Skip to main content
Log in

The connectome of the basal ganglia

  • Original Article
  • Published:
Brain Structure and Function Aims and scope Submit manuscript

Abstract

The basal ganglia of the laboratory rat consist of a few core regions that are specifically interconnected by efferents and afferents of the central nervous system. In nearly 800 reports of tract-tracing investigations the connectivity of the basal ganglia is documented. The readout of connectivity data and the collation of all the connections of these reports in a database allows to generate a connectome. The collation, curation and analysis of such a huge amount of connectivity data is a great challenge and has not been performed before (Bohland et al. PloS One 4:e7200, 2009) in large connectomics projects based on meta-analysis of tract-tracing studies. Here, the basal ganglia connectome of the rat has been generated and analyzed using the consistent cross-platform and generic framework neuroVIISAS. Several advances of this connectome meta-study have been made: the collation of laterality data, the network-analysis of connectivity strengths and the assignment of regions to a hierarchically organized terminology. The basal ganglia connectome offers differences in contralateral connectivity of motoric regions in contrast to other regions. A modularity analysis of the weighted and directed connectome produced a specific grouping of regions. This result indicates a correlation of structural and functional subsystems. As a new finding, significant reciprocal connections of specific network motifs in this connectome were detected. All three principal basal ganglia pathways (direct, indirect, hyperdirect) could be determined in the connectome. By identifying these pathways it was found that there exist many further equivalent pathways possessing the same length and mean connectivity weight as the principal pathways. Based on the connectome data it is unknown why an excitation pattern may prefer principal rather than other equivalent pathways. In addition to these new findings the local graph-theoretical features of regions of the connectome have been determined. By performing graph theoretical analyses it turns out that beside the caudate putamen further regions like the mesencephalic reticular formation, amygdaloid complex and ventral tegmental area are important nodes in the basal ganglia connectome. The connectome data of this meta-study of tract-tracing reports of the basal ganglia are available for further network studies, the integration into neocortical connectomes and further extensive investigations of the basal ganglia dynamics in population simulations.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Abbreviations

A:

All (all inputs and outputs)

AD:

Average degree

Ac:

Accumbens nucleus

AC:

Amygdaloid complex

AGl:

Lateral agranular prefrontal cortex

AGm:

Medial agranular prefrontal cortex

Aut:

Authoritativeness

AW:

Average weight

BC:

Betweenness centrality

BG:

Basal ganglia

C:

Circle

CC:

Cluster-coefficient

CE:

Closeness centrality

chain:

Chain pattern of a motif

CL:

Centrolateral thalamic nucleus

CM:

Central medial thalamic nucleus

CNS:

Central nervous system

CPu:

Caudate putamen

DG:

Degree

Dic:

Direct input from contralateral

Dii:

Direct input from ipsilateral

Dis:

Direct input from ipsi- and contralateral

DNN:

Direct neighbor network

Doc:

Direct output to contralateral

Doi:

Direct output to ipsilateral

Dos:

Direct output to ipsi- and contralateral

DR:

Dorsal raphe nucleus

EC:

Eigenvector centrality

Ent:

Entorhinal cortex

HIPP:

Hippocampus

Hub:

Hubness

I:

In (input to a region; used in tables)

in:

Symmetric input connection to a central node of a motif

INN:

Indirect neighbor network

L:

Laterality

LGP:

Lateral globus pallidus

LHb:

Lateral habenular nucleus

MDL:

Mediodorsal thalamic nucleus lateral part

MDM:

Mediodorsal thalamic nucleus medial part

MDS:

Metric multidimensional scaling

MGP:

Medial globus pallidus

MRF:

Mesencephalic reticular formation

O:

Out (Output of region; used in tables only)

out:

Symmetric output connection from a central node of a motif

PC:

Paracentral thalamic nucleus

PCA:

Principal component analysis

PF:

Parafascicular thalamic nucleus

Pir:

Piriform cortex

PL:

Path length

PRC:

Page rank centrality

Pub:

Number of articles

RADin:

Radiality of the input

RADout:

Radiality of the output

Rec:

Reciprocal

Rel:

Reliability

Sic:

Subtree input from contralateral

Sii:

Subtree input from ipsilateral

Sis:

Subtree input from ipsi- and contralateral

SG:

Subgraph centrality

SNC:

Substantia nigra compact part

SNR:

Substantia nigra reticular part

Soc:

Subtree output to contralateral

Soi:

Subtree output to ipsilateral

Sos:

Subtree output to ipsi- and contralateral

SP:

Length of shortest path

SPN:

Spiny neurons of the CPu

STh:

Subthalamic nucleus

VA:

Ventro anterior thalamic nucleus

VL:

Ventrolateral thalamic nucleus

VM:

Ventromedial thalamic nucleus

VTA:

Ventral tegmental area A10

References

  • Albert R, Barabasi AL (2002) Statistical mechanics of complex networks. Rev Mod Phys 74:47–97

    Article  Google Scholar 

  • Albin RL, Young AB, Penney JB (1989) The functional anatomy of basal ganglia disorders. Trends Neurosci 12:366–375

    Article  CAS  PubMed  Google Scholar 

  • Alexander GE, Crutcher MD (1990) Functional architecture of basal ganglia circuits: neural substrates of parallel processing. Trends Neurosci 13:266–271

    Article  CAS  PubMed  Google Scholar 

  • Archambault D, Munzner T, Auber D (2008) Grouseflocks: steerable exploration of graph hierarchy space. IEEE Trans Vis Comput Graph 14:900–913

    Article  PubMed  Google Scholar 

  • Bai X, Yu L, Liu Q, Zhang J, Li A, Han D, Luo Q, Gong H (2006) A high-resolution anatomical rat atlas. J Anat 209:707–708

    Article  PubMed  PubMed Central  Google Scholar 

  • Bakker R, Wachtler T, Diesmann M (2012) CoCoMac 2.0 and the future of tract-tracing databases. Front Neuroinform 6:1–6

    Article  Google Scholar 

  • Barbas H, Hilgetag CC, Saha S, Dermon CR, Suski JL (2005) Parallel organization of contralateral and ipsilateral prefrontal cortical projections in the rhesus monkey. BMC Neurosci 6:32

    Article  PubMed  PubMed Central  Google Scholar 

  • Bassett DS, Bullmore E (2006) Small-world brain networks. Neuroscientist 12:512–523

    Article  PubMed  Google Scholar 

  • Bennett BD, Bolam JP (1994) Synaptic input and output of parvalbumin-immunoreactive neurons in the neostriatum of the rat. Neuroscience 62:707–719

    Article  CAS  PubMed  Google Scholar 

  • Bevan MD, Booth PA, Eaton SA, Bolam JP (1998) Selective innervation of neostriatal interneurons by a subclass of neuron in the globus pallidus of the rat. J Neurosci 18:9438–9452

    CAS  PubMed  Google Scholar 

  • Bjaalie JG (2002) Localization in the brain: new solutions emerging. Nat Rev Neurosci 3:322–325

    Article  CAS  PubMed  Google Scholar 

  • Bohland J, Bokil H, Allen C, Mitra P (2009a) The brain atlas concordance problem: quantitative comparison of anatomical parcellations. PloS One 4:e7200

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  • Bohland JW, Wu C, Barbas H, Bokil H, Bota M, Breiter HC, Cline HT, Doyle JC, Freed PJ, Greenspan RJ, Haber SN, Hawrylycz M, Herrera DG, Hilgetag CC, Huang JZ, Jones A, Jones EG, Karten HJ, Kleinfeld D, Kötter R, Lester HA, Lin JM, Mensh BD, Mikula S, Panksepp J, Price JL, Safdieh J, Saper CB, Schiff ND, Schmahmann JD, Stillman BW, Svoboda K, Swanson LW, Toga AW, Van Essen DC, Watson JD, Mitra PP (2009b) A proposal for a coordinated effort for the determination of brainwide neuroanatomical connectivity in model organisms at a mesoscopic scale. PLoS Comput Biol 5:1–9

    Article  CAS  Google Scholar 

  • Bosch C, Mailly P, Degos B, Deniau JM, Venance L (2012) Preservation of the hyperdirect pathway of basal ganglia in a rodent brain slice. Neuroscience 215:31–41

    Article  CAS  PubMed  Google Scholar 

  • Bota M, Swanson L (2007a) The neuron classification problem. Brain Res Rev 56:79–88

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Bota M, Swanson L (2007b) Online workbenches for neural network connections. J Comp Neurol 500:807–814

    Article  PubMed  Google Scholar 

  • Bota M, Swanson L (2010) Collating and curating neuroanatomical nomenclatures: principles and use of the brain architecture knowledge management system (BAMS). Front Neuroinf 4:1–16

    Article  Google Scholar 

  • Bounova G, de Weck O (2012) Overview of metrics and their correlation patterns for multiple-metric topology analysis on heterogeneous graph ensembles. Phys Rev E 85:1–11

    Article  CAS  Google Scholar 

  • Brandes U, Erlebach T (2005) Network analysis. Springer, Berlin

    Book  Google Scholar 

  • Bullmore E, Sporns O (2009) Complex brain networks: graph theoretical analysis of structural and functional systems. Nat Rev Neurosci 10:186–198

    Article  CAS  PubMed  Google Scholar 

  • Burns G (1997) Neural connectivity of the rat: theory, methods and applications. PhD thesis, University of Oxford

  • Burns G, Cheng WC (2006) Tools for knowledge acquisition within the neuroscholar system and their application to anatomical tract-tracing data. J Biomed Discov Collab 1:10–16

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  • Burns G, Young M (2000) Analysis of the connectional organization of neural systems associated with the hippocampus in rats. Phil Trans R Soc Lond Ser B 355:55–70

    Article  CAS  Google Scholar 

  • Burns G, Cheng WC, Thompson R, Swanson L (2006) The NeuARt II system: a viewing tool for neuroanatomical data based on published neuroanatomical atlases. BMC Bioinf 7:531–550

    Article  CAS  Google Scholar 

  • Burns G, Cheng WC, Thompson R, Swanson L (2008a) The NeuARt II system: a viewing tool for neuroanatomical data based on published neuroanatomical atlases. LNCS 5151:9–18

    Google Scholar 

  • Burns G, Feng D, Hovy E (2008b) Intelligent approaches to mining the primary research literature: techniques, systems, and examples. In: Computational intelligence in medical informatics. Studies in computational intelligence, vol 85. Springer, Berlin, pp 17–50

  • Chida Y, Toyosawa K (1994) Laterality of theta waves recorded from the bilateral hippocampi in freely moving male rats. J Vet Med Sci 56:407–410

    Article  CAS  PubMed  Google Scholar 

  • Colizza V, Flammini A, Serrano MA, Vespignani A (2006) Detecting rich-club ordering in complex networks. Nat Phys 2:110–115

    Article  CAS  Google Scholar 

  • Costa LdF, Sporns O (2006) Correlating thalamocortical connectivity and activity. Appl Phy Lett 89:1–3

    Google Scholar 

  • DeLong MR (1990) Primate models of movement disorders of basal ganglia origin. Trends Neurosci 13:281–285

    Article  CAS  PubMed  Google Scholar 

  • Diesmann M, Gewaltig MO, Aertsen A (1999) Stable propagation of synchronous spiking in cortical neural networks. Nature 402:529

    Article  CAS  PubMed  Google Scholar 

  • Echtermeyer C, Costa L, Rodrigues F, Kaiser M (2011) Automatic network fingerprinting through single-node motifs. Plos One 6:1–9

    Article  CAS  Google Scholar 

  • Ercsey-Ravasz M, Markov NT, Lamy C, Van Essen DC, Knoblauch K, Toroczkai Z, Kennedy H (2013) A predictive network model of cerebral cortical connectivity based on a distance rule. Neuron 80:184–197

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Estrada E (2010) Quantifying network heterogeneity. Phys Rev E 82(066):102

    Google Scholar 

  • Estrada E, Hatano N (2008) Communicability in complex networks. Phys Rev E 77(036):111

    Google Scholar 

  • Fagiolo G (2007) Clustering in complex directed networks. Phys Rev E 76(026):107

    Google Scholar 

  • Felleman D, Essen DV (1991) Distributed hierarchical processing in the primate cerebral cortex. Cereb Cortex 1:1–47

    Article  CAS  PubMed  Google Scholar 

  • Feng G, Burton N, Hill B, Davidson D, Kerwin J, Scott M, Lindsay S, Baldock R (2005) JAtlasView: a Java atlas-viewer for browsing biomedical 3D images and atlases. BMC Bioinform 6:1–7

    Article  CAS  Google Scholar 

  • Foster KC, Muth SQ, Potterat JJ, Rothenberg RB (2001) A faster Katz status score algorithm. Comput Math Organ Theory 7:275–285

    Article  Google Scholar 

  • Freeman LC (1977) A set of measures of centrality based on betweenness. Sociometry 40:35–41

    Article  Google Scholar 

  • Gerfen C (2004) Basal ganglia. In: The rat nervous system. Elsevier Academic Press, San Diego, pp 455–508

  • Gerfen C, Bolam J (2010) The neuroanatomical organization of the basal ganglia. In: Handbook of basal ganglia structure and function: a decade of progress. Elsevier Academic Press, San Diego, pp 3–28

  • Gewaltig MO, Diesmann M (2007) Nest (neural simulation tool). Scholarpedia 2:1430

    Article  Google Scholar 

  • Gouws A, Woods W, Millman R, Morland A, Green G (2009) DataViewer3D: an open-source, cross-platform multi-modal neuroimaging data visualization tool. Front Neuroinf 3:1–18

    Article  Google Scholar 

  • Gustafson C, Tretiak O, Bertrand L, Nissanov J (2004) Design and implementation of software for assembly and browsing of 3D brain atlases. Comput Methods Programs Biomed 74:53–61

    Article  PubMed  Google Scholar 

  • Gustafson C, Bug W, Nissanov J (2007) Neuroterrain—a client–server system for browsing 3d biomedical image data sets. BMC Bioinform 8

  • Hagmann P, Cammoun L, Gigandet X, Meuli R, Honey CJ, Wedeen VJ, Sporns O (2008) Mapping the structural core of human cerebral cortex. PLoS Biol 6:1–15

    Article  CAS  Google Scholar 

  • Harriger L, van den Heuvel MP, Sporns O (2012) Rich club organization of macaque cerebral cortex and its role in network communication. PLoS One 7(e46):497

    Google Scholar 

  • Hilgetag CC, Grant S (2000) Uniformity, specificity and variability of corticocortical connectivity. Philos Trans R Soc Lond B Biol Sci 355:7–20

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Hjornevik T, Leergaard T, Darine D, Moldestad O, Dale A, Willoch F, Bjaalie J (2007) Three-dimensional atlas system for mouse and rat brain imaging data. Front Neuroinf 1:1–11

    Article  Google Scholar 

  • Honey C, Kötter R, Breakspear M, Sporns O (2007) Network structure of cerebral cortex shapes functional connectivity on multiple time scales. Proc Natl Acad Sci USA 104:10240–10245

  • Humphries M, Gurney K, Prescott T (2006) The brainstem reticular formation is a small-world, not scale-free, network. Proc R Soc B 273:503–511

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Izhikevich EM, Edelman GM (2008) Large-scale model of mammalian thalamocortical systems. Proc Natl Acad Sci USA 105:3593–3598

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Ju T, Warren J, Carson J, Bello M, Kakadiaris I, Chiu W, Thaller C, Eichele G (2006) 3d volume reconstruction of a mouse brain from histological sections using warp filtering. J Neurosci Methods 156:84–100

    Article  PubMed  Google Scholar 

  • Keinan A, Hilgetag CC, Meilijson I, Ruppin E (2004) Causal localization of neural function: the shapley value method. Neurocomputing 58–60:215–222

    Article  Google Scholar 

  • Kita H (1993) GABAergic circuits of the striatum. Prog Brain Res 99:51–72

    Article  CAS  PubMed  Google Scholar 

  • Kita H, Kitai ST (1994) The morphology of globus pallidus projection neurons in the rat: an intracellular staining study. Brain Res 636:308–319

    Article  CAS  PubMed  Google Scholar 

  • Kleinberg JM (1999) Authoritative sources in a hyperlinked environment. J ACM 46:604–632

    Article  Google Scholar 

  • Koos T, Tepper JM (1999) Inhibitory control of neostriatal projection neurons by GABAergic interneurons. Nat Neurosci 2:467–472

    Article  CAS  PubMed  Google Scholar 

  • Koslow S (1997) Neuroinformatic: an overview of the human brain project. Lawrence Erlbaum Associates, New Jersey

  • Koslow S (2005) Databasing the brain: from data to knowledge. Wiley, New York

  • Kötter R (2003) Neuroscience databases: a practical guide. Kluwer Academic Publishers, Dordrecht

  • Kötter R (2004) Online retrieval, processing, and visualization of primate connectivity data from the CoCoMac database. Neuroinformatics 2:127–144

    Article  PubMed  Google Scholar 

  • Kötter R (2007) Anatomical concepts of brain connectivity. Springer, Berlin

    Book  Google Scholar 

  • Kötter R, Reid AT, Krumnack A, Wanke E, Sporns O (2007) Shapley ratings in brain networks. Front Neuroinf 1:1–9

    Google Scholar 

  • Lanciego JL, Wouterlood FG (2011) A half century of experimental neuroanatomical tracing. J Chem Neuroanat 42:157–183

    Article  PubMed  Google Scholar 

  • Lee J, Munch K, Carlis J, Pardo J (2008) Internet image viewer (iiv). BMC Med Imaging 29:1–20

    Google Scholar 

  • Lima-Mendez G, van Helden J (2009) The powerful law of the power law and other myths in network biology. Mol BioSyst 5:1482–1493

  • Looi JC, Walterfang M, Nilsson C, Power BD, van Westen D, Velakoulis D, Wahlund LO, Thompson PM (2013) The subcortical connectome: hubs, spokes and the space between—a vision for further research in neurodegenerative disease. Aust N Z J Psychiatry EPub:1–4

  • Markov NT, Misery P, Falchier A, Lamy C, Vezoli J, Quilodran R, Gariel MA, Giroud P, Ercsey-Ravasz M, Pilaz LJ, Huissoud C, Barone P, Dehay C, Toroczkai Z, Van Essen DC, Kennedy H, Knoblauch K (2011) Weight consistency specifies regularities of macaque cortical networks. Cereb Cortex 21:1254–1272

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Markov NT, Vezoli J, Chameau P, Falchier A, Quilodran R, Huissoud C, Lamy C, Misery P, Giroud P, Ullman S, Barone P, Dehay C, Knoblauch K, Kennedy H (2014) Anatomy of hierarchy: feedforward and feedback pathways in macaque visual cortex. J Comp Neurol 522(1):225–259

    Article  PubMed  Google Scholar 

  • Martinez-Gonzalez C, van Andel J, Bolam JP, Mena-Segovia J (2013) Divergent motor projections from the pedunculopontine nucleus are differentially regulated in Parkinsonism. Brain Struct Funct 218 (epub ahead of print)

  • Martone ME, Gupta A, Ellisman MH (2004) E-neuroscience: challenges and triumphs in integrating distributed data from molecules to brains. Nat Neurosci 7:467–472

    Article  CAS  PubMed  Google Scholar 

  • Mascaro MB, Bittencourt JC, Casatti CA, Elias CF (2005) Alternative pathways for catecholamine action in oral motor control. Neurosci Lett 386:34–39

    Article  CAS  PubMed  Google Scholar 

  • Modha D, Singh R (2010) Network architecture of the long-distance pathways in the macaque brain. PNAS 107:13485–13490

  • Moore E, Poliakov A, Lincoln P, Brinkley J (2007) MindSeer: a portable and extensible tool for visualization of structural and functional neuroimaging data. BMC Bioinform 8(389):1–12

    Google Scholar 

  • Nambu A, Tokuno H, Takada M (2002) Functional significance of the cortico-subthalamo-pallidal ’hyperdirect’ pathway. Neurosci Res 43:111–117

    Article  PubMed  Google Scholar 

  • Newman M (2010) Networks: an introduction. Oxford University Press Inc., New York

    Book  Google Scholar 

  • Newman M, Girvan M (2004) Finding and evaluating community structure in networks. Phys Rev 69:1–15

    Google Scholar 

  • Newman M, Barabasi AL, Watts DJ (eds) (2006) The structure and dynamics of networks. Princeton University Press, New Jersey

  • Niggemann J, Gebert A, Schulz S (2008) Model formulation: modeling functional neuroanatomy for an anatomy information system. JAMIA 15:671–678

    PubMed  PubMed Central  Google Scholar 

  • Oh SW, Harris JA, Ng L, Winslow B, Cain N, Mihalas S, Wang Q, Lau C, Kuan L, Henry AM, Mortrud MT, Ouellette B, Nguyen TN, Sorensen SA, Slaughterbeck CR, Wakeman W, Li Y, Feng D, Ho A, Nicholas E, Hirokawa KE, Bohn P, Joines KM, Peng H, Hawrylycz MJ, Phillips JW, Hohmann JG, Wohnoutka P, Gerfen CR, Koch C, Bernard A, Dang C, Jones AR, Zeng H (2014) A mesoscale connectome of the mouse brain. Nature 508:207–214

    Article  CAS  PubMed  Google Scholar 

  • Ozik J, Hunt B, Ott E (2004) Growing networks with geographical attachment preference: emergence of small worlds. Phys Rev E 69(026):108

    Google Scholar 

  • Palombi O, Shin JW, Watson C, Paxinos G (2006) Neuroanatomical affiliation visualization-interface system. Neuroinformatics 4:299–317

    Article  PubMed  Google Scholar 

  • Parent A (1986) Comparative neurobiology of the basal ganglia. Wiley, New York

  • Parthasarathy K, Bhalla US (2013) Laterality and symmetry in rat olfactory behavior and in physiology of olfactory input. J Neurosci 33:5750–5760

    Article  CAS  PubMed  Google Scholar 

  • Passingham RE, Stephan KE, Kotter R (2002) The anatomical basis of functional localization in the cortex. Nat Rev Neurosci 3:606–616

    Article  CAS  PubMed  Google Scholar 

  • Paxinos G, Watson C (2004) The rat brain in stereotaxic coordinates, 5th edn. Elsevier Academic Press, London

    Google Scholar 

  • Paxinos G, Watson C (2007) The rat brain in stereotaxic coordinates, 6th edn. Elsevier Academic Press, Amsterdam

    Google Scholar 

  • Paxinos G, Watson C (2009) BrainNavigator. Academic Press Inc., San Diego

  • Press W, Olshausen B, Van Essen D (2001) A graphical anatomical database of neural connections. Phil Trans R Soc Lond B 356:1147–1157

    Article  CAS  Google Scholar 

  • Riddle DR, Purves D (1995) Individual variation and lateral asymmetry of the rat primary somatosensory cortex. J Neurosci 15:4184–4195

    CAS  PubMed  Google Scholar 

  • Rubinov M, Sporns O (2010) Complex network measures of brain connectivity: uses and interpretations. Neuroimage 52:1059–1069

    Article  PubMed  Google Scholar 

  • Sato F, Lavallee P, Levesque M, Parent A (2000) Single-axon tracing study of neurons of the external segment of the globus pallidus in primate. J Comp Neurol 417:17–31

    Article  CAS  PubMed  Google Scholar 

  • Schmitt O, Eipert P (2012) Neuroviisas: approaching multiscale simulation of the rat connectome. Neuroinformatics 10:243–267

    Article  PubMed  Google Scholar 

  • Schmitt O, Eipert P, Philipp K, Kettlitz R, Fuellen G, Wree A (2012) The intrinsic connectome of the rat amygdala. Front Neural Circuits 6:81

    Article  PubMed  PubMed Central  Google Scholar 

  • Schreiber F, Schwöbbermeyer H (2004) Towards motif detection in networks: frequency concepts and flexible search. In: Proceedings of the international workshop on network tools and applications in biology NETTAB04, pp 91–102

  • Schreiber F, Schwöbbermeyer H (2005) Mavisto: a tool for the exploration of network motifs. Bioinformatics 21:3572–3574

    Article  CAS  PubMed  Google Scholar 

  • Shanahan M, Wildie M (2012) Knotty-centrality: finding the connective core of a complex network. PLoS One 7(e36):579

    Google Scholar 

  • Shanahan M, Bingman VP, Shimizu T, Wild M, Güntürkün O (2013) Large-scale network organization in the avian forebrain: a connectivity matrix and theoretical analysis. Front Comp Neurosci 7(89):1–17

    Google Scholar 

  • Smith Y, Bevan MD, Shink E, Bolam JP (1998) Microcircuitry of the direct and indirect pathways of the basal ganglia. Neuroscience 86:353–387

    Article  CAS  PubMed  Google Scholar 

  • Sporns O (2002) Graph theory methods for the analysis of neural connectivity patterns. In: Kötter R (ed) Neuroscience databases. Kluwer Academic, Boston, pp 171–185

    Google Scholar 

  • Sporns O (2006) Small-world connectivity, motif composition, and complexity of fractal neuronal connections. Biosystems 85:55–64

    Article  PubMed  Google Scholar 

  • Sporns O (2011) Networks of the brain. The MIT Press, Cambridge

  • Sporns O, Kötter R (2004) Motifs in brain networks. PLoS Biol 2:1910–1918

    Article  CAS  Google Scholar 

  • Sporns O, Tononi G, Edelman GM (2000a) Connectivity and complexity: the relationship between neuroanatomy and brain dynamics. Neural Netw 13:909–922

    Article  CAS  PubMed  Google Scholar 

  • Sporns O, Tononi G, Edelman GM (2000b) Theoretical neuroanatomy: relating anatomical and functional connectivity in graphs and cortical connection matrices. Cereb Cortex 10:127–141

    Article  CAS  PubMed  Google Scholar 

  • Sporns O, Tononi G, Edelman GM (2002) Theoretical neuroanatomy and the connectivity of the cerebral cortex. Behav Brain Res 135:69–74

    Article  CAS  PubMed  Google Scholar 

  • Sporns O, Chialvo DR, Kaiser M, Hilgetag CC (2004) Organization, development and function of complex brain networks. Trends Cogn Sci 8:418–425

    Article  PubMed  Google Scholar 

  • Sporns O, Honey CJ, Kötter R (2007) Identification and classification of hubs in brain networks. PLoS One 2:1–14

    Article  Google Scholar 

  • Stam CJ, Reijneveld JC (2007) Graph theoretical analysis of complex networks in the brain. Nonlinear Biomed Phys 1:1–19

    Article  Google Scholar 

  • Stephan K, Kamper L, Bozkurt A, Burns G, Young M, Kötter R (2001a) Advanced database methodology for the collation of connectivity data on the macaque brain (CoCoMac). Philos Trans R Soc Lond B Biol Sci 356:1159–1186

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Stephan KE, Kamper L, Bozkurt A, Burns GAPC, Young MP, Kötter R (2001b) Advanced database methodology for the Collation of Connectivity data on the Macaque brain (CoCoMac). Philos Trans R Soc Lond Ser B 356:1159–1186

    Article  CAS  Google Scholar 

  • Steriade M (2004) Neocortical cell classes are flexible entities. Nat Rev Neurosci 5:121–134

    Article  CAS  PubMed  Google Scholar 

  • Sugar J, Witter M, van Strien N, Cappaert N (2011) The retrosplenial cortex: intrinsic connectivity and connections with the (para)hippocampal region in the rat. An interactive connectome. Front Neuroinf 5:1–13

    Article  Google Scholar 

  • Swanson L (2003) Brain maps: structure of the rat brain, vol 3. Elsevier, Amsterdam

  • Tepper JM, Abercrombie ED, Bolam JP (2007) Basal ganglia macrocircuits. Prog Brain Res 160:3–7

    Article  CAS  PubMed  Google Scholar 

  • Thibeault CM, Srinivasa N (2013) Using a hybrid neuron in physiologically inspired models of the basal ganglia. Front Comput Neurosci 7:88. (eCollection)

  • Tiesinga P, Fellous JM, Sejnowski TJ (2008) Regulation of spike timing in visual cortical circuits. Nat Rev Neurosci 9:97–107

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Towlson EK, Vértes PE, Ahnert SE, Schafer WR, Bullmore ET (2013) The rich club of the C. elegans neuronal connectome. J Neurosci 33:6380–6387

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • van den Heuvel MP, Sporns O (2011) Rich-club organization of the human connectome. J Neurosci 31:15775–15786

  • Van Horn JD, Gazzaniga MS (2013) Why share data? Lessons learned from the fMRIDC. Neuroimage 82:677–682

    Article  PubMed  Google Scholar 

  • Vlachos I, Aertsen A, Kumar A (2012) Beyond statistical significance: implications of network structure on neuronal activity. PLoS Comput Biol 8(e1002):311

    Google Scholar 

  • Wang Q, Sporns O, Burkhalter A (2012) Network analysis of corticocortical connections reveals ventral and dorsal processing streams in mouse visual cortex. J Neurosci 32:4386–4399

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Ware C (2013) Information visualization: perception for design. Morgan Kaufmann Publishers Inc., San Francisco

    Google Scholar 

  • Wong PC, Foote H, Chin G, Mackey P, Perrine K (2006a) Graph signatures for visual analytics. IEEE Trans Vis Comput Graph 12:1399–1413

    Article  PubMed  Google Scholar 

  • Wong PC, Foote H, Mackey P, Perrine K, Chin G (2006b) Generating graphs for visual analytics through interactive sketching. IEEE Trans Vis Comput Graph 12:1386–1398

    Article  PubMed  Google Scholar 

  • Yip AM, Horvath S (2007) Gene network interconnectedness and the generalized topological overlap measure. BMC Bioinform 8:22

    Article  CAS  Google Scholar 

  • Young M (1992a) Objective analysis of the topological organization of the primate cortical visual system. Nature 358:152–155

    Article  CAS  PubMed  Google Scholar 

  • Young MP (1992b) Objective analysis of the topological organization of the primate cortical visual system. Nature 358:152–155

    Article  CAS  PubMed  Google Scholar 

  • Young M (1993) The organization of neural systems in the primate cerebral cortex. Proc Biol Sci 252:13–18

    Article  CAS  PubMed  Google Scholar 

  • Young M, Scannell J, Burns G, Blakemore C (1994) Analysis of connectivity: neural systems in the cerebral cortex. Rev Neurosci 5:227–250

    Article  CAS  PubMed  Google Scholar 

  • Young M, Scannell J, Burns G (1996) The analysis of cortical connectivity. Springer, Berlin

  • Zhou S, Mondragn RJ (2004) The rich-club phenomenon in the internet topology. IEEE Commun Lett 8:180–182

    Article  Google Scholar 

Download references

Acknowledgments

The authors thank Klaus-Peter Schmitz (Department of Biomedical Engineering, University of Rostock) for the support of the neuroVIISAS project. We thank Frauke Winzer, Susanne Lehmann, Antje Schümann, Jennifer Meinhardt, Ann-Christin Klünker for their faithful work on extending the database and mappings. All work was supported by the Faculty of Mathematics and Natural Sciences and of the Faculty of Medicine of the University of Rostock.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Oliver Schmitt.

Appendices

Appendix A

The appendix contains in the first part figures, in the second part tables and the third part formal definitions of matrices, graph-theoretical parameters and randomization models.

See Figs. 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26.

Appendix B

Tables 7, 8, 9, 10, 11, 12, 13, 14, 15.

Fig. 10
figure 10

Ordinal categories of connection strengths are plotted on the abscissa and the hypothetical estimates of connection densities shown on the left ordinate which scales the blue straight line. The right ordinate is linearly scaled for the red graph. The largest slope lies between the semiquantitative values 3.5 and 4. This logarithmic transformation has been used for adapting the ordinal categories to a distribution of connection densities that was observed by others as well [Hilgetag and Grant (2000); Markov et al. (2011, 2014); Oh et al. (2014); Ercsey-Ravasz et al. (2013); Wang et al. (2012)]. The function for transforming ordinal categories of connection strengths to the logarithmic distributed connection densities is \(f(x)=e^{ln(10)x-4ln(10)}\)

Fig. 11
figure 11

The regions of the BG2 network are organized in a hierarchy. It is presented as a triangle hierarchy (top) and a region hierarchy (left). The sequence of regions of the hierarchy (from left to right) is the same as that in the adjacency matrix. All regions of the connectomes that were investigated are organized in a hierarchy

Fig. 12
figure 12

Chord diagram visualization of the bilateral BG1 network using CIRCOS. The diagram consists of 4 nested rings. The outer ring of arcs indicates the relative frequency of inputs and outputs to a particular region. The second ring is the input arc ring and the third ring is the output ring of arcs. The inner ring indicates the absolute numbers of connections followed by a thin layer of arcs with an input part and output part for each region. The extensive connectivity of the CPu with SNC and SNR is visualized prominently. Stronger contralateral connections can be easily found by the thicker red connections between the CPu and contralateral AGl

Fig. 13
figure 13

Chord diagram visualization of the bilateral BG2 network using CIRCOS. The amount of connections is relatively large. The strong connections of the CPu and AC are diverging into many different regions

Fig. 14
figure 14

The distance matrix of the ipsilateral BG2 network. The distances between regions correspond to the smallest number of connections that are necessary to connect pairs of regions. The thalamic paths are in most cases longer (lighter gray values) than paths of midbrain regions or MRF

Fig. 15
figure 15

The communicability matrix of the ipsilateral BG2 network. The number of shortest paths through a pair of nodes is presented by the communicability matrix (Estrada and Hatano 2008). Light gray values are indicating large communicability values. The CPu, AC, STh, PF, MRF and VTA are contributing to shortest paths of the BG2-network

Fig. 16
figure 16

The generalized topology overlap matrix (GTOM) of the ipsilateral BG2 network. The pairwise interconnectedness in relation to the number of neighbors that a pair of nodes share in common is presented by the GTOM. Large values indicate many similarities of neighbors and connections of a pair of nodes. E.g., the pair CL and CPu has a relatively large GTOM value like the pair STh and VTA

Fig. 17
figure 17

The variability of connection strengths of the bilateral BG2 network. Large-standard deviations of connection weights are indicated by light gray values. The variability of connection weights of the SNC has relatively smaller values in comparison to VTA, HIPP and AGl

Fig. 18
figure 18

The global parameters of the ipsilateral BG2 network. The first row (beginning with “Nodes”) presents the general parameters of the BG2 network. The line density is 52.333 %. If each regions would be connected with each other region then the line density is 100 %. Cy number of shortest cycles and CyC is the cycle coefficient, Avg-HD is the average hierarchical level of the regions in the BG2 network. Six-different randomizations have been performed. Each randomization was repeated 1,000 times using exactly the same number of regions and connections as in the real BG2 network. The global network parameters of the real BG network are shown in the first column followed by mean values of the randomizations. The rewiring randomization is showing the smallest differences with regard to global parameters when comparing with the real network

Fig. 19
figure 19

The local parameters of the ipsilateral BG2 network in parallel coordinates. Parallel coordinate visualization provides an overview of a high-dimensional parameter space. Parameters are sorted by similarity. The CPu, AGl and AGm are highlighted by dashed lines. The CPu has almost largest or smallest values with regard to centrality measures. For abbreviations of parameters see Appendix

Fig. 20
figure 20

The rich-club regions of the ipsi and contralateral BG2 network have a rich-club coefficient \(\phi > 0.8\) respectively a rank of \(\ge 44.\) The direct neighbors or feeders are shown in the upper arc. Their interconnections are presented by blue edges. The connection of rich nodes with feeder nodes is shown by yellow edges and the interconnectivity of rich nodes by black edges. The indices _L and _R are indicating the sides of the hemispheres

Fig. 21
figure 21

The regions of the ipsi- and contralateral BG2 network that belong to the knotty-center in an arc diagram visualization. The knotty-center regions are on the lower arc. Direct neighbors of knotty centers are shown in the upper arc. Their interconnections are presented by blue edges. The connections of the knotty-center with their direct neighbors are shown by yellow edges and the interconnectivity of rich-nodes by black edges. The indices _L and _R are indicating the sides of the hemispheres

Fig. 22
figure 22

The K-core analysis of the ipsilateral BG2 network. The connections of CPu are highlighted. In the outer circles are those regions which have the smallest number of connections and in the center are regions that have the largest number of connections

Fig. 23
figure 23

The K-core analysis of the ipsi- and contralateral BG2 network. The connections of the CPu are highlighted

Fig. 24
figure 24

Metric multidimensional scaling (MDS) applied to the ipsilateral BG2 network. Regions that are adjacent to each other have more similarities with regard to their connectivity than with regions that are further away from each other. The 8 thalamic regions that are located within the lower-left quadrant are relatively similar with regard to connectivity. AGm and AGl in the lower-right quadrant have also a large similarity. Interestingly, SNR is adjacent to MRF, however, the second major output node of the BG is MGP which is not to far from SNR. Limbic regions like Ac, Ent, HIPP and Pir are located in the upper left quadrant

Fig. 25
figure 25

MDS applied to the bilateral BG2 network. The major output regions SNR and MGP are located in the upper- and lower- left quadrant, however, not too far from each other. Most limbic and all thalamic regions are located in the upper-left quadrant (left hemisphere) or in the upper-right quadrant (right hemisphere)

Fig. 26
figure 26

Counts of cyclic connections or loops in the bilateral BG2 network. The logarithm (Log) of the frequency of contributions of regions to a cycle of size 2–7 is presented by the y-axis. Every second region (all regions of the left hemisphere) are labeled. For a cycle from MRF back to MRF passing 4 regions within the BG2 network there exist 8,255 possible paths or 41,191,973 paths when 7 regions are allowed

Table 7 Local parameters of the weighted-ipsilateral BG2 network
Table 8 Local parameters of the bilateral BG2 network
Table 9 Local parameters of the bilateral BG2 network
Table 10 Frequency of motifs of the ipsilateral BG2 network
Table 11 Global extrinsic connections of regions of the bilateral BG2 network
Table 12 Correlation coefficients of local parameters of the bilateral BG2 network
Table 13 The local parameters of the bilateral BG2 network were applied to the PCA
Table 14 The local parameters of the embedded BG1 network
Table 15 Local parameters of the ipsilateral BG2 network

Appendix C: Formal definitions of parameters, matrices and concepts

The definitions of expressions, parameters, matrices and simulation models (random graph models) used in this article are summarized in the following. More detailed descriptions, algorithms and proofs are provided elsewhere (Newman 2010; Rubinov and Sporns 2010; Newman et al. 2006; Kötter 2007; Brandes and Erlebach 2005; Kötter 2003).

Basic definitions

Node, vertex: The smallest subunit of a network. With regard to connectomes a node is a circumscribed or disjunctive region that contains neuron perikarya (sources of physiological action potential) and/or axonal terminals (targets of physiological action potentials).

Set of indexed nodes: The set of all indices of nodes is

$$\begin{aligned} N = \{ 1, 2, 3, \ldots , n\} \end{aligned}$$
(2)

Number of nodes: The number of nodes (regions, vertices) is

$$\begin{aligned} n=|N| \end{aligned}$$
(3)

Edge: A directed edge \((i, j) \in N \times N\) is the line that connects vertices i and j with source i and target j. The set of directed edges E is

$$\begin{aligned} E \subseteq N \times N\ \end{aligned}$$
(4)

Edges: The number of edges (connections, links) \(\epsilon \) is

$$\begin{aligned} \epsilon = |E| \end{aligned}$$
(5)

Set of edges:

$$\begin{aligned} L=\{ (i, j) \in E | i \ne j\} \end{aligned}$$
(6)

The set of all not self-referencing edges is

$$\begin{aligned} \ell = |L| \end{aligned}$$
(7)

Graph:

$$\begin{aligned} G = (N,E) \end{aligned}$$
(8)

Adjacency matrix: The adjacency matrix (connectivity matrix) A is

$$\begin{aligned} A= (a_{ij})_{i,j=1}^n \quad \hbox{where } a_{ij}= {\left\{ \begin{array}{ll} 1 & \text{if } (i,j) \in E\\ 0 & \text {else} \end{array}\right. } \end{aligned}$$
(9)

Weighted matrix: The weighted matrix W is

$$\begin{aligned} W= (w_{ij})_{i,j=1}^n \end{aligned}$$
(10)

whereas \(w_{ij}\) is the weight of the edge \((i, j)\) that connects i and j.

\(0 \le w_{ij} \le 1\).

Path: A sequence of vertices (\(v_1,\ldots v_k\)) is a path from (\(v_1\) to \(v_k\)) if \(\forall _i \in \{1,\ldots ,k-1\}:(v_i, v_{i+1}) \in E\). The length of a path \(v_1,\ldots ,v_k\) is \(k-1\).

Distance matrix: The distance matrix D is

$$\begin{aligned} D= (d_{ij})_{i,j=1}^{n} \end{aligned}$$
(11)

where

$$\begin{aligned} d_{ij}=d(i,j)= {\left\{ \begin{array}{ll} \hbox {length of the shortest path from } i \hbox{ to } j, & \hbox {if such a path exists}\\ \infty , & \text {else} \end{array}\right. } \end{aligned}$$
(12)

Generalized topology matrices (GTOM): Let \(N_m(i)\) be the \(m-step\) neighborhood of node i:

$$\begin{aligned} N_m(i)&= \{j \ne i | \min \{ d(i,j), d(j,i)\} \le m\}\end{aligned}$$
(13)
$$\begin{aligned} N^{\rm out}_m(i)&= \{j \ne i | d(i,j) \le m\}\end{aligned}$$
(14)
$$\begin{aligned} N^{\rm in}_m(i)&= \{j \ne i | d(j,i) \le m\} \end{aligned}$$
(15)

then the GTOM-matrix of step m is defined as

$$\begin{aligned} {\rm GTOM}(m)=(g^m_{ij})^n_{i,j=1} {\text with }\; g^m_{ij}= {\left\{ \begin{array}{ll} \frac{|N_m(i) \cap N_m(j) |+a_{ij} \cdot a_{ji}}{\min \{ |N_m(i)|,|N_m(j)| \}+1-a_{ij} \cdot a_{ji}}\\ & \hbox {if } i \ne j\\ 1 &{} \hbox {if } i = j\\ \end{array}\right. } \end{aligned}$$
(16)

The definitions of \({\rm GTOM}^{\rm in}_{(m)}\) and \({\rm GTOM}^{\rm out}_{(m)}\) with the directed \(m-step\) neighborhoods \(N^{\rm in}_m(i)\) and \(N^{\rm out}_m(i)\) are analog.

Degree all (\({\rm deg}_{{\rm all}}\), \({\rm DG}_{{\rm All}}\)): Self-references of nodes are not considered for all three degree measures.

\({\rm deg} (i) = {\rm deg}_{{\rm all}}(i)\)

$$\begin{aligned} {\rm deg} (i) = \sum \limits _{\begin{subarray}{l} j=1\\ j \ne i \end{subarray}}^n a_{ij} + a_{ji} \end{aligned}$$
(17)

Degree out:

$$\begin{aligned} {\rm deg} _{\rm out}(i) = \sum \limits _{\begin{subarray}{l} j=1\\ j \ne i\end{subarray}}^n a_{ij} \end{aligned}$$
(18)

Degree in:

$$\begin{aligned} {\rm deg} _{\rm in}(i) = \sum \limits _{\begin{subarray}{l} j=1\\ j \ne i \end{subarray}}^n a_{ji} \end{aligned}$$
(19)

Reciprocal edge count \({\rm Rec}(i)\): \({\rm Rec}(i)\) is the number of reciprocal edges adjacent to a node i.

$$\begin{aligned} {\rm Rec}(i) = \sum \limits _{\begin{subarray}{l} j \in N\\ i \ne j\end{subarray}} a_{ij} a_{ji} \end{aligned}$$
(20)

Laterality: Let \(N^{\rm IPSI} \subseteq N\) be a subset of nodes and \({\rm deg} ^{\rm IPSI}(i)\) the degree of the node i in the subset \(N^{\rm IPSI}\). Then the lateralities are defined as follows:

$$\begin{aligned}&{\rm Lat}_{{\rm all}}(i) = \frac{{\rm deg} ^{\rm IPSI}_{\rm all}(i)}{{\rm deg} _{\rm all}(i)}\end{aligned}$$
(21)
$$\begin{aligned}&{\rm Lat}_{\rm in}(i) = \frac{{\rm deg} ^{\rm IPSI}_{\rm in}(i)}{{\rm deg} _{\rm in}(i)}\end{aligned}$$
(22)
$$\begin{aligned}&{\rm Lat}_{\rm out}(i) = \frac{{\rm deg} ^{\rm IPSI}_{\rm out}(i)}{{\rm deg} _{\rm out}(i)} \end{aligned}$$
(23)

Laterality of the reciprocal edge count \({\rm Lat}_{{\rm Rec}}(i)\):

The laterality of the reciprocal edge count is the fraction of ipsilateral reciprocal edges. For a node \(i \in N^{\rm IPSI} \subseteq N\).

$$\begin{aligned} {\rm Lat}_{\rm Rec}(i) = \frac{1}{{\rm Rec}(i)} \sum \limits _{\begin{subarray}{l} j \in N^{\rm IPSI}\\ i \ne j\end{subarray}} a_{ij} a_{ji} \end{aligned}$$
(24)

Neighborhoods: Out-neighbors of i:

$$\begin{aligned} N^{\rm out}_i = \{j \in N \backslash \{i\} | a_{ij}=1\} \end{aligned}$$
(25)

In-neighbors of i:

$$\begin{aligned} N^{\rm in}_i = \{j \in N \backslash \{i\} | a_{ji}=1\} \end{aligned}$$
(26)

All neighbors of i:

$$\begin{aligned} N_i = N^{\rm out}_i \cup N^{\rm in}_i \end{aligned}$$
(27)
$$\begin{aligned} N^{+}_i = N_i \cup \{i\} \end{aligned}$$
(28)

Network parameters

Communicability matrix G:

$$\begin{aligned} G_{pq} = \sum _{k=0}^{\infty } \frac{({\mathbf{A}}^k)_{pq}}{k!} = (e^{\mathbf{A}})_{pq} \end{aligned}$$
(29)

Modularity measure: Let \(M=\{M_1,\ldots ,M_m\}\) be a partition of N. \(M_i\) is a group, module or cluster of vertices. With

$$\begin{aligned} e_i= \frac{1}{\ell } \sum \limits _{\begin{subarray}{l} j,k \in M_i \\ j < k\end{subarray}} (a_{jk} + a_{kj}), \end{aligned}$$
(30)

the fraction of edges that fall within group \(M_i \subseteq N\) and

$$\begin{aligned} a_i= \frac{1}{2\ell } \sum \limits _{j \in M} \sum \limits _{k \in N \backslash \{j\}} (a_{jk} + a_{kj}), \end{aligned}$$
(31)

the fraction of ends of edges that are attached to vertices in group \(M_i\), the Modularity

$$\begin{aligned} Q= \sum \limits ^m_{i=1} (e_i - a^2_i), \end{aligned}$$
(32)

whereas \(a^2_i\) is the fraction of edges that would connect vertices within group \(M_i\) if they were connected at random. A large modularity implies that the fraction of edges that fall within groups is larger than expected in the random case. The partition is generated by a “Greedy” optimization algorithm. Starting with a partition where every single node has its own group, stepwise those two groups are joined that increase Q most. The algorithm ends if there are no more such groups. The weighted case is similar, only the \(a_{ij}\) are replaced by \(w_{ij}\) and \(\ell \) is replaced by the sum of the edge weights

$$\begin{aligned} \ell ^{\overrightarrow{w}} = \sum \limits _{\begin{subarray}{l} i,j \in N \\ i \ne j\end{subarray}} w_{ij} \end{aligned}$$
(33)

The method of Newman and Girvan (2004) was used.

Global efficiency GE:

$$\begin{aligned} {\rm GE} = \frac{1}{n(n-1)} \cdot \sum \limits _{\begin{subarray}{l} i,j \in N \\ i \ne j\end{subarray}} \frac{1}{d(ij)} \end{aligned}$$
(34)

\({\rm GE}^\rightarrow \) and \({\rm GE}^{\overrightarrow{w}}\) analog with \(d^\rightarrow (i,j)\) and \(d^{\overrightarrow{w}}(i,j)\)

Directed global efficiency:

$$\begin{aligned} {\rm GE}^{\rightarrow } = \frac{1}{n(n-1)} \cdot \sum \limits _{\begin{subarray}{l} i,j \in N \\ i \ne j\end{subarray}} \frac{1}{d^{\rightarrow }(ij)} \end{aligned}$$
(35)

Harmonic mean HM:

$$\begin{aligned} {\rm HM} = \frac{1}{\rm GE} \end{aligned}$$
(36)

The directed and weighted versions use the directed and weighted global efficiencies.

Local efficiency: The local efficiency indicates how strong neighbors of nodes are interconnected. For each node i the inverse lengths of the shortest paths of the neighbors of i that are passing i are added. The local efficiency is this sum divided by the maximal possible sum of paths between neighbors that are connected with i. The efficiency of the network (global efficiency) is the average local efficiency of all nodes.

Directed local efficiency:

$$\begin{aligned} {\rm LE}^{\rightarrow } = \frac{1}{n} \sum \limits _{\begin{subarray}{l} i\in N \\ n_i > 1\end{subarray}} \frac{\sum \nolimits _{\begin{subarray}{l} j,k \in N_i \\ j \ne k\end{subarray}} d_{jk}(N_i)^{-1}}{n_i \cdot (n_i-1)} \end{aligned}$$
(37)

Weighted directed local efficiency:

$$\begin{aligned} {\rm LE}^{\overrightarrow{w}} = \frac{1}{n} \sum \limits _{\begin{subarray}{l} i\in N \\ n_i > 1\end{subarray}} \frac{\sum \nolimits _{\begin{subarray}{l} j,k \in N_i \\ j \ne k\end{subarray}} d^{\overrightarrow{w}}_{jk}(N_i)^{-1}}{n_i \cdot (n_i-1)} \end{aligned}$$
(38)

whereby \(n_i=|N_i|\) and \(d_{jk}(N_i)\), respectively, \(d^{\overrightarrow{w}}_{jk}(N_i)\) is the length of the shortest path between j and k that contains only neighbors of i.

Directed assortativity coefficient \(r^{\rightarrow }\):

$$\begin{aligned} r^{\rightarrow } = \frac{ \sum \nolimits _{(i,j)\in L} {\rm deg} _{\rm out}(i) \cdot {\rm deg} _{\rm in}(j) - \frac{1}{4\ell } \cdot \left[\sum \nolimits _{(i,j)\in L} ({\rm deg} _{\rm out}(i) + {\rm deg} _{\rm in}(j)) \right]^2}{\frac{1}{2} \cdot \sum \nolimits _{(i,j)\in L} ( {\rm deg} _{\rm out}(i)^2 + {\rm deg} _{\rm in}(j)^2 ) - \frac{1}{4\ell } \cdot \left[\sum \nolimits _{(i,j)\in L} ({\rm deg} _{\rm out}(i) + {\rm deg} _{\rm in}(j)) \right]^2} \end{aligned}$$
(39)

Directed and weighted assortativity coefficient \(r^{\overrightarrow{w}}\):

$$\begin{aligned} r^{\overrightarrow{w}} = \frac{\sum \nolimits _{(i,j)\in L} w_{ij} \cdot ( {\rm deg} ^w_{\rm out}(i) \cdot {\rm deg} ^w_{\rm in}(j) ) - \frac{1}{4\ell } \cdot \left[\sum \nolimits _{(i,j)\in L} w_{ij} \cdot ({\rm deg} ^w_{\rm out}(i) + {\rm deg} ^w_{\rm in}(j)) \right]^2}{\frac{1}{2} \cdot \sum \nolimits _{(i,j)\in L} w_{ij} \cdot ( {\rm deg} ^w_{\rm out}(i)^2 + {\rm deg} ^w_{\rm in}(j)^2) - \frac{1}{4\ell } \cdot \left[\sum \nolimits _{(i,j)\in L} w_{ij} \cdot ({\rm deg} ^w_{\rm out}(i) + {\rm deg} ^w_{\rm in}(j)) \right]^2} \end{aligned}$$
(40)

The correlation of the degrees of nodes that are connected: \(-1 \le r \le 1\). Large positive values imply that nodes are mainly connected to nodes with similar degrees. Large negative values imply that nodes with a large degree are mainly to nodes that have a small degree. If \(r \approx 0\) there is no relation detectable.

Average path length = characteristic path length \((\overline{d})\): With \(P=\{(i,j) \in N \times N | d(i,j) < \infty \}\), the set of paths.

$$\begin{aligned} \overline{d} = {\left\{ \begin{array}{ll} \frac{1}{|P|} \sum \limits _{(i,j) \in P} d(i,j), &{} P \ne \varnothing \\ 0, &{} P = \varnothing \end{array}\right. } \end{aligned}$$
(41)

In the weighted case the distances \(d(i,j)\) are replaced by the weighted distances.

Average directed degree \(\overline{{\rm deg} }\):

$$\begin{aligned} \overline{{\rm deg} }=\frac{2 \, \ell }{n} \end{aligned}$$
(42)

Heterogeneity VC: Coefficient of variation (VC) of the \({\rm degree}_{\rm all}\) parameter.

$$\begin{aligned} H_{\rm VC} = \frac{1}{\overline{{\rm deg} }} \cdot \sqrt{\sum \limits _{i \in N} ({\rm deg} _{\rm all}(i) - \overline{{\rm deg}})^2} \end{aligned}$$
(43)

If \(H_{\rm VC}=0\), all nodes have the same degree. The larger \(H_{\rm VC}\) the more diverse are the node degrees. In the weighted case the versions of the degrees are used. The heterogeneity measure of Estrada (2010) was not implemented because it is not defined for directed and weighted graphs.

Line density Ld:

$$\begin{aligned} {\rm Ld} = \frac{\ell }{n \cdot (n-1)} \end{aligned}$$
(44)

Without self-referencing edges.

Rich-club coefficient \(\phi (k)\): With \(N_k=\{i \in N | {\rm deg} (i) > k\}\) and \(E_k=\{(i,j) \in N_k \times N_k | (i,j) \in E\}\) \(G_k=(N_k, E_k)\) is the subgraph of \(G=(N,E)\) containing all vertices with a degree greater than k. The rich-club coefficient \(\phi (k)\) of a graph G is defined as the line density of the subgraph \(G_k\):

$$\begin{aligned} \phi (k) = {\rm Ld}(G_k) \end{aligned}$$
(45)

Diameter Diam:

$$\begin{aligned} {\rm Diam} = \max \lbrace d(i,j) | d(i,j)< \infty \rbrace \end{aligned}$$
(46)

Katz index: The Katz index (Katz status index, Katz centrality) is a measure for the direct and indirect input of a node (Foster et al. 2001).

$$\begin{aligned} C_{{\rm Katz}}(i) = \sum \limits _{k=1}^\infty \sum \limits _{j=1} \alpha ^k (A^k)_{ji} \end{aligned}$$
(47)

The attenuation factor \(\alpha \) has to be smaller than the reciprocal of the absolute value of the largest eigenvalue of A. For a better readability and comparability of the results, in neuroVIISAS the Katz centrality is multiplied by the mean of the quotient \(\frac{{\rm deg} _{\rm in}(i)}{C_{{\rm Katz}}(i)}\) of all nodes with \(C_{{\rm Katz}}(i) > 0\). Hence, the values lie in the same range as the indegrees.

Number of triangles:

$$\begin{aligned} t^{\rightarrow }(i) = \sum _{\begin{subarray}{l} j, k \in N \backslash \{i\} \\ j < k\end{subarray}} (a_{ij} +a_{ji})(a_{ik} +a_{ki})(a_{jk} +a_{kj}) \end{aligned}$$
(48)

The maximum number of possible triangles that can be deviated from a complete reciprocal triangle is 8.

Weighted number of triangles:

$$\begin{aligned} t^{\overrightarrow{w}}(i) = \sum _{\begin{subarray}{l} j, k \in N \backslash \{i\} \\ j < k\end{subarray}} (w_{ij}^{\frac{1}{3}} +w_{ji}^{\frac{1}{3}})(w_{ik}^{\frac{1}{3}} +w_{ki}^{\frac{1}{3}})(w_{jk}^{\frac{1}{3}} +w_{kj}^{\frac{1}{3}}) \end{aligned}$$
(49)

Instead of the sum of triangles (\(t^{\rightarrow }(i)\)) the sum of geometric means of edge weights of each triangle is calculated. The following example provides \((w_{ij} \cdot w_{jk} \cdot w_{ik})^{\frac{1}{3}}\) as the summand:

figure a

Directed transitivity: The general definition of transitivity (T) is the sum of number of triangles around all nodes divided by the maximum possible sum of triangles around all nodes.

$$\begin{aligned} T^{\rightarrow } = \frac{\sum \nolimits _{i\in N} t^{\rightarrow }(i)}{\sum \nolimits _{i\in N} t_{\max }(i)} \end{aligned}$$
(50)

Directed and weighted transitivity:

$$\begin{aligned} T^{\overrightarrow{w}} = \frac{\sum \nolimits _{i\in N} t^{\overrightarrow{w}}(i)}{\sum \nolimits _{i\in N} t_{\max }(i)} \end{aligned}$$
(51)

whereby \(t_{\max }(i)= {\rm deg} (i) \cdot ({\rm deg} (i)-1) - 2 \cdot \hbox {rec}(i)\) with \({\rm deg} (i)=\) number of adjacent edges of i and \(\hbox {rec}(i)=\) number of reciprocal edges of i (the two directions of one reciprocal edge are considered as one reciprocal edge).

The degree \({\rm deg} \) and the reciprocity \(\hbox {rec}\) are defined as:

$$\begin{aligned} {\rm deg} (i)&= \sum \limits _{j\in N \backslash \{i\}} a_{ij}+a_{ji}\end{aligned}$$
(52)
$$\begin{aligned} \hbox {rec}(i)&= \sum \limits _{j\in N \backslash \{i\}} a_{ij} \cdot a_{ji} \end{aligned}$$
(53)

For the directed and weighted case:

$$\begin{aligned} a_{ij}= {\left\{ \begin{array}{ll} 1 &{} w_{ij} > 0\\ 0 &{} \text {else} \end{array}\right. } \end{aligned}$$
(54)

Cluster coefficient (triangle based): The triangle based cluster coefficient (Fagiolo 2007) of a node n is the number of triangles around n divided by the maximum possible number. In this version of the cluster coefficient reciprocal edges to a neighbor of a node n can affect the cluster coefficient of node n. In the other version only edges between neighbors of n have an influence to the cluster coefficient of node n.

$$\begin{aligned} C^{\rightarrow }_T&= \frac{t^{\rightarrow }(i)}{t_{\max }(i)}\end{aligned}$$
(55)
$$\begin{aligned} C^{\overrightarrow{w}}_T&= \frac{t^{\overrightarrow{w}}(i)}{t_{\max }(i)} \end{aligned}$$
(56)

Cluster coefficient: Number of edges between the neighbors of a node divided by the maximum possible number. \(C^{\rightarrow }(i)\) refers to all neighbors of i.

$$\begin{aligned} C^{\rightarrow }(i) = \frac{1}{|N_i| \cdot (|N_i|-1)} \cdot \sum \limits _{\begin{subarray}{l} j,k \in N_i \\ j \ne k\end{subarray}} a_{jk} \end{aligned}$$
(57)

\(C^{\rightarrow }_{\rm out}(i)\) refers to the out-neighbors of i.

$$\begin{aligned} C^{\rightarrow }_{\rm out}(i) = \frac{1}{|N^{\rm out}_i| \cdot (|N^{\rm out}_i|-1)} \cdot \sum \limits _{\begin{subarray}{l} j,k \in N^{\rm out}_i \\ j \ne k\end{subarray}} a_{jk} \end{aligned}$$
(58)

\(C^{\rightarrow }_{\rm in}(i)\) refers to the in-neighbors of i.

$$\begin{aligned} C^{\rightarrow }_{\rm in}(i) = \frac{1}{|N^{\rm in}_i| \cdot (|N^{\rm in}_i|-1)} \cdot \sum \limits _{\begin{subarray}{l} j,k \in N^{\rm in}_i \\ j \ne k \end{subarray}} a_{jk} \end{aligned}$$
(59)

In the weighted case the \(a_{ij}\) are replaced by the \(w_{ij}\).

Average cluster coefficient:

$$\begin{aligned} C^{\rightarrow } = \frac{1}{n} \sum \limits _{i=1}^n C^{\rightarrow }_i \end{aligned}$$
(60)

and

$$\begin{aligned} C^{\overrightarrow{w}} = \frac{1}{n} \sum \limits _{i=1}^n C^{\overrightarrow{w}}_i \end{aligned}$$
(61)

Small worldness S:

$$\begin{aligned} S = \frac{\left( \frac{C}{C_{\rm rand}}\right) }{\left( \frac{\overline{d}}{\overline{d}_{\rm rand}}\right) } \end{aligned}$$
(62)

Centrality:

$$\begin{aligned} C_{\rm D} = \frac{\sum \nolimits _{i=1}^n {\rm deg} _{\max } -{\rm deg} (i)}{(n-1) \cdot (n-2)} = \frac{n \cdot {\rm deg} _{\max } - 2 \cdot \ell }{(n-1) \cdot (n-2)} \end{aligned}$$
(63)

This centrality (degree centrality) is defined for an undirected network based on undirected degrees. A directed or weighted version is not available yet. For the calculation the directed network is transferred to an undirected one.

Circle length LC:

$$\begin{aligned} {\rm LC}(i)= {\left\{ \begin{array}{ll} d(i,i), &{} d(i,i) < \infty \\ 0, &{} d(i,i) = \infty \end{array}\right. } \end{aligned}$$
(64)

Eccentricity out: Eccentricity out, the output eccentricity of the vertex i is the maximum distance from i to any vertex.

$$\begin{aligned} {\rm Ecc}^{\rm out}(i) = \max \lbrace d(i,j) | j \in N \rbrace \end{aligned}$$
(65)

Eccentricity in: Eccentricity in, the input eccentricity of the vertex i is the maximum distance from i to any vertex.

$$\begin{aligned} {\rm Ecc}^{\rm in}(i) = \max \lbrace d(j,i) | j \in N \rbrace \end{aligned}$$
(66)

Cluster-coefficient of second neighbors: The cluster-coefficient of second neighbors (Hierarchical directed cluster coefficient of second (indirect) neighbors) \(C_2(i)\) is the number of edges between the 2nd neighbors of node i, divided by the maximum possible number of edges. In the weighted case it is the sum of weights of the edges between the 2nd neighbors of node i, divided by the maximum possible sum. With

$$\begin{aligned} N_2(i)= \left( \bigcup _{j \in N_i} N_j\right) \left. \vphantom{\bigcup _{j \in N_i}}\right\backslash N^{+}_i, \end{aligned}$$
(67)

the set of second neighbors of node i is:

$$\begin{aligned} C_2(i)= {\left\{ \begin{array}{ll} \frac{1}{|N_2(i)| \cdot (|N_2(i)|-1)}, &{} \hbox {if } |N_2(i)| > 1\\ 0, & \text {otherwise} \end{array}\right. } \end{aligned}$$
(68)

In the weighted case the \(a_{ij}\) are replaced by \(w_{ij}\).

Average neighbor degree: The non-weighted average neighbor degree \({\rm NB}(i)\) of node i is

$$\begin{aligned} {\rm deg} {\rm NB}(i) = \frac{1}{|N_i|} \sum \nolimits _{j \in N_i} {\rm deg} _{\rm all}(j) \end{aligned}$$
(69)

Weighted average neighbor degree: The weighted average neighbor degree \({\rm NB}(i)\) of node i is

$$\begin{aligned} {\rm deg} {\rm NB}^{\overrightarrow{w}}(i) = \frac{1}{|N_i|} \sum \limits _{j \in N_i} {\rm deg} ^w_{\rm all}(j) \end{aligned}$$
(70)

Variation coefficient of neighbor degree:

$$\begin{aligned} {\rm VC}(i) = \frac{ \sqrt{\frac{1}{|N_i|} \sum \nolimits _{j \in N_i} ({\rm deg} _{\rm all}(j) - {\rm deg} {\rm NB}(i))^2}}{{\rm deg} {\rm NB}(i)} \end{aligned}$$
(71)

The weighted case is analogue.

Locality index of node i \(({\rm Loc}(i))\): The locality index of node i is the fraction of edges adjacent to nodes in \(N^{+}_i\) whose source and target lie in \(N^{+}_i\).

$$\begin{aligned} {\rm Loc}(i) = \frac{ \sum \nolimits _{j \in N^{+}_i} \sum \nolimits _{\begin{subarray}{l} k \in N^{+}_i \\ k \ne j\end{subarray}}a_{jk}}{\sum \nolimits _{j \in N^{+}_i} \sum \nolimits _{\begin{subarray}{l} k \in N \\ k \ne j \end{subarray}}a_{jk}} \end{aligned}$$
(72)

The weighted case is analogue. A value of 0 means that the node is isolated. The larger the value, the less edges connect the neighborhood of i to outside node. The maximum of one is reached if the neighborhood of i is not connected to outside nodes.

Closeness centrality out \({\rm CC}^{\rm out}(i)\): The closeness centrality out with indices of nodes from which node i can be reached (\({\rm RN}^{\rm OUT}(i)\))\({\rm RN}^{\rm OUT}(i)=\lbrace j \in N \backslash \{i\} | d (i,j) < \infty \rbrace \)

$$\begin{aligned} {\rm CC}^{\rm OUT}(i) = \frac{|{\rm RN}^{\rm OUT}(i)|}{\sum \nolimits _{j\in {\rm RN}^{\rm OUT}(i)} d(i,j)} \end{aligned}$$
(73)

Closeness centrality in \({\rm CC}^{\rm in}(i)\): The closeness centrality in with indices of nodes which can be reached from node i (\({\rm RN}^{\rm IN}(i)\))\({\rm RN}^{\rm IN}(i)=\lbrace j \in N \backslash \{i\} | d (j,i) < \infty \rbrace \)

$$\begin{aligned} {\rm CC}^{\rm IN}(i) = \frac{|{\rm RN}^{\rm IN}(i)|}{\sum \nolimits _{j\in {\rm RN}^{\rm IN}(i)} d(j,i)} \end{aligned}$$
(74)

Betweenness centrality BC:

$$\begin{aligned} {\rm BC}(i)= \frac{1}{(n-1)(n-2)} \cdot \sum \limits _{\begin{subarray}{l} j,k \in N \backslash \{i\}\end{subarray}} \frac{\rho _{j,k}(i)}{\rho _{j,k}} \end{aligned}$$
(75)

where \(\rho _{j,k}\) is the number of shortest paths from j to k and \(\rho _{j,k}(i)\) is the number of shortest paths from j to k that pass through i. The directed and weighted definitions are the same.

Knotty centrality KC: Let \(\hat{N} \subseteq N\), \(|\hat{N}| > 1\) be a subset of N. Then

$$\begin{aligned} {\rm KC}(\hat{N})= {\rm Ld}(\hat{N}) \cdot \frac{\sum \nolimits _{i \in \hat{N}} {\rm BC}(i)}{\sum \nolimits _{i \in N} {\rm BC}(i)} \end{aligned}$$
(76)

with the line density \({\rm Ld}(\hat{N})\) of the subgraph \(\hat{N}\). The knotty center of a graph G is a subset \(N_{\rm KC}\) of nodes with

$$\begin{aligned} {\rm KC}(N_{\rm KC})= \min _{\begin{subarray}{l} \hat{N} \subseteq N \\ |\hat{N}| > 1 \end{subarray}} \{ {\rm KC}(\hat{N}) \} \end{aligned}$$
(77)

\({\rm KC}(N_{\rm KC})\) is called the knotty-centerdness of the graph G.

The knotty-centrality of a node i is defined as

$$\begin{aligned} {\rm KC}(i)= {\left\{ \begin{array}{ll} 1, &{} i \in N_{\rm KC}\\ 0, &{} \hbox {else} \end{array}\right. } \end{aligned}$$
(78)

Stress S:

$$\begin{aligned} S(i) = \sum \limits _{ j,k \in N \backslash \{i\}} \rho _{j,k}(i) \end{aligned}$$
(79)

The directed and weighted definitions are the same.

Central point distance CPD:

$$\begin{aligned} {\rm CPD} = \frac{1}{n-1} \sum \limits ^n_{i=1} \frac{{\rm BC}_{\max } - {\rm BC}(i)}{{\rm BC}_{\max }} \end{aligned}$$
(80)

where \( {\rm BC}_{\max }=\max\nolimits_{i \in N} \lbrace {\rm BC}(i) \rbrace \) is the maximum Betweenness centrality. The directed and weighted versions use the directed and weighted Betweenness centralities.

Participation coefficient: The partition \(M=\{M_1,\ldots M_m\}\) is generated as described in the definition of modularity.

$$\begin{aligned} {\rm PC}^{\rightarrow }_x(i) = 1 - \sum \limits _{M_j \in M} \left( \frac{{\rm deg} _x(i,M_j)}{{\rm deg} _x(i)} \right) ^2 \end{aligned}$$
(81)

with \(x \in \{\rm in, out, all\}\) and

$$\begin{aligned} {\rm deg} _{\rm in}(i, M_j)= \sum \limits _{k \in M_j \backslash \{i\}} a_{ki} \end{aligned}$$
(82)

(Number of edges from vertices of \(M_j\) to i).

$$\begin{aligned} {\rm deg} _{\rm out}(i, M_j)= \sum \limits _{k \in M_j \backslash \{i\}} a_{ik} \end{aligned}$$
(83)

(Number of edges from i to vertices of \(M_j\)).

$$\begin{aligned} {\rm deg} _{\rm all}(i, M_j)= \sum \limits _{k \in M_j \backslash \{i\}} (a_{ik}+a_{ki}) \end{aligned}$$
(84)

(Number of edges between i and vertices of \(M_j\)).

$$\begin{aligned} {\rm PC}^{\overrightarrow{w}}_x(i) = 1 - \sum \limits _{M_j \in M} \left( \frac{{\rm deg} ^w_x(i,M_j)}{{\rm deg} ^w_x(i)} \right) ^2 \end{aligned}$$
(85)

with the same x and weighted definitions of degrees. One has \(0 \le {\rm PC}(i) \le 1\). If \({\rm PC}(i)=1\), the node i has no edges (in, out, all). If \({\rm PC}(i)=0\) all edges (in, out all) come from, go to or stay in the same cluster. The larger \({\rm PC}(i)\) the more clusters are involved in the edges of node i.

Z score/within module degree: Let \(M_i\) be the module containing node i. \({\rm deg} _x(i, M_i)\) \(x\in \{\rm in, out, all\}\) is defined in the participation coefficient.

$$\begin{aligned} \overline{{\rm deg} }_x(M_i)=\frac{1}{|M_i|} \sum \limits _{j \in M_i} {\rm deg} _x(j, M_i) \end{aligned}$$
(86)

is the mean and

$$\begin{aligned} \sigma _{{\rm deg} _x(M_i)} = \sqrt{\frac{1}{|M_i|} \left( \sum \limits _{j \in M_i} {\rm deg} _x(j, M_i) - \overline{{\rm deg} }_x(M_i) \right) ^2} \end{aligned}$$
(87)

the standard deviation of the within module \(M_i\) degree distribution. Then the Z score is defined as

$$\begin{aligned} Z^\rightarrow _x(i) = \frac{{\rm deg} _x(i, M_i) - \overline{{\rm deg} }_x(M_i)}{\sigma _{{\rm deg} _x(M_i)}} \end{aligned}$$
(88)

and analogous

$$\begin{aligned} Z^{\overrightarrow{w}}_x(i) = \frac{{\rm deg} ^w_x(i, M_i) - \overline{{\rm deg} ^w}_x(M_i)}{\sigma _{{\rm deg} ^w_x(M_i)}} \end{aligned}$$
(89)

with the weighted versions of the mean and standard deviation. A value above one or below minus one implies that a node has significantly more or less edges from, to or from and to nodes in its cluster than the average node in its cluster has.

Eigenvector centrality: The eigenvector centrality \({\rm EC}(i)\) is the i-th component of the eigenvector with the largest corresponding eigenvalue of the adjacency matrix resp. weight matrix.

Shapley rating \(\phi \): The Shapley rating is a measure that provides information about the loss of connectivity following the removal of a node.

$$\begin{aligned} {\rm SR}(i)= \sum \limits _{\hat{N} \subseteq N \backslash \{i\}} (|{\rm SCC}(\hat{N} \cup \{i\})| - |{\rm SCC}(\hat{N})|) \cdot \frac{(n - |\hat{N}| -1)! \cdot |\hat{N}|! }{n!} \end{aligned}$$
(90)

where \({\rm SCC}(\hat{N})\) is the set of strongly connected components of \(\hat{N}\). The smaller the value is, the more important is the node in the sense of connectivity of the graph. Because of the exponential number of subsets, this parameter can be approximated for large networks, only.

Radiality: The radiality of a node Rad is a measure of the distance of a node to all other nodes. Nodes that have a small radiality have larger distances to other nodes than those with a greater radiality.

Input radiality \({\rm Rad}_{\rm in}\): The input radiality of a node \({\rm Rad}_{\rm in}\) is

$$\begin{aligned} {\rm Rad}_{\rm in}(i)= \frac{1}{n-1} \sum \limits_{\begin{subarray}{l} j \in N \\ d(j,i) < \infty\end{subarray}} {\rm Diam} + 1 - d(j,i) \end{aligned}$$
(91)

In the weighted case the weighted distances are used.

Output radiality \({\rm Rad}_{\rm out}\): The output radiality of a node \({\rm Rad}_{\rm out}\) is

$$\begin{aligned} {\rm Rad}_{\rm out}(i)= \frac{1}{n-1} \sum \limits_{\begin{subarray}{l} j \in N \\ d(i,j) < \infty\end{subarray}} {\rm Diam} + 1 - d(i,j) \end{aligned}$$
(92)

In the weighted case the weighted distances are used.

Centroid value Cen: With \(g_{\rm out}(i,j)= |\{k \in N | d(i,k) < d(j,k) < \infty \}|\) and \(g_{\rm in}(i,j)= |\{k \in N | d(k,i) < d(k,j) < \infty \}|\) which are the number of nodes closer to node i than to node j with regard to In- and Out-distance, the centroid value is defined in the following.

Output centroid value \({\rm Cen}_{\rm out}\):

$$\begin{aligned} {\rm Cen}_{\rm out}(i)= \min \{ g_{\rm out}(i,j) - g_{\rm out}(j,i) | j \in N \backslash \{\ i \}\} \end{aligned}$$
(93)

Input centroid value \({\rm Cen}_{\rm in}\):

$$\begin{aligned} {\rm Cen}_{\rm in}(i)= \min \{ g_{\rm in}(i,j) - g_{\rm in}(j,i) | j \in N \backslash \{\ i \}\} \end{aligned}$$
(94)

In the weighted case the weighted distances are used. A value \({<}0\) implies, that there exists a node that is closer to most other nodes. A value \({\ge }0\) implies, that this node is most central in the network. A value \({=}0\) implies, that there are more than one most central nodes.

Page rank centrality PRC: \({\rm PRC}(i)= r_i\) where r is the solution of the linear system

$$\begin{aligned} (I-\alpha \cdot A^T \cdot B) \cdot r = \frac{1}{n} (1-\alpha ) \cdot \left( \begin{array}{l} 1 \\ \vdots \\ 1 \end{array} \right) \end{aligned}$$
(95)

with the damping factor \(\alpha =0.85\), the identity matrix I and the diagonal matrix B, whereby

$$\begin{aligned} b_{ii}= {\left\{ \begin{array}{ll} \frac{1}{{\rm deg} _{\rm out}(i)}, &{} {\rm deg} _{\rm out}(i) > 0\\ 0, & \hbox {otherwise} \end{array}\right. } \end{aligned}$$
(96)

In the weighted case the weight matrix W is used instead of A and the weighted version \({\rm deg} ^{\overrightarrow{w}}_{\rm out}(i)\) of the outdegree.

Flow coefficient FC: Number of paths of length 2 between neighbors of a node i that pass node i divided by the maximum possible numbers of sub paths.

$$\begin{aligned} {\rm FC}(i)=\frac{1}{|N_i| \cdot (|N_i| - 1)} \cdot \sum \limits_{\begin{subarray}{l} j,k \in N_i \\ j \ne k\end{subarray}} a_{ji} \cdot a_{ik} \end{aligned}$$
(97)

In the weighted case we define the flow coefficient as the sum of weights of paths of length 2 between neighbors of a node i that pass node i divided by the maximum possible sum.

$$\begin{aligned} {\rm FC}^{\overrightarrow{w}}(i)=\frac{1}{2 \cdot |N_i| \cdot (|N_i| - 1)} \cdot \sum \limits _{\begin{subarray}{l} j \in N_i \\ w_{ji} >0 \end{subarray}} \sum \limits _{\begin{subarray}{l} k \in N_i \backslash \{j\} \\ w_{ik} >0 \end{subarray}} (w_{ji} + w_{ik}) \end{aligned}$$
(98)

Average flow coefficient FC: Number of paths of length 2 between neighbors of a node i that pass node i divided by the maximum possible numbers of sub paths.

$$\begin{aligned} {\rm FC}&= \frac{1}{n} \sum \limits _{i=1}^n {\rm FC}(i)\end{aligned}$$
(99)
$$\begin{aligned} {\rm FC}^{\overrightarrow{w}}&= \frac{1}{n} \sum \limits _{i=1}^n {\rm FC}^{\overrightarrow{w}}(i) \end{aligned}$$
(100)

Subgraph centrality SC:

$$\begin{aligned} {\rm SC}(i)&= \sum \limits _{k=0}^{\infty } \frac{(A^k)_{ii}}{k!}\end{aligned}$$
(101)
$$\begin{aligned} {\rm SC}^{\overrightarrow{w}}(i)&= \sum \limits _{k=0}^{\infty } \frac{(W^k)_{ii}}{k!} \end{aligned}$$
(102)

The subgraph centrality of the network is the average subgraph centrality of its nodes.

$$\begin{aligned} {\rm SC}&= \frac{1}{n} \sum \limits _{i=1}^n {\rm SC}(i)\end{aligned}$$
(103)
$$\begin{aligned} {\rm SC}^{\overrightarrow{w}}&= \frac{1}{n} \sum \limits _{i=1}^n {\rm SC}^{\overrightarrow{w}}(i) \end{aligned}$$
(104)

Undirected cyclic coefficient CyclC: The undirected cyclic coefficient as published by (Kim et al. 2005) Cyclic topology in complex networks.

$$\begin{aligned} {\rm CyclC}(i)=\frac{2}{|N_i| \cdot (|N_i| - 1)} \cdot \sum \limits_{\begin{subarray}{l} (j,k) \in N_i \times N_i \\ j \ne k\end{subarray}} \frac{1}{2+{\rm dist}_i (j,k)} \end{aligned}$$
(105)

With

$$\begin{aligned} {\rm dist}_i(j,k)= {\left\{ \begin{array}{ll} \hbox {length of the shortest path from } j \hbox { to } k \hbox { that does not contains } i,\\ \hbox {if such a path exists}\\ \infty , \hbox {otherwise} \end{array}\right. } \end{aligned}$$
(106)

Directed cyclic coefficient \({\rm CyclC}^\rightarrow \): A publication about the directed cyclic coefficient is unknown. The directed cyclic coefficient is implemented here as follows:

$$\begin{aligned} {\rm CyclC}^{\rightarrow }(i)=\frac{1}{|N^{\rm out}_i| \cdot |N^{\rm in}_i| - |N^{\rm out}_i \cap N^{\rm in}_i|} \cdot \sum \limits_{\begin{subarray}{l} (j,k) \in N^{\rm out}_i \times N^{\rm in}_i \\ j \ne k\end{subarray}} \frac{1}{2+{\rm dist}_i (j,k)} \end{aligned}$$
(107)

Directed weighted cyclic coefficient \({\rm CyclC}^{\overrightarrow{w}}\): A publication about the directed weighted cyclic coefficient is unknown. The directed weighted cyclic coefficient is implemented here as follows:

$$\begin{aligned} {\rm CyclC}^{\overrightarrow{w}}(i)=\frac{1}{|N^{\rm out}_i| \cdot |N^{\rm in}_i| - |N^{\rm out}_i \cap N^{\rm in}_i|} \cdot \sum \limits _{\begin{subarray}{l} (j,k) \in N^{\rm out}_i \times N^{\rm in}_i \\ j \ne k\end{subarray}} \frac{1}{w_{ij}+w_{ki}+{\rm dist}^w_i (j,k)} \end{aligned}$$
(108)

with \({\rm dist}^w_i(j,k)\) is the weighted version of \({\rm dist}_i(j,k)\) with the weighted path length.

Cyclic network coefficient \({\rm CyclC}^\rightarrow \): The cyclic coefficient of the network is the average cyclic coefficient of its nodes:

$$\begin{aligned} {\rm CyclC}^{\rightarrow }=\frac{1}{n} \sum ^n_{i=1} {\rm CyclC}^{\rightarrow }(i) \end{aligned}$$
(109)

Hubness and authoritativeness: A hub is a node that points to many authorities and an authority is a node that has numerous input connections from many hubs (Kleinberg 1999; Sporns et al. 2007). The hubness \({\rm Hub}(i)\) of a node i is:

$$\begin{aligned} {\rm Hub}(i)=\sum \limits_{\begin{subarray}{l} j \in N^{\rm out}_i\end{subarray}} {\rm Auth}(j) \end{aligned}$$
(110)

with the authoritativeness \({\rm Auth}(i)\)

$$\begin{aligned} {\rm Auth}(i)=\sum \limits _{\begin{subarray}{l} j \in N^{\rm in}_i\end{subarray}} {\rm Hub}(j) \end{aligned}$$
(111)

An iterative algorithm is used to calculate a fixed point of these equations.

Vulnerability V: The vulnerability V is the maximum relative decrease of the global efficiency removing a single node.

$$\begin{aligned} V= \max \limits_{i \in N} \left\{ \frac{{\rm GE} - {\rm GE}(i)}{\rm GE} \right\} \end{aligned}$$
(112)

where \({\rm GE}(i)\) is the global efficiency of the graph (\(N \backslash \{i\}, \{(j,k) \in E | j \ne i \ne k \}\)) that originates by removal of node i and all edges adjacent to i. The weighted version is analog using the weighted global efficiencies.

Random models

The following random graph models are compared to the real network of the intrinsic amygdala connectivity. By comparing the average path length and the cluster coefficient of the models with the real network it is feasible to determine a model that is most similar to the real network.

Erdös Rényi graph:

$$\begin{aligned} G(n,p) \end{aligned}$$
(113)

where n is the number of vertices and p is the probability that an edge \((i, j)\) exists, for all \(i, j\). The degree distribution of the Erdös Rényi random graph is binomial in terms of

$$\begin{aligned} P({\rm deg} (v)=k)=\left( {\begin{array}{c}n-1\\ k\end{array}}\right) p^k(1-p)^{n-1-k} \end{aligned}$$
(114)

Watts–Strogatz graph: The small-world model of Watts–Strogatz is a random graph generation model that provides graphs with small-world properties. The network (initially it has a non-random lattice structure) is build by linking each node to its \(\langle k \rangle \) closest neighbors using a rewiring probability p. Hence, an edge has the probability p that it will be rewired as a random edge. The number of rewired links can be estimated by:

$$\begin{aligned} pE=pN \langle k \rangle / 2 \end{aligned}$$
(115)

Barabasi–Albert graph: The Barabasi–Albert graph is used to generate preferential attachments between nodes. The probability \(p_i\) that the new node is connected to node i is

$$\begin{aligned} p_i=\frac{k_i}{\sum _j k_j} \end{aligned}$$
(116)

The degree distribution of a Barabasi–Albert network is scale free following the power law distribution of the form:

$$\begin{aligned} P(k) \sim k^{-3} \end{aligned}$$
(117)

Eipert graph: The modified Eipert model (EN: Eipert network) is based on the Barabasi–Albert graph. However, the algorithm starts at a fixed number of nodes and edges are added iteratively.

Ozik–Hunt–Ott graph: The Ozik–Hunt–Ott model (OHO) (Ozik et al. 2004) is a small-world randomization approach that was modified for directed networks and a fixed number of edges. The OHO-model uses a growing mechanism in which all connections are made locally to topographical nearby regions.

Rewiring graph: The rewiring-models connects each target of an edge of a network to another target node.

Power law:

$$\begin{aligned} P(k) = \alpha \cdot k^{-\gamma } \end{aligned}$$
(118)

\(\Delta \) is the deviation (error) of an empirical distribution of degrees from the power law function. A small \(\Delta \) value means that the empirical distribution is similar with the power law function.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Schmitt, O., Eipert, P., Kettlitz, R. et al. The connectome of the basal ganglia. Brain Struct Funct 221, 753–814 (2016). https://doi.org/10.1007/s00429-014-0936-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00429-014-0936-0

Keywords

Navigation