Pharmacopsychiatry 2010; 43: S67-S81
DOI: 10.1055/s-0030-1252058
Original Paper

© Georg Thieme Verlag KG Stuttgart · New York

Network Effects of Synaptic Modifications

H. Liljenström1 , 2
  • 1Biometry and Systems Analysis Group, Energy & Technology, SLU, Uppsala, Sweden
  • 2Agora for Biosystems, Sigtuna, Sweden
Further Information

Publication History

Publication Date:
18 May 2010 (online)

Abstract

In this paper, we use computational models of varying complexity to investigate the role of synaptic modifications for cortical network properties. In particular, we study how the dynamics can be regulated by neuromodulators, intrinsic noise and chemical agents. We focus on the complex neurodynamics and its modulation, and how this is related to the neural circuitry, where connectivity enhancement and pruning is considered. The emphasis is on the overall network structures, with feedforward and feedback loops between excitatory and inhibitory neurons at several layers and distances, and less details at the synaptic level. Our models aim at linking processes at a molecular and cellular (microscale), with processes at a network level (mesoscale), which in turn are linked to the mental processes and cognitive functions (macroscale). We also discuss the relevance of these results for clinical and experimental neuroscience, with applications to learning, memory, arousal, and mental disorders.

References

  • 1 Altar CA, Laeng P, Jurata LW. et al . Electroconvulsive seizures regulate gene expression of distinct neurotrophic signaling pathways.  Neuroscience. 2004;  24 2667-2677
  • 2 Amit DJ, Tsodyks MV. Quantitative study of attractor neural networks retrieving at low spike rates: II. Low-rate retrieval in symmetric networks.  Network. 1991;  2 275-294
  • 3 Arbib MA. ed The handbook of brain theory and neural networks.. Cambridge, MA: MIT Press; 1995
  • 4 Arbib MA, Erdi P, Szentagothai J. Neural organization – Structure, function and dynamics.. Cambridge, MA: MIT Press; 1997
  • 5 Århem P, Liljenström H. Fluctuations in neural systems: From subcellular to network levels.. In: Moss F, Gielen S, editors Handbook of Biological Physics. 2001. 4: 83-129
  • 6 Århem P, Klement G, Blomberg C. Channel density regulation of firing patterns in a cortical neuron model.  Biophys J. 2006;  90 4392-4404
  • 7 Babloyantz A, Lourenco C. Brain chaos and computation.  Intl J Neural Systems. 1996;  7 461-471
  • 8 Benefiel-Kunkel AC, Greenough WT. Effects of experience and environment upon the developing and mature brain: Implications for laboratory animal housing.  ILAR Journal. 1998;  39 5-11
  • 9 Beyer JL, Weiner RD, Glenn MD. Electroconvulsive therapy.. Washington, DC: American Psychiatric Press; 1998
  • 10 Biedenbach MA. Effects of anesthetics and cholinergic drugs on prepyriform electrical activity in cats.  Exp Neurol. 1966;  16 464-479
  • 11 Bland BH. The physiology and pharmacology of hippocampal formation theta rhythms.  Prog Neurobiol. 1986;  26 1-54
  • 12 Bressler SL, Freeman WJ. Frequency analysis of olfactory system EEG in cat, rabbit, and rat.  Electroenceph Clin Neurophysiol. 1980;  50 19-24
  • 13 Crick F, Koch C. Towards a neurobiological theory of consciousness.  Seminars in the Neurosciences. 1990;  2 263-275
  • 14 Culic M, Gjoneska B, Hinrikus H. et al . Signatures of depression in non-stationary biometric time series.  Computational intelligence and neuroscience. 2009;  1-7 Article ID 989824, doi: 10.1155/2009/989824
  • 15 Eckhorn R, Bauer R, Jordan W. et al . Coherent Oscillations: A Mechanism of feature linking in the visual cortex.  Biol Cybern. 1988;  60 121-130
  • 16 Freeman WJ. Mass Action in the Nervous System.. New York: Academic Press; 1975
  • 17 Freeman WJ. Spatial properties of an EEG event in the olfactory bulb and cortex.  Elect Clin Neurophys. 1978;  44 586-605
  • 18 Freeman WJ. Nonlinear gain mediating cortical stimulus-response relations.  Biol Cybern. 1979;  33 237-247
  • 19 Freeman WJ. Neurodynamics – An exploration in mesoscopic brain dynamics.. Springer-Verlag; 2000
  • 20 Freeman WJ, Skarda CA. Spatial EEG patterns, non-linear dynamics and perception: the neo-Sherringtonian view.  Brain Res Rev. 1985;  10 147-175
  • 21 Freeman WJ. Societies of Brains – A Study in the Neuroscience of Love and Hate.. Hillsdale, NJ: Lawrence Erlbaum; 1995
  • 22 Gordon E. Editor Integrative neuroscience: Bringing together biological psychological and clinical models of the human brain.. Harwood Academic Press; 2000
  • 23 Gray CM. Synchronous oscillations in neuronal systems: mechanisms and function.  J Computat Neurosci. 1994;  1 11-38
  • 24 Gray CM, König P, Engel AK. et al . Oscillatory Responses in Cat Visual Cortex Exhibit Inter-columnar Synchronization Which Reflects Global Stimulus Properties.  Nature. 1989;  338 334-337
  • 25 Gu Y, Halnes G, Liljenström H. et al . A cortical network model for clinical EEG data analysis.  Neurocomputing. 2004;  58–60 1187-1196
  • 26 Gu Y, Wahlund B, Liljenström H. et al . Analysis of phase shifts in clinical EEG evoked by ECT.  Neurocomputing. 2005;  65–66 475-483
  • 27 Gu Y, Halnes G, Liljenström H. et al . Modelling ECT effects by connectivity changes in cortical neural networks.  Neurocomputing. 2006;  69 1341-1347
  • 28 Haberly LB, Bower JM. Olfactory cortex: model circuit for study of asociative memory.  Trends in Neurosci. 1989;  12 258-264
  • 29 Hagan J, Morris R. The cholinergic hypothesis of memory: A review of the animal experiments.. In: Iversen LL, Snyder SI, editors Psychopharmacology of the ageing nervous system. New York: Plenum Press; 1989: 237-324
  • 30 Halnes G, Liljenström H, Århem P. Density dependent neurodynamics.  BioSystems. 2007;  89 126-134
  • 31 Hasselmo M E. Acetylcholine and learning in a cortical associative memory.  Neural Computation. 1993;  5 32-44
  • 32 Hasselmo ME, Anderson BP, Bower JM. Cholinergic modulation of cortical associative memory function.  J Neurophysiol. 1992;  67 1230-1246
  • 33 Haykin S. Neural networks – A comprehensive foundation.. New York: Macmillan; 1999
  • 34 Hebb DO. The organization of behavior.. New York: John Wiley; 1949
  • 35 Hodgkin AL, Huxley AF. A quantitative description of membrane current and its application to conduction and excitation in nerve.  J Physiol. 1952;  117 500-544
  • 36 Hopfield JJ. Neural networks and physical systems with emergent collective computational abilities.  Proc. Natl Acad Sci USA. 1982;  79 2554-2558
  • 37 Hopfield JJ. Neurons with graded response have collective computational properties like those of two-state neurons.  Proc Natl Acad Sci USA. 1984;  81 3088-3092
  • 38 Huber MT, Braun HA, Krieg JC. Consequences of deterministic and random dynamics for the course of affective disorders.  Biol. Psychiatry. 1999;  46 256-262
  • 39 Huber MT, Braun HA, Krieg JC. Effects of noise on different disease states of recurrent affective disorders.  Biol Psychiatry. 2000;  47 634-642
  • 40 Ivanco TL, Greenough WT. Physiological consequences of morphologically detectable synaptic plasticity: Potential uses for examining recovery following damage.  Neuropharmacology. 2000;  39 765-776
  • 41 Johansson S, Århem P. Single-channel currents trigger action potentials in small cultured hippocampal neurons.  Proc Natl Acad Sci USA. 1994;  91 1761-1765
  • 42 Koch C. Biophysics of computation – Information processing in single neurons.. Oxford: Oxford University Press; 1999
  • 43 Koch C, Segev I. Methods of neuronal modeling: From synapses to networks.. Cambridge, MA: MIT Press; 1989
  • 44 Liljenström H. Modeling the dynamics of olfactory cortex using simplified network units and realistic architecture.  Int J Neural Systems. 1991;  2 1-15
  • 45 Liljenström H. Autonomous learning with complex dynamics.  Intl J Intelligent Systems. 1995;  10 119-153
  • 46 Liljenström H. Global effects of fluctuations in neural information processing.  Intl J Neural Systems. 1996;  7 497-505
  • 47 Liljenström H. Cognition and the efficiency of neural processes.. In: Århem P, Liljenström H, Svedin U, editors Matter matters? On the material basis of the cognitive activity of mind. Heidelberg: Springer; 1997: 177-213
  • 48 Liljenström H. Neural stability and flexibility – A computational approach.  Neuropsychopharmacology. 2003;  28 S64-S73
  • 49 Liljenström H, Århem P. Investigating amplifying and controlling mechanisms for random events in neural systems.. In: Bower JM, editor Computational Neuroscience New York: Plenum Press; 1997: 711-716
  • 50 Liljenström H, Hasselmo ME. Cholinergic modulation of cortical oscillatory dynamics.  J Neurophysiol. 1995;  74 288-297
  • 51 Liljenström H, Wu X. Noise-enhanced performance in a cortical associative memory model.  Intl J Neural Systems. 1995;  6 19-29
  • 52 Madison DV, Lancaster B, Nicoll RA. Voltage clamp analysis of cholinergic action in the hippocampus.  J Neurosci. 1987;  7 733-741
  • 53 Malenka RC, Nicoll RA. NMDA-receptor-dependent synaptic plasticity: multiple forms and mechanisms.  Trends Neurosci. 1993;  16 521-527
  • 54 Rowntree C, Bland BH. An analysis of cholinoceptive neurons in the hippocampal formation by direct microinfusion.  Brain Res. 1986;  362 98-113
  • 55 Saugstad LF. Age at puberty and mental illness.  British Journal of Psychiatry. 1989;  155 536-544
  • 56 Saugstad LF. Mental illness and cognition in relation to age puberty: a hypothesis.  Clinical Genetics. 1989;  36 156-167
  • 57 Saugstad LF. The maturational theory of brain development and cerbral excitability in the multifactorially inherited manic-depressive psychosis and schizophrenia.  International Journal of Psychophysiology. 1994;  18 189-203
  • 58 Shepherd GM. The synaptic organisation of the brain.. Oxford: Oxford University Press; 1998
  • 59 Sirosh J, Miikkulainen R. Self-organizing feature maps with lateral connections: modeling ocular dominance.. In Boulder, CO: 1993 Connectionist Models Summer School; 1994
  • 60 Skarda CA, Freeman WJ. How brains make chaos in order to make sense of the world.  Brain and Behavioral Sciences. 1987;  10 161-195
  • 61 Steyn-Ross DA, Steyn-Ross M. Editors Modeling phase transitions in the brain. Springer Series in Computational Neuroscience.. New York: Springer; 2010
  • 62 Szentagothai J. Local neuron circuits of the neocortex.. In: Schmitt FO, Worden FG, editors The Neurosciences 4th Study Program Cambridge, MA: MIT Press; 1979: 399-415
  • 63 Tateno T, Harsch A, Robinson HP. Threshold firing frequency-current relationships of neurons in rat somatosensory cortex: type 1 and type 2 dynamics.  J Neurophysiol. 2004;  92 2283-2294
  • 64 Tateno T, Robinson HP. Rate coding and spike-time variability in cortical neurons with two types of threshold dynamics.  J Neurophysiol. 2006;  95 2650-2663
  • 65 Thimm G, Fiesler E. Evaluating pruning methods.  Proceedings of the International Symposium on Artificial neural networks. 1995;  18-20
  • 66 Tsuda I. Chaotic itinerancy as a dynamical basis of hermeneutics in brain and mind.  World Futures. 1991;  32 167-184
  • 67 Wahlund B, von Rosen D. ECT of major depressed patients in relation to biological and clinical variables: a brief overview.  Neuropsychopharmacology. 2003;  28 S21-S26
  • 68 Wahlund B, Piazza P, von Rosen D. et al . Seizure (ictal) EEG characteristics in subgroups of depressive disorder in patients receiving electroconvulsive therapy (ECT) – A preliminary study and multivariate approach.  Computational Intelligence and Neuroscience 2009; Vol.. 2009;  Article ID 965209 DOI: doi:10.1155/2009/965209
  • 69 Wright JJ, Liley DTJ. Dynamics of the brain at global and microscopic scales. Neural networks and the EEG.  Behavioral and Brain Sciences. 1996;  19 285-320

Appendix A

Equations for the paleocortical network model with population units

The time evolution for a network of N neural units is given by a set of coupled nonlinear first order differential delay equations for all the N internal states, u. With external input, I(t), characteristic time constant, τi , and connection weight w ij between units i and j, separated with a time delay δ ij , we have for each unit activity, u i ,

Zoom Image

The input-output function, g i (u i ), is a continuous sigmoid function, experimentally determined by Freeman [18]:

Zoom Image

The gain parameter Q i determines the slope, threshold and amplitude of the curve for unit i. This gain parameter is associated with the level of arousal, or alternatively, the level of any particular neuromodulator. C is a normalization constant. Neuronal adaptation is implemented as an exponential decay of the output, proportional to the time average of previous output [50]. In such cases, C is not a constant, but instead denotes the adaptation function, and the input–output relation becomes,

Zoom Image

where <> T denotes the time average over the last T ms, and α is an adaptation parameter that is under neuromodulatory (cholinergic) control. Noise, or spontaneous neural activity is represented by a Gaussian noise function, ξ t (t), such that <ξ t (t)≥and <ξ t (t)ξ t (s)2Aδ(t–s).

The connection weights w ij are initially set and constrained by the general connectivity principles that have evolved for the three-layered cortex. To allow for learning and associative memory, the connection weights, w ij are incrementally changed, according to a learning rule of Hebbian type [34], adapted for the system dynamics. It takes into account that there is a conduction delay, δ ij , between the output (presynaptic) activity of one network unit and its (postsynaptic) effect on the receiving unit. With learning rate η the change in connection weight between unit j and i is given by

Zoom Image

where wmax is the maximum strength of an intrinsic synaptic connection.

Appendix B

Equations for the network model with Frankenhaeuser-Huxley neurons

The model neuron is described by the Frankenhaeuser-Huxley equations, with parameters based on recordings from small hippocampal interneurons [41]. The FH-model deviates slightly from the classical Hodgkin-Huxley formalism, since the FH-equations are based on recordings from vertebrate axons, whereas the HH-equations are based on recordings from squid axons. The FH-model is based on four differential equations:

Zoom Image

where v is the membrane potential, and m, h and n are permeability variables for P Na activation (m), P Na inactivation (h) and P K activation (n). The variation in the membrane potential is due to the stimulation current (I S ), the leak current (I L ), the initial transient current through Na channels (I Na ) and the delayed sustained current through K channels (I K ) divided by the membrane capacitance (C M ). The α's and, β's are rate constants for m, h, and n, as indicated by the suffices. A m is the membrane area, R the gas constant, F is Faraday's constant and T the absolute temperature. [Na] o and [Na] i are Na concentrations. Parameter values and the equations for I Na , I K , I L are given in [41] and in [6]. In our simulations, the only free parameters for the neuronal model are P* Na and P* K , the permeability values for fully open ion channels, also referred to as the channel densities. For simplicity, we skip the dimensions when referring to the permeability constants (or channel densities), always giving the values in units of 10−6 m/s.

Our network model consists of 6×6 neurons, arranged in a square lattice and connected in an all-to-all manner. We use a distance-dependent connectivity, with the connection strength decreasing with distance as w∼1/r (although in reality such effects may be negligible at this scale). In the cases when inhibition is included, we let six (out of 36) homogenously distributed network neurons be inhibitory (with periodical boundary conditions), relating to the fact that about 20% of the neocortical neurons in the mammalian brain are inhibitory.

We have previously studied the dynamical effects of changing the balance between excitation and inhibition in cortical neural networks [25]. In the current work, we also study the effects that different neuronal types (in terms of channel densities) could have on the network activity, by letting excitatory and inhibitory neurons having different properties. For simplicity, we let the post-synaptic activity be the same for inhibition as for excitation, except for opposite signs.

There is a synaptic delay of 1 ms in the signal transfer, while delays due to axonal propagation are neglected for the short distances considered here. The signal transfer is modelled in an all-or-nothing manner. A neuron is considered to elicit an action potential when the sodium activation exceeds a certain threshold (m>0.6). The effect of an action potential is a stereotyped postsynaptic current that is weighted and given as input to the receiving neurons. For the postsynaptic current, we follow the approach of Gerstner (2000), using an exponential pulse. A single action potential (f) elicited by neuron j thus results in a postsynaptic input to neuron i, as given by the equations:

Zoom Image

where I (f) ij (t) is the input to neuron i from neuron j at time t, as resulting from the action potential at time t (f) . t syn (1 ms) is the synaptic delay, and τ syn is the synaptic (membrane) time constant (30 ms). The connection weight (w ij ) for the signal transfer to neuron i from neuron j decreases proportionally with the distance (d ij ) between the neurons (normalized by use of the nearest neighbour distance d 0 =0.2 μm). w ij is a measure of the efficacy of the synapse between the neurons i and j, and has the unit of charge. Since the remaining factor of equation (6) is normalized, w ij can be identified with the charge deposited on the capacitor by a single presynaptic pulse from neuron j. q (=2.3*10−10 C) is just a scaling constant, chosen in order to have w=1 as a unitless reference. The weight factor, w, is a key parameter in our simulations, since it is regarded as a global measure of connection strength. At w=1, the initial postsynaptic current is about 7.75 pA, which is also a typical value for the stimulation current. Since the synaptic input is decaying with distance, it normally requires several AP inputs in order to generate an AP in the receiving neuron.

When a presynaptic neuron oscillates, the postsynaptic contributions from preceding action potentials will simply add up in the inputs. The total synaptic input to a neuron i will be the sum over action potentials (f) of all the neurons (j) that give an input to neuron i:

Zoom Image

The synaptic input will enter the single neural model as an additional input current. Equation (1) thus becomes:

Zoom Image

In the network simulations, we also include a white noise term (I G ), using a sampling rate of 0.1 ms and drawing the noise amplitude from a Gaussian distribution, with mean 0 and a variance equal to 5pA.

Appendix C

Equations for the neocortical network model with FitzHugh-Nagumo units

In our models of ECT EEG, the network dynamics are described by Eq. (1). The neurons are modelled as continuous output units of a Fitzhugh-Nagumo type, described by Eq. (2) and Eq. (3), which are essentially the same as in Ref. [4].

Zoom Image

In Equations (1–3), n is the number of neurons, ui is the postsynaptic potential of neuron i, vi is the membrane potential at the axon initial segment, and wi is an auxiliary variable. a, b, and c are positive constants, which guarantee the existence of an oscillatory interval.

Zoom Image

is a monotonically increasing (α>0), nonnegative (0≤mg<Mg ) function that describes the relation between the pre- and postsynaptic potentials of the neurons. cik describes the topology of the network. p + and p stand for the excitatory and inhibitory connection strengths. The neurons have time-constants τ ex and τ in . The total time delay, T ik , is due to a synaptic delay, Ts=1 ms, and the time a signal takes to travel through the axonal and dendritic trees between neuron k to neuron i with the propagation velocity 0.5 m/s. γ i denotes the synaptic membrane conductance of neuron i. We take the mean membrane potential of all excitatory neurons in each layer in the system as a model of the EEG readout.

Correspondence

Prof. H. Liljenström

Biometry and Systems Analysis Group

Energy & Technology, SLU

Box 7032

750 07 Uppsala

Sweden

Email: hans.liljenstrom@et.slu.se

    >