Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

REVIEW ARTICLE

published: 13 August 2012


COMPUTATIONAL NEUROSCIENCE doi: 10.3389/fncom.2012.00058

Computational models of neuron-astrocyte interaction


in epilepsy
Vladislav Volman 1,2,3*, Maxim Bazhenov 4 and Terrence J. Sejnowski 1,2,5
1
Computational Neurobiology Laboratory, Howard Hughes Medical Institute, The Salk Institute for Biological Studies, La Jolla, CA, USA
2
Center for Theoretical Biological Physics, University of California at San Diego, La Jolla, CA, USA
3
L-3 Applied Technologies/Simulation, Engineering, and Testing, San Diego, CA, USA
4
Department of Cell Biology and Neuroscience, University of California at Riverside, Riverside, CA, USA
5
Division of Biological Sciences, University of California at San Diego, La Jolla, CA, USA

Edited by: Astrocytes actively shape the dynamics of neurons and neuronal ensembles by affecting
Vincenzo Crunelli, Cardiff University, several aspects critical to neuronal function, such as regulating synaptic plasticity,
UK
modulating neuronal excitability, and maintaining extracellular ion balance. These pathways
Reviewed by:
for astrocyte-neuron interaction can also enhance the information-processing capabilities
Boris Gutkin, Institut Pasteur, France
Upinder S. Bhalla, National Center of brains, but in other circumstances may lead the brain on the road to pathological ruin.
for Biological Sciences, India In this article, we review the existing computational models of astrocytic involvement in
*Correspondence: epileptogenesis, focusing on their relevance to existing physiological data.
Vladislav Volman, Computational
Neurobiology Laboratory, The Salk Keywords: astrocytes, regulatory signaling, epileptogenesis, trauma, inflammation
Institute for Biological Studies,
10010 North Torrey Pines Road,
La Jolla, CA 92037, USA.
e-mail: volman@salk.edu

INTRODUCTION for up to one third of brain mass (Kandel, 1991). Because of


There has been an increasing effort to understand the etiology of their important role in glucose metabolism and maintenance
brain disorders using computational models of the neural circuits of extracellular ion homeostasis, astrocytes have been viewed as
that can predict the impact of changes in intrinsic neuronal prop- “support cells,” secondary to neurons and having little to do with
erties and interneuronal coupling on the dynamics of neuronal brain activity per se. However, growing experimental evidence
networks. The ability of biophysically detailed models to address [reviewed in Seifert et al. (2006); Giaume et al. (2010); Halassa
working hypotheses for brain disorders is especially important and Haydon (2010)] points to the possibility that glial cells (and
given the methodological and ethical concerns associated with the astrocytes in particular) are actively involved in the modulation
experimental studies on humans. of synaptic transmission and neural excitability. This has led
Models based on the biophysical underpinnings of brain researchers to propose that astrocytes can dynamically redefine
pathologies, focusing on neurons, and neuronal networks, are the functional architecture of neuronal networks (Nedergaard
useful in determining how different intrinsic and network prop- et al., 2003).
erties can result in pathological activity. For example, in models of Unlike the fast millisecond time scale for action potential
cortical networks, an imbalance in the ratio of fast excitation and and fast chemical transmission, astrocytes are regulated mainly
inhibition can lead to epileptic seizures; however, this approach through calcium signaling on a much slower time scale of seconds
is primarily based on the relatively fast neuronal dynamics and to minutes, consistent with the slow time scale associated with
thus cannot explain how the slow (>days) transition from nor- homeostatic regulation of neuronal activity. Furthermore, astro-
mal to pathological state occurs. In contrast, some brain disorders cytes respond to chronic changes in neuronal activity (such as
(e.g., epilepsy and schizophrenia to provide few examples) may those incurred by trauma) by releasing various pro-inflammatory
result from aberrations in slow, homeostatic, mechanisms that molecules that offset the perturbation. Thus, astrocyte-neuron
modulate the function of single neurons and properties of synap- signaling appears to be a plausible setting to explain at least
tic plasticity/connectivity. Thus, though useful, models of fast part of the etiology of brain diseases linked to the aberrations in
neuronal dynamics provide only limited insight into slowly devel- homeostatic regulation.
oping pathological states of the brain, and a paradigm shift is This review attempts to survey existing computational mod-
needed for computational approaches to succeed in revealing the els of astrocytic involvement in disorders of the nervous sys-
mechanisms underlying brain pathologies. tem, with particular emphasis on models of epileptogenesis.
Neurons are critical for information processing in the human The dynamical signature of epilepsy is hypersynchronization of
brain, but they comprise less than half of brain mass, which collective neuronal activity, suggesting a breakdown in homeosta-
also includes glial cells. Glial cells are further subdivided sis. We also provide a perspective on what is needed to better
into microglia, oligodendrocytes, Schwann cell, and astrocytes. understand how these non-neuronal cells contribute to brain
Astrocytes are the most numerous glial cell type and account pathologies.

Frontiers in Computational Neuroscience www.frontiersin.org August 2012 | Volume 6 | Article 58 | 1


Volman et al. Computational models of neuron-astrocyte interaction in epilepsy

MODULATION OF SYNAPTIC TRANSMISSION tripartite synapses not only in epilepsy, but perhaps also in other
Although hyperexcitability of individual neurons can significantly diseases of the nervous system that are based on aberrant synaptic
contribute to epileptogenesis, epilepsy is widely considered to communication (Halassa et al., 2007).
be a “network disease” driven by aberrant synaptic interaction The first, to our knowledge, computational model that
between neurons (McCormick and Contreras, 2001). Synaptic explored the potential involvement of astrocytes in synaptic
plasticity (in particular short-term synaptic depression and facil- mechanisms of epileptogenesis incorporated the basic notion of
itation) can critically mold the functional strength of synaptic astrocytic “eavesdropping” on neuronal activity and assumed that
coupling; thus, it is possible that specific properties of synaptic the sole effect of neuronal activity-dependent astrocytic calcium
transmitter release could predispose a network to seizures. elevations was to promote neuronal depolarization (Nadkarni
Astrocytes are ideally positioned to modulate synaptic con- and Jung, 2003; Appendix). This study concluded that astrocytes
tributions to epileptogenesis. Experimental data [reviewed in promote epileptogenesis through positive feedback. Similar con-
Araque et al. (1999)] shows that astrocytic processes are often clusions were reached in another modeling study that showed
apposed to synaptic junctions, giving rise to the notion of “tri- that glutamate released from astrocytes could be responsible for
partite synapse” consisting of presynaptic bouton, the astrocyte, paroxysmal depolarization shifts, often associated with epilep-
and the postsynaptic density. Stimulation of synaptic boutons and tic activity (Silchenko and Tass, 2008). However, gliotransmit-
subsequent release of neurotransmitter elicits complex calcium ter released from astrocytes can either up- or down-regulate
responses in nearby astrocytes, showing that these glial cells can the release of glutamate from synaptic boutons (Araque et al.,
“eavesdrop” on synaptic activity [reviewed in Haydon (2001)]. 1998; Zhang et al., 2003), and computational modeling studies
Elevation of astrocytic calcium culminates in the regulated release suggested that such negative feedback modulation of synaptic
of molecules such as glutamate and/or ATP which, by diffusing transmission could account for the regulation of spontaneous
in the extracellular space (ECS) and binding to dedicated recep- paroxysmal activity in neural cultures (Volman et al., 2007).
tors, can modulate the synaptic transmission (Araque et al., 1998) Thus, in the context of the tripartite synapse, astrocytic signaling
and the excitability of adjacent neurons (Parpura and Haydon, could have either positive or negative impact on neuronal activ-
2000; Reyes and Parpura, 2008) (Figure 1). It was shown recently ity, or perhaps a combination of both. Revealing the mechanisms
that NMDA-R mediated astrocytic modulation of postsynaptic behind this diversity of modulating effects could help in under-
neuronal excitability can promote epileptogenesis in neuronal standing the conditions for seizure suppression or promotion by
networks (Gomez-Gonzalo et al., 2010). This suggests the role for astrocytes.

FIGURE 1 | Astrocytes regulate presynaptic transmission and concentration. Calcium causes a release of astrocytic glutamate (a process
postsynaptic excitability at glutamatergic synapse. Following successful termed gliotransmission) which can bind mGlu receptors on presynaptic
vesicular fusion, a fraction of released glutamate can spill over and reach membrane and/or NMDA receptors on postsynaptic side, thus
metabotropic glutamate (mGlu) receptors on the membrane of modulating the probability of synaptic transmitter release (PRELEASE )
adjacent astrocyte. Glutamate activation of astrocytic mGlu receptors and/or postsynaptic excitability. For more quantitative details regarding the
activates a cascade of biochemical events that culminates in elevation of process of astrocyte modulation of synaptic transmission, see De Pitta et al.
inositol trisphosphate (IP3) and increased intra-astrocytic free calcium (2011).

Frontiers in Computational Neuroscience www.frontiersin.org August 2012 | Volume 6 | Article 58 | 2


Volman et al. Computational models of neuron-astrocyte interaction in epilepsy

Synaptic neurotransmitter release is a calcium-driven process; Unlike astrocytic sensing of synaptic activity (which can
thus, a change in presynaptic calcium levels through glial modula- in principle play a regulatory role in the flow of neural
tion (for example by astrocytic activation of presynaptic receptor information, as mentioned below), astrocytic regulation of
channels) could affect synaptic transmission. A computational extracellular ion concentration is mostly associated with neu-
modeling of a tripartite synapse supported the possibility that ropathology. Physiological experiments and theoretical argu-
astrocytes may optimize the synaptic transmission of informa- ments (Frankenhaeuser and Hodgkin, 1956) show that elevated
tion by affecting the levels of presynaptic calcium (Nadkarni extracellular potassium can significantly increase the reversal
et al., 2008). In another recent modeling study (De Pitta et al., potential of potassium channel conductance, thus rendering
2011), astrocytes either depressed or facilitated synaptic trans- potassium currents depolarizing. Cumulatively, these observa-
mission, which suggested that the ultimate effect of astrocytes on tions led to the “high K+ ” theory of epileptogenesis (Fisher
transmitter release probability was determined by the interplay et al., 1976; Lothman and Somjen, 1976), according to which
between the different mGlu-Rs (located presynaptically) and the activity-dependent elevation of extracellular potassium acts as
baseline synaptic vesicle release probability. Incorporating these a positive feedback mechanism promoting dynamical instabil-
pathways, the model was successful at explaining several contra- ity that culminates in seizure generation (Janigro et al., 1997)
dictory experimental observations regarding the role of astrocytes [reviewed in Frohlich et al. (2008b)]. Because astrocytes are
in the modulation of synaptic transmission. involved in setting potassium levels (Janigro et al., 1997), it is
imperative to understand how astrocytes contribute to “high K+ ”
REGULATION OF EXTRACELLULAR ION AND WATER LEVELS epileptogenesis.
Besides eavesdropping on synaptic activity, astrocytes also pos- Early computational models that explored the role of high
sess the means to “estimate” the level of neuronal activity in K+ in epileptogenesis (Kager et al., 2000; Bazhenov et al., 2004;
their proximity. The shape of a neuronal action potential is gov- Frohlich et al., 2006, 2010) focused on the dynamics of single
erned by ionic gradients across the membrane, but, in turn, neurons and neuronal networks and made simplistic assumptions
the action potential itself causes changes in local ion concentra- regarding the astrocytic mechanisms for potassium regulation
tions (Figure 2). The principal effect is a spike-activity-dependent (Appendix). More recent models have incorporated more diverse
accumulation of potassium ions in the ECS. Astrocytes maintain mechanisms of ion movement in single neurons (Cressman et al.,
potassium homeostasis by employing inward rectifying potas- 2009; Somjen et al., 2008; Krishnan and Bazhenov, 2011) and
sium channels to take up the excess K+ that accumulates fol- ion exchange between single neurons and astrocytes (Somjen
lowing action potential activity and redistributing K+ ions across et al., 2008; Oyehaug et al., 2011) (Figure 2). The interrelated
space compartments. Thus, astrocytes can monitor the level of dynamics of sodium and potassium currents during epilepto-
neuronal activity through changes in extracellular potassium genesis were examined in a network model that incorporated
concentration. both neuronal and glial cells (Ullah et al., 2009). Although,

FIGURE 2 | Astrocytes regulate extracellular ion concentrations and water channels and thus are in position to regulate the levels of extracellular potassium.
level. Neuronal excitability is defined in part by activity-dependent Nernst Through this regulation, astrocytes can also indirectly affect the levels of
potentials of sodium (Na+ ), potassium (K+ ), and chloride (Cl− ) ions. The intracellular chloride ions in neurons. The efficacy of extracellular potassium
intracellular and extracellular concentrations of these ions are critically regulation by astrocytes may be achieved by water-dependent regulation of
determined by the action of different ion transport mechanisms (sodium extracellular space, as explained in Text. For more quantitative details regarding
potassium pump and chloride co-transporters). Astrocytes possess potassium the process of ion level regulation by astrocytes, follow Ostby et al. (2009).

Frontiers in Computational Neuroscience www.frontiersin.org August 2012 | Volume 6 | Article 58 | 3


Volman et al. Computational models of neuron-astrocyte interaction in epilepsy

these studies highlighted different aspects of ion concentration multiple transitions between tonic and clonic states (Krishnan
regulation by astrocytes (pumps, transporters, etc.), they all con- and Bazhenov, 2011).
verged to the same conclusion—the failure of astrocytes to main- The impact of chloride on epileptogenesis primarily thus
tain proper ionic micro-environment in neuronal surrounding depends on the homeostasis of intracellular chloride concen-
can be critical for initiating epileptogenesis through the “high tration. Because astrocytes cannot directly access ion levels in
K+ ” mechanism. neurons, astrocytes are not expected to be involved in chloride
Several lines of evidence suggest that astrocytes could also mechanisms in epileptogenesis. However, chloride homeostasis
affect potassium-mediated epileptogenesis through regulation in neurons is affected by the opposing action of cation-chloride
of extracellular water levels (Simard and Nedergaard, 2004). co-transporters, NKCC1 (importing two ions of chloride along
Water movement in the brain is likely to involve aquaporin with one ion of sodium and one ion of potassium) and KCC2
4 (AQP4) channels, which are widely expressed in astrocytes (exporting one ion of chloride along with one ion of potas-
(Nagelhus et al., 2004). Interestingly, AQP4 channels in astro- sium) (Payne et al., 2003) (Figure 2). The relative expres-
cytes are co-localized with potassium Kir4.1 channels (Nagelhus sion of these different chloride co-transporters directly affects
et al., 2004). Furthermore, potassium clearance was significantly seizure generation (Glykys et al., 2009; Dzhala et al., 2010).
compromised in AQP4 knockout mice (Binder et al., 2006). In particular, KCC2 co-transporter can mediate an effective
In these AQP4−/− mice, the threshold for seizure generation coupling between extracellular potassium concentration and
was higher, but seizure duration was longer compared to con- intracellular chloride concentration: a higher extracellular potas-
trols (Amiry-Moghaddam et al., 2003; Binder et al., 2006). sium level leads to higher intracellular chloride concentration
This suggests that astrocytic regulation of extracellular water (Payne et al., 2003). Thus, a dysfunction in extracellular potas-
could contribute to epileptogenesis during intense neuronal sium uptake by astrocytes could indirectly facilitate the desta-
activity through the depolarization of astrocytes by the accu- bilizing effect of intracellular chloride on a neuron’s spiking
mulation of extracellular K+ ions. This depolarization would activity.
activate the electrogenic sodium bicarbonate cotransporter, favor-
ing the uptake of these ions by the astrocyte. The uptake of ASTROCYTES IN POST-TRAUMATIC EPILEPSY
sodium and bicarbonate ions by astrocytes would establish an Traumatic brain injury (TBI) (e.g., as a result of penetrative
osmolarity gradient (higher intracellular osmolarity) and thus wound) can increase the predisposition to epileptic seizures
would drive the water into glial cells through AQP4 channels, after a latent period following a traumatic event (Annegers
eventually contributing to astrocytic swelling and shrinkage of et al., 1998). Although there is a clear causal link between TBI
ECS. Activity-dependent shrinkage of ECS implies a smaller events and the later emergence of epileptic seizures, the etiol-
distribution volume for extracellular potassium, thus lowering ogy of “post-traumatic epilepsy” remains elusive (Agrawal et al.,
the seizure threshold. The implications of activity-dependent, 2006).
astrocyte-mediated shrinkage of ECS were recently explored in A possible immediate outcome of TBI is the massive death of
a computational model that incorporated a variety of astrocytic neuronal cells and damage to synaptic connectivity between neu-
ion exchange mechanisms and water regulation (Ostby et al., rons, which can create areas with chronic neuronal and synaptic
2009). inactivity. Evidence from in vitro studies suggests that chronic
Epileptogenesis has also been shown to be facilitated by the inactivity can modify several parameters of neuronal circuitry
intracellular accumulation of chloride in neurons (Dzhala et al., (e.g., synaptic connectivity, synaptic conductances, local neuronal
2005). The electrochemical gradient of chloride defines the rever- morphology, intrinsic neuronal excitability) to compensate for
sal potential of gamma-aminobutyric acid (GABA) receptor- the loss of activity incurred by trauma. As a general rule, chronic
mediated chloride currents, and thus determines the extent inactivity promotes the upregulation of depolarizing influences
to which GABAergic synaptic signaling (normally assumed to and downregulation of hyperpolarizing influences, while over-
have hyperpolarizing seizure-suppressing influence) is inhibitory. excitation leads to downregulation of depolarizing influences
During physiologically normal activity, the intracellular chlo- and upregulation of hyperpolarizing influences. This suggests
ride concentration is much lower than its extracellular levels, that homeostatic regulatory pathways may be activated after
keeping ECl low and thus ensuring hyperpolarizing effect of traumatic event to offset the perturbation in electrical activity
GABAergic synapse. The high levels of activity that occur dur- (Turrigiano, 1999). Regional homeostatic adjustments of neu-
ing seizures results in intracellular chloride accumulation and ronal excitability and synaptic transmission could contribute to
a higher GABA reversal potential, which can potentially lead breaching the excitation-inhibition balance (Fritschy, 2008), thus
to depolarizing GABA currents. In early development, intra- promoting seizure generation in traumatized networks (Timofeev
cellular chloride concentration is so high that GABA has a et al., 2010). Understanding the mechanisms of trauma-triggered
depolarizing effect on the neuronal membrane potential (Ben- homeostatic plasticity could therefore generate insights into the
Ari and Holmes, 2005). Increase of intracellular chloride con- etiology of post-traumatic epilepsy.
centration can also directly affect the resting potential and Several computational modeling studies have addressed the
excitability of neurons. Using model incorporating Na+ , K+ , role of homeostatic plasticity in post-traumatic epilepsy. In
and Cl− concentration dynamics, it was shown that an increase one model of trauma that included cell death, post-traumatic
of intracellular chloride concentration extends seizure duration axonal sprouting led to the emergence of seizure-like activity
making possible long-lasting epileptic activity characterized by in an otherwise non-seizing network (Santhakumar et al., 2004;

Frontiers in Computational Neuroscience www.frontiersin.org August 2012 | Volume 6 | Article 58 | 4


Volman et al. Computational models of neuron-astrocyte interaction in epilepsy

Schneider-Mizell et al., 2010). This is consistent with earlier could have important implications for the ability of astrocytes to
theories showing that network topology of synaptic connectiv- modulate synaptic function.
ity can critically determine its seizing propensity (Netoff et al., In experimental models of epilepsy, astrocytes undergo signif-
2004). In deafferentation models of cortical trauma (mimick- icant rapid morphological remodeling following the traumatic
ing loss of excitation and chronic reduction of neuronal activ- event (Oberheim et al., 2008), a process known as reactive
ity), homeostatic scaling of synaptic conductances could restore gliosis (Sofroniew, 2009). Specifically, astrocytes that are located
the “normal” firing rates, but this came at the expense of relatively close (200 microns) to the boundary between intact
paroxysmal bursting (Houweling et al., 2005; Frohlich et al., and injured cortical regions lose their trademark star shape and
2008a; Appendix). Interestingly, the rate of paroxysmal activ- elongate in the direction perpendicular to the trauma bound-
ity depended on both intrinsic neuronal parameters (Houweling ary (Oberheim et al., 2008). The implications of this remodel-
et al., 2005) and the spatial pattern of trauma (Volman et al., ing for network function remain largely unknown, but it was
2011b,c). Although these studies addressed different aspects of recently proposed that trauma-induced morphological reorgani-
post-traumatic reorganization, all supported the notion that zation of astrocytes could reduce the incidence of paroxysmal
trauma-triggered homeostatic regulation can contribute to post- activity by providing better functional segregation of synaptic
traumatic epileptogenesis. input from intact vs. injured neurons (Volman et al., 2011d).
Recent in vitro studies show that glial cells are critically Functional segregation of synaptic input could help localize
involved in the homeostatic scaling of synaptic conductances the release of TNFα to the regions of synaptic inactivity and
that follows prolonged synaptic inactivity (Beattie et al., 2002; thus prevent pathological over-excitation of relatively intact
Stellwagen and Malenka, 2006; Steinmetz and Turrigiano, 2010). areas.
Following chronic synaptic inactivity, astrocytes release tumor Modeling the microphysiology of astrocyte-neuron interac-
necrosis factor alpha (TNFα) which diffuses and acts on postsy- tions is a challenging problem without knowing the distribution
naptic neurons to scale up the number of AMPA/NMDA recep- of different proteins on the surfaces of astrocytes and the move-
tors and scale down the number of GABAa receptors (Figure 3). ment of ions across membranes. One promising approach is
Thus, signaling by TNFα could represent a homeostatic regula- based on Monte Carlo simulations of cellular microphysiology,
tory mechanism compensating for a chronic reduction in neural which can explore the impact of different astrocytic morphologies
excitability and could contribute to post-traumatic epileptogen- (Nadkarni et al., 2010).
esis. An observation in support of this hypothesis is that TNFα
is released by glial cells early (1–2 h) after trauma, as a part of NEURO-GLIAL NETWORKS
neuroinflammatory cascade. Computational models constructed Much of the information exchange between neurons is achieved
with variable “glia” to explore this hypothesis concluded that with fast chemical synapses (although some types of neurons
glial release and diffusion of TNFα could affect post-traumatic are also coupled by gap junctions). In contrast, astrocytes lack
paroxysmal activity (Savin et al., 2009). chemical synapses but are extensively coupled by gap junctions,
an “astrocytic syncytium” through which various substances
THE NEXT FRONTIER are transported (Scemes and Spray, 2004). Given the coupling
MODELING THE EFFECTS OF CELL MORPHOLOGY between astrocytic calcium and modulation of synaptic transmis-
In the first-generation computational models of astrocyte-neuron sion, and the role of synaptic communication in neuronal net-
interaction, astrocytes and neurons were usually assumed to be work dynamics, gap junction coupling between astrocytes could
point entities, neglecting completely their intracellular organi- affect pathological neuronal dynamics. Indeed, several studies
zation and dynamics. This simplifying assumption allowed the found that the levels of astrocytic gap junction protein are com-
modulation of basic neural functions by astrocytes to be explored. promised in experimental models of epilepsy (Lee et al., 1995;
However, the spatial aspects of cellular organization may also be Fonseca et al., 2002), but the exact implications of this aberration
important to understanding how glial cells affect brain patholo- are not clear yet.
gies. For example, the spatial co-localization of potassium and Experimentally, mice with deficient gap junctional coupling
water transport mechanisms in astrocytes is involved in the between astrocytes exhibit impaired potassium clearance and
regulation of traumatic edema and consequent post-traumatic higher extracellular potassium accumulation during synchro-
epilepsy, as mentioned above. nized neuronal firing (Wallraff et al., 2006); thus, astrocytic
Another elegant example of the importance of spatial organi- gap junctions could aid in potassium buffering. These early
zation is derived from energy considerations (Hertz et al., 2007). observations were re-validated in a recent study that also
The cell body of a typical astrocyte accounts for only ∼2% of revealed, in a hippocampal preparation, that gap junctions
its total volume, with larger, branching organelle-containing pro- between astrocytes are important regulators of synaptic trans-
cesses constituting ∼60% (Hertz et al., 2007). The remaining mission (Pannasch et al., 2011). These findings suggest that
∼40% are taken by tiny (∼3 microns length by ∼0.2 microns astrocytic networks might control the scale of activity in neu-
width) filopodia and lamellipodia which cannot accommodate ronal networks (and perhaps vice versa) through regulation
energy-efficient organelles such as mitochondria. This implies of synaptic transmission, ECS volume, and extracellular ion
that the terminal thin processes (which regulate synaptic func- homeostasis.
tion) and thicker organelle-containing branches use different Computational modeling of neuronal network dynamics sug-
energy strategies for their function (Hertz et al., 2007). This gests that gap junctions between neurons can either promote

Frontiers in Computational Neuroscience www.frontiersin.org August 2012 | Volume 6 | Article 58 | 5


Volman et al. Computational models of neuron-astrocyte interaction in epilepsy

FIGURE 3 | Astrocytes take part in neuronal homeostatic plasticity. In modulation aims to offset the decrease in neuronal activity by shifting the
response to prolonged neuronal and synaptic inactivity, as occurs following excitation-inhibition balance in favor of excitation. The participation of
trauma, astrocytes release tumor necrosis factor alpha (TNFα), which diffuses astrocytes in homeostatic plasticity can be further facilitated by slow (10s of
and binds to its dedicated receptors on pyramidal neurons causing a seconds) intercellular communication via gap junctions between these cells.
reduction in the number of postsynaptic GABA receptors and an increase in For more details regarding the role of astrocytes in homeostatic plasticity and
the number of postsynaptic AMPA and NMDA receptors. Such homeostatic intercellular communication, see Pannasch et al. (2011).

or suppress epileptic-like activity in a way that depends on gap junctions are likely to be important determinants of inter-
the topology and strength of the coupling and the level of cellular communication. This issue has been addressed study-
neuronal excitation. A surprising prediction of the model was ing the spatial characteristics of calcium waves in networks of
that an increase in gap junctions may reduce membrane input reconstructed astrocytes (Kang and Othmer, 2009). This study
resistance thus making a network of neurons less excitable confirmed that calcium waves in astrocytes are partially regen-
(Volman et al., 2011a). Whether or not the same principles erative; however, the model used 2-dimensional projections of
apply to calcium excitability in astrocytic networks is not clear. reconstructed astrocytes, and whether or not the same con-
Gap junctions are usually modeled as linear electrical coupling clusions hold in realistic 3-dimensional models remains to be
between a pair of cells that is constant, but in fact their per- shown.
meability to different ions can be modulated by several fac- Activation of astrocytic regulatory pathways, such as the
tors, including neural activity itself (Rouach et al., 2000). This release of gliotransmitters, is triggered by changes in neuronal
adds an additional level of complexity to intercellular signal- and synaptic activity. This in turn implies that the spatial dis-
ing between glial cells and implies that the “transfer function” tribution of neurons and the structure of synaptic connectivity
of a gap junction could depend non-linearly on the permeat- influence glial signaling. Indeed, neuronal activity can itself shape
ing ion. One modeling study addressed the role of non-linear the coupling properties of astrocytic syncytium (Rouach et al.,
gap junctional transport in the context of astrocytic calcium 2000). Existing network models of neuronal glial interaction, such
waves (Goldberg et al., 2010) but further studies that account for as the ring network model (Ullah et al., 2009), have circum-
their variable permeability and co-transport of different ions are vented this issue by assuming simplistic network topologies and
needed. connectivity. The implications of complex neuro-glial network
Challenges associated with realistic modeling of intercellu- topologies for normal and pathological dynamics remain to be
lar astrocytic communication depend on the microphysiology of studied.
the cell since calcium waves propagate intracellularly at a speed
of ∼15 microns/s but this can vary greatly throughout the cell ACKNOWLEDGMENTS
depending on the topology of endoplasmic reticulum in which Completion of this article was partially supported by NIH grants
most of the cellular calcium is stored (Yagodin et al., 1994); con- R01 NS059740 and R01 NS060870 and the Howard Hughes
sequently, the cellular morphology and the exact locations of Medical Institute.

Frontiers in Computational Neuroscience www.frontiersin.org August 2012 | Volume 6 | Article 58 | 6


Volman et al. Computational models of neuron-astrocyte interaction in epilepsy

REFERENCES PLoS Comput. Biol. 7:e1002293. doi: Glykys, J. C., Dzhala, V. I., Kuchibhotla, J. H. Schwartz, and T. M. Jessel
Agrawal, A., Timothy, J., Pandit, L., and 10.1371/journal.pcbi.1002293 K. V., Feng, G., Kuner, T., Augustine, (New York, NY: Elsevier Science
Manju, M. (2006). Post-traumatic Dzhala, V. I., Kuchibhotla, K. V., G. J., Bacskai, B. J., and Staley, K. J. Publishing), 18–32.
epilepsy: an overview. Clin. Neurol. Glykys, J. C., Kahle, K. T., Swiercz, (2009). Differences in cortical ver- Kang, M., and Othmer, H. G. (2009).
Neurosurg. 108, 433–439. W. B., Feng, G., Kuner, T., sus subcortical GABAergic signal- Spatiotemporal characteristics of
Amiry-Moghaddam, M., Williamson, Augustine, G. J., Bacskai, B. J., ing: a candidate mechanism of elec- calcium dynamics in astrocytes.
A., Palomba, M., Eid, T., de and Staley, K. J. (2010). Progressive troclinical uncoupling of neonatal Chaos 19, 037116.
Lanerolle, N. C., Nagelhus, E. A., NKCC1-dependent neuronal seizures. Neuron 63, 657–672. Krishnan, G. P., and Bazhenov, M.
Adams, M. E., Froehner, S. C., chloride accumulation during Goldberg, M., De Pitta, M., Volman, (2011). Ionic dynamics mediate
Agre, P., and Ottersen, O. P. (2003). neonatal seizures. J. Neurosci. 30, V., Berry, H., and Ben-Jacob, E. spontaneous termination of seizures
Delayed K+ clearance associated 11745–11761. (2010). Nonlinear gap junctions and postictal depression state. J.
with aquaporin-4 mislocaliza- Dzhala, V. I., Talos, D. M., Sdrulla, D. enable long-distance propaga- Neurosci. 31, 8870–8882.
tion: phenotypic defects in brains A., Brumback Ac, Mathews, G. C., tion of pulsating calcium waves Lee, S. H., Magge, S., Spencer, D. D.,
of alpha-syntrophin-null mice. Benke, T. A., Delpire, E., Jensen, F. in astrocyte networks. PLoS Sontheimer, H., and Cornell-Bell, A.
Proc. Natl. Acad. Sci. U.S.A. 100, E., and Staley, K. J. (2005). NKCC1 Comput. Biol. 6:e1000909. doi: H. (1995). Human epileptic astro-
13615–13620. transporter facilitates seizures in the 10.1371/journal.pcbi.1000909 cytes exhibited increased gap junc-
Annegers, J. F., Hauser, A., Coan, S. developing brain. Nat. Med. 11, Gomez-Gonzalo, M., Losi, G., tion coupling. Glia 15, 195–202.
P., and Rocca, W. A. (1998). A 1205–1213. Chiavegato, A., Zonta, M., Lothman, E. W., and Somjen, G.
population-based study of seizures Fisher, R. S., Pedley, T. A., Moody, W. Cammarota, M., Brondi, M., Vetri, G. (1976). Functions of primary
after traumatic brain injury. N. Engl. J., and Prince, D. A. (1976). The role F., Uva, L., Pozzan, T., de Curtis, afferents and responses of extra-
J. Med. 338, 20–24. of extracellular potassium in hip- M., Ratto, G. M., and Carmignoto, cellular K+ during spinal epilep-
Araque, A., Parpura, V., Sanzgiri, pocampal epilepsy. Arch. Neurol. 33, G. (2010). An excitatory loop tiform seizures. Electroencephalogr.
R. P., and Haydon, P. G. (1998). 76–83. with astrocytes contributes to Clin. Neurophysiol. 41, 253–267.
Glutamate-dependent astrocyte Fonseca, C. G., Green, C. R., and drive neurons to seizure thresh- McCormick, D. A., and Contreras, D.
modulation of synaptic transmis- Nicholson, L. F. B. (2002). old. PLOS Biol. 8:e1000352. doi: (2001). On the cellular and network
sion between cultured hippocampal Upregulation in astrocytic con- 10.1371/journal.pbio.1000352 bases of epileptic seizures. Ann. Rev.
neurons. Eur. J. Neurosci. 10, nexin 43 gap junction levels may Halassa, M. M., Fellin, T., and Haydon, Physiol. 63, 815–846.
2129–2142. exacerbate generalized seizures in P. G. (2007). The tripartite synapse: Nadkarni, S., Bartol, T. M., Sejnowski,
Araque, A., Parpura, V., Sanzgiri, R. P., mesial temporal lobe epilepsy. Brain roles for gliotransmission in health T. J., and Levine, H. (2010).
and Haydon, P. G. (1999). Tripartite Res. 929, 105–116. and disease. Trends Mol. Med. 13, Modeling vesicular release at hip-
synapses: glia, the unacknowl- Frankenhaeuser, B., and Hodgkin, A. L. 54–63. pocampal synapses. PLoS Comput.
edged partner. Trends Neurosci. 22, (1956). The after-effects of impulses Halassa, M. M., and Haydon, P. G. Biol. 6:e1000983. doi: 10.1371/
208–215. in the giant nerve fibres of loligo. J. (2010). Integrated brain circuits: journal.pcbi.1000983
Bazhenov, M., Timofeev, I., Steriade, Physiol. 131, 341–376. astrocytic networks modulate neu- Nadkarni, S., and Jung, P. (2003).
M., and Sejnowski, T. J. (2004). Fritschy, J. M. (2008). Epilepsy, E/I bal- ronal activity and behavior. Annu. Spontaneous oscillations of dressed
Potassium model for slow (2–3 Hz) ance and GABAa receptor plastic- Rev. Physiol. 72, 335–355. neurons: a new mechanism for
in vivo neocortical paroxysmal ity. Front. Mol. Neurosci. 1:5. doi: Haydon, P. G. (2001). GLIA: listening epilepsy? Phys. Rev. Lett. 91, 1–4.
oscillations. J. Neurophysiol. 92, 10.3389/neuro.02.005.2008 and talking to the synapse. Nat. Rev. Nadkarni, S., Jung, P., and Levine,
1116–1132. Frohlich, F., Bazhenov, M., and Neurosci. 2, 185–193. H. (2008). Astrocytes optimize
Beattie, E. C., Stellwagen, D., Morishita, Sejnowski, T. J. (2008a). Hertz, L., Peng, L., and Dienel, G. the synaptic transmission of
W., Bresnahan, J. C., Ha, B. K., Pathological effect of homeo- A. (2007). Energy metabolism information. PLoS Comput. Biol.
Von Zastrow, M., Beattie, M. S., static synaptic scaling on network in astrocytes: high rate of 4:e1000088. doi: 10.1371/journal.
and Malenka, R. C. (2002). Control dynamics in diseases of the cortex. oxidative metabolism and spa- pcbi.1000088
of synaptic strength by glial TNFα. J. Neurosci. 28, 1709–1720. tiotemporal dependence on Nagelhus, E. A., Mathiisen, T. M., and
Science 295, 2282–2285. Frohlich, F., Bazhenov, M., Iragui- glycolysis/glycogenolysis. J. Cereb. Ottersen, O. P. (2004). Aquaporin-
Ben-Ari, Y., and Holmes, G. L. (2005). Madoz, V., and Sejnowski, T. J. Blood Flow Metab. 27, 219–249. 4 in the central nervous system:
The multiple facets of gamma- (2008b). Potassium dynamics in Houweling, A. R., Bazhenov, M., cellular and subcellular distribu-
aminobutyric acid dysfunction in the epileptic cortex: new insights Timofeev, I., Steriade, M., and tion and coexpression with KIR4.1.
epilepsy. Curr. Opin. Neurol. 18, on an old topic. Neuroscientist 14, Sejnowski, T. J. (2005). Homeostatic Neuroscience 129, 905–913.
141–145. 422–433. synaptic plasticity can explain post- Nedergaard, M., Ransom, B., and
Binder, D. K., Yao, X., Sick, T. J., Frohlich, F., Bazhenov, M., and traumatic epileptogenesis in Goldman, S. A. (2003). New roles
Verkman, A. S., and Manley, G. T. Sejnowski, T. J. (2010). Network chronically isolated neocortex. for astrocytes: redefining the func-
(2006). Increased seizure duration bistability mediates spontaneous Cereb. Cortex 15, 834–845. tional architecture of the brain.
and slowed potassium kinetics transitions between normal Janigro, D., Gasparini, S., D’Ambrosio, Trends Neurosci. 26, 523–530.
in mice lacking aquaporin- and pathological brain states. J. R., McKhann, G., and DiFrancesco, Netoff, T. I., Clewley, R., Arno, S.,
4 water channels. Glia 53, Neurosci. 30, 10734–10743. D. (1997). Reduction of K+ uptake Keck, T., and White, J. A. (2004).
631–636. Frohlich, F., Bazhenov, M., Timofeev, in glia prevents long-term depres- Epilepsy in small-world networks. J.
Cressman, J. R., Ullah, G., Ziburkus, J., I., Steriade, M., and Sejnowski, T. sion maintenance and causes epilep- Neurosci. 24, 8075–8083.
Schiff, S. J., and Barreto, E. (2009). J. (2006). Slow state transitions tiform activity. J. Neurosci. 17, Oberheim, N. A., Tian, G. F., Han, X.,
The influence of sodium and of sustained neural oscillations by 2813–2824. Peng, W., Takano, T., Ransom, B.,
potassium dynamics on excitabil- activity-dependent modulation of Kager, H., Wadman, W. J., and Somjen, and Nedergaard, M. (2008). Loss of
ity, seizures, and the stability of intrinsic excitability. J. Neurosci. 26, G. G. (2000). Simulated seizures and astrocytic domain organization in
persistent states: I. Single neuron 6153–6162. spreading depression in a neuron the epileptic brain. J. Neurosci. 28,
dynamics. J. Comput. Neurosci. 26, Giaume, C., Koulakoff, A., Roux, model incorporating interstitial 3264–3276.
159–170. L., Holcman, D., and Rouach, space and ion concentrations. J. Ostby, I., Oyehaug, L., Einevoll, G.
De Pitta, M., Volman, V., Berry, H., N. (2010). Astroglial networks: Neurophysiol. 84, 495–512. T., Nagelhus, E. A., Plahte, E.,
and Ben-Jacob, E. (2011). A tale of a step further in neuroglial and Kandel, E. R. (1991). “Nerve cell Zeuthen, T., Lloyd, C. M., Ottersen,
two stories: astrocyte regulation of gliovascular interactions. Nat. Rev. and behavior,” in Principles of O. P., and Omholt, S. W. (2009).
synaptic depression and facilitation. Neurosci. 11, 87–99. Neural Science, eds E. R. Kandel, Astrocytic mechanisms explaining

Frontiers in Computational Neuroscience www.frontiersin.org August 2012 | Volume 6 | Article 58 | 7


Volman et al. Computational models of neuron-astrocyte interaction in epilepsy

neural-activity-induced shrink- induction by glia-mediated synap- Posttraumatic epilepsy: the roles of posttraumatic epileptogenesis. Soc.
age of extraneuronal space. PLoS tic scaling. J. R. Soc. Interface 6, synaptic plasticity. Neuroscientist Neurosci. Abstr. 250.11/V5.
Comput. Biol. 5:e1000272. doi: 655–668. 16, 19–27. Wallraff, A., Kohling, R., Heinemann,
10.1371/journal.pcbi.1000272 Scemes, E., and Spray, D. C. (2004). The Tsodyks, M. V., and Markram, H. U., Theis, M., Willecke, K., and
Oyehaug, L., Ostby, I., Lloyd, C. M., astrocytic syncytium. Adv. Mol. Cell (1997). The neural code between Steinhauser, C. (2006). The impact
Omholt, S. W., and Einevoll, G. T. Biol. 31, 165–179. neocortical pyramidal neurons of astrocytic gap junctional cou-
(2011). Dependence of spontaneous Schneider-Mizell, C. M., Parent, J. M., depends on neurotransmitter pling on potassium buffering in
neuronal firing and depolarization Ben-Jacob, E., Zochowski, M. R., release probability. Proc. Natl. Acad. the hippocampus. J. Neurosci. 26,
block on astroglial membrane and Sander, L. M. (2010) From net- Sci. U.S.A. 94, 719–723. 5438–5447.
transport mechanisms. J. Comput. work structure to network reorgani- Turrigiano, G. G. (1999). Homeostatic Yagodin, S. V., Holtzclaw, L., Sheppard,
Neurosci. 32, 147–165. zation: implications for adult neu- plasticity in neuronal networks: the C. A., and Russell, J. T. (1994).
Pannasch, U., Vargova, L., Reingruber, rogenesis. Phys. Biol. 7, 046008. more things change, the more they Nonlinear propagation of agonist-
J., Ezan, P., Holcman, D., Giaume, Seifert, G., Schilling, K., and stay the same. Trends Neurosci. 22, induced cytoplasmic calcium waves
C., Sykova, E., and Rouach, N. Steinhauser, C. (2006). Astrocyte 221–227. in single astrocytes. J. Neurobiol. 25,
(2011). Astroglial networks scale dysfunction in neurological disor- Ullah, G., Cressman, J. R., Barreto, 265–280.
synaptic activity and plasticity. ders: a molecular perspective. Nat. E., and Schiff, S. J. (2009). The Zhang, J. M., Wang, H. K., Ye, C. Q.,
Proc. Natl. Acad. Sci. U.S.A. 108, Rev. Neurosci. 7, 194–206. influence of sodium and potas- Ge, W., Chen, Y., Jiang, Z. L., Wu, C.
8467–8472. Silchenko, A. N., and Tass, P. A. sium dynamics on excitability, P., Poo, M. M., and Duan, S. (2003).
Parpura, V., and Haydon, P. G. (2000). (2008). Computational modeling of seizures, and the stability of per- ATP released by astrocytes mediates
Physiological astrocytic calcium paroxysmal depolarization shifts in sistent states: II. Network and glial glutamatergic activity-dependent
levels stimulate glutamate release neurons induced by the glutamate dynamics. J. Comput. Neurosci. 26, heterosynaptic suppression. Neuron
to modulate adjacent neurons. release from astrocytes. Biol. Cybern. 171–183. 40, 971–982.
Proc. Natl. Acad. Sci. U.S.A. 97, 98, 61–74. Volman, V., Ben-Jacob, E., and Levine,
8629–8634. Simard, M., and Nedergaard, M. H. (2007). The astrocyte as a
Conflict of Interest Statement: The
Payne, J. A., Rivera, C., Voipio, J., and (2004). The neurobiology of glia gatekeeper of synaptic informa-
authors declare that the research
Kaila, K. (2003). Cation-chloride in the context of water and ion tion transfer. Neural Comput. 19,
was conducted in the absence of any
co-transporters in neuronal com- homeostasis. Neuroscience 129, 303–326.
commercial or financial relationships
munication, development, and 877–896. Volman, V., Perc, M., and Bazhenov, M.
that could be construed as a potential
trauma. Trends Neurosci. 26, Sofroniew, M. V. (2009). Molecular dis- (2011a). Gap junctions and epilep-
conflict of interest.
199–206. section of reactive gliosis and glial tic seizures-two sides of the same
Reyes, R. C., and Parpura, V. (2008). scar formation. Trends Neurosci. 32, coin? PLoS ONE 6:e20572. doi:
Models of astrocytic Ca2+ dynam- 638–647. 10.1371/journal.pone.0020572 Received: 01 April 2012; accepted: 23 July
ics and epilepsy. Drug Discov. Today Somjen, G. G., Kager, H., and Wadman, Volman, V., Bazhenov, M., and 2012; published online: 13 August 2012.
Dis. Models 5, 13–18. W. J. (2008). Computer simula- Sejnowski, T. J. (2011b). Pattern Citation: Volman V, Bazhenov M and
Rouach, N., Glowinski, J., and Giaume, tions of neuron-glia interactions of trauma determines the thresh- Sejnowski TJ (2012) Computational
C. (2000). Activity-dependent mediated by ion flux. J. Comput. old for epileptic activity in a models of neuron-astrocyte interaction
neuronal control of gap-junctional Neurosci. 25, 349–365. model of cortical deafferentation. in epilepsy. Front. Comput. Neurosci.
communication in astrocytes. J. Cell Stellwagen, D., and Malenka, R. C. Proc. Natl. Acad. Sci. U.S.A. 108, 6:58. doi: 10.3389/fncom.2012.00058
Biol. 149, 1513–1526. (2006). Synaptic scaling medi- 15402–15407. Copyright © 2012 Volman, Bazhenov
Santhakumar, V., Aradi, I., and Soltesz, ated by glial TNFα. Nature 440, Volman, V., Sejnowski, T. J., and and Sejnowski. This is an open-access
I. (2004). Role of mossy fiber 1054–1059. Bazhenov, M. (2011c). Topological article distributed under the terms of the
sprouting and mossy cell loss in Steinmetz, C. C., and Turrigiano, G. G. basis of epileptogenesis in a Creative Commons Attribution License,
hyperexcitability: a network model (2010). Tumor necrosis factor-α sig- model of severe cortical trauma. J. which permits use, distribution and
of the dentate gyrus incorporating naling maintains the ability of corti- Neurophysiol. 106, 1933–1942. reproduction in other forums, provided
cell types and axonal topography. J. cal synapses to express synaptic scal- Volman, V., Bazhenov, M., and the original authors and source are
Neurophysiol. 93, 437–453. ing. J. Neurosci. 30, 14685–14690. Sejnowski, T. J. (2011d). A compu- credited and subject to any copyright
Savin, C., Triesch, J., and Meyer- Timofeev, I., Bazhenov, M., Avramescu, tational model of neuronal and glial notices concerning any third-party
Hermann, M. (2009). Seizure S., and Nita, D. A. (2010). homeostatic synaptic plasticity in graphics etc.

Frontiers in Computational Neuroscience www.frontiersin.org August 2012 | Volume 6 | Article 58 | 8


Volman et al. Computational models of neuron-astrocyte interaction in epilepsy

APPENDIX (Tsodyks and Markram, 1997):


MODELING TOOLBOX 1—GLIAL MEDIATED REGULATION OF
dx 
SYNAPTIC NEUROTRANSMITTER RELEASE AND NEURAL = R (1 − x) − uxδ(t − ti ) (5)
EXCITABILITY dt
i
In response to mechanical stimulation or stimulation of
du 
G-protein coupled receptors, astrocytes can release a variety of = −F u + U0 () (1 − u)δ(t − ti ) (6)
factors, the most documented ones being ATP and glutamate dt
i
(Araque et al., 1999; Zhang et al., 2003). Depending on the
localization of its glial release site, the released glutamate can In Equations (5)–(6), R is the rate of recovery from short-term
activate receptors on postsynaptic membrane (thus modulating synaptic depression, F is the rate of recovery from “astrocytic
neural excitability) or bind to its dedicated receptors on presynap- effects”, and u is proportional to the probability of vesicular
tic membrane (and thus modulate presynaptic neurotransmitter release. The modulating effect of astrocytes on synaptic transmis-
release) (see also Figure 1). sion is captured with the “gating function,” U0 (). The exact form
Modulation of neural excitability by glial glutamate is a rela- of this gating function depends on the specifics of synapse under
tively well understood process, as it involves activation of post- study [i.e., different physiological and ultra-structural parameters
synaptic NMDA receptors (Parpura and Haydon, 2000). Thus, of a synapse will result in different functional forms of U0 ()]. A
glial-generated NMDA current can be modeled as: detailed discussion of astrocytic modulation of presynaptic func-
tion is provided in the article by De Pitta et al. in this issue, and in
De Pitta et al. (2011).
gF (t) − gS (t)
NMDA
IASTRO (t) =   (Vm − ENMDA )
1 + 0.33 Mg2+ exp(−0.06 Vm ) MODELING TOOLBOX 2—GLIAL MEDIATED REGULATION OF ION
(1) CONCENTRATIONS
dgF,S −gF,S In physiological conditions, potassium and chloride cur-
= + GGLU
ASTRO (t) (2) rents generate hyperpolarizing influences on neuronal mem-
dt τF,S
brane potential and thus can be considered to be seizure-
suppressing; however, in pathological situations these cur-
rents become depolarizing due to the change in their reversal
ASTRO (t) represents the glutamate pulse due to glial glu-
where GGLU
potentials. Reversal potentials of potassium and chloride cur-
 the τF and the τS are the fast and slow NMDA
tamate release,
time scales, Mg2+ is the extracellular level of magnesium, and rents are determined by the electrochemical gradients of their
Vm is neuronal membrane potential. However, this detailed mod- constituent ions:
  
eling requires knowledge of GGLUASTRO (t) dependence on astrocytic
calcium, information that is not fully available at the moment. K+ OUT
EK = 26.64 · log  +  [ mV];
An alternative, phenomenological, way to model the effect of K IN
astrocytes on neuronal excitability was proposed by Nadkarni and   − 
Jung (2003): Cl
ECl = 26.64 · log  −  IN [ mV] (7)
Cl OUT
 
IASTRO = 2.11 log(y) log(y) (3) so the high extracellular (intracellular) concentration of potas-
  sium (chloride) shifts the reversal potential toward more positive
y = Ca2+ ASTRO − 196.69 (4)
values.
The dynamics of extracellular potassium concentration can be
In equations (3)–(4), (x) is the Heaviside function (taking captured with the following equation:
on value of 1 if the argument
 is greater than zero and value  
of 0 otherwise), Ca2+ ASTRO is the concentration of free cal- d K+ OUT    
cium (measured in nM) in astrocyte, and current is given in pA. = IK +  K+ DIFF − G (8)
dt
Fit parameters were derived from experimental data [details in
Nadkarni and Jung (2003)]. in which IK is the sum over all neuronal
  potassium
 currents
Astrocytic modulation of presynaptic function is more var- (from ion channels and/or pumps),  K+ DIFF is the con-
ied. A basic assumption, supported by experimental data, is that tribution from potassium diffusion in extracellular space, and
astrocytes modulate the properties of short-term synaptic plas- G is the term that represents buffering of extracellular potas-
ticity (both short-term depression and short-term facilitation). sium by astrocytes. Relatively detailed biophysical models exist
On semi-phenomenological level, short-term synaptic plasticity is to describe glial buffering of potassium, based on activation
well captured with Tsodyks–Markram use model which assumes of glial potassium channels (Oyehaug et al., 2011); however,
that at any given time, “synaptic resource” (proportional to the in many cases it is convenient to describe glial buffering in a
number of neurotransmitter vesicles) per synaptic bouton is lim- semi-phenomenological way by assuming the existence of glial
ited and only a fraction x(t) of it is available for release at time t buffer, [B], that binds free potassium ions. Glial contribution

Frontiers in Computational Neuroscience www.frontiersin.org August 2012 | Volume 6 | Article 58 | 9


Volman et al. Computational models of neuron-astrocyte interaction in epilepsy

can then be written as: (Krishnan and Bazhenov, 2011):

  τCl∞
−d[B] τCl = τBASE +
(14)
G= = kON K+ OUT [B] − kOFF ([B]T − [B]) (9) [ − ]I∞ −[K+ ]OUT
Cl
dt 1 + exp αKCl
kOFF
kON =
+ (10)
[ ]TH −[K+ ]OUT
K MODELING TOOLBOX 3—GLIAL MEDIATED HOMEOSTATIC
1 + exp αKG SYNAPTIC PLASTICITY
Homeostatic synaptic plasticity (HSP) allows neurons and net-
where kON and kOFF are forward (binding) and backward works to adjust synaptic conductance in order to offset the effect
(unbinding) rates, correspondingly, [B]T is the total buffer con- of long-term perturbation on their firing rates (Turrigiano, 1999).
centration (usually assumed to be very high), and [K+ ]TH is the Astrocytes contribute to neuronal HSP by releasing tumor necro-
threshold concentration of extracellular potassium above which sis factor alpha (TNFα) that binds to neuronal TNFα receptors
glial buffer is activated (∼10 mM) (Kager et al., 2000; Krishnan and causes insertion and removal of AMPA and GABA recep-
and Bazhenov, 2011). tors, correspondingly (Stellwagen and Malenka, 2006). Astrocytic
Because of the electrogenic nature of Na+ /K+ exchange pump release of TNFα was observed to occur following prolonged
bringing in 2 K+ ions in exchange for 3 Na+ ions, increase of the neuronal inactivity (Stellwagen and Malenka, 2006). Given that
extracellular K+ (or intracellular Na+ ) concentration leads to the astrocytes “sense” neuronal activity through glutamate levels, it
outward current that may hyperpolarize the cell. This current is plausible to assume that astrocytic involvement in HSP is trig-
can be modeled as (Kager et al., 2000; Krishnan and Bazhenov, gered by prolonged inactivity of pyramidal (PY) neurons. In the
2011): most general form, this information can be captured with the
following equations:
I pump = IK
pump pump
+ INa
pump
, IK = −2Imax A,
pump
INa = 3Imax A

(11) gAMPA = gAMPA + αGLU
HSP · f0 − f T,PY (15)


A = (1/(1 + (Koα /[Ko ])))2 (1/(1 + (Naiα /[Nai ]))) (12) gGABA = gGABA − αGABA
HSP · f0 − f T,PY (16)

where Koα and Naiα are baseline ionic concentrations. In Equations


(15)–(16), f0 is the preset target rate of PY neu-
The dynamics of intracellular chloride concentration can be rons, f T,PY is the averaged (over time and PY neurons) firing
modeled with an equation similar to equation (8): rate of PY neurons in “astrocytic domain” (Volman et al., 2011d),
and αGLU
HSP , αHSP are the rates of HSP convergence for specific
GABA
   −     type of synaptic receptor. Because astrocytic TNFα can only up
d Cl− IN  Cl IN − Cl− I∞
= ICl + (13) regulate synaptic conductance, Equations (15)–(16) can only be
dt τCl applied in the case when f T,PY ≤ f0 (i.e., recovery from pro-
longed inactivity). Homeostatic down regulation of neuronal
where the characteristic time τCl is the inverse of rate of chlo- excitability is likely achieved through other neuronal mechanisms
ride extrusion by KCC2 exporter. The KCC2 mediated effec- (Turrigiano, 1999). This “division of labor” between astrocytes
tive coupling between extracellular potassium concentration and neurons may result in non-trivial collective activity following
and intracellular chloride concentration can be modeled as perturbations such as, e.g., brain trauma (Volman et al., 2011d).

Frontiers in Computational Neuroscience www.frontiersin.org August 2012 | Volume 6 | Article 58 | 10

You might also like