybrid promoter CB6 (CMV enhancer, chicken β-actin promoter
and hybrid intron, Penn Vector Core) as well as a small stop cassette relying on an open reading frame shift
mechanism (“ORF shift” round corner empty box). The ORF shift encodes a Kozak sequence and translational
start (red solid arrow) to initiate an “out of frame” translation of the following mTFP. Upon Cre recombination,
this stop cassette is removed and mTFP, PhiYFP or TagRFPt is expressed. AAV made from this construct
displayed cre-independent expression of XFPs in wildtype mouse brains. Moreover, expression from the CB6
promoter was weaker than that from a EF1α promoter, tested in parallel.
(b) This series used a pair LoxP left and right element mutants (LE-RE) to construct a Cre dependent
“unidirectional spin” cassette. A Brainbow1 construct expressing Kusabira-Orange, EGFP and mCherry was
placed in inverted orientation between the LE-RE loxP sites. AAV generated from this construct exhibited no Creindependent expression in mouse brain. However, the proportions of the three XFPs expressed following
infection of Cre transgenic mice was greatly influenced by the level of Cre expression. When Cre levels were low,
Kusabira-Orange was expressed in most cells. In contrast, when Cre levels were high, very little expression of
Kusabira-Orange was detected, with most cells expressing mCherry and a minority expressing EGFP. This
imbalance led us to use the LE-RE elements exclusively (Fig. 2).
(c) This series incorporated a STOP cassette generated from a tandem Myc epitope tag and a bovine growth
hormone polyadenylation sequence, totaling 0.6kB. To accommodate the packaging size of AAV, only two XFPs
were incorporated into each construct. Infection with pairs of AAVs led to multicolor labeling (see d). As in
constructs shown in b, however, relative expression of the two XFPs in each vector varied among Cre lines.
Based on these tests, we designed new AAVs (Fig. 5a) incorporating EF1α promoter, LE-RE mutant sites,
farnesylated XFPs, a single WPRE sequence, and a 0.5kB human growth hormone polyadenylation sequence.
(d) Brain from a Thy1-cre mice infected with the two AAV shown in c. Native fluorescence of TagBFP in blue,
Kusabira-Orange and EGFP in green, mCherry in red.
Bar is 50μm.
Nature Methods: doi:10.1038/nmeth.2450
Supplementary Figure 9. Labeling color pallet changes at different distances to Brainbow AAV injection site.
(a) Sagittal section of PV-Cre mouse brain shows fluorescent labeling by Brainbow AAV infection. White arrow
indicates injection site. (b) Magnified view of blue boxed regions in a. (c, d) Magnified views of regions boxed in
b. Bars are 300μm in a, 50μm in b and 20μm in c, and d.
Nature Methods: doi:10.1038/nmeth.2450
Supplementary Figure 10. Chromatic aberration in Brainbow images.
(a,b) X-Y plane (a) and cross sections (b) show axial and lateral chromatic aberration, as well as laser
misalignment cause color offset in X-Y-Z directions. (c,d) Color-shift correction of images in a,b align the
channels and restore correct colors. Arrowheads indicate corresponding objects in the original and corrected
images. Bars are 3µm.
Nature Methods: doi:10.1038/nmeth.2450
Supplementary Figure 11. Noise reduction methods comparison.
(a) Original image. (b) Enlarged view of region boxed in a. (c) Guassian filter with radius of 1.0 pixel. (d) Median
filter with radius of 1.0 pixel. (e) Deconvolution by Huygens software. Bars are 10µm in a, 2µm in b-e.
Nature Methods: doi:10.1038/nmeth.2450
Supplementary Table 1. Fluorescent proteins screened in transgenic mice.
EBFP2
TagBFP
ECFP
Cerulean
mTFP1
EGFP
EYFP
PhiYFP *
383
399
433
433
462
488
516
525
448
456
475
475
492
507
529
537
Brightness
(EGFP=1)
0.55
1.00
0.33
0.82
1.63
1.00
1.48
1.58
Kusabira Orange
548
559
0.94
Bright; Aggregation
Fugia concinna
Rb : KuO
mOrange2
tdTomato
tagRFP-T
mCherry
mKate2
eqFP650
549
554
555
587
588
592
565
581
584
619
635
650
1.06
2.88
1.00
0.48
0.76
0.48
Bright; Aggregation
Bright
Bright
Dim; Aggregation
Bright
Dim; Far-red
Discosoma sp.
Discosoma sp.
Entacmaea quadricolor
Discosoma sp.
Entacmaea quadricolor
Entacmaea quadricolor
Rb : DsRed
Rb : DsRed
GP : mKate2
Rb : DsRed
GP : mKate2
GP : mKate2
Fluorescent Protein Ex (nm) Em (nm)
in vivo property
species
Antibody
Dim; Scattering
Bright; Scattering
Dim; High Expression
Bright; Photobleaching
Bright
Bright; High Expression
Bright; High Expression
Bright; PFA Quenching
Aequorea victoria
Entacmaea quadricolor
Aequorea victoria
Aequorea victoria
Clavularia sp.
Aequorea victoria
Aequorea victoria
Phialidium sp.
Ck : GFP
GP : mKate2
Ck : GFP
Ck : GFP
Rt : TFP
Ck : GFP
Ck : GFP
Rt : PhiYFP
Columns are as follows: (1) Proteins tested; see refs. 2-12. (2) Ex, excitation peak. (3) Em, emission peak. (4) Brightness calculated as the product of extinction
coefficient and quantum yield, with EGFP=1. (5) Global result from observation of brain tissue in transgenic mice. Main drawbacks encountered were low
fluorescence in vivo despite high intrinsic brightness, presumable because of low expression; high scattering of deep blue proteins; rapid photobleaching;
aggregation in neurons; and sensitivity to paraformaldehyde fixation. (6) Species from which XFP was derived. (7) Antibodies used to amplify signal: Ck:GFP,
chicken-anti-GFP; GP:mKate2, Guinea pig-anti-mKate2; Rt:TFP, rat-anti-mTFP1; Rt:PhiYFP, rat-anti-PhiYFP; Rb:KuO, rabbit-anti-Kusabira Orange; Rb:DsRed,
rabbit-anti-DsRed.
* The nonfluorescent mutant PhiYFP(Y65A) was used as stainable “STOP”.
Nature Methods: doi:10.1038/nmeth.2450
Supplementary Table 2. Brainbow 3 transgenic mouse lines.
Cell type labeled (Brightness*)
Expression
# line subset size
1st/2nd/3rd/4th
Spinal cord / Peripheral
Central Nervous System
A
medium
not tested
RGC(+), CTL(++), HPL(++), CBR(+)
B
large
not labeled
whole brain (+)
Brainbow-3.0
MN
(+++),
SIN
(++),
Of/Gf/Kf
D
large
RGC (++), RBP (++), CTL(+++), HPL(+++)
DRG (++), SSN (++)
F
medium
MN(++), DRG(+++)
CTL(++), HPL(++)
2
large
not tested
whole brain (+)
3
medium
MN (++)
CTL (++), HPL (+++), CBR (++), CPC (+++)
4
large
MN (+), DRG (++)
CTL(+), HPL(+++), CBR (++), HB (+)
10
small
not labeled
RGC (++), RAC (++), CTL (+), HPL (++)
11
small
MN (+), DRG (++)
RGC (++), RAC (++), CTL(++), HPL(+++)
Brainbow-3.1
16
small
not
labeled
CP (++), TH (++)
nØ/Of/Gf/Kf
17
very small
not labeled
CTL(+++), HPL(+++), CP (+++)
RGC (++), RAC (++), RBC (++), RHC (++),
18
medium
not tested
CTL(+++), CBR (+++), MB (+++), HB (+++)
21
small
not labeled
CTL (+), HPL (+)
23
medium
MN (+), DRG (+)
HPL (++), MF (+)
2
small
NM (+++), DRG (+++)
CTL (+++),HPL (+++)
RGC (+++), RAC (+++),
4
small
not labeled
CTL (++++),HPL (++++)
Brainbow-3.2
5
large
not tested
whole brain (+)
nØ/Ofw/Gfw/Kfw
6
medium
NM (+++), DRG (++)
CTL (+++),HPL (+++)
RGC(+++), RAC(+++), RBP (+++),
7
medium
SIN (++)
CTL (++++), HPL (++++), CGC (++++)
1
Large
NM (++), DRG (+)
CTL (++), HPL (++)
Autobow
2
very small
not labeled
CTL (++), HPL (++)
CreINT/Cer/P/K
3
medium
NM(+)
CTL (+), HPL (+)
1
medium
NM (+), DRG (+)
CTL (+), HPL (+)
2
large
not tested
whole brain (+)
3
large
DRG (+)
RGC (++), whole brain (+)
Flpbow-3.0
5
very small
not labeled
RGC (+)
T/P/C
RGC (++), CTL (++), HPL (++),
6
medium
NM (++), SIN (++)
CP (++), MB (++), HB (++)
7
large
NM (++)
CTL (++), HPL (++), MB (++), HB (++)
Flpbow-3.1
nØ/sOf/sGf/sKf
4
large
not tested
MF (+), HB (+)
5
6
small
large
not labeled
MN (+), RDG (+)
RGC (+), CTL (+), HPL (+), MF (+)
whole brain (+)
A total of 46 Brainbow 3 transgenic mouse lines were generated. 15 lines, not described here died, had, no detectable XFP
expression, or did not transmit the Brainbow transgene to their progeny. Lines available from Jackson Laboratory are indicated in
bold.
f, farnesylated XFP; n, nucleic localization sequence; w, WPRE; s, SUMO tag; CreINT, intron inserted Cre; O, monomeric Orange2; G,
EGFP; K, monomeric Kate2; P, PhiYFP; Ø, “dark” PhiYFP; T, tandem-dimer Tomato; Cer, Cerulean; C, ECFP; MN, motor neuron; SIN,
spinal cord interneuron; SSN, skin sensory neuron; DRG, dorsal root ganglion; RGC, retina ganglion cell; RAC, retina amacrine cell;
RBP, retina bipolar cell; RHC, retina horizontal cell; CBR, cerebellar neuron; CPC, cerebellar Purkinje cell; CGC, cerebellar granule cell;
CTL, cortical neuron; HPL, hippocampal neuron; CP, caudate putamen; TH, thalamus; MB, mid brain; HB, hindbrain; MF, Mossy fiber.
* Specific antibodies effectively enhanced the brightness of all XFPs in all lines.
§ Autobow line 1 and 2, founders were screened; line 3, founder as well as the first and the 6th generations were screened.
Nature Methods: doi:10.1038/nmeth.2450