Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Prefronto-cortical dopamine D1 receptor sensitivity can critically influence working memory maintenance during delayed response tasks

  • Melissa Reneaux ,

    Contributed equally to this work with: Melissa Reneaux, Rahul Gupta

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Resources, Software, Validation, Visualization, Writing – original draft, Writing – review & editing

    meliss14_sit@jnu.ac.in (MR); rahul35_sit@jnu.ac.in (RG)

    Affiliation School of Computational and Integrative Sciences, Jawaharlal Nehru University, New Delhi, India

  • Rahul Gupta

    Contributed equally to this work with: Melissa Reneaux, Rahul Gupta

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Resources, Software, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing

    meliss14_sit@jnu.ac.in (MR); rahul35_sit@jnu.ac.in (RG)

    Affiliation School of Computational and Integrative Sciences, Jawaharlal Nehru University, New Delhi, India

Abstract

The dopamine (DA) hypothesis of cognitive deficits suggests that too low or too high extracellular DA concentration in the prefrontal cortex (PFC) can severely impair the working memory (WM) maintenance during delay period. Thus, there exists only an optimal range of DA where the sustained-firing activity, the neural correlate of WM maintenance, in the cortex possesses optimal firing frequency as well as robustness against noisy distractions. Empirical evidences demonstrate changes even in the D1 receptor (D1R)-sensitivity to extracellular DA, collectively manifested through D1R density and DA-binding affinity, in the PFC under neuropsychiatric conditions such as ageing and schizophrenia. However, the impact of alterations in the cortical D1R-sensitivity on WM maintenance has yet remained poorly addressed. Using a quantitative neural mass model of the prefronto-mesoprefrontal system, the present study reveals that higher D1R-sensitivity may not only effectuate shrunk optimal DA range but also shift of the range to lower concentrations. Moreover, higher sensitivity may significantly reduce the WM-robustness even within the optimal DA range and exacerbates the decline at abnormal DA levels. These findings project important clinical implications, such as dosage precision and variability of DA-correcting drugs across patients, and failure in acquiring healthy WM maintenance even under drug-controlled normal cortical DA levels.

Introduction

Working memory (WM) is a crucial asset of cognitive facility during delayed-response tasks. It is comprised of many subprocesses, namely, attentional control system, retention of cue-induced information over a brief delay interval (WM maintenance), and other executive functions performing manipulation as well as retrieval of cue-specific information at the end of the delay period. These processes concertedly guide the goal-directed response. However, WM maintenance lies at the core of these various cognitive operations [1]. Sustained/persistent-firing activity in the cortices of human as well as non-human primate brains during delay is the proposed neural correlate of WM maintenance [2]. Although participation of various regions of the cortex, including prefrontal cortex (PFC), posterior parietal cortex (PPC) and inferior temporal cortex (ITC), has been observed in WM maintenance [3], the PFC is known to play a pivotal role.

The neurochemical dopamine (DA) exerts a strong modulating effect on WM. Although the effect of DA is mediated through the activation of D1 receptors (D1Rs) as well as D2 receptors (D2Rs) present locally in the cortical region, it is suggested that the effect on WM maintenance is predominantly mediated through the activation of D1Rs whereas D2Rs are primarily involved in the WM updating and executive functions [4, 5]. The computational studies [610] and experimental studies [1115] have brought immense growth in our understanding of the dopaminergic modulation of WM maintenance. These attempts have led to the well-known DA hypothesis of cognitive deficit observed under various neuropsychiatric conditions, such as ageing [16, 17], stress [18, 19], and schizophrenia [15, 20]. According to this hypothesis, too low or too high extracellular DA concentration in the PFC can severely impair the WM maintenance during delay period. Thus, there exists only an optimal range of DA where the WM-associated sustained-firing activity in the cortex possesses optimal firing frequency as well as robustness against noisy distractions.

However, several experimental studies [16, 17, 2127] have also reported alterations even in the cortical D1R density and reactivity of DA-binding sites on individual D1Rs under various neuropsychiatric conditions. Together, these factors critically regulate the efficiency of the local cortical network for detecting changes in the extracellular DA content and, thus, define the D1R-sensitivity of the cortical region. The D1R-sensitivity is experimentally measured in terms of binding potential (BP) of D1Rs in the PFC [17, 23, 26]. Accordingly, the alteration in D1R-sensitivity appears as an additional important factor to be considered in conjunction with the alteration in cortical DA content. However, the impact of alteration in D1R-sensitivity on the WM maintenance has still remained unaddressed.

The present study addresses this issue by employing a quantitative neural mass model of the prefronto-mesoprefrontal system, which is comprised of the reciprocal interaction between the PFC and the cortical-projecting DA neurons residing in the ventral tegmental area (VTA) in the midbrain [28]. Particularly, the effects of D1R-sensitivity on the firing frequency and robustness of the cortical persistent activity during delay are observed. Moreover, the mesocortical scale of the framework facilitates quantitative observation on the variation in modulation-associated extracellular DA under different conditions of the sensitivity. The findings suggest that cortical D1R-sensitivity critically governs the range of cortical DA level underlying the modulation of WM maintenance in the physiological scenario. Interestingly, this regulation is a consequence of the feedback control of cortical D1R-sensitivity on the dynamics of DA release from VTA-residing DA neurons during delay. Accordingly, increase in D1R-sensitivity causes shrinking of the optimal DA range and shift of the range to lower concentrations. This essentially curtails the safe DA range of efficient WM maintenance in the PFC in the presence of physiological fluctuations in the cortical DA. Furthermore, besides exacerbating the decline in WM-robustness at abnormal DA levels, increased sensitivity is characterized with lesser robustness of the persistent cortical activity even within the optimal DA range.

Methods

The particular subset of the larger prefronto-mesoprefrontal system modeled here includes interactions between a local population of cortical neurons in the dorsolateral prefrontal cortex (DLPFC) extending corticomesencephalic glutamatergic projections [29, 30] to a subpopulation of DA neurons in the VTA, which in turn sends mesocortical dopaminergic projections [31] to the cortical population. In this way, the reciprocal interaction (Fig 1A) gives rise to the mesocortical circuit. The DLPFC is a cortical region within the PFC and has been observed to be actively involved in many visuospatial WM tasks [3234]. The mathematical model [35] employed here adopts a neural mass approach where the population-averaged activities of the different kinds of neuronal populations constituting the circuit dynamics are considered. The present model provides quantitative profiles of the various measurable entities of the mesocortical dynamics in close association with their experimentally known estimates. Further, a stochastic formulation of the mass model [36] is utilized to gain features of robustness of the WM maintenance during delay under the physiologically-relevant situation of noisy mesocortical dynamics.

thumbnail
Fig 1. Model of the closed-loop mesocortical circuit.

(A) A three-dimensional minimal rendering of the human brain essentially featuring the anatomical localization of the two brain regions, DLPFC and VTA, whose reciprocal interaction constitutes the mesocortical circuit. (B) A simplified illustration of the synaptic contact made by a terminal of the dopaminergic afferent projections onto a pyramidal neuron or GABAergic interneuron in the cortex. The DA-releasability (RDA) and D1R-sensitivity (D1Rsens) are the presynaptic and postsynaptic factors, respectively, which crucially regulate the transmission at dopaminergic synapses. (C) In the neural mass model of the mesocortical circuit, the cortical neurons are broadly categorized into the populations of excitatory pyramidal neurons and inhibitory GABAergic interneurons. The excitatory population, on receiving cue input, self excites itself (with the synaptic efficacy WPP) and also excites the population of inhibitory neurons in the cortex (WPI) as well as DA neurons in midbrain (WPD). On excitation, the inhibitory population inhibits excitatory population (WIP) as well as itself (WII) whereas the DA neuron population releases DA in the cortex (RDA) through dopaminergic projections and causes accumulation of the cortical DA pool, [DA].

https://doi.org/10.1371/journal.pone.0198136.g001

Modeling the dynamics of local cortical network

Glutamate-releasing excitatory pyramidal neurons and GABA-releasing inhibitory interneurons are the most abundant neurons in the PFC. The layer V-VI (deep-layer) neurons are the subject of interest here as they have been found, to be mainly associated with the recurrent sustained firing activity during WM-tasks [37]. The superficial layers are mainly involved in receiving afferent stimuli from various parts of the brain, such as thalamus and intercortical regions, and transmit them to the deep layers. Delayed-response tasks, such as spatial tasks, have demonstrated different cortical neurons to be specifically tuned to firing in response to a characteristic stimuli presented [38, 39]. Therefore, there exists local clusters of cortical neurons which fire maximally towards a specific external stimuli, such as orientation in space in the spatial tasks, than the others.

Under the present neural mass framework, the excitatory and inhibitory neurons in a local cortical network are pooled into distinct populations and the interactions among them are considered at the population-level. Accordingly, DLPFC activity is comprised of the local population activity of excitatory pyramidal neurons (aPN). The pyramidal population self-excites itself with the synaptic efficacy WPP (feed-forward excitation) and excites the population of local GABAergic interneurons with the synaptic efficacy WPI. In turn, the activity of interneuron population (aIN) exerts inhibition on aPN with the synaptic efficacy WIP (feed-back inhibition) as well as suppresses itself with the synaptic efficacy WII. This interplay between the feed-forward excitation and the feed-back inhibition leads to the establishment of sustained-firing activity in the DLPFC, which represents the formation and maintenance of WM during delay period. (1) (2) where, and . The and corresponds to the basal spontaneous activity level in the pyramidal and GABAergic interneuron populations, respectively, in the local cortical network in the PFC.

The activation function, fx) where Δx(t) ∈ {ΔaPN(t), ΔaIN(t), ΔaDN(t), Δ[DA](t)} signifies a biophysically-imposed finite saturating limit to which the different variables may rise during their activation and is given by, (3) here, C denotes the constants c1, c2, c3, c4 associated with the tanh function of ΔaPN(t), ΔaIN(t), ΔaDN(t), Δ[DA](t), respectively.

The first term on the right-hand side of Eqs 1 and 2 denotes the excitability of the population of pyramidal neurons and interneurons characterized by the specific time constants τPN and τIN, respectively. A large time constant implies a greater excitability of the neurons constituting a population. The second term in Eqs 1 and 2 represents the recurrent excitation of the pyramidal neurons and excitation of interneurons by the pyramidal activity, respectively, with the corresponding synaptic efficacies WPP and WPI. The last term in these equations represents the inhibition of pyramidal population by interneuron population and self-inhibition of interneuron population, respectively, with the corresponding synaptic efficacies WIP and WII.

Modeling the dynamics of cortical DA regulation

According to the standing literature, there still exist numerous elements of confusion regarding the regulation of cortical DA during WM maintenance. Particularly, a definitive conclusion could not be drawn yet regarding the pertinent roles of tonic vs. phasic release of DA in the cortex as well as the associated tonic and phasic activities of the cortical-projecting sub-population of DA neurons residing in the midbrain region.

In vivo single-cell recordings of midbrain dopamine (DA) neurons in monkeys have demonstrated a continuous occurrence of basal-level spontaneous activity in the DA neurons [40, 41]. This tonic activity of DA neurons leads to the tonic release of DA at various target sites in the brain innervated by dopaminergic projections [42]. Accordingly, the tonic activity of cortical-projecting DA neurons in the VTA serves as the source for the stable and volume-wide basal extracellular DA concentration in the PFC. In addition, there also occurs heightened phasic burst-activity in the DA neurons mainly at two instances during delayed-response tasks [40, 41]. First, it occurs at the initial instance of cue presentation and serves to alert subject’s attention towards the external stimuli of salience for correctly performing the task. Secondly, it occurs at the eventual moment of making a motor response in the expectation of reward. These evoked phasic activities lead to a sudden excessive release of DA from the afferent dopaminergic terminals at the target sites [4244]. However, synaptic as well as extrasynaptic rapid uptake of DA by the local dopamine uptake transporters (DATs) and catechol-O-methyl transferase (COMT)-based degradation of DA lead to only a transient high amplitude pulse-like increase in DA concentration in very close vicinity of the release sites [42, 44]. This manner of DA release is commonly referred to as the phasic mode of DA release [42]. Accordingly, phasic DA release does not considerably affect the extracellular DA concentration across a wide volume [42, 44].

Interestingly, sustained activity in DA neurons has not been observed during delay periods in primates VTA undergoing delayed alternation tasks [40]. Although later experiments revealed increase in DA activity during delay [41], this modulation in DA activity was mainly attributed to the intensity of reward probability and uncertainty, rather than to the sustained-firing activity in the PFC. Accordingly, it is implied that the tonic activity of the VTA-residing DA neurons does not change during the delay interval. It has been often suggested that the phasic cortical DA release at the instance of cue presentation may underlie the dopaminergic modulation of WM maintenance during delay. This possibility immediately connects to the gating hypothesis of the dopaminergic modulation [45, 46]. It suggests that phasic activity of the DA neurons at cue presentation initially gates the input stimuli associated with the WM updation and later facilitates WM maintenance by restraining the entry of distracting stimuli [12] during delay. Although the phasic DA release is transient and its influence is spatially-restricted within close vicinity of the DA release sites, the slow intracellular DA signaling [13] and the presence of statistically-significant population of dopaminergic synapses closely-apposed to the asymmetric excitatory synapses in the cortical region [47] may strongly support the gating hypothesis.

However, besides gating of the input, various experimental [11, 13, 18] and theoretical studies [6, 7, 9, 10] have also suggested dopaminergic modulation of the intrinsic excitability and robustness of the cortical neural networks to distracting stimuli during delay intervals, which is also the prime objective of the present modeling study. This would require a volume-wide stable change in cortical DA, as D1Rs are mostly located extrasynaptic to the site of DA release [48]. In fact, in the behaving rhesus monkeys correctly performing in the delayed alternation tasks, a noticeable increase over the basal DA level in the DLPFC has been reported through in vivo microdialysis [49]. Such a change is certainly beyond the capacity of phasic DA release. Although the sparse presence of DATs in the prefrontal cortex [5052] had been doubted to enable phasic DA release for causing a volume-wide change in DA level, the experimental observations on the sparse DA projections as well as DA release sites [53] in the PFC relative to the striatum and the uptake of DA by the norepinephrine uptake transporters present in high density on the local norepinephrinergic afferent projections [5456] again seem to decline such a possibility. In fact, a recent detailed computational study by Spühler and Hauri [57] of the spatiotemporal features of DA release in macaque prefrontal cortex has also demonstrated lack of a volume-wide stable change in cortical DA level due to the phasic activities in cortical-projecting DA neurons.

These observations suggest that the involvement of tonic DA release is indispensable to the volume-wide modulation of network excitability and robustness during delay. In the context of striatum, Grace [58] has proposed a plausible mechanism for the change in the local tonic DA release without any change in the tonic activity of the striatum-projecting DA neurons. It suggests that local activity-dependent change in the extracellular glutamate concentration can regulate the tonic DA release through ionotropic AMPA and NMDA receptors located at the afferent dopaminergic terminals. However, the electron microscopic investigations [59, 60] of the distribution of immunofluorescently-labelled ionotropic glutamate receptors in the rodent striatum did not demonstrate a statistically-significant presence of these receptors on the dopaminergic terminals. Further, it has been experimentally observed that there occurs increase in extracellular DA concentration but no change in the extracellular glutamate concentration in the DLPFC during delayed alternation tasks performed by healthy rhesus monkeys [61].

Eventually, it appears that delay-associated change in the tonic activity of cortical-projecting DA neurons may underlie the change in tonic DA release in the cortex. In fact, application of NMDA and AMPA agonist (antagonist) in rodent VTA has been shown to cause increase (decrease) in the extracellular DA level in the PFC [62]. Therefore, change in the local glutamate concentration in the VTA through VTA-projecting cortical neurons may influence the tonic activity of DA neurons during delay. However, it also demands a reconsideration of the abovementioned experimental observations on the lack of sustained activity in the midbrain DA neurons during delay. In this regard, it must be noted that the DA neurons recorded during delayed tasks in these studies [40, 41] were not specific to cortical-projecting DA neuron sub-population in the VTA. Rather, DA neurons belonging to a wide range of projection areas were collectively sampled in the VTA as well as the substantia nigra pars compacta. In contrast, the closed-loop mesocortical circuit addressed here involves the specific DA neuron sub-population which receives excitatory signals from the PFC as well as project back to cortical region.

Moreover, not all the mesocortical DA neurons in the VTA fire under basal resting conditions [42]. Rather, a significant proportion of these neurons remain in the hyperpolarized inactive state. However, during delay period of WM tasks, the increased glutamate level in the VTA due to sustained activity in the PFC may lead to activation of more fractions of inactive DA neurons. It is thought that early activation of DA neurons from their inactive state leads to tonic mode of Poissonian firing in the DA neurons [42, 63]. Accordingly, a larger fraction of VTA-residing DA neurons will acquire tonic activity. Moreover, the firing frequency of a fraction of tonically-firing DA neurons may also rise as well as the fraction of burst DA firing neurons may also increase during delay period. These different processes would together be responsible for the stable (tonic) increase in population-averaged activity of mesocortical DA neuron sub-population in the VTA in response to increase in sustained-firing activity in the PFC. Therefore, it is strongly possible that the tonic activity of the DA sub-population strictly involved in the closed-loop mesocortical circuit may increase due to the sustained-firing activity in the PFC. Consequently, it may lead to enhanced tonic DA release in the cortex and underlies WM maintenance during delay period.

Accordingly, the variations in the population-averaged activity of mesocortical DA neurons, aDN, in the VTA and the cortical bulk or volume-averaged extracellular DA concentration or content, [DA], under the mesoencephalic excitation are modeled here as, (4) (5) Where, and Δ[DA](t) = [DA](t) − [DA]basal. The and [DA]basal corresponds to the basal activity of mesocortical DA neurons and the basal extracellular DA concentration, respectively, in the PFC under resting conditions.

The first term on the right-hand side of Eq 4 denotes the excitability of the population of DA neurons characterized by the specific time constant τDN. A large time constant implies a greater excitability of the DA neurons. The second term in Eq 4 represents the excitation of DA neurons by the cortical pyramidal activity aPN with the glutamatergic synaptic efficacy WPD. Further, the first term in Eq 5 represents the uptake and degradation of DA in the extracellular region in the PFC with the characteristic time constant τDA whereas the second term signifies the release of DA by the excited DA neuron population with the efficiency parameter, RDA. RDA denotes the DA-releasability of the dopaminergic projections and critically relies on the intrinsic DA metabolism and release probability of the DA-containing vesicles at the axonal terminals of mesocortical projections (Fig 1B).

Anatomical and electrophysiological studies have shown that there also exists a population of GABAergic neurons in the VTA which receives glutamatergic inputs from the cortical areas and acts as a brake system to suppress the excess activity of the DA neuron population [64]. The present model does not incorporate an explicit dynamics of GABA population in the VTA. Rather, the magnitudes of the parameters WPD for excitation of DA neurons by cortical projections and τDN for the self-decay of DA population activity have been adjusted in a manner so that the putative effects of VTA-inhabiting GABA population could be accounted for. Somatodendritic D2 autoreceptors are generally known to play a crucial role in lateral inhibition of DA neuron activity in the VTA. However, the sub-population of DA neurons in the VTA extending mesocortical projections stands as an exception to this phenomenon of somatodendritic lateral inhibition [65]. Furthermore, the cortical DA content has been assumed here as a single entity or a pool which varies according to DA neuron’s activity. An explicit consideration of synaptic release of DA and its volume diffusion in the cortical area is ignored to satisfy the neural mass framework of the model.

Modeling the effect of D1R activation on cortical excitability and synaptic transmission

In the presence of extracellular DA in the PFC, D1R activation causes modulation of the activity of several voltage-gated and ligand-gated ionotropic receptors [66] located on the cortical neurons. Consequently, this leads to the modulation of neuronal excitability of the pyramidal neurons [67] and GABAergic interneurons [68] as well as the modulation of the excitatory [69] and inhibitory [70] synaptic efficacies in the local cortical network (Fig 1C). However, the resultant level of cortical D1R stimulation in response to cortical DA level further depends on the parameter, D1R-sensitivity [26]. It signifies how efficiently the cortical network perceives any change in DA content and, hence, depends collectively on the cortical D1R density and the reactivity of DA-binding sites on individual D1Rs (Fig 1B). Therefore, the resultant level of D1R activation or stimulation, D1Ract, in the presence of cortical DA content [DA] is modeled here as, (6) where, D1Rsens signifies the D1R-sensitivity of the cortical neurons to cortical DA pool. Further, the dopaminergic modulation of the neuronal excitability and the synaptic efficacies in the cortical neuronal populations in response to D1R stimulation is given by, (7) (8) (9) where, , and are the basal magnitudes of the respective parameters. Notably, the strengths of the parameters WPP, WPI and τIN are modeled here to linearly increase with the increase in D1Ract.

D1R stimulation leads to increase in the excitability of GABAergic interneurons by causing decrease in the potassium channel conductance [68]. Therefore, increase in τIN with the increase in D1Ract leads to slower spontaneous decay of the activity of interneuron population and reflects increase in the population excitability. Further, at excitatory synapses, D1R stimulation causes increase in the conductance and decay time constant of the NMDA receptors whereas it leads to slight reduction in the AMPA receptor-mediated postsynaptic currents [69]. In fact, this effect on NMDA receptors is pivotal to the robust sustained-firing activity in the cortical network [71, 72]. As mentioned above, WPP and WPI, both are the strengths of excitatory synapses involved in the recurrent excitation of pyramidal neurons and the excitation of inhibitory interneurons, respectively. It is evident that these synaptic efficacies as such do not differentiate between the AMPA and NMDA receptor-mediated synaptic currents. However, the increase in WPP and WPI with rising D1Ract is meant to achieve the increase in excitatory synaptic transmission naturally occurring due to the prolonged charge transfer under the increased NMDA receptor conductance as well as time constants of the NMDA receptor-mediated currents. This efficiently leads to the enhancement of the self-excitation of pyramidal population and the excitation of interneuron population, which engenders sustained-firing activity in the present modeling framework. Accordingly, the increase in the synaptic efficacies with the increase in D1R stimulation manifests into the form of synaptic plasticity [13]. Nonetheless, D1R stimulation also causes increase in the excitability of pyramidal neurons by decreasing the threshold of depolarization by the persistent sodium current (INaP) and simultaneously reduces the inactivating potassium currents (IK+) [13]. However, contrary to the case of interneuron excitability, the parameter τPN, representing pyramidal population excitability, has not been conceived here to increase with increase in D1Ract. Rather, this effect is compensated through an appropriate magnification of WPP. Owing to the fact that the pyramidal population has a term of self-amplification of their activity, decrease in spontaneous decay of its population activity under high excitability can be conceived through relatively stronger recurrent excitation and an additional term of D1Ract-dependence could be dropped for the tractability of the model.

Therefore, the present model not only considers the direct modulation of pyramidal neurons through D1Rs located on them but also indirect modulation through GABAergic transmission. A glossary of the key variables and the free parameters of the model is available in Table 1.

thumbnail
Table 1. The definitions of the key dynamical variables and the free parameters of the closed-loop mesocortical model.

https://doi.org/10.1371/journal.pone.0198136.t001

The magnitudes of the various parameters in the model are available in Table 2. The parameters of the cortical dynamics have been computed by establishing equivalence of the system of coupled differential equations (Eqs 1 and 2) for cortical neuronal populations to the set of differential equations for population-activities described in the mean-field approach by Brunel and Wang [7]. The remaining parameters of the dynamics of DA neuron population, DA release and D1R stimulation are calibrated in a trial-based manner to acquire the modulation output of the cortical activities known during delay [7] and of the associated empirical observations of cortical DA level [49].

thumbnail
Table 2. List of parameters present in the mathematical model and its stochastic framework along with their values.

The parameters with values in bold font are the free parameters varied in the present study.

https://doi.org/10.1371/journal.pone.0198136.t002

Equilibrium analysis and WM-robustness

The delay-associated state of the mesocortical dynamics is characterized by its global steady or equilibrium state, which is defined as (10) where and represents the set of state-variables. In this regard, the nullcline plots of the state-variables aPN and D1Ract in the aPN-D1Ract state-space are obtained first (S1 Fig). The intersection points of the aPN- and D1Ract-nullclines define the operational points of mesocortical dynamics during the delay period for a given set of parameters RDA and D1Rsens. Accordingly, the nullcline analysis facilitates the obtainment of the bifurcation plots of the state-variables by varying RDA under a fixed D1Rsens (Fig 2).

thumbnail
Fig 2. The delay-associated state of the mesocortical dynamics is characterized by the global equilibrium state of its various dynamical elements.

(A) Given a fixed value of D1R-sensitivity D1Rsens (here D1Rsens = 3, normal control), the bifurcation profiles of the dynamical elements are shown with DA releasability RDA as the bifurcation parameter. Critical RDA, and the corresponding critical cortical dopamine content [DA] and D1R stimulation level D1Ract, mark the beginning of bistable regime favoring the working memory maintenance during delay period. The higher stable states of the bifurcation profiles are together associated with the sustained-firing state of the cortical dynamics whereas the lower stable states together signify the basal spontaneous-activity state. The ranges of [DA] and D1Ract spanned by their higher stable states represent the spans or windows of cortical DA content and D1R stimulation, respectively, underlying the entire modulation profile of the cortical dynamics. The maximum limit to which [DA] or D1Ract may may increase with increase in RDA marks the saturation level. The cue-threshold in the aPN bifurcation profile signifies the minimum excitation of the pyramidal population by cue input, which causes switching to the sustained-firing state. (B) Alteration in D1Rsens further affects the bifurcation profiles. Most prominently, increase in D1Rsens causes leftward shift of the bifurcation region.

https://doi.org/10.1371/journal.pone.0198136.g002

Further, the mesocortical dynamics is constantly affected by the various natural sources of noise in the neural system [73]. Therefore, the stochastic framework of the mesocortical dynamics [36] is given by, (11) (12) (13) (14) Here, represent the right-hand sides of the Eqs (1), (2), (4) and (5), respectively. {dWi (t), t ≥ 0}, (i = 1, 2, 3, 4), denotes the Wiener process increment to each state-variable during their noisy temporal-evolution and σi, (i = 1, 2, 3, 4), represents the corresponding noise-intensity. The magnitudes of the noise-intensities applied here are available in the Table 2 and are kept conserved throughout the study. The noise causes the state of the system to diffuse around its deterministic response and the state-variables are essentially characterized by their statistical distributions in the state-space.

To gain insight into the WM-robustness during delay period, a global potential landscape of the stochastic mesocortical dynamics is constructed. For this, the steady-state marginal probability distributions of the state-variables Pst(x), where x ∈ {aPN, aIN, aDN, [DA], D1Ract}, are obtained from the numerical simulation of the stochastic mesocortical dynamics using the Euler-Maruyama scheme [74] with a fixed time-step Δt. Consequently, a joint probability distribution Pst(aPN, D1Ract) is obtained over the state-space aPN-D1Ract and the global potential landscape, U(aPN, D1Ract) of the stochastic mesocortical dynamics is constructed as [36], (15)

The landscape is comprised of two basins of attractions associated with the spontaneous-activity state and the sustained-firing state of the mesocortical dynamics. The robustness of the WM-associated circuit dynamics is analyzed based on the two physical measures, potential barrier (PB) and the signal-to-noise ratio (SNR) of the pyramidal activity aPN, related with the geometry of the basin associated with the sustained-firing state. PB signifies the depth of the basin from the crest potential separating the two basins of attraction in the landscape and can be directly obtained from the U(aPN, D1Ract). However, the SNR is affected by the girth of the basins and is given by, (16) where, μ denotes the mean of the aPN distribution and corresponds to the deterministic equilibrium magnitude of aPN associated with the sustained-firing state and σ denotes the standard deviation of the noisy fluctuations around the mean aPN. The mathematical analysis and numerical simulations have been performed in MATLAB (The MathWorks). The scripts for the nullcline analysis, the bifurcation profiles and the numerical simulation of the stochastic dynamics are available on the ModelDB, https://senselab.med.yale.edu/modeldb/ShowModel.cshtml?model=240382.

Results

Features of mesocortical dynamics facilitating WM maintenance during delay

The parameters DA-releasability (RDA) and D1R-sensitivity (D1Rsens) of the model serve as the free parameters or handles for realizing here the alterations in cortical DA content and sensitivity, respectively. Notably, RDA signifies the volume-averaged rate or efficiency of DA influx from dopaminergic projections into the cortical extracellular space. Although change in DA-releasability has indeed been observed to affect cognitive performance in the earlier studies involving administration of psychostimulant drugs such as amphetamine and phencyclidine [11, 75], the exact quantification of this rate of DA influx could not have been possible. Accordingly, RDA is varied here within a range of 0.00–0.05nM.ms−1, which is found suitable to capture the experimentally-observed profile of DA-dependent modulation of cortical persistent activity [7] within the present model framework.

Similarly, D1Rsens regulates the sensing-end of the process of dopaminergic transmission. Although D1R-sensitivity is experimentally measured in terms of BP (a dimensionless quantity), alteration in D1Rsens has been scaled here to an integer interval of 2–10. It must be noted that alteration in D1R-sensitivity does not generally imply alterations in the intracellular signalling of D1R activation. Therefore, the parameters in Eqs 79, which govern the excitability of neuronal populations and excitatory or inhibitory synaptic efficacies in response to a given D1R stimulation level (D1Ract), remain unaffected when D1Rsens is varied.

The study begins here with noting the salient features of the delay-associated responses of the various quantifiable variables embedded in the proposed dynamical framework towards change in RDA, while the other free parameter D1Rsens is kept fixed at a particular value. This configuration would physiologically correspond to the change in cortical DA content and the associated changes in the cortical as well as VTA neuron dynamics, under a control normal D1R-sensitivity of the cortex within the present modeling framework. Fig 2A shows the firing frequencies of different neuronal populations (aPN for pyramidal neurons, aIN for interneurons and aDN for DA neurons), extracellular cortical DA level ([DA]) and level of cortical D1R stimulation (D1Ract) during delay period at different values of RDA, for the D1Rsens = 3. The profile of each quantity exhibits a bifurcation behavior. The set of lower values provide the basal magnitude of the quantity associated with spontaneous-activity state in the cortex whereas that of the higher values provide the magnitude associated with sustained-firing activity. The monostable region is characterized by a single stable equilibrium state associated with spontaneous-activity in the cortex. Hence, for the values of RDA within the monostable region, sustained-firing in the cortex is biophysically not feasible. Only in the bistable region, sufficiently strong cue stimulus can cause the switching of the mesocortical dynamics to the sustained-firing state. Therefore, the initiation point of bifurcation signifies the critical RDA, which marks the boundary of phase transition from a region devoid of sustained firing to that of WM maintenance. Accordingly, [DA] and D1Ract associated with the critical RDA indicate the critical DA content and D1R stimulation level required to commence the regime of sustained firing. Notably, RDA naturally comes forth as the bifurcation parameter because its variation, under a fixed D1Rsens, leads to change in [DA] and associated D1Ract, which eventually causes modulation of the neuronal activities during delay.

Remarkably, the increase in RDA does not lead to an unlimited increase in the sustained firing-associated [DA] and D1Ract during delay. The maximum level to which they may rise is marked by their unique saturation levels (Fig 2A). This limitation is of purely functional nature imposed by the mesocortical dynamics during steady-state of the sustained-firing activity in cortex. Therefore, together with the critical [DA] and D1Ract, the corresponding saturation levels define the spans or windows of cortical DA content and D1R stimulation, respectively, which underlie the entire dopaminergic modulation profiles of the neuronal activities in the bistable region.

Nonetheless, in the bistable region, the modulation profile of sustained aDN activity remains in phase with that of the aPN (Fig 2B) as it is the pyramidal activity which directly governs the excitation of DA neuron subpopulation in the VTA within the present mesocortical framework (Fig 1C). However, there exists a phase-lag between the modulation profiles of sustained aPN and aIN activities. In fact, this has also been noted in the earlier studies [33, 76] and the increase in the interneuron excitability by D1R stimulation has been proposed to lag behind that of the pyramidal neurons with respect to increase in cortical DA content and D1R stimulation level.

The levels of spontaneous and sustained activities of the various types of neuronal populations involved here closely resemble their empirically-known estimates during delay. aPN and aIN display spontaneous activities at 3Hz and 9Hz, respectively, during delay (Fig 2A), which are of the order of the average spontaneous activities of pyramidal neurons and fast-spiking GABAergic interneurons observed in the experiments carried out by Wilson et al. [77] on monkeys performing oculomotor tasks. Similarly, the modulation profiles of the sustained-firing activities (the higher stable states) in these neuronal populations span the frequency ranges 13–25Hz and 10–13Hz, respectively, which are in concordance with the earlier computational studies by Compte et al. [78] and Brunel and Wang [7] involving detailed neural network simulations. Moreover, the experimental study by Tsujimoto and Sawaguchi [79] involving delayed WM tasks also provides a similar range of these modulation profiles.

DA neurons in the VTA have been experimentally recorded to fire tonically at an approximate frequency of 3–4Hz under basal or resting condition in delayed-response tasks [40, 42, 63]. Accordingly, the firing rate of the spontaneous activity in DA population aDN is obtained here at 3Hz. As argued above, the tonic firing activity in the VTA-residing DA neuron subpopulation closely associated with a local cortical network may increase in response to the sustained activity in the DLPFC during delay. However, it is also demanded that this increase should remain under the bound of the maximum tonic frequency of 10Hz noted earlier [63]. Therefore, the modulation profile of the sustained tonic aDN is observed here to span a frequency range limited by 10Hz (Fig 2A).

The basal DA concentration in the spontaneous-activity state is obtained here as [DA] = 0.2nM (Fig 2A), which is close to the basal DA concentrations observed in the micro-dialysis studies performed by Watanabe et al. [49] (0.098 ± 0.013nM) and Jedema et al. [75] (0.31 ± 0.03nM) on primates during resting conditions. [DA] associated with sustained-firing activity in the cortex during delay (higher stable state) is observed to increase with rise in RDA (Fig 2A). In this regard, Watanabe et al. [49] reported approximately 17% increase in the DA concentration in the DLPFC of healthy monkeys performing more than 98% successful trials during delayed alternation tasks. This increase in DA characterizes an optimum WM maintenance, which is also found to be associated with optimum strength or frequency of sustained-firing activity in the cortex during delay interval [80]. Accordingly, the peak aPN sustained-activity coincides here with [DA] = 0.234nM (Fig 2A), equivalent to the DA increase reported by Watanabe et al. [49] under optimum performance, only for D1Rsens = 3. Therefore, the corresponding RDA = 0.0058nM.ms−1 and D1Rsens = 3 together portray a normal healthy control in terms of the free parameters of the present model framework. Any increase or decrease in these values of RDA and D1Rsens would represent an altered condition of DA-releasability and D1R-sensitivity, respectively. Subsequently, the effects of alteration in the cortical D1R-sensitivity on WM maintenance are observed through the effects on the abovementioned features of the mesocortical dynamics.

Effects of variation in D1R-sensitivity on cortical DA level and modulation of neuronal activities

Variation in D1Rsens significantly affects the bifurcation plots (Fig 2B). Its increase causes leftward shift of the profiles to lower DA-releasability (RDA). Consequently, increase in D1Rsens leads to a considerable decrease in the critical RDA and the critical [DA] (Fig 3A & 3B). Notably, the variations in critical RDA and [DA] follow a strong positive correlation (Fig 3C) depicting a tight causality-relationship between them. For the control D1Rsens = 3, the critical [DA] = 0.207nM. However, the critical [DA] decreases by 30% when D1Rsens is increased to 10 whereas increases by 50% across unit reduction in the control D1Rsens. This leftward shift of the bifurcation profiles is due to the enhanced sensitivity of D1Rs to respond even to a less amount of DA in the surrounding medium, a consequence also hypothesized earlier for increased D1R density [23, 24, 26], and signifies possibility of WM-associated sustained activity even at lower cortical DA levels. Interestingly, the amount of leftward shift observed by increasing D1Rsens from the normal control level of 3 to 10 is equivalent to that of the rightward shift occurring through only a unit decrease in D1Rsens from the control level. It suggests that even a slight decrease in the cortical D1R-sensitivity may mark a stronger impact on the WM maintenance than a relatively significant increase in the sensitivity.

thumbnail
Fig 3. Effects of variation in D1R-sensitivity on the critical DA releasability, on the critical as well as saturations levels of cortical DA content, and on the modulation-associated windows of DA content and D1R stimulation.

Increase in D1Rsens causes significant decrease in the critical RDA (A) and [DA] (B) marking an early beginning of the bifurcation regime. The variations in critical RDA and [DA] (C) exhibit a strong positive correlation. Moreover, the saturation level of [DA] (D) significantly decreases with increase in D1Rsens, causing the modulation-associated window of DA (E) to shift to lower values as well as shrinks in its span. However, the modulation-associated window of D1R stimulation (F) does not vary with change in D1Rsens.

https://doi.org/10.1371/journal.pone.0198136.g003

Further, increase in D1Rsens significantly reduces the [DA] saturation level (Figs 2B and 3D). As a result, due to the concomitant decrease in the critical as well as saturation levels of [DA], the DA window underlying the entire modulation phenomenon in the bistable region shifts to lower values and also shrinks in its span (Fig 3E) with rise in D1R-sensitivity. With respect to the control D1Rsens = 3, there occurs almost 27% decrease in the size of modulation-associated [DA] window when D1Rsens is increased to 10. At the same time, the amount of shift of the window to lower [DA] is almost 30%, which is the percentage decrease in the critical [DA] mentioned above. However, the window size increases by almost 200% (i.e. doubles in size) when D1Rsens is reduced to 2.

These observations clearly describe the impact of cortical D1R density on the regulation of DA release under the local administration of pyschostimulants in the cortical region studied by Tanaka and Okada [81]. Their study shows that, when the cortical D1R density is upregulated, the DA release is significantly reduced owing to the declined pyramidal activity. As a result, this does not allow the psychostimulants to cause any increase in the cortical DA content. Therefore, the cortical region intrinsically tends to attain a hypodopaminergic situation, which is illustrated here as the shift of modulation profiles to lower RDA and the shift of modulation-associated DA window to lower DA levels.

The critical and the saturation levels of D1Ract remain unaffected (Fig 2B) from changing D1Rsens. Therefore, the D1R stimulation window underlying the entire modulation phenomenon in the bistable region remains completely unaffected (Fig 3F) from D1Rsens alterations. Instead, it only influences how sharply the D1Ract responds to the change in [DA] associated with variation in RDA and reaches its saturation level (Fig 2B).

The observed effects of D1R-sensitivity on the modulation-associated DA and D1R windows further noticeably influences the modulation profiles of delay-associated sustained activities in the different neuronal populations. Owing to the invariant D1Ract, the respective ranges of magnitude spanned by the modulation profiles of sustained activities aPN, aIN and aDN, viz. 13–25Hz, 10–13Hz and 6–10Hz, respectively, remain conserved with the variation in D1Rsens (Fig 2B). In fact, D1R stimulation level is the immediate driver of the modulation of these neuronal activities. However, besides the leftward shift of the profiles towards lesser RDA and [DA] noted above, the sharpness of the modulation profiles of sustained activities in response to change in RDA considerably increase at higher D1Rsens across all neuronal populations. Moreover, the phase-lag between the peak sustained aPN and aIN activities in terms of [DA] significantly decreases with increase in D1Rsens (Fig 4A) and indicates lesser difference in the cortical DA required for D1R-mediated enhancement of the pyramidal and interneuron excitability. This decrease in phase-lag essentially emanates from the observed shrinkage in the DA span underlying modulation (Fig 3E) at higher D1Rsens. However, the phase-lag with respect to D1Ract remains unaffected (Fig 4B), again due to the absence of effect of D1Rsens on modulation-associated D1Ract span. This decrease in the phase-lag and increase in the sharpness of the modulation profiles of sustained-activities in the neuronal populations at higher D1R-sensitivity critically affect the optimal range of cortical DA content underlying optimal WM maintenance, as described below.

thumbnail
Fig 4. Effects of variation in D1R-sensitivity on the phase-lag between the dopaminergic modulation profiles of sustained pyramidal and interneuron activities.

(A) The phase-lag between the peak aPN and the peak aIN activities with respect to the associated [DA] levels is seen to considerably decrease with increase in D1Rsens signifying a steeper modulation of the neuronal activities with unit change in [DA]. (B) However, the phase-lag with respect to the associated D1Ract levels does not vary.

https://doi.org/10.1371/journal.pone.0198136.g004

In the earlier studies involving D1R agonists and antagonists, it has been noted that the strength of sustained-firing activity [32, 82] and WM performance [11, 12] both exhibit inverted-U shaped profile with variation in the level of D1R stimulation. Accordingly, both are highly correlated such that a poor performance is often associated with poor persistent activity in the PFC [33]. Recent studies have provided strong evidences for a linear relationship between them [80, 83, 84]. Accordingly, a symmetric span around the peak sustained aPN activity in the modulation profile (Fig 5A) is chosen such that activity greater than or equal to 80% of the peak activity is assumed to facilitate sound WM maintenance during delay. Hence, the range of cortical DA facilitating this span of optimal sustained pyramidal activity signifies the ‘optimal DA window’. It is observed that the optimal DA window substantially shrinks with rise in D1Rsens (Fig 5B). In the present study, this optimal DA window shrinks to 30% of the normal control with increase in D1Rsens to 10.

thumbnail
Fig 5. Effects of variation in D1R-sensitivity on the range of optimal DA facilitating optimal WM maintenance.

(A) An illustration for the concept of optimal DA range or window associated with the region of optimal sustained aPN activity. It is assumed here that the sustained pyramidal activity above 80% of the peak activity in the modulation profile facilitates efficient WM maintenance. (B) The optimal DA window is seen to considerably shrink and shift to lower values as the D1Rsens is increased.

https://doi.org/10.1371/journal.pone.0198136.g005

Shrinking of optimal DA window demonstrates a smaller range of cortical DA content over which optimal WM maintenance could be acquired. Therefore, even weak natural fluctuations in DA-releasability and the resulting extracellular DA would be able to shift the dynamics to poor maintenance and may have dramatic effects on the cognitive ability. This observation supports the earlier hypothesis [23, 24, 26] that alteration in cortical D1R density has been suggested as a potential factor affecting the optimal region of WM maintenance in schizophrenia. Although the estimate (> = 80%) set here for the boundary of optimal sustained pyramidal activity is merely for the purpose of demonstration, the observation regarding narrowing of the optimal DA window with increase in D1R-sensitivity will remain unaffected regardless of the different estimates one may choose.

Effects of variation in D1R-sensitivity on the robustness of WM maintenance

The global potential landscape (Fig 6) is procured from the steady-state of the noisy mesocortical dynamics (Eqs 1114). The features of WM-robustness under different conditions of DA-releasability (RDA) and D1R-sensitivity (D1Rsens) are derived simultaneously from two physical measures related with the geometry of WM-associated basin of attraction. First, the potential barrier (PB) emanates from the depth of the basin and restricts the noise-induced transition of circuit dynamics from the sustained-firing state to the spontaneous-activity state. Second, the signal-to-noise ratio (SNR) of sustained-firing activity manifests from the girth of the basin and illustrates the strength of the sustained-firing activity relative to its noise content. WM-robustness is directly proportional to both these measures.

thumbnail
Fig 6. The global potential landscape of the noisy mesocortical dynamics.

For the normal control parameters DA-releasability (RDA = 0.0058nM.ms−1) and D1R-sensitivity (D1Rsens = 3) of the mesocortical dynamics, the global potential landscape is shown over the aPN-D1Ract plane, along with its contour projection onto the plane. The system in sustained-firing state is depicted by a ball sitting in the corresponding basin of attraction whose depth provides the potential barrier (PB) restricting the noise-induced transition of the system to spontaneous-activity state. The contour projection illustrates the fluctuation size in the system state around its mean point, which governs the signal-to-noise ratio (SNR) of the cortical sustained activity facilitating WM maintenance.

https://doi.org/10.1371/journal.pone.0198136.g006

We begin with exploring the features of WM-robustness which remain intact despite alterations in D1Rsens. This involves examining the effect of varying RDA on the WM-robustness and essentially projects the impact of alteration in cortical DA content [DA] on WM-robustness mediated through change in the underlying level of D1R stimulation D1Ract during delay. Variations in PB and SNR along the modulation profile of sustained-firing activity aPN in the bistable region (Fig 2A) always depict a concave profile of WM-robustness (Fig 7A & 7C), similar to the shape of the modulation profile itself. Accordingly, it illustrates a tight relationship between the firing frequency of sustained-firing activity and WM-robustness. Moreover, if the aPN modulation profile is partitioned into two sections, the pre-peak set and the post-peak set (including the peak sustained activity), the average PB and SNR of the post-peak set are substantially higher than that of the pre-peak set (Fig 7B & 7D). Notably, this suggests that the post-peak set which involves higher D1Ract as well as an inhibition-dominated cortical dynamics is much more robust than the pre-peak set involving relatively lesser D1Ract and an excitation-dominated cortical dynamics (Fig 2A). These observations have a remarkable similarity with that obtained in the earlier theoretical studies [6, 7, 9, 36, 85] involving change in the D1R stimulation level assumed to occur through alteration in cortical DA content during delay. However, the present investigation further shows that these specific features also remain identically conserved across alterations in D1R-sensitivity. Furthermore, sustained aPN activity more than or equal to 80% of the peak sustained activity in the aPN modulation profile noticeably share high levels of WM-robustness (Figs 8A & 8B and 9A & 9B). This suggests that the optimal region in the aPN modulation profile associated with the optimal DA window is not only defined by its optimal levels of sustained-firing activity but also by the optimal WM-robustness during delay.

thumbnail
Fig 7. The conserved features of WM-robustness across variation in D1R-sensitivity.

(A) For the different D1Rsens, PB for the sampled levels of sustained pyramidal activities along the aPN-modulation profile always follows a concave profile. The sampled activities from the pre-peak side of the aPN-modulation profile are marked with color-filled squares and that from the post-peak side are shown in color-filled circles. The sampled activities, 90% (cyan), 80% (blue), 70% (magenta), 60%(red), are percentage activities with respect to the peak 100% (green) sustained activity. (B) The average PB of the post-peak set of sustained activities (including the peak activity) in the aPN-modulation profile is always higher than that of the pre-peak set for every D1Rsens. (C) Similarly, SNR for the sampled levels of sustained pyramidal activities always follows a concave profile under different condition of D1Rsens. (D) Moreover, the average SNR of the post-peak set of sustained activities is always higher than that of the pre-peak set for all values of D1Rsens.

https://doi.org/10.1371/journal.pone.0198136.g007

thumbnail
Fig 8. Effects of variation in D1R-sensitivity on the WM-robustness in terms of potential barrier (PB).

Increase in D1Rsens causes a consistent decrease in the PB of any individual level of sustained activity either sampled from the pre-peak (A) or from the post-peak (B) set of the modulation profile of cortical sustained aPN activity. The percentage activities are with respect to the peak (100%) sustained activity. (C) The percent decrease in the average PB of pre-peak and post-peak sets across increase in D1Rsens shows higher vulnerability of the pre-peak set to change in D1R-sensitivity.

https://doi.org/10.1371/journal.pone.0198136.g008

thumbnail
Fig 9. Effects of variation in D1R-sensitivity on the WM-robustness in terms of signal-to-noise ratio (SNR).

Similar to the PB, increase in D1Rsens causes a consistent decrease in the SNR of any individual level of sustained activity either sampled from the pre-peak (A) or from the post-peak (B) set of the modulation profile of cortical sustained aPN activity. The percentage activities are with respect to the peak (100%) sustained activity. (C) The percent decrease in the average SNR of the pre-peak and post-peak sets across increase in D1Rsens indicates higher vulnerability of the pre-peak set to change in D1R-sensitivity.

https://doi.org/10.1371/journal.pone.0198136.g009

Next, we examine the D1Rsens-sensitive features of WM-robustness. The entire concave profile of robustness, either in terms of PB (Fig 7A) or SNR (Fig 7C), exhibits a downward shift to lower levels when the D1Rsens is increased. This is also seen through a consistent decrease in PB (Fig 8A & 8B) and SNR (Fig 9A & 9B) of the individual sustained-firing activities of different firing strengths sampled across the aPN modulation profile. Consequently, average PB and SNR of the pre-peak as well as the post-peak set of sustained activities in the aPN modulation profile also decrease. Together, these observations illustrate a concomitant rise in instability of the WM maintenance during delay with increase in D1Rsens. Nonetheless, the amount of decrease in the average PB and SNR (Figs 8C and 9C) is higher for the pre-peak set in comparison to the post-peak set. This differential response to increase in D1Rsens immediately indicates that the robustness of sustained-firing activities during delay resulting from lower D1Ract and excitation-dominated excitation-dominated cortical dynamics is more vulnerable to alteration in D1Rsens. However, sustained activities associated with higher D1Ract and inhibition-dominated cortical dynamics is more resistant to decrease in robustness inflicted by increase in D1Rsens.

As noted above, D1Rsens does not affect the span of D1Ract which underlies the modulation of sustained neuronal activities during delay (Figs 2A and 3f). Therefore, the observed effects of D1Rsens on the WM-robustness is certainly not mediated through the known conventional mechanisms involving the D1R stimulation level [6, 7, 9]. However, what varies across the D1Rsens is the [DA] underlying underlying the conserved D1Ract. Therefore, the observed effects on the WM-robustness appears to be essentially mediated through the impact of D1R-sensitivity on the modulation-associated DA window (Fig 3E). More specifically, it appears to arise from the shift of modulation-associated DA window to lower levels with rising sensitivity. Immediately, a completely new role of cortical DA content in shaping the WM-robustness is realized under the conditions of varying D1R-sensitivity, where sustained-firing activity acquired at a particular level of D1R stimulation but at lower cortical DA content would be lesser robust than that acquired at the same D1R stimulation but at higher cortical DA content.

Discussion

Using the neural mass model of the prefronto-mesoprefrontal system, the present study provides a mechanistic description of how cortical D1R-sensitivity may critically influence WM maintenance during delay and manners in which altered sensitivity may harm the cognitive ability. The two striking features of D1R-sensitivity are its tight control over the level as well as size of optimal DA span facilitating optimum strength of sustained cortical activity during delay and the resulting impact on the robustness of sustained-firing activity against annihilation due to noisy perturbations. An important point to be noted is that the model takes into account only functional alterations in the dopaminergic synaptic transmission and does not consider anatomical alterations in the circuit’s connectivity.

Significance and limitations of the neural mass model of the mesocortical dynamics

Highly detailed cortical network models [6, 7, 10, 71] are already available to elaborate the D1R-dependent dopaminergic modulation of cortical persistent activity during WM maintenance. These studies investigate the role of every minute component of the neuronal excitability as well as synaptic transmission in the network’s firing activity. By taking into account the empirical observations on the effect of D1R stimulation on these components [6668, 70], the theoretical studies have laid down the fundamental picture of the biophysical driving forces behind the dopaminergic modulation in the PFC. However, the present issue with D1R-sensitivity naturally demands consideration of a more comprehensive prefronto-mesoprefrontal machinery, which involves the additional dynamics for regulating the cortical DA content in close association with the prefrontal activity. However, many quantitative intricacies of the dynamics within VTA, regulation of DA neurons activity by prefrontal cortex, involvement of tonic versus phasic activity of DA neurons and local cortical regulation of DA content are still sufficiently missing to construct an appreciably detailed network model of the mesocortical circuitry. Moreover, a detailed network model for this large a system would not only be cumbersome for computation but also be intractable for understanding its consequent dynamics.

Under such circumstances, a neural mass model of the mesocortical circuit may prove an effective framework. Such models conceptualize the bare essentials behind a system’s dynamics distilled out from its natural complexity [86, 87]and the details may be carefully amalgamated into minimal factors required for capturing the system’s original dynamics. Nonetheless, the physical quantities of interest in the present inquiry, viz. intensity of sustained-firing activity during delay and its robustness, are the functional features shaped at the population-level, instead of single independent neurons of the local network. Therefore, the mass model approach fits in well for addressing these issues. As a result, despite its simplicity and tractability, the model effectively captures salient attributes of the phenomenon of dopaminergic modulation as observed in the earlier experimental and theoretical studies.

The bifurcation profile observed in the earlier computational study by Brunel and Wang [7] (see Fig 10 of the study) exhibits loss of bistability at higher levels of D1R stimulation and, hence, is different from the bifurcation profile of cortical pyramidal activity obtained here with respect to the releasability parameter RDA (Fig 2A). The former study considers an isolated cortical module and an independent parametric variation in the level of D1R stimulation. However, the level of D1R stimulation is not an independent parameter in the present study. Rather, it is shown to be strictly regulated by the extracellular cortical DA content, which is further governed by the cortical firing activity in a feed-back manner. Since the interactions at the mesocortical scale results into saturation in the cortical DA content with parametric increase in the RDA, there exists a limit to which stimulation of D1Rs can increase. Within this limit, the bistable bifurcation profile does not vanish. However, the nullcline analysis of the isolated cortical dynamics in the present dynamical model involves independent variation in D1R stimulation level (S1 Fig). Accordingly, the nullcline profile of the pyramidal activity obtained here indeed shows the loss of bistability at higher D1R stimulation, akin to the Brunel and Wang observation.

The earlier experimental studies on the dopaminergic modulation of cortical persistent activity have also applied controlled variation in the local cortical concentrations of D1R agonist and antagonist using iontophoretic techniques [11, 12, 14, 32]. This is similar to the parametric variation in the D1R stimulation in the earlier computational studies as well as the nullcline analysis for the pyramidal activity performed here. Consequently, loss of bistability at higher D1R stimulation has also remained profusely evident in the empirical observations. However, the present study suggests that such a bifurcation profile may not exist in vivo as there is a functional limitation on the rise of cortical D1R stimulation level under physiological conditions.

In reality, there exists multiple local populations of cortical neurons within a small region of the cortex such that each population is tuned to exhibit persistent activity for a specific feature or information of the stimulus presented in a WM task, such as spatial orientations in visuospatial WM tasks [39]. Therefore, there exists simultaneously multiple attractor states. However, while dealing with a single local population of cortical neurons, the present model projects a single attractor state with persistent activity. From the viewpoint of network’s dynamics, failures at the end of the delay period recorded for the behavioral performance of a subject may occur either through the premature collapse of the sustained-firing activity to the spontaneous-activity state or the transition of the firing state to another attractor during delay [84, 88, 89]. As far as the former route is concerned, the present study directly elaborates the ways in which the anomalies in dopaminergic modulation may affect WM-robustness and the behavioral performance. However, it also paves a way to explain the latter route to some extent. The observed increase in the shallowness of the basin of attraction associated with the sustained-firing state causing declined robustness under anomalous conditions of dopaminergic modulation is a biophysical property of the local cortical dynamics. Therefore, the various attractor states relying on the similar dynamical principles of sustained-firing activity would together get shallower and lesser robust under such conditions. In a way, the entire global potential landscape of sustained-firing attractors becomes shallower. Although the intensity of instability may not be identical for all the attractor states, it would be reasonable to envisage that transitions from one sustained-firing attractor to the other would become easier as well as frequent. Accordingly, the WM-robustness shown here in terms of PB and SNR is merely indicative of the WM performance.

Nonetheless, the ongoing discourse regarding the exact role of the sustained-firing activity as a neural correlate of WM maintenance in the PFC is worth considering. It is still under intense debate whether the sustained-firing activity itself stores relevant information about a presented stimulus [88, 89] or it serves as a top-down biasing control over other areas of the cortex, such as posterior parietal cortex (PPC) or inferior temporal cortex (ITC), to aid them in encoding the salient features in their local persistent activities [90, 91]. To some extent, it is quite apparent that the spatial location of the cue presentation in the visuospatial WM tasks is at least encoded by the PFC circuitry whereas other features of the visual stimulus have been suggested to be represented in the PPC, which normally responds towards these specific sensory stimulus besides WM tasks [89, 90]. Accordingly, complex visuospatial WM demands may simultaneously involve storage of information as well as top-down biasing by the sustained-firing activity in the PFC. However, the quality of sustained-firing activity in the PFC in terms of its mean firing frequency and robustness against noisy fluctuations is essential for the eventual behavioral performance of a subject undergoing WM task, regardless of which route it takes to shape the WM maintenance during delay. In fact, some studies have shown that robustness is a unique feature of PFC microcircuitry which avoids the loss of goal-directed memory in the presence of distractors whereas the other cortical areas lack this attribute [89]. Additionally, the robustness of persistent activity in the PFC has been observed as a requirement for the biasing control over the stable representations in PPC. Therefore, the observations made here are equally applicable to both the ways through which sustained-firing activity in the PFC may be involved in WM maintenance.

Another important fact is that, in the case of primates, DLPFC as well as medial prefrontal cortex (mPFC) both have been observed to show sustained-firing activity in WM tasks [3]. In this regard, domain-specific hypothesis proposed by Goldman-Rakic [38] suggests that spatial WM features are dealt by DLPFC whereas non-spatial features are dealt by mPFC. Contrastingly, a process-specific model by Petrides [92] hypothesizes that mPFC retrieves information from PPC and DLPFC does the job of monitoring the information. Although the present model involves only DLPFC, it is recognized that the present study is not limited to DLPFC but can also be applied to WM maintenance-associated persistent activity in mPFC. This is due to the fact that both these regions share similar cortical microcircuitry to some extent [37] and, thus, involve a common physical mechanism for the establishment of sustained-firing activity.

Notably, despite the overwhelming evidences for the role of cortical persistent activity in WM maintenance, there are ample empirical observations which also suggest that persistent activity during the entire span of delay is not necessary at all for the WM maintenance [2]. Across trials in a variety of WM tasks [80, 93, 94], it has been observed that a long initial span of the delay after the presentation of cue stimulus sometimes lacks a persistent activity and is rather characterized by a spontaneous activity state. Only in the response preparation phase of the delay, immediately before the response, the persistent activity rapidly appears and leads to a successful trial. This is indeed surprising, unless the spontaneous activity during delay itself stores the goal-directed information. A proposed dynamic coding model of WM maintenance [2] suggests that the heightened activity in a local cortical network at the instance of cue presentation can temporarily energize a hidden activity state of the network through short-term plasticity or coherence. This hidden state possesses a specific pattern of activity which can carry the desired WM information but generally stays in the network as an activity-silent state. At the instance of its cue-induced emergence, it can be transiently adopted by the spontaneous activity state of the network. Accordingly, this mode of WM maintenance during delay is referred to as the activity-silent WM maintenance [2]. It seems that the occurrence of activity-silent mode across several trials differs with the nature of WM tasks and depends particularly on the demand of parallel attention and processing [80]. However, the present model is not equipped with the essential framework to accommodate the activity-silent mode of WM maintenance during delay.

Clinical implications in ageing and schizophrenia

The observed dependence of the various essential features of dopaminergic modulation on DA-releasability and D1R-sensitivity carries potential clinical implications. In the case of ageing, there occurs a substantial decrease in the cortical D1R-sensitivity [16, 17, 21]. Bäckman et al. [17] using PET study estimated a 14% average age-related loss of D1Rs BP per decade in DLPFC. In another PET study, Suhara et al. [21] using [11C]-SCH23390, a highly selective ligand for D1Rs, reported a 39% decrease in D1Rs BP in the frontal cortex with age. Keyser et al. [16] also observed a significant decrease in D1R density and reactivity of their high affinity sites in the frontal cortex with age. Interestingly, decrease in D1R-sensitivity is observed here to be associated with wider range of optimal DA content and relatively higher robustness of WM maintenance. However, these benefits are strongly counteracted by large shifts of the WM regime of cortical dynamics to higher DA levels. Here, with decrease in D1R-sensitivity, the associated optimal range of DA appears more and more unapproachable by the normal levels of DA-releasability of the mesocortical projections. The situation becomes more severe as the DA-releasability also exhibits a decline in ageing [16, 21, 36, 95]. Therefore, ageing may end up either in a complete loss of WM maintenance or a poor WM maintenance depending on the severity of D1Rs depletion as well as decline in the DA-releasability.

A contrary situation is witnessed in the case of schizophrenia where a chronic hypodopaminergic state of DLPFC leads to a substantial upregulation of cortical D1R density. PET studies by Abi-Dargham et al. [23, 24, 26] observed that [11C]NNC112 BP was significantly elevated in the DLPFC of unmedicated schizophrenic patients. A postmortem study performed by Knable et al. [22] also reported a significant increase in the BP of [3H]-SCH23390 in the prefrontal cortex of schizophrenic patients as compared to normal controls. Accordingly, it demonstrates the situation of elevated cortical D1R-sensitivity. It is observed here that high D1R-sensitivity causes the WM regime of cortical dynamics to shift to very low levels of DA. At first, it seems a homeostatic mechanism so that WM could be formed even under hypodopaminergic state, as has also been suggested earlier [23, 24, 26]. But this rescue doesn’t seem to be eventually much useful as the schizophrenic patients indeed show impairment of WM maintenance. The present observations suggest that too much responsiveness of cortical dynamics to even a slight change in cortical DA content makes it difficult to stay within the optimal range of DA under the conditions of natural fluctuations in the cortical DA content. This is aided by the fact that the optimal DA window also considerably shrinks with increase in D1R-sensitivity. Moreover, the associated WM-robustness also decreases under such conditions. Further, if there occurs an uncontrolled increase in DA content due to the administration of DA elevating drugs [11] or due to the heavy demand of a WM task [26], the cortical dynamics would easily shift to the very far sections of the post-peak region in the bifurcation profile, which may again lead to poor WM maintenance.

Currently, no well-defined protocol of medication exists for the cognitive deficit associated with DA-dysfunction [15, 96], owing to the limited knowledge of the several factors involved in the dopaminergic modulation of cortical activity. Yet, two genres of drugs are being examined for their medicinal potency: (a) drugs which are pharmacologically D1R agonists and antagonists [15, 20] (b) drugs which modulate the DA release probability of the afferent dopaminergic projections to cortex [25, 97]. The former has a direct role in regulating the cortical D1R stimulation whereas the latter does it indirectly via regulating the dopaminergic condition of cortex. Moreover, an efficient use of these drugs requires a trial-based estimation of the appropriate drug-combination and drug-dosage, which exhibit a huge unpredictable variability across the patients suffering from the same neuropsychiatric disorder [98].

It is shown here that one of the neglected aspects in the current clinical diagnosis, i.e. alteration in D1R-sensitivity, has a strong deterministic contribution to the otherwise unpredictable variability in response to DA-correcting drugs across patients. Features such as critical DA-releasability and cortical DA content required to capacitate cortical circuitry for WM function, modulation-associated DA window, the sharpness of the modulation profiles of neuronal activities, the optimal region of modulation and the associated optimal DA window, are significantly affected by alterations in D1R-sensitivity. This suggests that the drug-mediated tuning of cortical DA content to improve the cortical D1R stimulation based only on the knowledge of dopaminergic condition of the cortex is not sufficient. It should also be accompanied by the diagnosis of the intensity of alteration in D1R-sensitivity inflicted by the pathological condition. In fact, the precision of DA-tuning substantially varies according to the intensity of alteration in D1R-sensitivity and so is the effective drug-dosage [98].

Other clinically important aspects demonstrated here stem from the features of WM-robustness. It is shown that the optimal region of the modulation does not manifest only from the optimal levels of cortical sustained activity but also from the optimal levels of robustness during delay. Moreover, the effect of alteration in D1R-sensitivity on the robustness suggests that even if an optimal cortical sustained activity is achieved by retrieving an optimal cortical DA content, the associated robustness cannot be identically gained if the alteration in cortical D1R-sensitivity is not equally improved. This further indicates that there is no substitute of a remedy for altered D1R-sensitivity condition. The manner in which antipsychotics impact D1R-sensitivity [99] is unclear and therefore, its effect is not under appropriate clinical control. A perfect medication of cognitive deficits emanating from DA-dysfunction would necessarily require an amalgam of strategies which can together alleviate the anomalies in cortical DA content as well as D1R-sensitivity.

Supporting information

S1 Fig. The nullclines of the excitatory population activity, aPN, and the cortical D1R stimulation, D1Ract.

(A) For a given D1Rsens, the solid black curve is the aPN-nullcline and the grey lines are the D1Ract-nullclines for the different % values of DA-releasability, RDA, relative to RDA = 0.0058nM.ms−1. As evident, increase in RDA causes a rightward shift in the D1Ract-nullcline. The point(s) at which a D1Ract-nullcline for a given value of RDA intersects the aPN-nullcline together defines the corresponding operating point(s) of the mesocortical system, where a point marked with solid circle represents the stable state and that marked with open circle represents the unstable state of the system. (B-C) As D1Rsens is increased, the rate of rightward shift in the D1Ract-nullcline in response to variation in RDA considerably increases, which illustrates a heightened response of the mesocortical system to variation in the cortical DA content.

https://doi.org/10.1371/journal.pone.0198136.s001

(TIF)

Acknowledgments

MR acknowledges the research fellowship provided by CSIR, India grant 09/263(0991)/2013-EMR-I. RG acknowledges the research fellowship provided by CSIR, India grant 09/ 263(1057)/2015-EMR-I. We thank Professor Karmeshu for critical discussions on the broad area of stochastic processes and nonlinear dynamical systems.

References

  1. 1. Miller EK. The ‘working’of working memory. Dialogues Clin Neurosci. 2013;15: 411–418. pmid:24459408
  2. 2. Stokes MG. ‘Activity-silent’working memory in prefrontal cortex: a dynamic coding framework. Trends Cogn Sci. 2015;19: 394–405. pmid:26051384
  3. 3. Christophel TB, Klink PC, Spitzer B, Roelfsema PR, Haynes J-D. The distributed nature of working memory. Trends Cogn Sci. 2017;21: 111–124. http://dx.doi.org/10.1016/j.tics.2016.12.007. pmid:28063661
  4. 4. Wang M, Vijayraghavan S, Goldman-Rakic PS. Selective D2 receptor actions on the functional circuitry of working memory. Science. 2004;303: 853–856. pmid:14764884
  5. 5. Floresco SB, Magyar O. Mesocortical dopamine modulation of executive functions: beyond working memory. Psychopharmacology. 2006;188: 567–585. pmid:16670842
  6. 6. Durstewitz D, Kelc M, Güntürkün O. A neurocomputational theory of the dopaminergic modulation of working memory functions. J Neurosci. 1999; 19: 2807–2822. pmid:10087092
  7. 7. Brunel N, Wang X-J. Effects of neuromodulation in a cortical network model of object working memory dominated by recurrent inhibition. J Comput Neurosci. 2001;11: 63–85. pmid:11524578
  8. 8. Cohen JD, Braver TS, Brown JW. Computational perspectives on dopamine function in prefrontal cortex. Curr Opin Neurobiol. 2002;12: 223–229. pmid:12015241
  9. 9. Deco G, Rolls ET. Attention and working memory: a dynamical model of neuronal activity in the prefrontal cortex. Eur J Neurosci. 2003;18: 2374–2390. pmid:14622200
  10. 10. Loh M, Rolls ET, Deco G. A dynamical systems hypothesis of schizophrenia. PLoS Comput Biol. 2007;3(11): e228. pmid:17997599
  11. 11. Murphy BL, Arnsten AFT, Goldman-Rakic PS, Roth RH. Increased dopamine turnover in the prefrontal cortex impairs spatial working memory performance in rats and monkeys. Proc Natl Acad Sci USA. 1996;93: 1325–1329. pmid:8577763
  12. 12. Zahrt J, Taylor JR, Mathew RG, Arnsten AF. Supranormal stimulation of D1 dopamine receptors in the rodent prefrontal cortex impairs spatial working memory performance. J Neurosci. 1997;17: 8528–8535. pmid:9334425
  13. 13. Seamans JK, Yang CR. The principal features and mechanisms of dopamine modulation in prefrontal cortex. Prog Neurobiol. 2004;74: 1–58. pmid:15381316
  14. 14. Williams GV, Castner SA. Under the curve: critical issues for elucidating D1 receptor function in working memory. Neuroscience. 2006;139: 263–276. pmid:16310964
  15. 15. Arnsten AFT, Girgis RR, Gray DL, Mailman RB. Novel dopamine therapeutics for cognitive deficits in schizophrenia. Biol Psychiatry. 2017;81: 67–77. pmid:26946382
  16. 16. de Keyser J, De Backer JP, Vauquelin G, Ebinger G. The effect of aging on the D1 dopamine receptors in human frontal cortex. Brain Res. 1990;528: 308–310. pmid:2148706
  17. 17. Bäckman L, Karlsson S, Fischer H, Karlsson P, Brehmer Y, Rieckmann A, et al. D1 receptors and age differences in brain activation during working memory. Neurobiol Aging. 2011;32: 1849–1856. pmid:19962789
  18. 18. Arnsten AF, Goldman-Rakic PS. Noise stress impairs prefrontal cortical cognitive functions in monkeys: evidence for a hyperdopaminergic mechanism. Arch Gen Psychiatry. 1998;55: 362–368. pmid:9554432
  19. 19. Qin S, Hermans EJ, van Marle HJF, Luo J, Fernández G. Acute psychological stress reduces working memory-related activity in the dorsolateral prefrontal cortex. Biol Psychiatry. 2009;66: 25–32. pmid:19403118
  20. 20. Goldman-Rakic PS, Castner SA, Svensson TH, Siever LJ, Williams GV. Targeting the dopamine D1 receptor in schizophrenia: insights for cognitive dysfunction. Psychopharmacology. 2004;174: 3–16. pmid:15118803
  21. 21. Suhara T, Fukuda H, Inoue O, Itoh T, Suzuki K, Yamasaki T, et al. Age-related changes in human D1 dopamine receptors measured by positron emission tomography. Psychopharmacology (Berl). 1991;103: 41–45.
  22. 22. Knable MB, Hyde TM, Murray AM, Herman MM, Kleinman JE. A post-mortem study of frontal cortical dopamine D1 receptors in schizophrenics, psychiatric controls, and normal controls. Biol Psychiatry. 1996;40: 1191–1199. pmid:8959283
  23. 23. Abi-Dargham A, Mawlawi O, Lombardo I, Gil R, Martinez D, Huang Y, et al. Prefrontal dopamine D1 receptors and working memory in schizophrenia. J Neurosci. 2002;22: 3708–3719. pmid:11978847
  24. 24. Abi-Dargham A, Moore H. Prefrontal DA transmission at D1 receptors and the pathology of schizophrenia. Neuroscientist. 2003;9: 404–416. pmid:14580124
  25. 25. Slifstein M, Kolachana B, Simpson EH, Tabares P, Cheng B, Duvall M, et al. COMT genotype predicts cortical-limbic D1 receptor availability measured with [11C]NNC112 and PET. Mol Psychiatry. 2008;13: 821–827. pmid:18317466
  26. 26. Abi-Dargham A, Xu X, Thompson JL, Gil R, Kegeles LS, Urban N, et al. Increased prefrontal cortical D1 receptors in drug naive patients with schizophrenia: a PET study with [11C]NNC112. J Psychopharmacol. 2012;26: 794–805. pmid:21768159
  27. 27. Slifstein M, van de Giessen E, Van Snellenberg J, Thompson JL, Narendran R, Gil R, et al. Deficits in prefrontal cortical and extrastriatal dopamine release in schizophrenia: a positron emission tomographic functional magnetic resonance imaging study. JAMA Psychiatry. 2015;72: 316–324. pmid:25651194
  28. 28. Peters Y, Barnhardt NE, O’Donnell P. Prefrontal cortical up states are synchronized with ventral tegmental area activity. Synapse. 2004;52: 143–152. pmid:15034920
  29. 29. Kornhuber J, Kim JS, Kornhuber ME, Kornhuber HH. The cortico-nigral projection: reduced glutamate content in the substantia nigra following frontal cortex ablation in the rat. Brain Res. 1984;322: 124–126. pmid:6151416
  30. 30. Sesack SR, Carr DB. Selective prefrontal cortex inputs to dopamine cells: implications for schizophrenia. Physiol Behav. 2002;77: 513–517. pmid:12526992
  31. 31. Björklund A, Lindvall O. Handbook of chemical neuroanatomy: Classical transmitters and transmitter receptors in the CNS, Part II. Amsterdam (North Holland): Elsevier; 1984.
  32. 32. Williams GV, Goldman-Rakic PS. Modulation of memory fields by dopamine D1 receptors in the prefrontal cortex. Nature. 1995;376: 572–575. pmid:7637804
  33. 33. Goldman-Rakic PS, Muly EC III, Williams GV. D1 receptors in prefrontal cells and circuits. Brain Res Brain Res Rev. 2000;31: 295–301. pmid:10719156
  34. 34. Arnsten AFT, Wang MJ, Paspalas CD. Neuromodulation of thought: flexibilities and vulnerabilities in prefrontal cortical network synapses. Neuron. 2012;76: 223–239. pmid:23040817
  35. 35. Tanaka S. Dopaminergic control of working memory and its relevance to schizophrenia: a circuit dynamics perspective. Neuroscience. 2006;139: 153–171. pmid:16324800
  36. 36. Reneaux M, Gupta R, Karmeshu . Stochastic mesocortical dynamics and robustness of working memory during delay-period. PLoS ONE. 2015;10(12): e0144378. pmid:26636712
  37. 37. Douglas RJ, Martin KAC. Neuronal circuits of the neocortex. Annu Rev Neurosci. 2004;27: 419–451. pmid:15217339
  38. 38. Goldman-Rakic PS. Topography of cognition: parallel distributed networks in primate association cortex. Annu Rev Neurosci. 1988;11: 137–156. pmid:3284439
  39. 39. Goldman-Rakic PS. Regional and cellular fractionation of working memory. Proc Natl Acad Sci USA. 1996;93: 13473–13480. pmid:8942959
  40. 40. Schultz W, Apicella P, Ljungberg T. Responses of monkey dopamine neurons to reward and conditioned stimuli during successive steps of learning a delayed response task. J Neurosci. 1993;13: 900–913. pmid:8441015
  41. 41. Fiorillo CD, Tobler PN, Schultz W. Discrete coding of reward probability and uncertainty by dopamine neurons. Science. 2003;299: 1898–1902. pmid:12649484
  42. 42. Grace AA. Dysregulation of the dopamine system in the pathophysiology of schizophrenia and depression. Nat Rev Neurosci. 2016;17: 524–532. pmid:27256556
  43. 43. Floresco SB, West AR, Ash B, Moore H, Grace AA. Afferent modulation of dopamine neuron firing differentially regulates tonic and phasic dopamine transmission. Nat Neurosci. 2003;6: 968–973. pmid:12897785
  44. 44. Dreyer JK, Herrik KF, Berg RW, Hounsgaard JD. Influence of phasic and tonic dopamine release on receptor activation. J Neurosci. 2010;30: 14273–14283. pmid:20962248
  45. 45. Braver TS, Barch DM, Cohen JD. Cognition and control in schizophrenia: a computational model of dopamine and prefrontal function. Biol Psychiatry. 1999;46: 312–328. pmid:10435197
  46. 46. Dreher JC, Burnod Y. An integrative theory of the phasic and tonic modes of dopamine modulation in the prefrontal cortex. Neural Netw. 2002;15: 583–602. pmid:12371514
  47. 47. Smiley JF, Williams SM, Szigeti K, Goldman-Rakic PS. Light and electron microscopic characterization of dopamine-immunoreactive axons in human cerebral cortex. J Comp Neurol. 1992;321: 325–335. pmid:1506472
  48. 48. Smiley JF, Levey AI, Ciliax BJ, Goldman-Rakic PS. D1 dopamine receptor immunoreactivity in human and monkey cerebral cortex: predominant and extrasynaptic localization in dendritic spines. Proc Natl Acad Sci USA. 1994;91: 5720–5724. pmid:7911245
  49. 49. Watanabe M, Kodama T, Hikosaka K. Increase of extracellular dopamine in primate prefrontal cortex during a working memory task. J Neurophysiol. 1997;78: 2795–2798. pmid:9356427
  50. 50. Sesack SR, Hawrylak VA, Matus C, Guido MA, Levey AI. Dopamine axon varicosities in the prelimbic division of the rat prefrontal cortex exhibit sparse immunoreactivity for the dopamine transporter. J Neurosci. 1998;18: 2697–2708. pmid:9502827
  51. 51. Ciliax BJ, Drash GW, Staley JK, Haber S, Mobley CJ, Miller GW, et al. Immunocytochemical localization of the dopamine transporter in human brain. J Comp Neurol. 1999;409: 38–56. pmid:10363710
  52. 52. Lewis DA, Melichitzky DS, Sesack SR, Whitehead RE, Auh S, Sampson A. Dopamine transporter immunoreactivity in monkey cerebral cortex: regional, laminar, and ultrastructural localization. J Comp Neurol. 2001;432: 119–136. pmid:11241381
  53. 53. Martin KAC, Spühler IA. The fine structure of the dopaminergic innervation of area 10 of macaque prefrontal cortex. Eur J Neurosci. 2013;37: 1061–1071. pmid:23331617
  54. 54. Yamamoto BK, Novotney S. Regulation of extracellular dopamine by the norepinephrine transporter. J Neurochem. 1998;71: 274–280. pmid:9648875
  55. 55. Moró n JA, Brockington A, Wise RA, Rocha BA, Hope BT. Dopamine uptake through the norepinephrine transporter in brain regions with low levels of the dopamine transporter: evidence from knock-out mouse lines. J Neurosci. 2002;22: 389–395.
  56. 56. Valentini V, Frau R, Di Chiara G. Noradrenaline transporter blockers raise extracellular dopamine in medial prefrontal but not parietal and occipital cortex: differences with mianserin and clozapine. J Neurochem. 2004;88: 917–927. pmid:14756813
  57. 57. Spühler IA, Hauri A. Decoding the Dopamine Signal in Macaque Prefrontal Cortex: A Simulation Study Using the Cx3Dp Simulator. PLoS ONE 2013;8(8): e71615. pmid:23951205
  58. 58. Grace AA. Phasic versus tonic dopamine release and the modulation of dopamine system responsivity: a hypothesis for the etiology of schizophrenia. Neuroscience. 1991;41: 1–24. pmid:1676137
  59. 59. Bernard V, Somogyi P, Bolam JP. Cellular, subcellular and subsynaptic distribution of AMPA-type glutamate receptor subunits in the neostriatum of the rat. J Neurosci. 1997;17: 819–833. pmid:8987803
  60. 60. Bernard V, Bolam JP. Subcellular and subsynaptic distribution of the NR1 subunit of the NMDA receptor in the neostriatum and globus pallidus of the rat: co-localization at synapses with the GluR2/3 subunit of the AMPA receptor. Eur J Neurosci. 1998;10: 3721–3736. pmid:9875351
  61. 61. Kodama T, Hikosaka K, Watanabe M. Differential changes in glutamate concentration in the primate prefrontal cortex during spatial delayed alternation and sensory-guided task. Exp Brain Res. 2002;145: 133–141. pmid:12110952
  62. 62. Takahata R, Moghaddam B. Glutamatergic regulation of basal and stimulus-activated dopamine release in the prefrontal cortex. J Neurochem. 1998;71: 1443–1449. pmid:9751176
  63. 63. Grace AA, Bunney BS. The control of firing pattern in nigral dopamine neurons: single spike firing. J Neurosci. 1984;4: 2866–2876. pmid:6150070
  64. 64. Bourdy R, Barrot M. A new control center for dopaminergic systems: pulling the VTA by the tail. Trends Neurosci. 2012;35: 681–690. pmid:22824232
  65. 65. Ford CP. The role of D2-autoreceptors in regulating dopamine neuron activity and transmission. Neuroscience. 2014;282: 13–22. pmid:24463000
  66. 66. Lachowicz JE, Sibley DR. Molecular characteristics of mammalian dopamine receptors. Pharmacol Toxicol. 1997;81: 105–113. pmid:9335067
  67. 67. Yang CR, Seamans JK, Gorelova N. Developing a neuronal model for the pathophysiology of schizophrenia based on the nature of electrophysiological actions of dopamine in the prefrontal cortex. Neuropsychopharmacol. 1999;21: 161–194.
  68. 68. Gorelova N, Seamans JK, Yang CR. Mechanisms of dopamine activation of fast-spiking interneurons that exert inhibition in rat prefrontal cortex. J Neurophysiol. 2002;88: 3150–3166. pmid:12466437
  69. 69. Wang J, O’Donnell P. D1 dopamine receptors potentiate NMDA-mediated excitability increase in layer V prefrontal cortical pyramidal neurons. Cereb Cortex. 2001;11: 452–462. pmid:11313297
  70. 70. Seamans JK, Gorelova N, Durstewitz D, Yang CR. Bidirectional dopamine modulation of GABAergic inhibition in prefrontal cortical pyramidal neurons. J Neurosci. 2001;21: 3628–3638. pmid:11331392
  71. 71. Durstewitz D, Seamans JK, Sejnowski TJ. Dopamine-mediated stabilization of delay-period activity in a network model of prefrontal cortex. J Neurophysiol. 2000;83: 1733–1750. pmid:10712493
  72. 72. Wang M, Yang Y, Wang CJ, Gamo NJ, Jin LE, Mazer JA, et al. NMDA receptors subserve persistent neuronal firing during working memory in dorsolateral prefrontal cortex. Neuron. 2013;20: 736–749.
  73. 73. Faisal AA, Selen LPJ, Wolpert DM. Noise in the nervous system. Nat Rev Neurosci. 2008;9: 292–303. pmid:18319728
  74. 74. Kloeden PE, Platen E. Numerical solution of stochastic differential equations. Berlin (Heidelberg): Springer-Verlag; 1992.
  75. 75. Jedema HP, Narendran R, Bradberry CW. Amphetamine-induced release of dopamine in primate prefrontal cortex and striatum: striking differences in magnitude and timecourse. J Neurochem. 2014;130: 490–497. pmid:24749782
  76. 76. Muly EC III, Szigeti K, Goldman-Rakic PS. D1 receptor in interneurons of macaque prefrontal cortex: distribution and subcellular localization. J Neurosci. 1998;18: 10553–10565.
  77. 77. Wilson FAW, O’Scalaidhe SP, Goldman-Rakic PS. Functional synergism between putative γ-aminobutyrate-containing neurons and pyramidal neurons in prefrontal cortex. Proc Natl Acad Sci USA. 1994;91: 4009–4013. pmid:8171027
  78. 78. Compte A, Brunel N, Goldman-Rakic PS, Wang X-J. Synaptic mechanisms and network dynamics underlying spatial working memory in a cortical network model. Cereb Cortex. 2000;10: 910–923. pmid:10982751
  79. 79. Tsujimoto S, Sawaguchi T. Properties of delay-period neuronal activity in the primate prefrontal cortex during memory- and sensory-guided saccade tasks. Eur J Neurosci. 2004;19: 447–457. pmid:14725639
  80. 80. Watanabe K, Funahashi S. Neural mechanisms of dual-task interference and cognitive capacity limitation in the prefrontal cortex. Nat Neurosci. 2014;17: 601–611. pmid:24584049
  81. 81. Tanaka S, Okada S. Functional prefrontal cortical circuitry for visuospatial working memory formation: A computational model. Neurocomputing. 1999;6–27: 891–899.
  82. 82. Vijayraghavan S, Wang M, Birnbaum SG, Williams GV, Arnsten AFT. Inverted-U dopamine D1 receptor actions on prefrontal neurons engaged in working memory. Nat Neurosci. 2007;10: 376–384. pmid:17277774
  83. 83. Wang M, Gamo NJ, Yang Y, Jin LE, Wang XJ, Laubach M, et al. Neuronal basis of age-related working memory decline. Nature. 2011;476: 210–213. pmid:21796118
  84. 84. Wimmer K, Nykamp DQ, Constantinidis C, Compte A. Bump attractor dynamics in prefrontal cortex explains behavioral precision in spatial working memory. Nat Neurosci. 2014;17: 431–439. pmid:24487232
  85. 85. Okimura T, Tanaka S, Maeda T, Kato M, Mimura M. Simulation of the capacity and precision of working memory in the hypodopaminergic state: relevance to schizophrenia. Neuroscience. 2015;295: 80–89. pmid:25818554
  86. 86. Deco G, Jirsa VK, Robinson PA, Breakspear M, Friston KJ. The dynamic brain: from spiking neurons to neural masses and cortical fields. PLoS Comput Biol. 2008;4(8): e1000092. pmid:18769680
  87. 87. Buice MA, Chow CC. Beyond mean field theory: statistical field theory for neural networks. J Stat Mech. 2013;P03003. pmid:25243014
  88. 88. Constantinidis C, Klingberg T. The neuroscience of working memory capacity and training. Nat Rev Neurosci. 2016;17: 438–449. pmid:27225070
  89. 89. Riley MR, Constantinidis C. Role of prefrontal persistent activity in working memory. Front Syst Neurosci. 2016;9: 181. pmid:26778980
  90. 90. Lara AH, Wallis JD. Executive control processes underlying multi-item working memory. Nat Neurosci. 2014;17: 876–883. pmid:24747574
  91. 91. D’Esposito M, Postle BR. The cognitive neuroscience of working memory. Annu Rev Psychol. 2015;66: 115–142. pmid:25251486
  92. 92. Petrides M. Specialized systems for the processing of mnemonic information within the primate frontal cortex. Philos Trans R Soc Lond B Biol Sci. 1996;351: 1455–1461. pmid:8941957
  93. 93. Watanabe K, Funahashi S. Prefrontal delay-period activity reflects the decision process of a saccade direction during a free-choice ODR task. Cereb Cortex. 2007;17: i88–i100. pmid:17726006
  94. 94. Barak O, Tsodyks M, Romo R. Neuronal population coding of parametric working memory. J Neurosci. 2010;30: 9424–9430. pmid:20631171
  95. 95. Goldman-Rakic PS, Brown RM. Regional changes of monoamines in cerebral cortex and subcortical structures of aging rhesus monkeys. Neuroscience. 1981;6: 177–187. pmid:6111765
  96. 96. Lett TA, Voineskos AN, Kennedy JL, Levine B, Daskalakis ZJ. Treating working memory deficits in schizophrenia: a review of the neurobiology. Biol Psychiatry. 2014;75: 361–370. pmid:24011822
  97. 97. Schacht JP. COMT val158met moderation of dopaminergic drug effects on cognitive function: a critical review. Pharmacogenomics J. 2016;16: 430–438. pmid:27241058
  98. 98. Arnsten AFT, Wang M. Targeting prefrontal cortical systems for drug development: potential therapies for cognitive disorders. Annu Rev Pharmacol Toxicol. 2016;56: 339–360. pmid:26738476
  99. 99. Lidow MS, Elsworth JD, Goldman-Rakic PS. Down-regulation of the D1 and D5 dopamine receptors in the primate prefrontal cortex by chronic treatment with antipsychotic drugs. J Pharmacol Exp Ther. 1997;281: 597–603. pmid:9103549