Diplomarbeit
Transcription
Diplomarbeit
Mathematisch-Naturwissenschaftliche Fakultät I Institut für Biologie Diplomarbeit ZUM E RWERB DES AKADEMISCHEN G RADES D IPLOM -B IOPHYSIKER Automated Optimization of a Reduced Layer 5 Pyramidal Cell Model Based on Experimental Data vorgelegt von: Armin Bahl Matr.-Nr. 504502 geb. am 3. Juni 1983 in Berlin Stand der Arbeit: 30. August 2009 1. Gutachter: Prof. Dr. Andreas Herz1 2. Gutachter: Prof. Dr. Hanspeter Herzel2 1. Betreuer: Dr. Arnd Roth3 2. Betreuer: Dr. Martin Stemmler1 1 BCCN Munich, Ludwig-Maximilians-Univ. Munich, Munich D-80539, Germany Berlin, Institute for Theoretical Biology, Invalidenstraße 43, 10115 Berlin, Germany 3 Wolfson Institute for Biomedical Research, University College London, London WC1E 6BT, UK 2 Humboldt-Univ. Deutsche Zusammenfassung Genaue Modelle für Pyramidenzellen im Neokortex werden benötigt um realistische Simulationen der Informationsverarbeitung in kortikalen Schaltkreisen durchzuführen, allerdings fehlen geeignete Modelle in der Literatur. Gängige biophysikalische Modelle von Nervenzellen sind entweder zu komplex, ihre Parameter nur schwer einzustellen, und sie machen daher nur qualitativ korrekte Aussagen. Andere Modelle sind zu einfach und behandeln nur eine bestimmte biophysikalische Fragestellung. In dieser Arbeit wird eine systematische Herangehensweise beschrieben, um automatisch ein vereinfachtes Kompartimentmodell einer Pyramidenzelle der fünften Schicht des Neokortex zu konstruieren. Die Optimierung des Modells erfolgt auf der Grundlage experimenteller Daten. Das resultierende Modell gibt das Verhalten von Pyramidenzellen quantitativ genau wieder. Um eine geeignete Geometrie für unser Modell zu finden, präsentieren wir zunächst eine neuartige Methode, um eine realistische Morphologie zu vereinfachen und dabei die passiven Antworteigenschaften der Zelle möglichst wenig zu verändern. Um dann im nächsten Schritt die ionischen Leitfähigkeiten zu schätzen, hatten wir bereits zuvor vorgeschlagen, die Daten aus einer kürzlich veröffentlichten Arbeit (Bekkers & Häusser, 2007), in welcher die Dendriten physisch verschlossen wurden (Pinching), als Zieldaten für eine schrittweise “Parameter-Schälung” zu nutzen (Roth & Bahl, 2009). Wir präsentieren hier einen vergleichbaren Ansatz, nutzen aber eine Multiple-Richtungen Optimierungsstrategie (Deb et al., 2002; Druckmann et al., 2007) um die 18 freien aktiven und passiven Modellparameter in einem einzelnen Optimierungsdurchlauf zu schätzen. Unser Modell reproduziert einige der experimentellen Messkurven und generalisiert. Wir nutzen dann das automatisch generierte Modell, um den Einfluss der dendritschen Leitfähigkeiten auf das Ruhepotential und die Form von zurücklaufenden Aktionspotentialen zu untersuchen. Ebenfalls sehen wir , dass die zurücklaufenden Aktionspotentiale die somatische Hyperpolarisierung verändern können und dass unser Modell den schnellen somatischen Aktionspotentialbeginn reproduziert. Diese Ergebnisse stimmen gut mit experimentellen Studien überein. Das resultierende leitfähigkeitsbasierte Kompartimentmodell einer Pyramidenzelle ist unseres Wissens das erste, bei dem mehrere Zellregionen gleichzeitig und automatisch an experimentelle Daten angepasst wurden. iii Abstract Accurate models of pyramidal neurons are desperately needed to perform network simulations of cortical information processing but models of this type are lacking at present in the literature. Current models are either too complex, hard to constrain and give only qualitative results or they are too simple and focus only on one specific biophysical question. In this study we present a systematic approach to automatically create a reduced model of a layer 5 pyramidal neuron based on experimental data. The model reproduces the response properties of pyramidal neurons in a quantitatively exact manner. To obtain a reasonable geometry for our model we first present a novel approach to simplify a realistic morphology while maintaining the neuron’s passive response properties. In order to estimate the ionic conductances in our model we have suggested previously to use the data from a recent publication (Bekkers & Häusser, 2007) in which the dendrites were physically occluded (Pinching) as a target data set for a stepwise “parameter-peeling” optimization (Roth & Bahl, 2009). Here we present a comparable approach, but we use a multi-objective optimization strategy (Deb et al., 2002; Druckmann et al., 2007) to estimate the 18 free active and passive model parameters in a single optimization run. Our model is able to reproduce several experimental recordings and does also generalize. We use the automatically constrained neuron model to study the influence of the dendritic conductances on the resting potential and the shape of backpropagating action potentials. We see how backpropagating action potentials modulate the somatic afterhyperpolarization as well as that the model can reproduce the sharp somatic action potential onset. These modelling results are in acceptable agreement with experimental findings. To our knowledge the resulting conductance based multi-compartment model of a pyramidal neuron is the first neuron model that has been optimized in several cellular regions at once and automatically to experimental data. v Acknowledgments The work described in this diploma thesis was done in two different laboratories. I started with the first part in the group around Michael Häusser (UCL, London) and finished my work in the group around Andreas Herz (LMU, Munich). Both places and groups were very stimulating and it was very interesting to compare how neuroscientific questions are asked by experimentalists and theoreticians. Primarily I want to thank Andreas Herz who – almost three years ago – offered me the opportunity to join his group when he was still in Berlin. I thank him that he allowed me to play around in the field of neuroscience and that he has shown me how joyful scientific work can be. I want to thank Martin Stemmler for the many important questions and suggestions that helped me to adjust my focus. During my time in Berlin I developed a special interest in the detailed biophysics of neuronal information processing and decided to go to London to get closer to experimental data. The following month were highly efficient and I am thankful to Arnd Roth that he has made that exchange possible, for his many brilliant ideas and his supervision. I am grateful to John Bekkers for providing the experimental data, to Shaul Druckmann, Idan Segev, Michael London and Hermann Cuntz for discussions in the very early phase of this project and to Arnd Roth and Martin Stemmler for many very helpful comments on this manuscript. I want to thank Andreas Herz and Arnd Roth for organizing the financial support during the whole period. Finally I want to thank all the many colleagues, friends and my family who enriched my time when I did not think about this thesis. vii Contents Contents ix List of Figures xiii List of Tables xv List of Abbreviations 1 Introduction 1 1.1 Layer 5 Pyramidal Neurons . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.2 Compartmental Modelling of Neurons . . . . . . . . . . . . . . . . . . . 5 1.2.1 Membrane Properties . . . . . . . . . . . . . . . . . . . . . . . . 5 1.2.2 Single-Compartment Models . . . . . . . . . . . . . . . . . . . . 6 1.2.3 Multi-Compartment Models . . . . . . . . . . . . . . . . . . . . 9 1.3 Geometry Reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 1.4 Problems Constraining Neuron Models . . . . . . . . . . . . . . . . . . . 13 1.4.1 Experimental Uncertainties . . . . . . . . . . . . . . . . . . . . . 13 1.4.2 Constraining Parameters by Hand . . . . . . . . . . . . . . . . . 15 Automatic Fitting Strategies . . . . . . . . . . . . . . . . . . . . . . . . 15 1.5.1 16 1.5 2 xvii A Brief Review of Earlier Studies . . . . . . . . . . . . . . . . . Methods 21 2.1 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 2.2 High-Resolution Alignment of APs . . . . . . . . . . . . . . . . . . . . . 21 2.3 Modelling in NEURON . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 2.3.1 Using Python to Control NEURON . . . . . . . . . . . . . . . . 25 Multi-Objective Optimization using EAs . . . . . . . . . . . . . . . . . . 26 2.4.1 Evolutionary Algorithms . . . . . . . . . . . . . . . . . . . . . . 27 2.4.2 Multi-Objective Sorting . . . . . . . . . . . . . . . . . . . . . . 31 2.4.3 Parallelization . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 2.4 ix CONTENTS 3 The Cell Model 35 3.1 Neuronal Morphology . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 3.1.1 Geometry Reduction . . . . . . . . . . . . . . . . . . . . . . . . 35 3.1.2 Axon Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 3.1.3 Segmentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 Ion Channel Kinetics and Distribution . . . . . . . . . . . . . . . . . . . 40 3.2.1 Hyperpolarization-Activated Cation Channel . . . . . . . . . . . 40 3.2.2 Transient Sodium Channel . . . . . . . . . . . . . . . . . . . . . 42 3.2.3 Fast Potassium Channel . . . . . . . . . . . . . . . . . . . . . . 43 3.2.4 Slow Potassium Channel . . . . . . . . . . . . . . . . . . . . . . 44 3.2.5 Persistent Sodium Channel . . . . . . . . . . . . . . . . . . . . . 44 3.2.6 Muscarinic Potassium Channel . . . . . . . . . . . . . . . . . . . 45 Defining the Static and the Free Parameters . . . . . . . . . . . . . . . . 45 3.2 3.3 4 Results 47 4.1 Experimental Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 4.2 Fitting Strategy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 4.2.1 Checking Response Properties . . . . . . . . . . . . . . . . . . . 48 4.2.2 Distance Functions . . . . . . . . . . . . . . . . . . . . . . . . . 49 4.2.3 Combining Intact and Pinching Data . . . . . . . . . . . . . . . . 51 4.2.4 Selection of the Optimal Solution . . . . . . . . . . . . . . . . . 52 Fitting Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 4.3.1 Surrogate Data Optimization . . . . . . . . . . . . . . . . . . . . 53 4.3.2 Experimental Data Optimization . . . . . . . . . . . . . . . . . . 59 4.3.3 Generalization for Other Input Currents . . . . . . . . . . . . . . 65 Model Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 4.4.1 Resting Potential . . . . . . . . . . . . . . . . . . . . . . . . . . 67 4.4.2 AP Backpropagation . . . . . . . . . . . . . . . . . . . . . . . . 67 4.4.3 Currents Shaping the Somatic AP Waveform . . . . . . . . . . . 69 4.3 4.4 5 Discussion 73 5.1 Neuronal Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 5.1.1 Geometry Reduction . . . . . . . . . . . . . . . . . . . . . . . . 73 5.1.2 Passive Influence of the Basal Dendrite . . . . . . . . . . . . . . 75 5.1.3 Axonal Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . 75 5.1.4 Segmentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 Ion Channel Composition . . . . . . . . . . . . . . . . . . . . . . . . . . 76 5.2.1 Choice of Ion Channel Models . . . . . . . . . . . . . . . . . . . 76 5.2.2 Ion Channel Distribution . . . . . . . . . . . . . . . . . . . . . . 76 Fitting Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 5.2 5.3 x CONTENTS 5.4 5.5 5.3.1 Choosing the Free Parameters . . . . . . . . 5.3.2 Surrogate Data Optimization . . . . . . . . . 5.3.3 Experimental Data Optimization . . . . . . . 5.3.4 AP Initiation . . . . . . . . . . . . . . . . . 5.3.5 Effects of Pinching . . . . . . . . . . . . . . Model Evaluation . . . . . . . . . . . . . . . . . . . 5.4.1 Resting Potential . . . . . . . . . . . . . . . 5.4.2 Rapid AP Onset . . . . . . . . . . . . . . . 5.4.3 AP Backpropagation . . . . . . . . . . . . . 5.4.4 Currents Shaping the Somatic AP Waveform Outlook . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 79 80 81 81 83 83 83 84 84 85 87 xi List of Figures 1.1 Golgi Staining and Some of Cajal’s Drawings . . . . . . . . . . . . . . . 2 1.2 The Neuron as an RC-Circuit . . . . . . . . . . . . . . . . . . . . . . . . 6 1.3 An RC-circuit with Hodgkin-Huxley Like Voltage-Dependent Conductances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 1.4 Circuit Representation of a Three-Compartment Model . . . . . . . . . . 10 1.5 Comparison of the Squared-Distance Measure with LeMasson and Maex’ Distance Measure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 2.1 High-Resolution Alignment of APs . . . . . . . . . . . . . . . . . . . . . 22 2.2 Example NEURON-Python Simulation Results . . . . . . . . . . . . . . 26 2.3 Illustration of a Simple Two-Objective Optimization Problem . . . . . . . 27 2.4 Flowchart of the Working Principle of an EA . . . . . . . . . . . . . . . 28 2.5 Visualization of the Crossover and Mutation Operators . . . . . . . . . . 30 2.6 Illustration of the Ranking-Concept using Pareto Fronts and the Crowding Distance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 3.1 Morphology and Passive Properties for the Complex and the Reduced Model 38 3.2 Comparison of the Voltage Traces in the Complex and Reduced Model in Response to Noisy Input Current . . . . . . . . . . . . . . . . . . . . . . 39 3.3 Geometry of the Axon for the Reduced Model . . . . . . . . . . . . . . . 39 3.4 Ion Channel Gating Particles Used in This Study . . . . . . . . . . . . . 41 4.1 Experimental Recordings before and after Pinching . . . . . . . . . . . . 48 4.2 Best Solution of the Initial Random Population before the Surrogate Data Optimization, Trial 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 4.3 Best Solution after the Surrogate Data Optimization, Trial 1 . . . . . . . . 56 4.4 Evolution of the Four Objective-Distance Functions and of the Total-Error Value during the Surrogate Data Optimization, Trial 1 . . . . . . . . . . . 57 4.5 Best Solution after the Surrogate Data Optimization, Trial 2 . . . . . . . . 58 4.6 Best Solution of the Initial Random Population before the Experimental Data Optimization, Trial 1 . . . . . . . . . . . . . . . . . . . . . . . . . 60 xiii LIST OF FIGURES 4.7 4.8 4.9 4.10 4.11 4.12 4.13 4.14 4.15 Best Solution after the Experimental Data Optimization, Trial 1 . . . . . . Evolution of the Four Objective-Distance Functions and of the Total-Error Value during the Experimental Data Optimization, Trial 1 . . . . . . . . . Best Solution after the Experimental Data Optimization, Trial 2 . . . . . . Best Solution after the Experimental Data Optimization, Trial 3 . . . . . . Model Prediction of Firing Frequency . . . . . . . . . . . . . . . . . . . Model Prediction of Detailed AP shape and Spiketrain in Response to Another Input Current . . . . . . . . . . . . . . . . . . . . . . . . . . . . Resting Potential and the Ionic Conductances as a Function of Distance to the Soma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Analysis of BAPs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Currents Shaping the Somatic AP Waveform before and after Pinching . . xiv 61 62 63 64 65 66 68 70 71 List of Tables 3.1 3.2 4.1 4.2 Optimal Geometrical and Passive Parameters for the Reduced Model after Simplification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Free Parameters in the Reduced Model . . . . . . . . . . . . . . . . . . . 36 46 Target Parameters and Best Parameter Combinations after the Surrogate Data Optimizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Best Parameter Combinations after the Experimental Data Optimizations . 54 59 xv List of Abbreviations τmem membrane time constant τp time constant for gating particle p gbarx maximal specific ionic conductance for ion channel x Cm effective membrane capacitance cm specific membrane capacitance Ex reversal potential for ion x gpas specific leak conductance p∞ steady state value for gating particle p ra intracellular resistivity Rm effective membrane resistance rm specific membrane resistance tadj temperature adjustment factor AHP afterhyperpolarization AP action potential BAP backpropagating action potential EA Evolutionary Algorithm fMRI functional magnetic resonance imaging HCN-channel hyperpolarization-activated cyclic nucleotide-gated cation channel iseg axon initial segment Kfast-channel fast potassium channel xvii LIST OF TABLES Km-channel muscarinic potassium channel Kslow-channel slow potassium channel MOO multi-objective optimization Nap-channel persistent sodium channel Nat-channel transient sodium channel PF Pareto front xviii CHAPTER 1 Introduction that the elementary units of our brain are single metabolically distinct cells and that our brain is the main information processing center for everything we do. With these cells, the brain receives sensory information, computes, stores, and accesses memory and directs our muscles how to move to accomplish our wish. We know that the autonomic nervous system maintains the functions of our organs, although we are not consciously aware of it. All this appears natural to us today, but this was not always that clear. T ODAY WE KNOW Around 350 B.C., Aristotle developed his Cardiocentric Hypothesis, which states that the brain solely exists for cooling the blood in the human body and that the heart is the central organ of mind and emotion. On the other hand Plato and other philosophers argued that the mind must be found in the brain. But only with the pioneering work of Herophilus and Erasistratus (around 300 B.C.), detailed anatomical and functional studies revealed that mind and brain are linked. Another milestone was set by Galen (around 180 A.D.). He wrote that the brain “receives all sensations, produces images and understands thoughts”, and that only with the help of rigorous anatomical methods one would be able to prove the Encephalocentric Hypothesis (Crivellato & Ribatti, 2007). Influenced by the work of the early Greek philosophers and anatomists, following generations of scientists accumulated an enormous amount of knowledge about the function, dysfunction and the anatomy of the brain and its parts, but still it remained unclear which were the intrinsic biological mechanisms for brain function. At the end of the 18th century, Luigi Galvani studied the electricity in the nervous system in dissected frog legs and thereby laid the foundation for a new science, electrophysiology (Piccolino, 1997). Emil Heinrich Du Bois-Reymond further developed the technique. Around 1850 he found that potentials in the tissue can change rapidly when a nerve stimulus elicits an “action” in the muscle, and thereby formed the idea of the action potential (AP) (Pearce, 2001). At the end of the 19th century, with the advanced microscopy techniques and Camille Golgi’s new silver staining method, the Spanish physician Santiago Ramón y Cajal started 1 CHAPTER 1. INTRODUCTION to study the microscopic structure of the brain. Cajal who originally wanted to become an artist created hundreds of fine drawings of the stained materials that are still admired today (Fig. 1.1). During his rigorous work on different brain tissues, he realized that the a b c Figure 1.1: G OLGI S TAINING AND S OME OF C AJAL’ S D RAWINGS a) Photomicrographs from Cajal’s preparations of the cerebral cortex showing neurons impregnated by the Golgi stain. b) Cajal’s drawing of the superficial layers of the human frontal cortex shows the characteristic morphology of pyramidal neurons. c) Drawing of the cerebellar cortex with a detailed view of Purkinje cells (A).1 structures he was drawing were separate units interconnected within the large network of the nervous system. This result was the proof for the “neuron doctrine" formulated by Wilhelm Waldeyer in 1891 and the single units were finally named neurons (Greek: string, wire). Moreover, Cajal formulated “The Law of Dynamic Polarization” stating that neurons are polarized, receiving information on their cell bodies and dendrites (Greek: tree) and transmit information to other distant neurons through axons (Greek: axis). 1 Finally Julius Bernstein realized that the electrical phenomena in all living tissue are a property of membrane currents and that also APs have an ionic basis and can be explained with a rapid membrane conductance change (Bernstein, 1906). In the 1950s Alan Lloyd Hodgkin and Andrew Fielding Huxley quantitatively measured the membrane currents during APs in the giant axon of the squid Loligo. With 1 Life and Discoveries of Santiago Ramón y Cajal: http://nobelprize.org/nobel_prizes/ medicine/articles/cajal/index.html 2 a new technique called space clamp they kept the potential along the entire axon spatially uniform. By voltage clamping the membrane to different potentials and the use of pharmacological agents to block various currents, Hodgkin and Huxley were able to dissect the membrane current into its constitutive components. The detailed analysis of these currents led Hodgkin and Huxley to a simple mathematical model with some fictive “gating particles” that explained the generation of APs with remarkable accuracy (Hodgkin & Huxley, 1952a,b). In the 1970s Erwin Neher and Bert Sakmann were able to isolate the gating particles predicted by Hodgkin and Huxley. They carried out further voltage-clamp studies with fine electrodes on small patches of membrane. Interestingly the currents in these membrane pieces were not constant for a given voltage and their amplitude was random and quantized. The result of these patch clamp experiments led Neher and Sakmann to the conclusion that in the tiny pieces of a membrane of only several µm2 there were only a few pores that were either open or closed, which proved the existence of stochastic ion channels (Sakmann & Neher, 1984). Today we know that the nervous system is formed of more than 1010 densely packed neurons connected to an intricate network. We can admire the geometry and the beauty of neurons in a detail Cajal could only dream of. Due to atomic-resolution crystal structures and fluorescence distance measurements we have understood the molecular details of ion channels, how they conduct ions and why they are selective and voltage-dependent. We know that dendritic spines can appear and disappear during learning and synaptic connections can become stronger or weaker. We know that the Hodgkin-Huxley model can be used to explain general currents in different parts of the neuron in different brain areas also for mammals. We can also observe the activity of larger brain areas with advanced imaging techniques and with functional magnetic resonance imaging (fMRI) even the activity of the whole brain during human behaviour. All these wonderful experimental discoveries during the last century have led to a very detailed picture of our brain and the neuron, but still, we have only vague ideas about the details of neural computation. It is the task of theoreticians and today’s philosophers to put the knowledge together and to extract the fundamental principles of how our brain works! One way to understand brain function is a bottom-up approach, by understanding and modelling the elementary unit of our brain, the single neuron. If we can describe the function of a single neuron with a mathematical model, we can create artificial networks of neurons and observe the network behaviour in a detail we could never do in an experiment. 3 CHAPTER 1. INTRODUCTION 1.1. Layer 5 Pyramidal Neurons Pyramidal neurons are some of the best studied cells in the brain and are present in virtually all mammals (Spruston, 2008). They are found primarily in structures that are asccociated with advanced cognitive functions like the hippocampus or the cerebral cortex. Here we focus on cortical layer 5 pyramidal neurons, but the general geometrical and many response properties are similar for the other pyramidal neuron types. Pyramidal neurons have an elaborate dendritic tree and axonal structure. The relatively short basal dendrites connect directly to the soma whereas the large apical dendrite connects the soma with the distal tuft. The oblique dendrites emanate proximally from the main apical dendrite. The extent of the dendritic tree of layer 5 pyramidal neurons is around 1 mm in the adult rat (Fig. 3.1). These large neurons can therefore extend their dendritic structures into different layers of the cortex to receive thousands of synaptic connections. Some inhibitory inputs are specifically targeting the soma and axon while most of the excitatory synaptic drive arrives via the dendrites. Distal dendritic regions receive inputs from higher cortical areas whereas local sources of synaptic input project to the proximity of the soma. It is thought that the geometry of pyramdial neurons might be designed for coincident detection of inputs from the tuft and the proximal dendrites (Cauller & Connors, 1994). Alternative hypotheses suggest that synapses at the tuft could control the responsivness to more proximal inputs (Larkum et al., 2004). Like the soma and the axon, also the dendrites are enriched with a variety of voltagegated ion channels that have an important influence on the integration of synaptic input (Johnston et al., 1996). It was shown that for clustered inputs the dendritic tree can initiate local dendritic spikes (Ariav et al., 2003; Golding & Spruston, 1998) which could eliminate the problem of distance-dependent synaptic efficacy (Katz et al., 2009). When considered mild or distributed input currents it is thought that AP initiation occurs in the axon initial segment (iseg) and that APs can then propagate actively into two directions, downwards the axon, targeting other neurons, but also antidromically through the soma into the dendritic tree (Stuart & Sakmann, 1994). These backpropagating action potentials (BAPs) give rise to an interesting backward-forward “ping-ping” communication between the axonal spike initiation zone and dendritic post-synaptic locations and are thought to be an important mechanism for synaptic plasticity (Segev & London, 2000). In order to better understand these mechanisms, their interactions and functional network implications one approach is to create detailed or reduced compartmental models of pyramidal neurons. 4 1.2. COMPARTMENTAL MODELLING OF NEURONS 1.2. Compartmental Modelling of Neurons In compartmental models of neurons the membrane and its intrinsic conductances are regarded as the central elements of the model. Depending on the level of complexity and the purpose of modelling the model can consist of only a single compartment or involve more than thousand compartments. 1.2.1. Membrane Properties The membrane of the neuron consists of a densely packed ≈ 4 nm thick bilayer of phospolipids with embedded proteins that make it almost impermeable for water, big molecules or any charged ion. However some of the membrane proteins act as “gates” or “channels” that open and close depending on voltage or by modification with some ligand. We can define the membrane potential V as the difference between the intracellular and the extracellular potentials: V = Vi −Ve . (1.1) Early experiments have shown that the membrane potential for neurons at rest (Vrest ) is almost always negative and lies between -90 mV and -30 mV (Koch, 2004, p. 6). The origin of the negative resting potential lies in the differential distribution of open ion channels across the membrane and the unequal concentration of ions with different charge inside and outside the cell at rest. We can summarize all ionic conductances per membrane area at rest into a single specific leak conductance (gpas ) or specific membrane resistance (rm ) with: 1 . (1.2) gpas = rm Units for gpas and rm are mostly given in S/cm2 and Ω · cm2 respectively. The insulating membrane also keeps charges apart and acts therefore as a capacitor. Thus the membrane potential allows the capacitance to build up a specific charge (qm ) on both sides of the membrane: qm = cm ·V . (1.3) The specific membrane capacitance (cm ) describes the capacitance per membrane area and should only depend on the biochemical composition of the membrane. Its units are mostly reported in µF/cm2 . Frequently used values for cm and rm lie around 1 µF/cm2 (Gentet et al., 2000) and 15000 Ω · cm2 respectively. 5 CHAPTER 1. INTRODUCTION 1.2.2. Single-Compartment Models 1907, inspired by the idea that the neural membrane can be described by simple electrical RC-circuit elements, Louis Lapicque modeled the neuron using a parallel capacitance, a resistor as well as a battery to account for the resting potential Vrest (Abbott, 1999; Lapicque, 2007). Let us neglect the spatial structure of the neuron so that we can model it using an isopotential membrane sphere with surface area A. The effective membrane resistance (Rm ) and effective membrane capacitance (Cm ) are then simply given by Rm = rm · A (1.4) Cm = cm · A . (1.5) Any input current Iin j to the cell can be simply represented with a current flowing directly into the circuit (Fig. 1.2). Using Ohm’s law we can calculate the current over the Figure 1.2: T HE N EURON AS AN RCC IRCUIT The single-compartment is modeled using elementary electrical circuit elements: A capacitance Cm , a resistor Rm and a battery Vrest . Resistance and capacitance depend on the membrane area. The capacitance determines how much charge can accumulate on both sides of the membrane for a given voltage, the resistor defines the leakconductance of the membrane at rest, while the battery maintains the resting potential. The membrane potential is defined as the difference between the potential inside (Vi ) and outside (Ve ) of the neuron. A current Iin j can be represented as a current flowing directly into the circuit. Ve Rm Cm IC Iinj IR Vrest Vi resistance and the battery: Vi = Ve − Rm · IR +Vrest ⇒ V = −Rm · IR +Vrest Vrest −V . ⇒ IR = Rm (1.6) (1.7) (1.8) Whenever the membrane voltage changes, a capacitive current will flow: IC = dQm dV = Cm · . dt dt (1.9) Due to Kirchhoff’s law of current conservation we can write down: IC = IR + Iin j 6 (1.10) 1.2. COMPARTMENTAL MODELLING OF NEURONS and hence dV dt dV ⇒ Cm · Rm dt Cm · = Vrest −V + Iin j Rm = −V +Vrest + Rm · Iin j . (1.11) (1.12) Introducing the membrane time constant τmem = Rm ·Cm leads to the membrane equation: dV (t) = −V (t) +Vrest + Rm · Iin j (t) . (1.13) dt This is an inhomogeneous ordinary differential equation that can easily be solved for a general input current. τmem · Hodgkin-Huxley Model Hodgkin & Huxley (1952b) extended the simple RC-circuit by introducing further resistors to describe the conductivity of the membrane for various currents (Fig. 1.3). They Figure 1.3: A N RC- CIRCUIT WITH H ODGKIN -H UXLEY L IKE VOLTAGE D EPENDENT C ONDUCTANCES Hodgkin and Huxley added two resistors to the circuit to account for the sodium Gna = 1/Rna and potassium Gk = 1/Rk conductances they found in the squid axonal membrane. These conductances are voltage-dependent and are therefore symbolized with a variable resistor. The reversal potential of each conductance (Ena and Ek ) as well the resting potential (Vrest ) are represented as a battery in the circuit. Hodgkin and Huxley developed a general framework in order to mathematically describe the voltage dependencies of the currents. Today many other ionic currents can be added to the circuit in the same fashion. Ve Rna Rk Rm Cm Iinj Ena Ek Vrest Vi found out that the membrane of the squid axon conducts mainly sodium and potassium ions and that the passive leak current is mainly due to chloride ions. Therefore the total current flowing through the membrane is Iion = Ina + Ik + I pas . (1.14) They also realized that the sodium and potassium currents are variable and time- and voltage-dependent and can be modeled using a relatively simple formalism: 7 CHAPTER 1. INTRODUCTION A membrane current Ix was described with the following differential equations: dp = (1 − p) · α p (V ) − p · β p (V ) dt dq = (1 − q) · αq (V ) − q · βq (V ) dt Gx = Gbarx · pi · q j Ix = Gx · (V − Ex ) (1.15) (1.16) (1.17) (1.18) where Gbarx = 1/RMax is the maximal membrane conductance for ion x. To describe the m dynamics of the conductance Gx , Hodgkin and Huxley introduced several fictive gating particles p and q. Each gating particle can be in one of two possible states, open or closed. In order for the current to flow, all gating particles need to be open simultaneously. The transition between open and closed for each gating particle was described with the help of the voltage-dependent transition functions α(V ), β (V ). The general form and parameters of these functions could be estimated by directly fitting the conductance changes during several voltage steps. To obtain better fits Hodgkin and Huxley also included the exponents i and j which can be related to the oligomeric nature of ion channels today. Finally, the current Ix depends on the driving force V − Ex . Ex is the reversal potential and depends on the ratio of the intra- and extracellular concentration of the ion. Hodgkin and Huxley modeled the sodium conductance Gna with two kinds of gating particles m and h. m was thought to be the activation particle while h was responsible for the inactivation of the current. On the other hand the potassium conductance was modeled using a single activation particle n and the leak current was voltage-independent. For simplification of the mathematics and for illustration of the kinetics of a gating particle p it is often useful to calculate its steady state value (p∞ ) and time constant (τ p ): p∞ (V ) = α p (V )/ (α p (V ) + β p (V )) (1.19) τ p (V ) = 1/(α p (V ) + β p (V )) dp = (p∞ − p)/τ p dt (1.20) (1.21) which is equivalent to Eqn. 1.15. Based on these detailed descriptions of the currents, we can rewrite the membrane equation: Cm · dV dt = −Gbarna · m3 · h · (V − Ena ) − Gbark · n4 · (V − Ek ) − G pas · (V −Vrest ) + · Iin j (t) (1.22) (1.23) where each of the gating particles m, h and n is described with a differential equation of the form 1.15 and its own voltage-dependent transition functions α(V ) and β (V ). The 8 1.2. COMPARTMENTAL MODELLING OF NEURONS exponents over the gating particles were found to produce the best fit for the experimental data. With that extended single-compartment model, Hodgkin and Huxley were able to elucidate the fundamental principles of AP generation based on a realistic description of ionic conductances. The Hodgkin-Huxley model is widely regarded as the cornerstone of quantitative modelling of nerve cell excitability and is seen as one of the greatest achievements of 20th -century biophysics. It should be remembered that at the time the model was suggested, the existence of stochastic ion channels was not known yet. The formalism developed by Hodgkin and Huxley was later generalized and used to create a variety of kinetic models for many different ionic conductances in other neuron types. The models of ionic conductances in pyramidal neurons that we are using in this study are explained in detail in sec. 3.2. Nevertheless, there are recent approaches to develop new kinetic schemes based on detailed kinetic transitions (Baranauskas & Martina, 2006; Gurkiewicz & Korngreen, 2007). In addition to the description of AP initiation and propagation, single-compartment Hodgkin-Huxley-type models have led to an understanding of many fundamental dynamics of spiking neurons. For example, they were used to explore the role of calcium currents and calcium dynamics in bursting neurons (Amini et al., 1999) or they were used to suggest mechanisms for spike frequency adaptation (Benda & Herz, 2003). Furthermore, based on a separation of time-scales, it was possible to reduce the four dimensional singlecompartment Hodgkin-Huxley model to models with only two dimensions where phaseplane analysis or bifurcation theory could be used to study neuronal excitability (Fitzhugh, 1961; Morris & Lecar, 1981; Nagumo et al., 1962). There is still a vigorous and controversial debate over whether the Hodgkin-Huxley model is sufficient for explaining the sharp AP onset in pyramidal neurons (McCormick et al., 2007; Naundorf et al., 2006; Yu et al., 2008). 1.2.3. Multi-Compartment Models Most neurons cannot be represented as a single-compartment as they have a complex morphology with different membrane and ion channel properties in different regions of the cell. However, we can discretize a neuron into multiple small pieces of membrane cylinders and each of these compartments can then be represented as a simple RC-circuit. The membrane pieces are connected via the intracellular solution and current can flow from one piece to the other. Hereby the electrolyte solution in the cable acts like a resistance. Normalizing this resistance to the length and the diameter of the cylinder we can define the intracellular resistivity (ra ). A common value used for pyramidal neuron models lies around ra = 100 Ω · cm. The compartments need not necessarily have the same diameter d and length L, hence 9 CHAPTER 1. INTRODUCTION each compartment will have its own effective membrane resistance Rm and capacitance Cm and the effective axial resistance Ra between two compartments will depend on the geometry of both: Rim = rm · π · di · Li (1.24) Cmi = cm · π · di · Li Lj ra ra Li Riaj = 2 · + 2 · . 2 2 d π · d2i π · 2j (1.25) (1.26) A model with three compartments of different sizes and its circuit representation are illustrated in Fig. 1.4. By means of compartmentalization even a complex morphology can L2 a L1 L3 d1 d3 d2 b R12 a 1 Cm 1 Rm Ra23 2 Rm 2 Cm Vrest 3 Cm Vrest 3 Rm Vrest Figure 1.4: C IRCUIT R EPRESENTATION OF A T HREE -C OMPARTMENT M ODEL a) A fictive simplified neuronal geometry is shown. The neuron model consists of three cylindrical compartments, each with different diameter d and length L. b) The compartments are represented as RC-circuits. The effective resistance Rim and capacitance Cmi of each unit depend on the surface area of the cylinder while the battery is the same in each circuit. The connection between two compartments can be described with an axial resistance Riaj . In each case the axial resistance depend on the diameter and length of two adjacent cylinders. be approximated by a set of membrane equations coupled through axial resistive currents: Cmi · dV i (t) Vrest −V i (t) V j (t) −V i (t) i = + I (t) + . ∑ in j ij dt Rim Ra j (1.27) Mostly, this system must be solved numerically. Choosing the right spatial discretization is a recurring practical problem in neuronal modelling and the size of one compartment will obviously depend on the morphological complexity of the neuron and of course on the question that is being asked. Finding the resting membrane potential in a simple passive geometry, for example, will require a rather rough discretization, while the analysis of 10 1.2. COMPARTMENTAL MODELLING OF NEURONS burst firing might demand much finer spatial resolutions. Reduced Models Reduced multi-compartment models consist of a few compartments only but their complexity is sufficient to gain insights into the complicated spatiotemporal interactions in the neuron. A two-compartment model was used, for example, to investigate the interplay of fast somatic sodium and potassium conductances with the dendritic slow calcium currents and their role in bursting behaviour (Pinsky & Rinzel, 1994). It was recently also possible to create reduced models that could explain BAPs in pyramidal neurons (Keren et al., 2009). In addition to their immense value for understanding mechanisms like that, reduced multi-compartment models are also computationally effective. They were therefore used in large-scale network studies with more than 3000 cells and demonstrated, for example, that gap junctions are instrumental for cortical gamma oscillations (Traub et al., 2005). Detailed Models Detailed multi-compartment models are based on exact anatomical reconstructions. To represent an entire morphology with sufficient discretization, often more than 1000 compartments are necessary that are connected via axial resistors. In each compartment several ion channels can be modeled to account for the local conductances. This might lead to a very complex system of more than 10000 coupled differential equations which rules out the chance of any analytical solution and even makes “by-hand” numerical simulation a daunting task. Therefore tools are needed that keep track of the neuronal properties and can create and efficiently solve the large number of equations automatically. Several of these tools have been developed during the last decades, for example GENESIS (Bower & Beeman, 1998), NEURON (Carnevale & Hines, 2006) or very recently MOOSE (Ray & Bhalla, 2008). The level of detail used for these models can lead to insights into the biophysical mechanisms and the role of the neuron’s spatial structure involved in neuronal information processing that today’s experimental methods could not give. The computer model however might suggest a new experimentally testable hypothesis. The new experimental result could than be used to further optimize the model parameters and mechanisms and then, with the improved model, suggest other experiments. Hence, detailed models, combined with experiments are powerful tools for the exploration of the complex biophysics in neurons. Detailed models were used for the study of axonal AP initiation (Kole et al., 2008) or voltage attenuation in dendrites (Stuart & Spruston, 1998). They were used to explore how the dendritic geometry determines somatic firing patterns (Mainen et al., 1995). Sev11 CHAPTER 1. INTRODUCTION eral models were used to examine the computational capabilities of dendrites (reviewed by London & Hausser, 2005; Segev & London, 2000). For example, it was suggested that mechanisms in the dendritic tree could explain translation-invariant tuning of visual complex cells (Mel et al., 1998). Another study analyzed the integration of synaptic input in a detailed model of a CA1 pyramidal cell and suggested that the neuron could be represented as a two-layer cascade (Poirazi et al., 2003). Even dendritic spines can be modeled but often this is simply done by increasing the dendritic leak conductance and the capacitance by a certain spinefactor to account for the additional spine membrane area. It was recently found experimentally that spine size is scaled along the dendrites. These findings were combined with a detailed model of a CA1 pyramidal neuron which led to further evidence for a two-layer integration of dendritic input (Katz et al., 2009). Although detailed models are computationally expensive they are widely considered to be useful for large-scale network simulations. The hope is to model a part of the brain as realistically as possible to understand brain function and dysfunction through detailed computer simulations and possibly to suggest new pharmaceutical treatments. The Blue Brain Project, for example, attempts to create a model of a neocortical column (Markram, 2006). Other studies are currently creating detailed models of a thalamocortical region (Izhikevich & Edelman, 2008) whereas others perform detailed network simulations of the primary visual cortex, in software as well as in silico.1 1.3. Geometry Reduction If a reduced multi-compartment model should be created, it is not clear how the simplified geometry must look like and which diameters, length and membrane parameters should be used for the cylinders. To get an idea about the structure of the model it is therefore reasonable to start with a detailed reconstruction and to simplify its dendritic geometry while maintaining the neuron’s passive response properties. For homogeneous passive cables the Linear Cable Theory was developed (Koch, 2004; Segev, 1994) which made it possible to study the voltage distribution in long cables. It was also shown that a small subset of neuronal morphologies can be collapsed into an equivalent single cylinder (Rall, 1962) which allowed the application of the Linear Cable Theory to study the voltage spread in complex dendritic geometries. Rall’s theory is based on the following assumptions for the dendritic tree: First, the membrane resistance Rm and the axial resistance Ra are the same in all branches of the dendritic tree. Second, the electrotonic distance from the soma to each dendritic terminal should be equivalent. Third, the branch points must follow the 3/2 power rule, meaning that for the diameter d0 of a parent branch and the diameters d1 , d2 of its daughter branches 1 Colamn-Project: http://gow.epsrc.ac.uk/ViewGrant.aspx?GrantRef=EP/C010841/1 12 1.4. PROBLEMS CONSTRAINING NEURON MODELS the following condition must hold: 3/2 d0 3/2 3/2 = d1 + d2 . (1.28) To overcome these constrains several authors suggested alternative methods to construct simple structures from arbitrarily branched dendritic trees (Bush & Sejnowski, 1993; Lindsay et al., 2003). A simple and intuitive way was suggested by Destexhe (2001). He divided the dendritic tree of a layer 6 pyramidal neuron into several functional subunits, namely the soma with proximal dendrites, the basal and the distal dendrites. Each of these functional subunits was represented by a single cylinder in the simplified model. The length of the equivalent compartment was chosen to be similar to the typical physical length of its associated functional region. The diameter of the cylinder was adjusted such that its total membrane area was the same as the subset of dendrites it represents. Then the intracellular resistivities were fitted so that the simplified model showed similar voltage attenuation like the complex model. 1.4. Problems Constraining Neuron Models Neuron models, in particular multi-compartment conductance-based models normally come with a large number of free parameters. Many of these parameters cannot be directly determined experimentally with the technique available today. Even if experimental data exists, we must be very careful using it without a detailed evaluation of the experimental protocol as there are several uncertainties that arise during electrophysiological recordings. 1.4.1. Experimental Uncertainties For example the measured membrane potential is often shifted in respect to the real value of sometimes more than 10 mV (Barry, 1994; Barry & Diamond, 1970). This is due to an insufficient compensation for the Liquid Junction Potential which occurs when two solutions of different concentration are in contact with each other. Yet many electrophysiological studies neglect this. Thus these measurements do not only result in a wrong estimate of the absolute membrane potential, but also lead to a failure when modelling ion channel kinetics since the fraction of open channels depends on the absolute voltage. Ion channel densities are mostly estimated using the cell-attached patch clamp configuration. All ion channels but the one of interest are blocked by application of various blocking agents into the extra- or intracellular solution (for a review see Catterall, 1995). Then a fine glass electrode is pressed against the cell membrane until a high resistance seal can establish (a gigaseal). Then the neuron is voltage clamped and isolated currents 13 CHAPTER 1. INTRODUCTION can be measured. By measuring the exact size of the pipette tip, it is possible to calculate the ion channel conductance per area. Ion channel densities for several ion channel-types have been suggested by this method and many ion channel models were published.2 However, one might question the quality of the results. For example it is not clear that the blocking agent really did block all ion channels, but the single one we are interested in and therefore the assumed pure current might be a mixture of several currents. Next, it was shown that cell-attached patch clamp recordings might underestimate ion channel density per se. For example, there is experimental evidence that sodium channels in the axon initial segment are anchored via the cytoskeleton to the inner neuronal membrane. Thus a pipette attached to the outer surface cannot record these currents and the effective conductance appears to be low (Kole et al., 2008). However qualitative statements about relative channel densities or the density distribution within a single neuron can be made and are very useful for modelling (for example Keren et al., 2009; Kole et al., 2006). To build a detailed model, the neuron is often filled with Biocytin after the recording and the slice is fixed. The neuronal morphology is then reconstructed for example via the manual reconstruction software Neurolucida (MicroBrightField, Williston, VT) under a microscope. There is currently development to automate the procedure based on liveimaging data (for example Losavio et al., 2008). Most of the detailed reconstructions available today are not accurate enough. This is due to the human-made error during the reconstruction procedure and due to tissue shrinkage of approximately 10 % during the fixation (Weaver & Wearne, 2008). In addition to these uncertainties about the experimental procedures themselves, we cannot obtain a full set of parameters for one neuron based on experiments. It is obviously not possible to obtain data for all ion channels densities and kinetics when the others were blocked and biocytin filling is sometimes not possible after long-lasting experimental procedures. Furthermore patch clamping is not easy and only a few successful patches can be obtained per neuron. It is particularly hard to obtain good recordings from the dendritic tree. Therefore the results from many different neurons are combined and hence all detailed models are based on dozens of experiments in different parts of the neuron, different cells, animals and even species, temperatures and recording conditions. We also need to make assumptions about the regions that are not accessible experimentally at all. It is often assumed, for example, that ion channel kinetics once obtained from somatic recordings are the same everywhere in the neuron. But all kinds of different subunits are expressed in different parts and ion channels can change their properties due to local modification. It was shown, for example, that the sodium activation and inactivation curves in the axon are shifted ≈ -7 mV to more hyperpolarized potentials (Colbert & Pan, 2002; Kole & Stuart, 2008). 2 ModelDB: http://senselab.med.yale.edu/ModelDB 14 1.5. AUTOMATIC FITTING STRATEGIES 1.4.2. Constraining Parameters by Hand Summarizing, the list of uncertainties about the parameters of multi-compartment conductance-based neuronal models is endless. But if we cannot properly constrain these parameters via experiments, we need to find other strategies. One common way is to only consider a subset of all possible parameters and to set the remaining parameters to some standard values. The free parameters are then adjusted by hand via trial and error to minimize the distance between some model response and an experimental data set, like the somatic voltage. This procedure however is tedious for the modeller and requires an extensive experience. The parameters have highly nonlinear interactions and a slight modification of one parameter might require a change of the others with unpredictable amplitudes. Nevertheless there are many detailed models that were tuned by hand which are very successful in explaining qualitatively many observations in single neurons (see above). But even if the model reproduces qualitatively the experimental data well one might doubt that the adjusted parameters represent biological reality and possibly many distinct parameter combinations will lead to similar good results. For example Prinz et al. (2004) have shown that similar network dynamics can arise with distinct circuit and neuron parameters. Moreover an extension of the model with some further mechanisms might require a completely new parameter search. Therefore it is necessary to automate the search strategy. 1.5. Automatic Fitting Strategies There are three parts of a good optimization: First, we need a model that could in principle fit the experimental data. Second, we need a good distance function between model and data. Third, we need an efficient optimization algorithm that using the distance function can evaluate different parameters for the model and thereby find better solutions, until an optimal parameter set is reached. The first point is the hardest and requires biophysical knowledge about the complex mechanisms responsible for information processing in single neurons. Even if the distance function and the optimization algorithm work well, the algorithm will definitely never find a good solution if the model will not be able to represent the experimental data. For example, the data might show spike frequency adaptation, but the model might not have such a mechanism implemented. To circumvent the search for the right model for a given experimental data set, the optimization algorithm and the distance functions are often evaluated fitting the model to so called surrogate data that were generated by some parameter set of the same model. In this case a perfect solution exists for sure and it is 15 CHAPTER 1. INTRODUCTION only a question of the search strategy whether the original parameter set can be found again. However it is not guaranteed that the optimization algorithm will also fit experimental data even if it has performed well in the model-to-model fit (Druckmann et al., 2008). But this is neglected by most studies developing strategies to fit a model to surrogate data and it could be that these algorithms cannot be directly used to fit experimental data. However these studies are useful to develop advanced optimization algorithms that can eventually be applied successfully to experimental data. 1.5.1. A Brief Review of Earlier Studies Let us briefly review some automatic parameter constraining procedures that were developed during the last decade (detailed reviews were published by Druckmann et al., 2008; Van Geit et al., 2008). One of the first studies to automatically fit neuronal models was published by Vanier & Bower (1999). They realized the need for a rigorous comparison of different search strategies to constrain a large parameter set. Different strategies were evaluated, namely gradient descent, genetic algorithms, simulated annealing and stochastic search to fit models of different complexities. Interestingly they found out that for simple models simulated annealing showed the highest performance, while detailed models with a large number of parameters were best fitted by genetic or evolutionary algorithms. However the comparison was only performed on surrogate data and thus their conclusions might not directly transfer to strategies fitting experimental data. Prinz et al. (2003) were able to constrain an 8 dimensional single-compartment model of a lobster stomatogastric neuron to experimental data. They did not directly fit the model to a single data set, but first created a database of many possible model responses. For each parameter 6 discrete values between a lower and an upper boundary were allowed and the parameters were varied individually in a grid-like manner and the corresponding model response was evaluated and stored. Once the database was created, it was possible to filter only those data sets that mimicked an experimental target data best. One may also use the data for many statistical studies, like the analysis of the role of a certain parameter in producing some specific spiking behaviour. However the grid-search strategy will become too costly when more parameters are involved, as in multi-compartment models, especially if the grid resolution becomes higher. Achard & De Schutter (2006) created a framework to fit an entire detailed model of a Purkinje neuron (De Schutter & Bower, 1994) to surrogate data. The model consists of 1600 compartments and 24 ion channel densities were set as free parameters. Fitting was performed using a distance measure based on LeMasson and Maex’ phase-plane analysis to overcome a problem with spike-timing (LeMasson & Maex, 2001). The problem with 16 1.5. AUTOMATIC FITTING STRATEGIES a direct mean-square comparison of a target and test spiking trace is that the resulting error value is strongly dependent on spike-timing. If, for example, target and test traces are almost similar, but the test solution shows slightly faster spike frequency adaptation, then the final error value will be huge anyhow. Moreover the error value will be smaller when the test solution is not spiking at all (Fig. 1.5a,b). Therefore such a distance measure does not represent the quality of the model. In contrast, the phase-plane distance measure is independent of the precise spike-timing. For the target and test spiketrain the voltage derivative dV /dt is calculated. Then the matrix of M(V, dV /dt) is binned and for each bin the number of points is determined. The differences of points in each bin for the target and test histogram are then summed to the final error value (Fig. 1.5c,d). Achard & De Schutter (2006) were able obtain good fits the entire Purkinje neuron model c 60 40 20 0 -20 -40 -60 -80 dV/dt (V/s) V (mV) a 0 100 200 300 400 500 600 Time (ms) d 60 40 20 0 -20 -40 -60 -80 dV/dt (V/s) V (mV) b V (mV) 0 100 200 300 400 500 600 Time (ms) V (mV) Figure 1.5: C OMPARISON OF THE S QUARED -D ISTANCE M EASURE WITH L E M ASSON AND M AEX ’ D ISTANCE M EASURE We chose a target parameter set to calculate the target spiketrain (black) for a neuron model. To compare the squared distance measure with LeMasson and Maex’ distance measure, we chose two test parameters sets and determined their spiketrains (red). The first test parameter set was similar to the target parameter set, but each parameter was changed by 1% leading to a similar spiketrain with small differences in spike timing (a). The second parameter set was also like the target parameter set but with much less sodium channels to disable spiking (b). a,b) The squared distance measure has a large value (39492 mV2 · ms) for the first test solution while the second test solution is given a smaller value (23583 mV2 · ms). The second test solution would therefore be preferred. c,d) Using LeMasson and Maex’ phase-plane distance measure the first test solution has more data points in the bins of the target solution than the second test solution. Therefore the first test solution obtains the smaller distance value and will be preferred over the second. reproducing even tiny details of the complex firing patterns. Interestingly the resulting 17 CHAPTER 1. INTRODUCTION parameter sets leading to similar firing patterns were distinct. Two important conclusions were made by the authors: First, the originally hand-tuned model is only one of many good models and therefore the channel densities cannot be regarded as representing the original channel distribution in the real neuron. Second, if similar firing patterns can be reproduced with different ion channel densities, then there might also be some homeostasis during development adjusting the neuron’s parameters in order to reproduce the complex firing patterns of real Purkinje neurons. Unfortunately this study has not been able to fit experimental data yet and it appears that the phase-plane distance measure is not sufficient for this task as it overestimates the errors in spiking traces below threshold (Druckmann et al., 2008). Another distance measure which is independent of spike-timing was used by Weaver & Wearne (2006). The model was a single-compartment model with several currents and calcium dynamics. For one parameter set they calculated the target spiketrain and simulated annealing was used to fit the model to this data. As for the error function, each AP of the target trace was aligned with its corresponding AP of the test trace and a mean-square distance was calculated. This value was combined with an error in firing rate. They found good fits for the whole spiketrain. Although the model might be too simple to fit experimental data, this study shows that the spike shape contains a lot of information about the ion channel distribution involved in AP generation. To constrain the parameters of a reduced multi-compartment model of a layer 5 pyramidal neuron Keren et al. (2005) tested different error functions and a genetic algorithm to fit surrogate data. They found out that one single error function is not sufficient if only one somatic voltage recording is available, but that several error functions need to be combined. In particular for constraining an entire neuron, they suggested that several voltage recordings from different locations in the neuron are needed. Although these suggestions were based on surrogate data, Keren et al. (2009) performed a preceding study aiming to fit the ion channel distribution in the soma and apical dendrite of a simplified neuron model to experimental data. Interestingly the combined error function of a somatic and apical dendritic recording initially failed to constrain the model. Therefore they modified the functions describing the decay or growth of the ion channel density in the dendrite which finally led to a successful fitting of the neuron model. Besides the fact that this study was one of the first that managed to automatically fit a model to experimental data, they also introduced a “parameter-peeling” strategy to reduce the number of free parameters per optimization step. This means that the passive parameters are fitted first. These parameters are then fixed and the remaining active parameters are fitted in following steps using different ion channel blockers. They also realized that passive and active parameters are not completely independent and suggested a way to estimate both sets of parameters in spite of these difficult dependencies. Druckmann et al. (2007) introduced a multi-objective optimization (MOO) strategy 18 1.5. AUTOMATIC FITTING STRATEGIES using Evolutionary Algorithms (EAs) (Deb, 2001) into computational neuroscience. Unlike previous studies using EAs minimizing a single error function (which might be a combination of several weighted separate error functions) this approach can minimize multiple error functions independently. Therefore it was possible to extract several meaningful features from a given spiketrain, like the spike height, rate or width and optimize the model parameters in respect to each feature without an arbitrary weighting. This approach was robust and overcame the problem that the model might be insufficient to perfectly fit experimental data in all features. Druckmann et al. (2007) could fit the somatic conductances and the passive membrane parameters of a detailed model of a nest basket cell to experimentally recorded spiketrains. Summarizing we observe constant progress in the field of automatic fitting algorithms for neuron models and we see that a key to success is a good distance measure and a powerful search strategy. We have also seen that it is useful to separate the parameter space to fit the subsets of parameters independently of each other. As today’s computer power available to the researcher is constantly increasing further elaborate studies with even more parameters and error functions will be possible. 19 CHAPTER 2 Methods 2.1. Experiment Experiments were performed by Bekkers & Häusser (2007). Briefly, Sprague-Dawley or Wistar (17- to 25-days-old) rats were anesthetized with Isoflurane and rapidly decapitated. Slices (300 µm thick) were prepared from the somatosensory cortex and maintained at 32 ◦ C to 34 ◦ C. A MultiClamp 700A amplifier (Molecular Devices, Union City, CA) was used to obtain whole-cell recordings from the somata of visually identified large layer 5 pyramidal neurons. Recordings were low-pass filtered at 10 kHz and sampled at 50 kHz. In current-clamp recordings pyramidal neurons were allowed to remain at their resting potential (≈ -67 mV). Voltages have not been corrected for the liquid junction potential (≈ -7 mV). Electrophysiological recordings were performed under two conditions: First, in the intact neuron. Then the apical dendrite and the soma were separated by a method called Pinching and the recording protocol was repeated. Pinching was performed by attaching two pincer pipettes to the proximal site of the apical dendrite (≈ 20 µm from the soma) and moving them slowly against each other. In successful experiments it was tested whether the initial properties of the cell membrane were restored by releasing the pinch to ensure that pinching had not destroyed the cell. Pincer pipettes resembled sharp intracellular electrodes with a shallow taper and very fine tips. 2.2. High-Resolution Alignment of APs In electrophysiological experiments the voltage is normally recorded with less than 100 kHz. Higher recording frequencies would require more elaborate equipment and lead to larger data files. However, in most experiments recording frequencies of 10 kHz or less are sufficient to observe the desired properties. APs in cortical pyramidal neurons have a rise time of approximately 0.2 ms, hence even with 100 kHz recording frequency we obtain only about 20 data points in the region between onset and peak. Additionally due to different sources of noise, each recorded AP is slightly different from the other. 21 CHAPTER 2. METHODS Therefore a single recorded AP can only give an approximate estimate of the AP shape. Thus, to obtain a high-resolution AP it is necessary to average data from many APs. a V 2 dV/dt 2 d V/dt 5 V/ms 30 mV 2 200 V/s b 0.3 ms Figure 2.1: H IGH -R ESOLUTION A LIGNMENT OF AP S a) The data from three APs from the same spike train as well as their first dV /dt and second d 2V /dt 2 derivatives are shown (black dots). The dashed lines are cubic spline interpolations fitted to V (t) (but not to the derivatives) of each AP onset. It can be seen that the position of the peak of each interpolated AP is slightly different. The first and second derivatives of the interpolations are even more variable and follow only barely their corresponding data points. b) The interpolated APs were peak-aligned. It can be seen that each interpolated AP is different, notably the first and second derivatives. Only the average of the peak-aligned interpolated APs provides a detailed picture of the AP onset shape including its derivatives (red lines). It was shown that a precise and high-resolution alignment of APs can be achieved without high recording frequencies (Wheeler & Smith, 1988): For each low-resolution AP the position of the peak is determined. This peak is not the real peak of that AP but lays somewhat nearby. Now a cubic spline interpolation is applied around that position which allows a prediction of the real AP shape. This is done with every AP. These interpolated APs are then peak-aligned and averaged. The result is a high-resolution average of many low-resolution APs (Fig. 2.1). 22 2.3. MODELLING IN NEURON The experimental data we use in this study were recorded with 50 kHz. We downsampled the recording frequency to different frequencies and tested whether the method could recover the detailed shape of the AP. For sampling frequencies below 25 kHz we did not observe a prominent biphasic AP onset anymore as seen in the first and second derivative averages in Fig. 2.1. We therefore conclude that at least recording frequencies of 25 kHz are necessary to analyze detailed AP onset shapes. 2.3. Modelling in NEURON The creation and analysis of detailed neuronal morphologies with many compartments and ion channels is very simple in NEURON (Carnevale & Hines, 2006). A neuron is described by a set of sections that are connected to each other. A section is a continuous length of unbranched cylindrical cable with its own anatomical and biophysical properties. Each section is automatically divided into several compartments. It is easy to introduce several point mechanisms, like synapses or current electrodes in each compartment separately. Ion channel equations can be set up externally in so called .mod-files. These files need to be compiled via nrnivmodl (Linux) before the mechanisms can be inserted into the sections. Based on the resulting properties of each compartment NEURON automatically creates the system of differential equations representing the neuron and applies efficient solving strategies. The neuron model can also be split into several components to be calculated on multi-core architectures (Eichner et al., 2009; Hines et al., 2008). During the calculation it is possible to visualize or record any variable in the cell. NEURON has its own programming language (hoc) to describe, analyze and control the models but also offers a graphical user interface. To give an example let us create a “simple” model with 25 compartments: // example.hoc nrn_load_dll("../../results/channels/x86_64/.libs/libnrnmech.so") load_file("stdrun.hoc") celsius = 37 create soma create dend soma { diam = 20 L = 20 23 CHAPTER 2. METHODS nseg = 5 insert nat insert kfast ena = 50 ek = -80 gbar_nat = 3000 gbar_kfast = 500 } dend { diam = 5 L = 500 nseg = 20 } forall { insert pas e_pas = -70 g_pas = 1./15000 } connect soma(1),dend(0) objref stim soma stim = new IClamp(0.5) stim.del = 100 stim.dur = 5 stim.amp = 0.5 At the beginning we must load the compiled ion channel mechanisms libnrnmech.so which are located elsewhere. We also need to load the main hoc-routines for NEURON (stdrun.hoc) before we can specify the model. We set the overall temperature to 37 ◦ C as ion channel kinetics are highly sensitive to temperature and our goal is to model in vivo conditions. We create a somatic (soma) and a dendritic (dend) section. The somatic cylinder is 20 µm long and has a diameter of 20 µm. The dendrite is 500 µm long and has a diameter of 5 µm. The soma is divided into 5 and the dendrite into 20 compart24 2.3. MODELLING IN NEURON ments. The somatic section contains transient sodium (nat) and fast potassium channels (kfast) (for detailed ion channel description see sec. 3.2). We set the reversal potential for sodium Ena = 50 mV and for potassium Ek = -80 mV. A passive leak conductance (pas) is inserted in all compartments. The reversal potential for the leak current is set to Epas = -70 mV and the specific membrane resistance is set to rm = 15000 Ω · cm2 . A current electrode (IClamp) is placed in the middle of the soma. A constant current of 0.5 nA is switched on after 100 ms for 5 ms. It would have been a huge amount of work to implement this simple model in MATLAB or C++, for example. The source code would be extremely error-prone, especially if one decides to change something in the model, like the number of compartments for the dendrite. NEURON extremely simplifies the modelling of complex neuron geometries and biophysics. 2.3.1. Using Python to Control NEURON The recent versions of NEURON provide an interface to Python1 (Hines et al., 2009). This offers the opportunity to control the cell models and the NEURON-simulation from Python, but do all data analysis and specific model modification in Python. Let us continue our “simple” example: # example.py from neuron import h import pylab as P h.load_file("example.hoc") h(""" v_init = -70 tstop = 150 objref time = rec1 = rec2 = time, rec1, rec2 new Vector() new Vector() new Vector() time.record(&t) rec1.record(&soma.v(0.5)) rec2.record(&dend.v(0.9)) 1 The Python Programming Language: www.python.org 25 CHAPTER 2. METHODS """) h.init() h.run() P.plot(h.time, h.rec1, ’b-’) P.plot(h.time, h.rec2, ’b--’) P.show() We import the neuron module as well as the Python-plotting package pylab. Now we load the cell model example.hoc. In the following we prepare the simulation using the hoc-interface. We set the initial voltage v_init = -70 mV and set the overall simulation time tstop = 150 ms. We create three vectors, one to record the time and the other two to record the somatic and dendritic voltage. We initialize the model and finally run the simulation. NEURON performs the evaluation of the model and records the variables. These variables are immediately available to Python and can be plotted (Fig. 2.2) or analyzed further using any scientific Python package. soma.v(0.5) dend.v(0.9) Figure 2.2: E XAMPLE NEURON-P YTHON S IMULATION R ESULTS The simple model of a soma and dendrite was loaded into NEURON and controlled with Python. We recorded the voltages in the middle of the soma (soma.v(0.5)) and at the end of the dendrite (dend.v(0.9)). The model initiates a somatic AP in response to the short current step. The voltage spreads and decays into the dendrite. 10 mV -70 100 1 ms 2.4. Multi-Objective Optimization using EAs The most significant number of real world problems involve more than one single objective and therefore several error values need to be minimized during an optimization procedure. Classical optimization algorithms only minimize one single error function and if multiple error functions are taken into account they are weighted and summed into a single one. But the choice of weights, especially for objectives with different units is highly arbitrary. Furthermore conflicts between the different objectives are neglected and the final optimal solution might not represent the desired trade-off. To overcome these problems we use a multi-objective optimization (MOO) strategy that optimizes several 26 2.4. MULTI-OBJECTIVE OPTIMIZATION USING EAS Harmful Gases Produced objectives simultaneously (Deb et al., 2002; Druckmann et al., 2007). This way of sorting leads to a number of trade-off optimal solutions that are optimal in each objective separately, but also contain mixed solutions. An illustration of a two-objective optimization problem is shown in Fig. 2.3. 1 0.8 0.6 0.4 0.2 0 0 0.2 0.4 0.6 0.8 Production Cost 1 Figure 2.3: I LLUSTRATION OF A S IMPLE T WO -O BJECTIVE O PTIMIZATION P ROBLEM Let us consider a car factory that while producing cars creates emissions. Most of the car production strategies will lead to solutions that are bad in both objectives (gray shaded area). The owners of the factory want to reduce the cost of car production and at the same time it is desired to minimize the amount of harmful gases produced. But minimizing the amount of produced gases will lead to high production costs and minimizing the production costs will strongly increase the amount of harmful gases produced and hence both objectives cannot be minimized at the same time. A multi-objective optimization strategy could find the optimal trade-off solutions (black line) while classical single error function minimization would only find one single particular solution (red dot). Minimization is often done with simple gradient-based methods: A random starting point in the search space defines an initial parameter set. Around this point the model is evaluated and the direction of parameter change is applied such that the improvement of the solutions is maximal. This works very well when the error surface is smooth but most of real-world problems cannot be described by smooth error functions and such a strategy would get stuck in local minima. One can improve the gradient-based methods by introducing random parameter fluctuations (simulated annealing) but if the search space becomes more complex the randomization of parameters is not sufficient to lead the optimization algorithm to the global minimum. Here evolutionary strategies have proven to be very effective. In this study, we have implemented a very general Python-framework for the MOOstrategy using EAs which can be easily used for any optimization problem. 2.4.1. Evolutionary Algorithms EAs mimic natural evolutionary principles like selection, crossover and mutation. One parameter set is called an individual and all individuals form the population. Each individual is associated with a fitness value that describes its ranking in the population and 27 CHAPTER 2. METHODS only good individuals are transferred into a new generation. Parameter combinations among individuals are exchanged during evolution and the whole population is guided to better solutions (Deb, 2001, chap. 4). EAs can be extended to multi-objective problems and can be easily used on a parallel machine. In the following we will describe the essential steps of this optimization strategy (Fig. 2.4). Begin Figure 2.4: F LOWCHART OF THE W ORK ING P RINCIPLE OF AN EA A random population is created and the selection, mutation and crossover operators are applied. Then the population is evaluated and each individual is associated with a unique fitness value. If several objectives are considered the sorting is done by the multi-objective-sorting strategy. After sorting only a subset of all individuals is transferred into the new generation. The evolution is repeated until a stop-criterion is fulfilled. Create Population; gen = 0 No? Cond? gen = gen + 1 Selection Yes? Stop Crossover Evaluate Individuals Mutation Set Fitness for Individuals Multi-Objective Sorting Create Population The search space is filled with a set of random solutions that form the initial population of size N. To simplify and generalize the following steps, each parameter is normalized between 0 and 1. The individuals are evaluated based on the chosen error functions to determine their fitness. Selection In the selection step a so called mating pool is formed. The population is shuffled and individuals are pairwise compared and only the individual with the higher fitness is copied into the mating pool. This step is done twice, so that the mating pool contains N individuals again. Thus, no matter what the order of individuals was, better individuals will have a higher chance to be transferred. The best individual will be definitely found twice in the mating pool as it is always compared with a worse one and the worst individual will definitely not reach the mating pool as it can only be compared with a better one. Hence the mating pool accumulates good solutions that are ready to “mate” in order to exchange their parameters in the next step. Crossover Two individuals (parents) are randomly selected from the mating pool and the crossover operator is applied to produce two offspring. We have chosen the Simulated Binary Crossover, but there are other possible operators (Deb, 2001, chap. 4). The crossover 28 2.4. MULTI-OBJECTIVE OPTIMIZATION USING EAS operator is applied to each parameter independently: First a random number ui between 0 and 1 is drawn from a uniform distribution. Then the value (2 · ui ) ηc1+1 if ui ≤ 0.5 ; βqi = (2.1) 1 1 ηc +1 , otherwise 2 · (1−ui ) is calculated. After obtaining βqi from the above probability distribution, the offspring parameters are calculated as follows: y1i = 0.5 · (1 + βqi ) · xi1 + (1 − βqi ) · xi2 y2i = 0.5 · (1 − βqi ) · xi1 + (1 + βqi ) · xi2 . (2.2) (2.3) Here y1i is the ith parameter of the first and y2i of the second offspring. xi1 and y2i are the ith parameters of the parents respectively. A large value of ηc gives a higher probability for creating “near-parent” solutions and a small value of ηc allows distant solutions to be selected as offspring. In this study we use ηc = 10. If an offspring parameter was obtained that was outside the allowed parameter range the value was repositioned: y → 0 i y → 1 i if yi < 0 ; if yi > 1 . (2.4) The crossover operator is applied until the total size of the parent and the produced offspring population have reached a certain capacity C. Offspring and parent population are combined to a new temporal population. The effect of the crossover operator is shown in Fig. 2.5a. After crossover we find many individuals that contain parameters that lie between the parent parameters. But we have not created entirely new individuals. This is done in the next step. Mutation To increase the diversity of individuals we apply the mutation operator to the temporal population. We use Polynomial Mutation, but again many other operators are available (Deb, 2001, chap. 4). Each parameter for each individual is mutated independently. To determine the strength of mutation we first draw a random variable ri from a uniform distribution between 0 and 1. Then the value δi is calculated: (2 · r ) ηm1+1 − 1 if ri < 0.5 ; i δi = (2.5) 1 1 − [2 · (1 − r )] ηm +1 , otherwise . i 29 CHAPTER 2. METHODS ηc = 10 ηc = 50 b Crossover ηm = 20 ηm = 100 Occurence in % Occurence in % a 0.8 0.6 0.4 0.2 0 0 Mutation 0.8 0.6 0.4 0.2 0 0.2 0.4 0.6 0.8 1 Normalized Parameter 0 0.2 0.4 0.6 0.8 1 Normalized Parameter Figure 2.5: V ISUALIZATION OF THE C ROSSOVER AND M UTATION O PERATORS To visualize the effects of the crossover and mutation operators we created a large population of N = 10000 individuals each with one single parameter x. We observe how the parameter distribution looks like after applying the operators. a) Half of the population parameters were set to x = 0.4 and the other half to x = 0.6 (red dots). The crossover operator was applied until a capacity of C = 20000 was reached. Then the histogram of the resulting parameter distribution was calculated. We see that the majority of solutions lies near one of the parents Using a small ηc = 5 (black line) leads to a large variability of parameters, while a large ηc = 55 (blue line) causes almost no change in the parameter distribution. b) All parameters are set to x = 0.5 (red dot). Then the mutation operator is applied to the population. Only 20 % of the individuals were mutated. We calculated the histogram of the resulting parameter distribution. It can be seen that the mutation operator leads to a broad distribution if ηm = 20 is small while the distribution remains relatively narrow if ηm = 100 is large. The ith parameter is then changed by δi : xi → xi + δi . (2.6) The value of ηm defines the strength of the mutation. Using a small ηm it is likely to lead to “distant mutations”, while the mutation operator with a large ηm has only minor effects. In this study, we use ηm = 20. We also defined a mutation probability pc = 0.2/k, where k is the number of parameters of the optimization problem. This means that on average one parameter per individual will be changed in 20 % of all individuals. If a parameter change led to a value outside the allowed parameter range it was repositioned: x → 0 i x → 1 i if xi < 0 ; (2.7) if xi > 1 . The effect of the mutation operator is shown in Fig. 2.5b. After we have applied the selection, crossover and mutation operators, we are left with a temporal diverse population that has accumulated the parameter combinations that lead to good solutions, but also contains entirely new parameter sets. After these randomization procedures the population needs to be sorted in order that each individual can obtain its ranking. This is done in the next two steps. 30 2.4. MULTI-OBJECTIVE OPTIMIZATION USING EAS Evaluation For each individual the model is evaluated based on the given parameter set and the result is compared with the target data. If M objectives are considered, this leads to a set of M error values that are associated with the individual. Set Fitness Finally the individuals are given a fitness value to describe their ranking in the population. In the case of a single error function this can be done by sorting all individuals in respect to their associated error value. If several independent error functions are considered a multi-objective sorting concept can be used, which is explained below. New Generation The temporal population of size C is sorted by the fitness of its individuals and only the first N individuals are transferred into the new generation. The generation counter is incremented and the optimization loop starts again. It is possible to define any stop-criterion. In this study, we simply stop after having reached a certain number of generations. 2.4.2. Multi-Objective Sorting If several independent error functions are to be minimized the sorting procedure can be replaced by the concept of nondomination. One solution (x1 ) dominates the other (x2 ) if it is better in at least one objective ( fk ), but not worse in any other ( f j ): f j (x1 ) ≤ f j (x2 ) for all j = 1...M ; fk (x1 ) < fk (x2 ) for at least one k = 1 . . . M . (2.8) (2.9) Analyzing the domination relation in the whole population leads to a set of solutions that are not dominated by any other solution (first Pareto front (PF), nondomination front). The PF of the rest of solutions is ranked second and so on (Fig. 2.6a). In this study, we use an efficient algorithm (NSGA-II) to calculate the PFs that was developed by Deb et al. (2002). The PFs give the individuals a first ranking in the population. However within such a front it is not possible to call one individual better than the other without making further specification. Nevertheless within a PF the individuals must also be ranked. The solutions that have the smallest error in one of the objectives are immediately ranked best in the front. The remaining solutions are ranked using the crowding distance measure (Deb, 2001, p. 248). For a solution i the crowding distance dmi in respect to the objective m, is defined as the 31 CHAPTER 2. METHODS distance between its direct neighbors in the same PF: dmi = fmi+1 − fmi−1 (2.10) where fm is the objective value for objective m. The total crowding distance d i for solution i is the weighted sum of the crowding distances for each objective: di = ∑ m dmi . fmmax − fmmin (2.11) Therefore very “lonely” solutions obtain higher values, whereas solutions in very “crowded” regions of the error space get smaller values (see Fig. 2.6b). By this means each individual in a PF obtains a further ranking and hence a unique fitness for every individual in the population can be determined. a b 200 100 100 i-1 f2 f2 150 150 50 PF 1 PF 2 PF 3 50 0 0 50 100 f1 d 2i i d1 i i+1 0 150 200 0 20 40 60 f1 Figure 2.6: I LLUSTRATION OF THE R ANKING -C ONCEPT USING PARETO F RONTS AND THE C ROWDING D ISTANCE For two parameters x, y ∈ [−10, 10] we calculated two objective functions f1 = 0.125 · (20 + x + y)2 and f2 = 100 + x · y. It can be seen from the formulas that both objectives are conflicting and cannot be minimized at the same time. MOO will help to find the best trade-off solutions for this problem. a) The objectives for the initial random population N = 50 are shown (black dots). The first, second and third Pareto front (PF) was calculated (colored dots). The PFs give the individuals a first ranking in the population. Black crosses symbolize the first PF after an evolution of 10 generations and show the desired trade-off solutions. b) A magnification of the first PF of the initial population is shown (red dots). Using the crowding distance measure each individual can obtain a further ranking. The measure is illustrated for a solution i (black dot) which has the highest crowding distance. We calculate the crowding distance d i of i using the distance between the two nearest neighbors i − 1 and i + 1 in respect to each objective (dashed lines). As solution i is located in a very “lonely” region of the first PF it obtains a very good fitness (boundary individuals are given the best fitness). Using both concepts, the Pareto-optimal ranking and the crowding distance measure it is ensured that the optimization minimizes all error functions at the same time and that it directs individuals into unexplored regions of the error space to obtain a good spread of solutions. 32 2.4. MULTI-OBJECTIVE OPTIMIZATION USING EAS 2.4.3. Parallelization As the principle of an EA is the evaluation of independent individuals it is very easy to evaluate the population via multiple processors. Parallelization in Python is very simple via the package mpi4py (Dalcin et al., 2008) and only needs a few lines of code. All evolutionary steps are done on the master-node. Then the master-node splits the population into equally sized packages and sends the list of individuals to the slave-nodes. The slave-nodes start the evaluation once they receive their package of individuals and submit the resulting error values to the master-node. For the optimization of complex models the evaluation of one parameter set is the most time consuming step and parallelization can reduce the time of population evaluation by a factor of the number of processors used. In this study, we use a Beowulf cluster with 44 processors.2 2 Cluster-access was kindly provided by the Colamn-Project: ViewGrant.aspx?GrantRef=EP/C010841/1 33 http://gow.epsrc.ac.uk/ CHAPTER 3 The Cell Model 3.1. Neuronal Morphology We created a reduced morphology of a cortical layer 5 pyramidal neuron. In order to obtain reasonable parameters for the geometry, we started with a detailed reconstruction and simplified the geometry while maintaining the neuron’s passive response properties. After this was done, we appended an axon. The final morphology consists of 7 functional sections and is described with a 39-compartment model. 3.1.1. Geometry Reduction Bekkers & Häusser (2007) did not reconstruct the cells after the experiments. Thus we needed to choose a detailed morphological reconstruction of another cortical layer 5 pyramidal neuron from a rat with similar age (Stuart & Spruston, 1998). We removed all ion channels, including the HCN-channels from the complex model and set the passive membrane parameters constant over the whole morphology (rm = 15000 Ω · cm2 , cm = 1 µF/cm2 , Epas = -70 mV, ra = 100 Ω · cm). We divided the complex neuron model into four functional sections (soma, basal dendrites, apical dendrite, tuft) (Fig. 3.1a) and determined the membrane area of each. They were Asoma = 1682 µm2 , Abasal = 7060 µm2 , Aapical = 9312 µm2 and Atuft = 9434 µm2 . In order to obtain a simplified geometry with passive response properties similar to those of the complex model, we modified the simplification strategy suggested by Destexhe (2001): For a given length Lx of a reduced section x the diameter dx of that section was always adjusted such that the resulting membrane area of the cylinder matched the area of the subset of dendrites it represents (Ax ): dx = Ax . π · Lx 35 (3.1) CHAPTER 3. THE CELL MODEL For the somatic cylinder we set: r dsoma = Lsoma = Asoma . π (3.2) Beside the uncertainty about the precise length of such a reduced section, it is unclear what the magnitude of its intracellular resistivity (ra ) has to look like after simplification. However the specific passive membrane parameters (rm , cm ) and the leak-reversal potential Epas were set to the same values as in the complex model. Hence, the entire reduced passive neuron model can be described with 8 geometrical parameters, 4 parameters for the intracellular resistivities as well as 3 membrane parameters, but only 7 parameters were free and used to optimize the model (Tab. 3.1). Parameter ra soma Lbasal ra basal Lapical ra apical Ltuft ra tuft dsoma Lsoma dbasal dapical dtuft rm global Epas global cm global Result 82.0 257.0 734.0 500.0 261.0 499.0 527.0 23.1 23.1 8.7 5.9 6.0 15000.0 -70.0 1.0 LS Bound 80.0 170.0 700.0 500.0 150.0 400.0 500.0 - US Bound 200.0 280.0 2000.0 800.0 300.0 600.0 1200.0 - Unit Ω · cm µm Ω · cm µm Ω · cm µm Ω · cm µm µm µm µm µm Ω · cm2 mV µF/µm2 Table 3.1: O PTIMAL G EOMETRICAL AND PASSIVE PARAMETERS FOR THE R EDUCED M ODEL AFTER S IMPLIFICATION The final set of parameters for the reduced neuron model is shown that was obtained after optimizing its passive response properties. Only 7 parameters were free and used to constrain the model: The length L and ra for each cylindrical section as well as ra for the soma. The lower search bound (LS Bound) and the upper search bound (US Bound) defined the allowed region in parameter space for each parameter during the search. The parameters for the initial random population were uniformly distributed in that region. The section diameters were not free and always adjusted using Eqn. 3.1. The diameter and the length of the soma were calculated using Eqn. 3.2. The remaining membrane parameters were the same as in the complex model and not changed. All values were rounded. It was already shown in earlier studies that an adequate estimate of the passive parameters can be found by optimizing the neuron’s input impedance and phase-shift (Borst & Haag, 1996). Injection of oscillating input current with a certain frequency f leads to an oscillation of the membrane potential with the same frequency. We describe the relative amplitudes between the sinusoidal input current and the membrane voltage 36 3.1. NEURONAL MORPHOLOGY with an impedance Z( f ) and the shift of the oscillation phase is described with a phase-shift θ ( f ). The membrane time constant determines the time that is necessary to charge the membrane. In case of fast input oscillation the membrane does not have enough time to accumulate charge before the next oscillation phase begins. Therefore the input current starts to hyperpolarize the membrane before an equilibrium potential is reached. The higher the input frequency the stronger this effect and the lower the membrane oscillation amplitude. This analysis was only done with injections and recordings at the soma. Nevertheless the impedance and phase-shift curves depend on dendritic geometry and can therefore help us to optimize the passive properties of an entire neuron. To obtain additional information about the voltage distribution in the neuron, we also injected a constant input current into the soma and measured the steady state voltage distribution. Distance Functions We optimized the simplified geometry such that the four objectives somatic steady state voltage (Vs (0)), voltage attenuation (Vs (x)), somatic input impedance (Zsoma ( f )) and somatic phase-shift (θsoma ( f )) mimicked those objectives in the complex model. To obtain a better resolution for the steady state voltage distribution, we divided each of the sections into 20 compartments during the optimization. For the calculation of the four error functions (e1 , e2 , e3 , e4 ) we used the squared distance measure: 2 e1 = 0.5 · Vscomplex (0) −Vsreduced (0) 2 e2 = 0.5 · ∑ Vscomplex (x) −Vsreduced (x) (3.3) (3.4) x∈X e3 = 0.5 · ∑ 2 complex reduced Zsoma ( f ) − Zsoma (f) (3.5) ∑ 2 complex reduced θsoma ( f ) − θsoma (f) (3.6) f ∈F e4 = 0.5 · f ∈F where X is the vector of distances at different locations in the neuron and F the vector of chosen frequencies. In order to minimize all four error functions independently, we applied the multi-objective optimization (MOO) strategy (sec. 2.4) to automatically and systematically explore the parameter space for good solutions. The population size was N = 400, the population capacity was C = 800 and the evolution was performed for 100 generations. Mutation and crossover parameters were as described (p. 27). Search boundaries and the final parameter results are given in Tab 3.1. Voltage attenuation, somatic input impedance and the phase-shift for the complex and the optimized simplified model are compared in Fig. 3.1. The passive response properties of the simplified neuron 37 CHAPTER 3. THE CELL MODEL model match well with those of the complex neuron model. c b SS Voltage (mV) a tuft Complex Model Reduced Model -100 -110 -120 -130 -140 d 100 !m apical Soma Phase-Shift (rad) e Soma Impedance (M!) -300 0 300 600 900 Distance to soma (µm) soma basal 70 60 50 40 30 20 10 0 0 200 400 600 800 1000 f (Hz) 0 200 400 600 800 1000 f (Hz) 0 -0.4 -0.8 -1.2 Figure 3.1: M ORPHOLOGY AND PASSIVE P ROPERTIES FOR THE C OMPLEX AND THE R EDUCED M ODEL a) The detailed reconstruction of a layer 5 pyramidal neuron taken from Stuart & Spruston (1998). We divided the complex morphology into 4 functional sections: The soma, the basal dendrites, the apical dendrites and the tuft. The oblique dendrites are taken to be part of the apical dendrites. b) An illustration of the reduced model (not to scale). The model consists of 4 cylinders that represent the sections described in (a). Each of the cylinder was divided into 20 compartments to increase the number of data points for plotting and geometrical optimization. c) A constant current (-1 nA) was injected into the somata of both neurons and the steady state voltage distribution was determined for the complex model (black dots) and the optimized reduced model (red dots). d) An oscillating input current with low amplitude was injected into the somata. The amplitude of the resulting membrane potential oscillation was used to calculate the somatic input impedance for the complex model (black line) and the optimized reduced model (red line). e) We calculated the somatic phase-shift between the input current and membrane potential oscillation for the complex model (black line) and the optimized reduced model (red line). Noise Test However it should also be tested whether the reduced model is really a good representation of the passive properties of the complex model. To do this we injected white-noise input current into the soma of the complex neuron model and the same noisy input into the somatic cylinder of the simplified model. For both models we recorded the voltage in the soma and in a distal location in the apical dendrite (Fig. 3.2). The voltage traces in the 38 3.1. NEURONAL MORPHOLOGY soma and in the distal dendrite are very similar for both models and we therefore conclude that we have found an adequate simplification of the complex dendritic geometry that we can safely use in the following steps. Complex Model Reduced Model Apical Distal 5 mV Soma 5 mV 2 nA 0 50 100 150 Time (ms) 200 250 300 Figure 3.2: C OMPARISON OF THE VOLTAGE T RACES IN THE C OMPLEX AND R EDUCED M ODEL IN R ESPONSE TO N OISY I NPUT C URRENT To test whether the reduced model is a good approximation of the complex model, we analyzed the response to white noise current injection in both models. The same random current was injected into the somata of the models (green trace). For both models the somatic voltage trace as well the voltage trace distally (≈ 425 µm from the soma) were recorded. The traces of the complex model (black) and of the reduced model (red dashed line) are almost indistinguishable. 3.1.2. Axon Geometry To address the question of AP initiation we appended an axon to the simplified morphology. The axonal geometry is based on a detailed reconstruction of a cortical layer 5 pyramidal neuron (Zhu, 2000). We represented that axon by three sections with different starting diameters and length (Fig. 3.3). The axonal parameters ra , rm , cm , Epas were the same as in the soma. We did not model axonal nodes of Ranvier or segments of myelin. soma hillock axon iseg 39 Figure 3.3: G EOMETRY OF THE A XON FOR THE R EDUCED M ODEL An illustration of the axonal geometry is shown (not to scale). The axon hillock is directly connected to the soma and starts with a diameter of 3.5 µm. Within 20 µm the hillock tapers to 2 µm where the axon initial segment begins. The axon initial segment had a length of 25 µm and the diameter tapers to 1.5 µm. The rest of the axon had a uniform diameter of 1.5 µm and a length of 500 µm. CHAPTER 3. THE CELL MODEL 3.1.3. Segmentation The reduced model should be able to show correct AP initiation and propagation. When APs are travelling in multi-compartment models, it is important to use a sufficient high number of compartments within one segment (Carnevale & Hines, 2006). In our model the axon initial segment and the axon hillock consist of 5 compartments each. The apical dendrite was divided into 16 and the tuft into 10 compartments. The soma, the basal dendrite and the axon were represented by a single compartment each. Thus the simplified model consists of 39 compartments in total. 3.2. Ion Channel Kinetics and Distribution All ionic currents in this study are modeled in the standard Hodgkin-Huxley style (for a general overview see sec. 1.2.2) and based on published ion channel models1 . In this study we are using the unit pS/µm2 for all specific ionic conductances while Hodgkin and Huxley reported the units in mS/cm2 (Hodgkin & Huxley, 1952b). As NEURON requires the unit mA/cm2 for the specific ionic current and the membrane potential is given in mV we need to include the factor 10−4 in the equation for the specific ionic currents (Eqn. 1.18). It is assumed that all ion channels of the same type have the same single channel conductance γ if they are open. Fast sodium channels for example were reported to have γ ≈ 14 pS (Koch, 2004, p. 197). We use deterministic ion channel models to describe the conductance of an ensemble of ion channels. Thus for an ion channel type x the maximal specific ionic conductance (gbarx ) is proportional to the ion channel density (ηx ). Both terms are often used equivalently. Ion channels are normally faster and their conductance increases with higher temperatures. We are using a temperature of celsius = 37 ◦ C in our modelling study but the ion channel models were created under different conditions. To properly adjust the rate constants and the peak conductances a temperature adjustment factor (tadj ) was introduced (Hodgkin & Huxley, 1952b). The steady state values and time constants for the gating variables in this study are illustrated in Fig. 3.4 and explained in the following sections. We did not model calcium channels or any other calcium related mechanisms in our model. 3.2.1. Hyperpolarization-Activated Cation Channel The hyperpolarization-activated cyclic nucleotide-gated cation channel (HCN-channel) gives rise to the h-current. The kinetic scheme and parameters for the channel were taken 1 ModelDB: http://senselab.med.yale.edu/ModelDB 40 3.2. ION CHANNEL KINETICS AND DISTRIBUTION τp (ms) p∞ 1 HCN 0.75 100 0.5 10 1 0.25 0.1 0 0.01 -100 -50 0 50 1 Nat 100 0.5 10 50 m h 1 0.01 -50 0 50 1 -100 -50 100 0.5 10 0 50 n 1000 0.75 1 0.25 0.1 0 0.01 -100 -50 0 50 1 -100 -50 0.75 100 0.5 10 0 50 a b b1 1000 1 0.25 0.1 0 0.01 -100 -50 0 50 1 -100 -50 100 0.5 10 0 50 m 1000 0.75 1 0.25 0.1 0 0.01 -100 -50 0 50 1 Km 0 0.1 -100 Nap -50 1000 0 Kslow -100 0.75 0.25 Kfast q 1000 -100 -50 100 0.5 10 50 m 1000 0.75 0 1 0.25 0.1 0 0.01 -100 -50 0 50 -100 -50 0 50 Clamped Voltage (mV) Figure 3.4: I ON C HANNEL G ATING PARTICLES U SED IN T HIS S TUDY The voltagedependent steady state values and time constants for the gating particles used to describe the ion channel kinetics in this study are shown. The first column shows the steady state values for the gating particles p∞ . The second column shows on a logarithmic scale the corresponding time constants. It can be seen that there is a huge range between the time constants from less than 0.1 ms to more than 3 s. Only physiologically relevant voltages are shown. 41 CHAPTER 3. THE CELL MODEL from Kole et al. (2006): α (v) = 0.001 · 6.43 · (V + 154.9) / (exp ((V + 154.9) /11.9) − 1) β (V ) = 0.001 · 193 · exp(V /33.1) τq (V ) = 1/ αq (V ) + βq (V ) q∞ (V ) = αq (V )/ αq (V ) + βq (V ) dq = (q∞ − q)/τq dt Ih = 10−4 · gbarHCN · q · (V − Eh ) . (3.7) (3.8) (3.9) (3.10) (3.11) (3.12) Only one gating particle q is required to describe the time- and voltage-dependent activation of the channel. The reversal potential was set to Eh = -47 mV. It was shown experimentally that HCN-channels in pyramidal neurons are mainly located in the tuft (Berger et al., 2001; Kole et al., 2006), but our data suggests that this channel type is also found in the soma. We therefore inserted the HCN-channels into the somatic and into the tuft section with a homogeneous density, but not into the apical dendrite. 3.2.2. Transient Sodium Channel We used a recent model of the transient sodium channel (Nat-channel) (Kole et al., 2006). The model equations are the following: Ṽ = V − vshift − vshift2 (3.13) αm (Ṽ ) = 0.182 · (Ṽ + 28)/(1 − exp(−(Ṽ + 28)/9)) (3.14) βm (Ṽ ) = −0.124 · (Ṽ + 28)/(1 − exp((Ṽ + 28)/9)) (3.15) αh (Ṽ ) = 0.024 · (Ṽ + 50)/(1 − exp(−(Ṽ + 50)/5)) (3.16) βh (Ṽ ) = −0.0091 · (Ṽ + 50)/(1 − exp((Ṽ + 50)/5)) (3.17) tadj = 2.3(celsius−23)/10 τm (Ṽ ) = 1/ tadj · αm (Ṽ ) + βm (Ṽ ) m∞ (Ṽ ) = αm (Ṽ )/ αm (Ṽ ) + βm (Ṽ ) τh (Ṽ ) = 1/ tadj · αh (Ṽ ) + βh (Ṽ h∞ (Ṽ ) = 1/(1 + exp((Ṽ + 55)/6.2)) dh = (h∞ − h)/τh dt INat = 10−4 · gbarNat · m3 · h · (V − Ena ) . (3.18) (3.19) (3.20) (3.21) (3.22) (3.23) (3.24) Here, an activation m and an inactivation h gating variable are required to describe 42 3.2. ION CHANNEL KINETICS AND DISTRIBUTION the channel’s opening state. The differential equations for the gating variables depend on a shifted membrane potential Ṽ . The first vshift was included because of uncertainties about the absolute voltage when channel kinetics were optimized based on experimental recordings (sec. 1.4). The vshift2 was added to have the opportunity to specify another voltage-shift precisely to one compartment. As done in previous studies (Keren et al., 2009; Mainen et al., 1995), Nat-channels were distributed in the soma, in the axon hillock and axon initial segment, as well in the apical dendrite and tuft. The sodium channel density in the apical dendrite and tuft is described with a linear decay: gbarNat (x) = f (gbarsoma Nat − decayNat · x) x if x ≥ 0 ; f (x) = 0 otherwise . (3.25) (3.26) x is the distance to the soma (in µm) and decayNat describes the slope of the decay. The rectifying function f (x) is needed in order that for any parameter combination channel densities cannot become negative. The reversal potential for sodium was set to Ena = 55 mV for the entire neuron. 3.2.3. Fast Potassium Channel The kinetics of the fast potassium channel (Kfast-channel) (Kole et al., 2006) are described by the following equations: αn (V ) = 0.02 · (V − 25)/(1 − exp(−(V − 25)/9)) (3.27) βn (V ) = −0.002 · (V − 25)/(1 − exp((V − 25)/9)) (3.28) tadj = 2.1(celsius−23)/10 (3.29) τn (V ) = 1/ tadj · (αn (V ) + βn (V )) (3.30) n∞ (V ) = αn (V )/(αn (V ) + βn (V )) dn = (n∞ − n)/τn dt IKfast = 10−4 · gbarKfast · n · (V − Ek ) . (3.31) (3.32) (3.33) As in previous studies, this channel type was distributed in the soma and in the apical dendritic tree (Keren et al., 2009). The channel density in the apical dendritic tree is described by an exponential function of the distance x (in µm) to the soma: gbarKfast (x) = gbarsoma Kfast · exp (−x/decayKfast ) . (3.34) decayKfast describes the distance within the channel density decays to a fraction of 1/e. The reversal potential for potassium is set to Ek = -80 mV. 43 CHAPTER 3. THE CELL MODEL 3.2.4. Slow Potassium Channel The kinetics of the slow potassium channel (Kslow-channel) (Korngreen & Sakmann, 2000) depend on three gating particles a, b and b1. The activation a is described by standard Hodgkin-Huxley kinetics: αa (V ) = 0.0052 · (V − 11.1)/(1 − exp(−(V − 11.1)/13.1)) (3.35) βa (V ) = 0.01938 · exp(−(V + 1.27)/71) − 0.0053 (3.36) τa (V ) = 1/(αa + βa ) (3.37) a∞ (V ) = αa /(α + βa ) da = (a∞ − a)/τa . dt (3.38) (3.39) The channel inactivation however was bi-exponential, therefore two inactivation particles were used: τb (V ) = 360 + (1010 + 23.7 · (V + 54)) · exp(−((V + 75)/48)2 ) (3.40) τb1 (V ) = 2350 + 1380 · exp(−0.01118 ·V ) − 210 · exp(−0.0306 ·V ) (3.41) b∞ (V ) = 1/ (1 + exp ((V + 58) /11)) db = (b∞ − b)/τb dt db1 = (b∞ − b1)/τb1 . dt (3.42) (3.43) (3.44) These gating particles are equally weighted in the final current equation: tadj = 2.3(celsius−21)/10 (3.45) IKslow = 10−4 ·tadj · gbarKslow · a2 · (0.5 · b + 0.5 · b1) · (V − Ek ) . (3.46) The slow potassium channel was inserted into the soma and into the apical dendritic tree. The channel density decays with distance x (in µm) to the soma: gbarKslow (x) = gbarsoma Kslow · exp (−x/decayKslow ) . (3.47) decayKslow describes the distance within the channel density decays to a fraction of 1/e. 3.2.5. Persistent Sodium Channel The persistent sodium channel (Nap-channel) was taken from Traub et al. (2003). This sodium channel activates very rapidly but does not show inactivation. It can, therefore, 44 3.3. DEFINING THE STATIC AND THE FREE PARAMETERS alter the overall excitability of the neuron. The kinetic equations are the following: m∞ (V ) = 1/(1 + exp(−(V + 48)/10)) 0.025 + 0.14 · exp((V + 40)/10) if v < −40 ; τm (V ) = 0.02 + 0.145 · exp(−(V + 40)/10) otherwise dm = (m∞ − m)/τm dt INap = 10−4 · gbarNap · m · (V − Ena ) . (3.48) (3.49) (3.50) (3.51) The Nap-channel was only present in the soma in our model. 3.2.6. Muscarinic Potassium Channel The muscarinic potassium channel (Km-channel) gives rise to the m-current. It is a noninactivating voltage-dependent very slow potassium channel. As it reaches its steady state opening in a timescale of seconds, it can lead to spike frequency adaptation. Kinetics were taken from Winograd et al. (2008): tadj = 2.3(celsius−36)/10 (3.52) m∞ (V ) = 1/(1 + exp(−(V + 35)/10)) (3.53) τm (V ) = 1000/ tadj · ((3.3 · exp((V + 35)/20) + exp(−(V + 35)/20))) (3.54) dm = (m∞ − m)/τm (3.55) dt Im = 10−4 · gbarKm · m · (V − Ek ) . (3.56) We used Km-channels only in the soma. 3.3. Defining the Static and the Free Parameters We have found a reasonable geometry for the reduced model, we have defined a set of ion channels that need to be present in our model and we have realized that approximately 39 compartments are needed to describe the interactions between dendrites, soma and axon. The model has therefore a very long list of parameters that could be modified. In order to obtain any reasonable solutions, we need to define those parameters that are less important or can safely be set to an experimentally secured value. The remaining parameters are those parameters that are crucial for the neuronal response properties but cannot be directly deduced from experimental data. In our reduced model we set the following as static: The optimized reduced morphology and the intracellular resistivities as well as all ion channel kinetics. The set of parameters that remain uncertain and need to be constrained are summarized in Tab. 3.2. 45 CHAPTER 3. THE CELL MODEL Table 3.2: F REE PARAMETERS IN THE R EDUCED M ODEL Eighteen parameters in our reduced model cannot be constrained directly by experiments but are crucial for the neuron’s active and passive response properties. We are uncertain about the leak reversal potential Epas but we assume that it is uniform in the whole model. The specific membrane resistance (rm ) and capacitance (cm ) are only adjusted directly in the axosomatic region (soma, basal, axon initial segment, hillock, axon). The spinefactor defines a factor describing the extension of the membrane area due to dendritic spines. rm and cm for the apical dendrite and the tuft are therefore indiapical, tuft axosomatic /spinefactor and rectly given through rm = rm apical, tuft axosomatic · spinefactor. We are also uncertain cm = rm about the ion channel densities in the soma as well as about the functions describing their gradients in the dendrites. The ion channel densities in the axon hillock and axon initial segment are also unknown and therefore parameterized. Moreover, we introduced two parameters describing the shift of the activation and inactivation curves for the transient sodium channel. The parameter vshift is applied globally as a general parameter for the kinetic model. The parameter vshift2 is only applied to the axon initial segment in order to introduce some additional local voltage shift as found experimentally (Colbert & Pan, 2002). 46 Parameter Unit global Epas axosomatic rm caxosomatic m mV spinefactor 1 gbarsoma Nat pS/µm2 gbarsoma Kfast pS/µm2 gbarsoma Kslow pS/µm2 gbarsoma Nap pS/µm2 gbarsoma Km pS/µm2 gbarsoma HCN pS/µm2 gbartuft HCN pS/µm2 decayNat pS/µm3 decayKfast µm decayKslow µm gbarhillock Nat pS/µm2 iseg gbarNat Ω · cm2 µF/cm2 pS/µm2 vshiftNat global mV iseg vshift2Nat mV CHAPTER 4 Results 4.1. Experimental Data The somatic voltage traces of layer 5 pyramidal neurons were recorded before and after pinching in response to the same current stimulus protocol. The protocol consisted of current steps with different amplitudes. The current was switched on for 1000 ms after a delay of 100 ms. The amplitudes were increases from -0.1 nA to 1 nA in 0.05 nA steps. When we compare the data before and after pinching several significant changes become obvious. Most notably, the input resistance increases upon pinching, leading to increased firing rates. The spike onset shape, spike height and the strength of the AHP do change after pinching as well (Fig. 4.1). 4.2. Fitting Strategy As outlined in chap. 3, we have created a model with a reduced morphology of 39 compartments and defined a biologically realistic channel composition based on experimental data. We have chosen a subset of 18 parameters that determine the neuron’s response properties but that are not fully constrained by experimental evidence. The experimental recordings before and after pinching allow us to observe the neuron’s behaviour under two different conditions, with and without an apical dendrite, which might give us enough information to constrain the free parameters indirectly. Once we will have constrained our model, we should be able to quantitatively explain the effects of pinching and use the model to explore other mechanisms. To optimize our model to the given data we apply the multi-objective optimization (MOO) strategy (sec. 2.4). To obtain any reasonable parameter combinations during the stochastic search, we introduced search boundaries for each of the 18 parameters. These boundaries are set relatively wide in order to put the least amount of knowledge into the values, but are in the biologically realistic range (Tab. 4.1). In the following optimization steps we will use a population size of N = 500, a population capacity of C = 1000 and the 47 CHAPTER 4. RESULTS Data Figure 4.1: E XPERIMENTAL R ECORDINGS BEFORE AND AFTER P INCHING The experimental recordings before (black lines) and after pinching (blue lines) are shown. Several significant changes are obvious. a) The AP onset and the repolarizing phase (offset) are overlayed before and after pinching. The voltage threshold for AP initiation is shifted to more hyperpolarized levels and the afterhyperpolarization (AHP) is enhanced after pinching. Also the spike height (measured from onset to the peak) increases. b) The first 600 ms of the spiketrains under both conditions are compared. The frequency increases after pinching. c) Four traces in response to subtreshold current injections before and after pinching are shown. From the stronger voltage responses it can be concluded that the input resistance is increased after pinching. We also see that the resting potential slightly, but significantly drops and that there is still a sag-response after pinching. d) The current amplitudes used for the stimulation. The same current protocol was used before and after pinching (green lines). Before Pinching After Pinching a 30 mV 0.2 ms 5 ms b 30 mV -70 c -70 5 mV d 0.4 nA 0 100 200 300 400 Time (ms) 500 600 evolution will be performed for 1000 generations. Crossover and mutation parameters are as explained (p. 27). On our cluster one complete evolution needed about two days and produced approximately 100 MB of data. To lead any optimization algorithm to good solutions we need to define the distance between the model response and the target data. However it is not straight forward to define a single reasonable distance function for spiking traces that is a reflection of the quality of the model responses (see Fig. 1.5). Our approach to define useful distance functions is explained in the following. 4.2.1. Checking Response Properties Pyramidal neurons operate in at least two different modes. Below a certain current injection amplitude they only show passive subthreshold responses. Then, if the current amplitude is increased pyramidal neurons start to elicit APs. For many parameter combinations the neuron model however will not spike for any current injection or it will spontaneously elicit APs under current injections where only a passive response is ex48 4.2. FITTING STRATEGY pected. Therefore we first need to check whether a given parameter set leads to a model that shows “good” spiking behaviour. An AP time is defined as the time after the voltage has crossed a threshold value of θ = -20 mV from below. A parameter set is defined as “good” when the following conditions hold: • The model neuron should not elicit an AP when the target data shows only passive responses. The model response should show APs when the target data does. If spiking is expected, further conditions must hold to check whether the given voltage trace is a regular spiketrain without bursts: • The model should elicit at least 6 spikes. • The spike width, defined as the time between the two points where the voltage crosses the threshold θ = -20 mV from below and from above, should not exceed 3 ms. • The absolute spike heights from the third to the next to last spike should not change by more than 20 %. • The voltage minimum between the third and the fourth spike compared to the voltage minimum between the next to last and the last spike should not change by more than 10 %. • There should not be any interspike interval below 15 ms. If one of these conditions is not fulfilled, the neuron’s response properties are considered as “bad”. Hence, the responsible model parameter set must be “punished” by associating it with large error values. This will tell the optimization algorithm to avoid this position in parameter space during the subsequent search. 4.2.2. Distance Functions If the conditions defining a “good” spiketrain are fulfilled, we can specify more detailed objectives that are elementary for quantitatively describing the spiking behaviour of pyramidal neurons. We have chosen four objectives that we consider as being good representations of the neuron’s response properties and that have a correlation with our set of free parameters. These objectives were used to calculate four error functions to define the distance between the response of the reduced model and the target data. All timedependent functions were interpolated such that the step size between two time points (dt) was always the same, the integrales described below are therefore sums over discrete time points. The objective-functions are illustrated in Fig. 4.2. The error functions are explained in the following: 49 CHAPTER 4. RESULTS 1. The passive subthreshold response traces should help use to constrain the passive membrane parameters in the neuron as well as the HCN-channel density. For each model (t) and passive trace i, the squared distance between the model response Vi,passive target the target data Vi,passive (t) is calculated only in the time window between 50 ms and 400 ms. The contribution of four non-spiking traces were summed into the passiveerror value: 4 ẽpassive = 0.5 · ∑ i=1 Z400 2 target model Vi,passive (t) −Vi,passive (t) dt . (4.1) 50 2. The voltage trace average of an AP (Vspike (t)) was determined by a high-resolution alignment (sec. 2.2). The first two spikes and the last spike from a trace were excluded from averaging to circumvent conflicts with adaptive currents and spike shape changes when the stimulating current is switched off. The time dependency for the mean-spike was shifted such that for Vspike (0) the first derivative dVspike/dt (0) was at its maximum. This occurs slightly before the actual AP peak. Doing this, the averaged APs obtained from the target data and the model response could be aligned at t = 0. We have found that choosing the peak of the first derivative as an alignment time was a better choice than the actual voltage peak of the AP. The shape of the peak was variable in different data sets and some of these peaks could not by reproduced with our model. The detailed spike onset contains information about the sodium channel distribution in the soma and axon. The squared distance between the averaged model and target AP onset was calculated from t = -0.5 ms to t = 0.1 ms. This value was combined with the squared distance between the first derivatives in the same time window to obtain the final onset-error value: Z0.1 2 target model Vspike (t) −Vspike (t) + ẽonset = 0.5 · (4.2) −0.5 0.01 · target dVspike /dt (t) − 2 /dt (t) dt . model dVspike (4.3) 3. The detailed spike repolarization (offset) should tell us about the potassium ion channel composition in the soma and the sodium and potassium channel gradients in the apical dendrite. Therefore we calculated the squared distance between the model and target APs in the time window t = 0.1 ms to t = 14 ms. We also included 50 4.2. FITTING STRATEGY the distance between the first derivatives and obtained the offset-error value: ẽoffset = 0.5 · Z14 2 target model Vspike (t) −Vspike (t) + (4.4) 0.1 0.01 · target dVspike /dt (t) − 2 /dt (t) dt . model dVspike (4.5) 4. Finally, the interspike intervals will constrain the parameters responsible for the adaptive currents and the overall excitability of the cell. The difference between the ith interspike interval of the model and target spiketrain was also calculated using the squared distance measure. All interspike interval distances were then summed and normalized by n leading to the isis-error. Here, n is the number of interspike intervals of the spiketrain with less interspike intervals: ẽisis = 2 0.5 n · ∑ ISIstarget (i) − ISIsmodel (i) . n i=0 (4.6) 4.2.3. Combining Intact and Pinching Data All four error functions were calculated for the target data and the model responses under two conditions. First in the intact neuron (ẽintact ) and then in the neuron with an occluded x pinch apical dendrite (ẽx ). These values were then combined into the four final error functions: pinch epassive = 2 · ẽintact passive + ẽpassive (4.7) eonset = ẽintact onset + ẽonset pinch (4.8) pinch (4.9) eoffset = ẽintact offset + ẽoffset eisis = pinch ẽintact isis + ẽisis . (4.10) The passive-error of the intact neuron were weigthed more to equalize the influence of the error before and after pinching. Pinching in the model neuron was simulated by increasing the intracellular resistivity and decreasing the diameter of the most proximal compartment pinch of the apical dendrite (ra = 1000000 Ω · cm, d pinch = 0.1 µm). We think that these four error functions are a good representation of the response properties of pyramidal neurons. They are determined from a combination of 8 passive responses, the detailed AP shape and precise interspike intervals from 2 spiketrains under different recording conditions in the same neuron. Thus, these error functions are putting heavy constraints on our reduced model. We also think that introducing these distance functions will help us to find a good trade-off solution for our reduced model approximating the experimental data, although it might not be possible to optimize all error functions 51 CHAPTER 4. RESULTS at the same time. During our studies we have seen that the nonlinear interactions between the different parameters are stronger than expected. For example the passive responses also depend on the sodium channel density and on the other hand the spike shape depends on the passive membrane parameters as well. It is therefore necessary to minimize all four error functions at the same time. 4.2.4. Selection of the Optimal Solution During the optimization we save each of the 1000 generations in order to make a final selection of a single optimal solution after evolution. Thus we get a solution matrix M(i, j) of the size 1000 × 500. A solution can be referred by S(i, j) and its corresponding error values by ex (i, j) where i is the generation number, j the index of the individual in generation i and x the name of the distance function. While during the search we do not put any weight on the four distance functions, we must introduce a weighting at the final step of selection. But how should such a weighting look like without being arbitrary? As we expect from our knowledge about the optimization algorithm the minimal distance value for each independent objective decays while the evolution proceeds and only rarely a single objective-optimal solution is lost due to a mutation (Fig. 4.4, 4.8). The minimal distances in the last generation can be used to introduce a reasonable weighting: wpassive = 1/min(epassive (1000, j)) (4.11) wonset = 1/min(eonset (1000, j)) (4.12) woffset = 1/min(eoffset (1000, j)) (4.13) wisis = 1/min(eisis (1000, j)) . (4.14) j j j j We can now normalize the distance functions by these weights. Thereby we do not only cancel the units, but also equalize the magnitudes of the error values. Summing the weighted distance functions leads to a total-error value for each generation and individual: etotal (i, j) = wpassive · epassive (i, j) + wonset · eonset (i, j) + woffset · eoffset (i, j) + wisis · eisis (i, j) . (4.15) (4.16) The fact that the ’best total-error value did not always decay’ can be explained by referring to the fact that the weighting is a posteriori – the minimization was not performed on this error function (Figs. 4.4, 4.8). To select a single optimal solution we therefore decided not to take the best individual from the last generation, but instead to choose that individual 52 4.3. FITTING RESULTS associated with the lowest total error value ever found during the evolution: (i0 , j0 ) = argmin (etotal (i, j)) (i, j) 0 0 optimal_solution = S(i , j ) . (4.17) (4.18) 4.3. Fitting Results 4.3.1. Surrogate Data Optimization In order to test the described fitting strategy we first analyzed the performance of the algorithm on data that were generated by the model itself (surrogate data). This model-tomodel fit guarantees that a perfect solution exists, namely the solution with exact the same parameters as the target parameters. Thus we circumvent the problem that the defined model might not be adequate for representing experimental data. Our choice of target parameters can be seen in Tab. 4.1. With these parameters the target data were created and the search algorithm was started to search for parameters that reproduce these data well. We performed the search twice under exact the same conditions, yet with another random initial population (Trial 1 and Trial 2). The two resulting best parameter sets can also be seen in the table. The corresponding optimal model responses before and after pinching can be compared with the target data in Fig. 4.3 and Fig. 4.5. For the first optimization trial we also plotted the evolution of the minimum of each of the four objective-distance functions together with the best totalerror value (Fig. 4.4). In order to visualize the improvement we have made using the MOO-strategy, we also plotted the best solution of the initial random population where no evolution has been performed yet (Fig. 4.2). Both optimization trials lead to very good model solutions that mimic the surrogate data very well and present good fits for all objectives. We also see that a significant improvement is made during the evolution from the best initial random solution to the final optimal one. 53 CHAPTER 4. RESULTS Parameter global Epas axosomatic rm caxosomatic m spinefactor gbarsoma Nat gbarsoma Kfast gbarsoma Kslow gbarsoma Nap gbarsoma Km gbarsoma HCN gbartuft HCN decayNat decayKfast decayKslow gbarhillock Nat iseg gbarNat global vshiftNat iseg vshift2Nat Target -78.0 12000.0 1.5 1.0 450.0 45.0 370.0 2.9 14.0 39.0 65.0 0.5 66.0 27.0 17000.0 16000.0 7.0 -5.0 Result 1 -76.0 19521.0 1.6 1.5 591.1 48.4 392.2 1.2 14.2 40.7 36.3 0.8 77.4 23.2 17555.2 19479.0 7.7 -6.2 Result 2 -74.0 15446.2 1.7 1.7 1043.8 56.5 339.4 1.0 12.3 31.6 43.0 1.9 78.5 26.3 13459.7 18820.3 4.2 -1.4 LS Bound -85.0 10000.0 0.6 0.8 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1.0 1.0 5000.0 5000.0 0.0 -15.0 US Bound -60.0 30000.0 3.0 3.0 1500.0 300.0 1000.0 5.0 15.0 50.0 150.0 2.0 100.0 100.0 20000.0 20000.0 10.0 0.0 Unit mV Ω · cm2 µF/cm2 1 pS/µm2 pS/µm2 pS/µm2 pS/µm2 pS/µm2 pS/µm2 pS/µm2 pS/µm3 µm µm pS/µm2 pS/µm2 mV mV Table 4.1: TARGET PARAMETERS AND B EST PARAMETER C OMBINATIONS AFTER THE S URROGATE DATA O PTIMIZATIONS The target parameters (Target) used for creating the target surrogate data as well as the optimal parameter sets found after two separate searches (Result 1, Result 2) are shown. The lower search bound (LS Bound) and the upper search bound (US Bound) defined the region in parameter space that was explored for good solutions during the optimization. The parameters for the initial random population were uniformly distributed in that region. The two optimal solutions are different from the target parameters and are different among each other. All reported values were rounded. 54 4.3. FITTING RESULTS After Pinching Before Pinching Target Data Model a Target Data Model 30 mV 30 mV 0.2 ms 0.2 ms 5 ms 5 ms b 30 mV -70 30 mV -70 c -70 5 mV d 0.4 nA -70 5 mV 0.4 nA 0 0 100 200 300 400 Time (ms) 500 600 100 200 300 400 Time (ms) 500 600 Figure 4.2: B EST S OLUTION OF THE I NITIAL R ANDOM P OPULATION BEFORE THE S UR ROGATE DATA O PTIMIZATION , T RIAL 1 All objectives used for the optimization are illustrated and shown for the target data and the model solution before pinching (black and red lines) and after pinching (blue and orange lines): a) The detailed shape of the AP onsets and offsets used to determine the objective onset and offset. b) Only the first 600 ms of the spiketrains resulting from a current injection of 0.4 nA are shown to better visualize the interspike intervals. The interspike intervals for the entire spiketrains were used to calculate the objective isis. c) The 4 passive subthreshold traces used to calculate the objective passive. d) The 5 different current amplitudes used for stimulation (-0.10 nA, -0.05 nA, 0.00 nA, 0.05 nA, 0.4 nA) (green lines). 55 CHAPTER 4. RESULTS After Pinching Before Pinching Target Data Model a Target Data Model 30 mV 30 mV 0.2 ms 0.2 ms 5 ms 5 ms b 30 mV -70 30 mV -70 c -70 5 mV d 0.4 nA -70 5 mV 0.4 nA 0 0 100 200 300 400 Time (ms) 500 600 100 200 300 400 Time (ms) 500 600 Figure 4.3: B EST S OLUTION AFTER THE S URROGATE DATA O PTIMIZATION , T RIAL 1 56 4.3. FITTING RESULTS 1 Passive Onset Offset ISIs Total Normalized Error 0.1 0.01 0.001 0.0001 0 100 200 300 400 500 600 700 800 900 1000 Generation Figure 4.4: E VOLUTION OF THE F OUR O BJECTIVE -D ISTANCE F UNCTIONS AND OF THE T OTAL -E RROR VALUE DURING THE S URROGATE DATA O PTIMIZATION , T RIAL 1 All values are shown on a logarithmic scale and were normalized by their maximal value and hence the relative improvement during the optimization can be seen. The distance describing the passive response error epassive (long-dashed line) was minimized by a factor of more than 1400 by our algorithm. The distance for the objective onset eonset (medium-dashed lines) could be optimized more than 100×. The distance for the objective offset eoffset (short-dashed line) as well as the error for objective isis eisis (long-dash-short-dashed line) were optimized by a factor of more than 2500. However the total-error value etotal (red line) could only be minimized about 50× during the optimization procedure. The optimal solution we have selected at the end of the evolution was found in generation 854 in this optimization trial. 57 CHAPTER 4. RESULTS After Pinching Before Pinching Target Data Model a Target Data Model 30 mV 30 mV 0.2 ms 0.2 ms 5 ms 5 ms b 30 mV -70 30 mV -70 c -70 5 mV d 0.4 nA -70 5 mV 0.4 nA 0 0 100 200 300 400 Time (ms) 500 600 100 200 300 400 Time (ms) 500 600 Figure 4.5: B EST S OLUTION AFTER THE S URROGATE DATA O PTIMIZATION , T RIAL 2 58 4.3. FITTING RESULTS 4.3.2. Experimental Data Optimization After the surrogate data optimization has shown that the fitting strategy is able to constrain our model to several separate active and passive target traces, we will now replace the surrogate data by real experimental recordings. We performed the search three times (Trial 1, Trial 2 and Trial 3). The three resulting optimal parameter sets can be seen in Tab. 4.2. The corresponding optimal model responses before and after pinching can be compared with the experimental target data in Figs. 4.7, 4.9 and 4.10. In order to show that the optimization algorithm leads to better solutions than the selection of the best solution of a random population we also plotted the evolution of the minimal error values during the optimization (Fig. 4.8). An illustration of the best individual from the initial population before any optimization has been performed is shown in Fig. 4.6. Parameter global Epas axosomatic rm caxosomatic m spinefactor gbarsoma Nat gbarsoma Kfast gbarsoma Kslow gbarsoma Nap gbarsoma Km gbarsoma HCN gbartuft HCN decayNat decayKfast decayKslow gbarhillock Nat iseg gbarNat global vshiftNat iseg vshift2Nat Result 1 -78.4 20494.5 2.0 1.2 780.6 47.7 396.8 1.3 11.8 32.0 53.0 1.4 82.7 10.7 7574.9 12727.9 8.2 -7.9 Result 2 -77.8 11609.7 1.8 0.8 380.8 50.0 400.5 2.8 15.0 23.1 60.2 0.5 76.0 1.6 12561.8 16152.3 8.5 -7.5 Result 3 -75.8 19975.8 2.1 1.1 796.7 56.3 392.7 0.8 12.3 31.0 31.9 1.4 52.9 63.5 11199.6 15932.1 5.9 -5.8 LS Bound -85.0 10000.0 0.6 0.8 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1.0 1.0 5000.0 5000.0 0.0 -15.0 US Bound -60.0 30000.0 3.0 3.0 1500.0 300.0 1000.0 5.0 15.0 50.0 150.0 2.0 100.0 100.0 20000.0 20000.0 10.0 0.0 Unit mV Ω · cm2 µF/cm2 1 pS/µm2 pS/µm2 pS/µm2 pS/µm2 pS/µm2 pS/µm2 pS/µm2 pS/µm3 µm µm pS/µm2 pS/µm2 mV mV Table 4.2: B EST PARAMETER C OMBINATIONS AFTER THE E XPERIMENTAL DATA O PTI MIZATIONS The optimal parameter sets found after three separate searches (Result 1, Result 2, Result 3) are shown. The lower search bound (LS Bound) and the upper search bound (US Bound) defined the region in parameter space that was explored for good solutions during the optimization. The parameters for the initial random population were uniformly distributed in that region. The three optimal solutions are different among each other. All reported values were rounded. The MOO-strategy was able to constrain our model to experimental data. The spike onset and offset shapes as well as the interspike intervals and the subtreshold responses are reproduced well for the intact and pinching condition for all three trials. Only the spike onset and the height of the AP after pinching are not reproduced satisfactorily. 59 CHAPTER 4. RESULTS The solutions found are trade-off solutions between the optima of the four objectives. We have seen that the model can perform better in each objective alone but that such a solution was obtained only at the cost of the other objectives (data not shown). We see that the best solution of the initial random population does not present good fits. Although the decay of the distance functions is less significant than in the surrogate data optimization (Fig. 4.4), the results show that the optimization algorithm works well and helps to constrain the parameters better than a random choice does. After Pinching Before Pinching Target Data Model a Target Data Model 30 mV 30 mV 0.2 ms 0.2 ms 5 ms 5 ms b 30 mV 30 mV -70 -70 c -70 5 mV d 0.4 nA -70 5 mV 0.4 nA 0 0 100 200 300 400 Time (ms) 500 600 100 200 300 400 Time (ms) 500 600 Figure 4.6: B EST S OLUTION OF THE I NITIAL R ANDOM P OPULATION BEFORE THE E X PERIMENTAL DATA O PTIMIZATION , T RIAL 1 60 4.3. FITTING RESULTS After Pinching Before Pinching Target Data Model a Target Data Model 30 mV 30 mV 0.2 ms 0.2 ms 5 ms 5 ms b 30 mV 30 mV -70 -70 c -70 5 mV d 0.4 nA -70 5 mV 0.4 nA 0 0 100 200 300 400 Time (ms) 500 600 100 200 300 400 Time (ms) 500 600 Figure 4.7: B EST S OLUTION AFTER THE E XPERIMENTAL DATA O PTIMIZATION , T RIAL 1 61 CHAPTER 4. RESULTS 1 Passive Onset Offset ISIs Total Normalized Error 0.1 0.01 0.001 0.0001 0 100 200 300 400 500 600 700 800 900 1000 Generation Figure 4.8: E VOLUTION OF THE F OUR O BJECTIVE -D ISTANCE F UNCTIONS AND OF THE T OTAL -E RROR VALUE DURING THE E XPERIMENTAL DATA O PTIMIZATION , T RIAL 1 All values are shown on a logarithmic scale and were normalized by their maximal value and hence the relative improvement during the optimization can be seen. The distance describing the passive response error epassive (long-dashed line) was minimized by a factor of more than 50 by our algorithm. The distance for the objective onset eonset (medium-dashed lines) could be optimized only about 3×. The distance for the objective offset eoffset (short-dashed line) as well as the error for the objective isis eisis (long-dash-short-dashed line) were optimized by a factor of approximately 20. The total-distance value etotal (red line) could be minimized about 25× during the optimization procedure. The optimal solution we have selected at the end of the evolution was found in generation 741 in this optimization trial. 62 4.3. FITTING RESULTS After Pinching Before Pinching Target Data Model a Target Data Model 30 mV 30 mV 0.2 ms 0.2 ms 5 ms 5 ms b 30 mV 30 mV -70 -70 c -70 5 mV d 0.4 nA -70 5 mV 0.4 nA 0 0 100 200 300 400 Time (ms) 500 600 100 200 300 400 Time (ms) 500 600 Figure 4.9: B EST S OLUTION AFTER THE E XPERIMENTAL DATA O PTIMIZATION , T RIAL 2 63 CHAPTER 4. RESULTS After Pinching Before Pinching Target Data Model a Target Data Model 30 mV 30 mV 0.2 ms 0.2 ms 5 ms 5 ms b 30 mV 30 mV -70 -70 c -70 5 mV d 0.4 nA -70 5 mV 0.4 nA 0 0 100 Figure 4.10: T RIAL 3 200 300 400 Time (ms) 500 600 100 200 300 400 Time (ms) 500 600 B EST S OLUTION AFTER THE E XPERIMENTAL DATA O PTIMIZATION , 64 4.3. FITTING RESULTS 4.3.3. Generalization for Other Input Currents Frequency (Hz) After having found several model solutions that represent the experimental data well in all objectives, we also need to test whether a selected model is able to reproduce experimental data that has not been used for the optimization algorithm, that is, whether the model generalizes. The optimization has been performed with 4 subthreshold currents and one suprathreshold current with an amplitude of 0.4 nA. In order to test whether our model generalizes we can change the amplitudes of the suprathreshold current injections and check whether we are able to predict the experimental spiking responses. We first tested how our model is able to predict the experimental mean firing frequency in response to current injections between 0 nA and 1.1 nA under intact and pinching conditions (Fig. 4.11). It can be seen that the model predicts the firing frequency of the experimental data well for many currents. Moreover the current thresholds for spike initiation under both conditions are predicted well by our model. Only for higher current injections above 0.8 nA the prediction becomes worse. 60 50 40 30 20 10 0 Data Before Pinching Data After Pinching Model Before Pinching Model After Pinching 0 0.2 0.4 0.6 0.8 Input Current (nA) 1 Figure 4.11: M ODEL P REDICTION OF F IRING F REQUENCY The experimentally measured and the predicted firing frequencies in response to current injections from 0 nA to 1.1 nA are shown for the two conditions before pinching (black and red dots) and after pinching (blue and orange dots). Furthermore we challenge the model not only to predict the firing frequencies, but also to reproduce the detailed spiketrains that were not used for the optimization. We changed the suprathreshold current injection from 0.4 nA to 0.6 nA and compared the predicted spiking responses with the experimentally measured ones (Fig. 4.12). The stronger current injection leads to a change in spike rate of ≈ 10 Hz for the experimental data and predicting these spiketrains must therefore be an accomplishment. However, it can be seen that the spiketrains and the detailed spike shapes before and after pinching are predicted well by our model. Remarkably our model also predicts the first experimentally measured interspike intervals, which are much shorter than the following ones. Thus, in conclusion, the model generalizes well. 65 CHAPTER 4. RESULTS After Pinching Before Pinching Target Data Model a Target Data Model 30 mV 30 mV 0.2 ms 0.2 ms 5 ms 5 ms b 30 mV 30 mV -70 -70 c -70 5 mV d 0.6 nA -70 5 mV 0.6 nA 0 0 100 200 300 400 Time (ms) 500 600 100 200 300 400 Time (ms) 500 600 Figure 4.12: M ODEL P REDICTION OF D ETAILED AP SHAPE AND S PIKETRAIN IN R E SPONSE TO A NOTHER I NPUT C URRENT We checked whether our optimized reduced model is able to predict the voltage trace of a spiketrain in response to a higher current injections (0.6 nA, (d)) that has not been used for the optimization. The objectives ((a),(b),(c)) can be compared for the model prediction and the experimental recordings before (black and red lines) and after pinching (blue and orange lines). 66 4.4. MODEL EVALUATION 4.4. Model Evaluation We have selected the best solution after the experimental data optimization, Trial 1 for all following analysis. This model showed very good fit results and good generalization for other input currents, and thus we will use this model to study mechanisms that would be hard or impossible to be explored experimentally. The model will be referred simply as “our model” in the following. 4.4.1. Resting Potential We have measured the resting potential in the dendritic tree as a function of the distance to the soma and observed a change of about 4 mV in our model (4.13a). HCN-channel are open at rest and therefore the resting potential in a specific location depends on the local HCN-channel density. But the resting potential could also be altered by sodium and potassium channels that are slightly open at rest as well. However larger densities will be needed in order that these channels will have a significant effect on the resting potential. As it is not completely clear which conductances lead the observed change of the resting potential we plotted the steady state conductance for each ion channel present in the apical dendritic tree as a function of distance to the soma (Fig. 4.13b,c,d,e). It can be seen that the effect of sodium channels is negligible, but that indeed potassium and HCN-channels have a significant influence on the resting potential. The potassium channel density decays quickly with distance to the soma. Thus the hyperpolarizing potassium conductance is only present in the somatic region and hence the distal dendrite is more depolarized than the soma at rest. On the other hand, we find a higher HCN-channel density in the apical tuft than in the soma. Therefore, these channels introduce a depolarizing conductance that is stronger in the tuft than in the soma and hence the distal depolarization is even more enhanced. 4.4.2. AP Backpropagation We were also interested in whether the automatically created reduced model would show AP backpropagation. Thus we recorded the voltage in different locations in the neuron when a single spike was observed in the soma in response to a short 20 ms current stimulation. We could see that the AP is generated in the axon initial segment and then travels antidromically through the soma into the dendrites (Fig. 4.14a). To further quantify the AP shape in different locations we calculated its absolute peak (Fig. 4.14b) and determined its half-width (Fig. 4.14c). The half-width was defined as the width at halfway from -60 mV to the AP peak. It can be seen that the AP peak decays slowly in the apical dendrite and quickly when it reaches the tuft (500 µm). The half-width increases linearily with distance to the soma. When it reaches the tuft the slope of this change is enhanced. 67 a Resting Potential (mV) CHAPTER 4. RESULTS -65 -66 -67 -68 -69 -70 0 c 800 700 600 500 400 300 200 100 0 gNat (fS/µm2) gh (fS/µm2) b 0.4 0.3 0.2 0.1 200 400 600 800 1000 Distance to Soma (µm) 0 200 400 600 800 1000 Distance to Soma (µm) 0 200 400 600 800 1000 Distance to Soma (µm) e gKslow (fS/µm2) 50 gKfast (fS/µm2) 0.5 0 0 d 200 400 600 800 1000 Distance to Soma (µm) 40 30 20 10 0 400 300 200 100 0 0 200 400 600 800 1000 Distance to Soma (µm) Figure 4.13: R ESTING P OTENTIAL AND THE I ONIC C ONDUCTANCES AS A F UNCTION OF D ISTANCE TO THE S OMA All conductance units are reported in fS/µm2 . a) The resting potential is more depolarized in the distal regions than in the proximity of the soma. b) The conductance of the HCN-channels is stronger in the tuft than in the soma. As we did not insert HCN-channels in the apical dendrite, the conductance is zero in the apical region. c) The negligible sodium conductance decays linearly with distance. d) The fast potassium channel introduces a hyperpolarizing conductance at rest mainly in the soma as the ion channel density decays exponentially with distance. e) The slow potassium channel density decays quickly with distance and thus the conductance at rest introduces a strong hyperpolarization only in the soma. 68 4.4. MODEL EVALUATION In order to understand this behaviour we determined which where the ionic mechanisms involved into AP backpropagation. We measured the conductances during an AP peak in different locations along the apical dendrite and tuft (Fig. 4.14c,d,e). We can see that the sodium conductance near the soma has large values, hence this conductance appears to be important for the local AP generation. However the sodium conductance drops distally which explains the decay of the voltage-peak. Both potassium conductances also decay, hence the repolarizing current is missing in the regions with less or no potassium conductance and therefore the AP becomes wider. However it appears that active backpropagation in our model only occurs until the AP reaches the tuft. From then on there are neither sodium nor potassium channels active and the voltage spreads passively only. 4.4.3. Currents Shaping the Somatic AP Waveform We stimulated the soma with a short current pulse of 20 ms to elicit a single AP and analyzed the influence of different currents on its shape. We recorded the sodium and potassium currents as well as the axial currents from the axon and from the apical dendrite. We did this analysis before and after pinching (Fig. 4.15). Axial currents were determined by the voltage difference between two adjacent compartments and the local effective resistance (Eqn. 1.26). If the apical dendrite is accessible a large amount of the axonal and somatic sodium current quickly escapes into the apical dendrite where it leads to a dendritic depolarization and BAPs (see above). Then the somatic potassium conductances set in and the repolarization of the AP is initiated. However due to BAPs the dendritic tree has undergone a massive depolarization. This current can now flow back into the soma and introduce a long lasting depolarizing current that reduces the effect of the local potassium channels and therefore leads to a reduced AHP. After pinching the AP height becomes slightly higher. The axonal and somatic sodium current cannot escape into the apical dendrite anymore and enhance the local onset depolarization. Next, if the dendritic tree is not available to deliver current to compensate the hyperpolarization produced by the somatic potassium conductances, the AHP becomes stronger. 69 CHAPTER 4. RESULTS a 35 0 -35 317 b 318 35 0 319 320 321 Time (ms) c 7 6 5 4 3 2 1 0 322 f 324 Iseg soma Apical (161 µm) Apical (295 µm) Apical (495 µm) Tuft (686 µm) 323 0 200 400 600 800 1000 Distance to Soma (µm) 4 3 2 1 0 0 200 400 600 800 1000 Distance to Soma (µm) 70 -35 -70 e 20 15 10 5 0 Half-Width (ms) 0 200 400 600 800 1000 Distance to Soma (µm) gKslow (pS/µm2) 0 200 400 600 800 1000 Distance to Soma (µm) gKfast (pS/µm2) -70 800 600 400 200 0 0 200 400 600 800 1000 Distance to Soma (µm) Peak Voltage (mV) Figure 4.14: A NALYSIS OF BAP S We analyzed the shape transformation and the underlying ionic conductances of BAPs in our model. a) APs are shown at 6 locations in the neuron: In the initial segment (Iseg, dark red line), in the soma (red line), and in the apical dendrite and tuft (light red lines). The absolute peak voltage of the AP decays with distance to the soma, while the half-width (width at halfway from -60 mV to the AP peak) increases. The initiation of the AP occurs in the initial segment. It can also be seen that the voltage threshold for spike initiation decays with distance to the spike initiation zone. b,c) Quantification of the absolute voltage peak and the half-width of the AP as a function of distance to the soma. d) The decaying sodium conductance at the AP peak is shown. e,f) The conductances of the fast and slow potassium conductance decay with the distance to the soma. d Voltage (mV) gNat (pS/µm2) 4.4. MODEL EVALUATION a Before Pinch 50 After Pinch 50 30 mV -60 30 mV -60 b 1 ms 1 ms 0 0 IaxHillock 5 nA 5 nA IaxApical INat IKfast IKslow Figure 4.15: C URRENTS S HAPING THE S OMATIC AP WAVEFORM BEFORE AND AFTER P INCHING a) The voltage traces of the APs before (red line) and after pinching (orange line) are shown. It can be seen that the height of the AP is slightly increased after pinching and that the AHP without an apical dendrite is strongly enhanced. b) The somatic currents during the APs before and after pinching are overlayed: Axial current from the axon (black line), axial current from the apical dendrite (black dashed line) as well as the ionic currents of sodium (blue line), and potassium (green and cyan lines) are shown before and after pinching. Negative currents are somatic inward currents. 71 CHAPTER 5 Discussion We have presented a systematic strategy to automatically construct a reduced compartmental model of a cortical layer 5 pyramidal neuron. First, we developed an approach to reduce a detailed pyramidal neuron morphology to a simpler one, and found that our simplification strategy preserves the neuronal passive response properties very well. Then we selected a set of ion channel models as well as their spatial distribution in our model and defined its set of 18 free parameters. These parameters were then fitted with a multiobjective optimization (MOO) strategy (sec. 2.4). To test the optimization procedure, we used target data that had been generated by the model itself. We found solutions that fit these surrogate data very well, but we did not find a unique parameter set. We then repeated the optimization but replaced the target data with experimental recordings from pyramidal neurons. The different optimal parameter sets we have obtained suggest general trends for and homeostatic adjustments of conductance densities in pyramidal neurons. We used the optimized model to investigate the conductances underlying the resting potential and BAPs as well the currents shaping the somatic AP. These results have helped to quantitatively explain the effects of pinching. 5.1. Neuronal Geometry 5.1.1. Geometry Reduction In order to simplify a detailed morphological reconstruction of a layer 5 pyramidal neuron by Stuart & Spruston (1998), we were not interested in an elaborate mathematical analysis of how a reduction could be performed theoretically (for example Lindsay et al., 2003), but we simply aimed to obtain approximate parameter values for the reduced model that preserve the passive response properties well. We considered the suggestion by Destexhe (2001) as suited for our purpose. His reduced model consisted of several cylinders each representing a subset of the dendritic geometry of the detailed model. The intracellular resistivities were then fitted by him 73 CHAPTER 5. DISCUSSION to optimize the neuron’s voltage attenuation and passive responses. In addition to maintaining these objectives we additionally aimed to optimize the somatic impedance and phase-shift curves. Therefore we did not only fit the intracellular resistivity values in the neuron, but at the same time the dendritic geometry (Tab. 3.1). Using the MOO-strategy we obtained good fits (Fig. 3.1). As done by Destexhe (2001) we challenged the method by injecting noisy current into the somata of both neuron models. To our astonishment we could see that the voltage responses in the soma and distal dendrite were almost indistinguishable (Fig. 3.2). Our simplification strategy is straightforward, intuitive and precise and we therefore consider it as more useful than those currently found in the literature. There is a large heterogeneity among neurons in the cortex and the size of the neuronal morphology we have chosen might be different from the size of the neuron our electrophysiological data came from. However, we assume that the general proportions between different sections are similar in all layer 5 pyramidal neurons in a rat with similar age. Then the overall size of the neuron will play a lesser role as channel densities can compensate for variations in size. It was relatively easy to obtain good solutions for the chosen free parameter set with only a few generations of the evolutionary optimization method. Unlike our evaluation of the fitting procedure during the optimization of the ionic conductances (sec. 4.3), in this first step of model creation we have not tested yet how the evolution of the solution quality looks like or whether different optimal solutions can be obtained with other optimization runs. Therefore further investigation of the principles underlying neuronal geometry reduction is needed to explore why these results are so convincing and so easy to obtain. Application to Experiments In this study a detailed model produced the target data for the optimization of our reduced model. But the strategy also transfers directly to the creation of a simplified geometry based on direct experimental recordings. Voltage recordings with high time resolution are currently only possible by patch clamp experiments. But those recordings were only rarely successful in three (for example Larkum et al., 2001) and are still not possible in more locations of the dendritic tree. However the steady state voltage distribution can be obtained from the entire dendritic geometry via voltage sensitive dyes (Loew, 1996) that have time constants in the order of ms. In combination with a detailed dendritic reconstruction these data would give a perfect target for our simplification strategy. It would be very interesting to check whether the dendritic specific membrane leak conductance and capacitance will need to be higher than in the soma in order to fit the data. This would give us an idea about the membrane extension due to dendritic spines and could therefore be used to estimate the spine density in the dendritic tree. 74 5.1. NEURONAL GEOMETRY 5.1.2. Passive Influence of the Basal Dendrite Our reduced model contains a single-compartment passive basal dendrite connected to the soma. It could be argued that this compartment is not necessary and might be collapsed into the soma. We have tested a model without a basal dendrite but with a larger soma and failed to obtain good results with biologically realistic parameters. To produce realistic spiking a certain match between the ionic conductances and the capacitance is needed in the soma. However optimizing these conductances led to combinations of the capacitance and the membrane-leak that produced time constants that did not match those of the experimentally measured passive responses. Furthermore the value for the specific capacitance was lower than standard values. Therefore we think that it is helpful to add an additional passive capacitor to the soma that has only minor effects on the spiking behaviour, but can introduce an additional capacitive current to adjust the time constants for the passive responses. 5.1.3. Axonal Geometry The geometry of our axon (Fig. 3.3) was adjusted by hand with only approximate geometrical estimates based on a detailed reconstruction (Zhu, 2000). We have observed that the diameters and length of the hillock and axon initial segment had an influence on the AP shape. We should further investigate the influence of the axon geometry on our fitting results. 5.1.4. Segmentation In order to obtain a sufficiently high resolution for the segmentation of a neuron model it was suggested to start with a certain number of compartments and to determine the spiking responses. Then the number of compartments should be increased and if the response properties do not change the model can be accepted whereas if a change occurs the resolution needs to be refined (Carnevale & Hines, 2006). As for our model we had started with an apical dendrite with 10 and a tuft with 5 compartments. We fitted the free model parameters to the experimental data and obtained good results and could also study BAPs. However we then tested whether the responses would remain the same if we increased the number of compartments in the optimized model. We observed small changes in the AHP. Thus we increased the number of compartments of the apical dendrite to 15 and of the tuft to 10 and repeated the optimization. However we have not tested whether in our new optimized model a further increment of compartments would alter the spiking responses again. We should do this in order to ensure that the compartmentalization is sufficient now. It should also be tested whether the number of compartments used in the other sections of our reduced model is sufficient. 75 CHAPTER 5. DISCUSSION 5.2. Ion Channel Composition 5.2.1. Choice of Ion Channel Models Many ion channel models are available in databases and are ready to be implemented in a neuron model.1 However the kinetics of the ion channels available were measured under different conditions and often for a specific modelling purpose. It was therefore not easy and one of the most time consuming parts of this study to make a selection of the right ion channels to be combined in our model. We have tested a variety of different sodium and potassium channel models, we also used different HCN-channels and calcium channels. It was especially interesting to see that realistic AP onsets could only be obtained with the model of the Nat-channel by Kole et al. (2006) although this model is based on kinetic measurements that were performed around 20 years ago (Hamill et al., 1991; Huguenard et al., 1988) while recent models of Nat-channels (for example Baranauskas & Martina, 2006; Gurkiewicz & Korngreen, 2007) failed. As for the repolarizing phase we tested several potassium channel combinations but only the choice of the Kfast-channel (Kole et al., 2006) and Kslow-channel (Korngreen & Sakmann, 2000) led to satisfying results. Furthermore this was the only potassium channel combination we were able to produce BAPs with. Differences in the results due to the other HCN-channel models tested were not significant. We also introduced calcium channels, calcium dynamics and calcium-dependent potassium channels. These conductances however did not significantly improve the fitting results, but led to bursting behaviour for larger current inputs which was not seen in the experimental recordings. Therefore we decided not to model calcium channels and related mechanisms although they are present in pyramidal neurons (Schiller et al., 1997). It was very useful to have a fitting algorithm available in order to automatically test whether the insertion or modification of an ion channel model improves or worsens the fitting results. In a subsequent study we might build a framework to easily quantify the performance of any ion channel combination in order to rank them. This would surely help many modelling studies to choose the best ion channel composition for their purpose. 5.2.2. Ion Channel Distribution We inserted ion channel models in the axon initial segment, hillock, soma, apical dendrite and tuft. The basal dendrite and the rest of the axon were left passive. 1 ModelDB: http://senselab.med.yale.edu/ModelDB 76 5.2. ION CHANNEL COMPOSITION Potassium Channels in the Axon Hillock and axon initial segment only contained Nat-channels. We did not insert potassium channels although they are present in this region and their influence on shaping the axonal AP was studied recently. The experimentally determined repolarization in the axon initial segment (Kole et al., 2007, Fig. 1A) is faster than in our model (Fig. 4.14a). We tested the influence of potassium channels in the axon but they did not significantly improve the fitting results. We think that somatic potassium channels can take over the work of the missing ion channels in the axon to optimize the somatic AP repolarization. Active Conductancs in the Basal Dendrites We did not insert sodium and potassium channels in the basal dendrites of our model although it is known that active propagation of APs occurs in these dendrites (Nevian et al., 2007; Polsky et al., 2004). It would indeed be very interesting to model AP propagation in basal dendrites, but due to their small diameters reliable experimental data are rare. We think that modelling sodium and potassium channels in the basal dendrites could lead to a similar mechanism like that of BAPs in the apical dendrite. Thus also the basal dendrites might act as a current source after the somatic AP and thus influence the repolarization. Apical Sodium and Potassium Gradients As for the ion channel density gradients in the apical dendrite and tuft we first tested constant densities but were not successful in obtaining good fitting results. Only with a linear decay of the Nat- and exponential decays of the Kfast- and Kslow-channel densities as suggested by Keren et al. (2009) we could obtain satisfying results. HCN-Channel Distribution Keren et al. (2009) have also described the HCN-channel density in the apical dendrite with a function that depends on distance from the soma and increases sigmoidally but this distribution did not lead to good fitting results for the passive neuron responses in our study. We explain this failure with our need to have a certain number of HCN-channels available also in the soma as we observe a sag-response in the experimental data after pinching as well (Fig. 4.1). With the described function the HCN-channel density directly increased in the apical dendrite and hence the total number of proximal HCN-channels was high. This produced a sag-response that was much stronger than the experimentally measured one before pinching. We therefore decided not to model HCN-channels in the apical region but only in the soma and tuft. It would be interesting to test another scenario with two sigmoidal density functions, one increasing towards the distal part of the basal dendrite and one increasing towards 77 CHAPTER 5. DISCUSSION the distal part of the apical dendrite. If such a distribution is reasonable the somatic density can be zero now and the proximal apical dendrite would contain less HCN-channels producing a good sag-response for the intact neuron. On the other hand the basal HCNchannels would lead to the right sag-response after pinching. That study however would require several additional free parameters and would need more elaborate distance functions but could exploit the pinching data for the study of HCN-channel distribution in the basal dendrite. Further Somatic Conductances We inserted the Nap- and Km-channel only in the soma. These channels were needed to adjust the interspike intervals for our model neuron as the experimental data shows slight spike frequency adaptation. The Km-channel activation reaches its steady state value only within seconds (Fig. 3.4) and is therefore suited for modulating the spike frequency. However it was not possible to get the right interspike intervals only with this channel alone, but we needed to include the Nap-channel as a counterpart to increase the overall excitability which was reduced by the Km-channel. AP Propagation in the Axon In the axon, only the hillock and the axon initial segment were able to initiate APs, the rest of the axon was passive and homogeneous. This is obviously not realistic as axons of pyramidal neurons contain nodes of Ranvier and segments of myelin. It would be important to check whether the existence of nodes of Ranvier with high sodium channel densities would alter the somatic response properties or the shape of the somatic AP. Nevertheless in order to use our reduced model neuron for the creation of realistic network connections via synapses the axon does not necessarily need to reproduce propagating APs as postsynaptic locations can also be activated with a certain time delay instead. 5.3. Fitting Results 5.3.1. Choosing the Free Parameters The right choice of the free parameters for a complex model is very difficult as it is not clear which of all possible model parameters need to be free in order to fit the target data. It is obvious that the more parameters will be free the easier it will be for the model to reproduce that data, but at the same time it will become harder for any algorithm to find the optimal point in parameter space. Therefore there exists a certain set of free parameters that offer a good trade-off between the ability to fit the target data and to be identifiable with the chosen fitting strategy. 78 5.3. FITTING RESULTS For a given list of free parameters we checked whether we would be able to fit our experimental data well enough and whether the optimized model would be able to generalize. Depending on the result we resized the list of free parameters and repeated the optimization. Thus the choice of the 18 free parameters we are using in this study (Tab. 3.2) is the result of a long trial and error process. We consider these parameters to be important for shaping the neuronal response properties. At the same time we are uncertain about their precise values as it is hard to constrain them via direct experiments. 5.3.2. Surrogate Data Optimization A first test for any optimization strategy is to challenge the algorithm with data that have been generated by the model itself (surrogate data) and to check whether it can find the original target parameters again (Druckmann et al., 2008). We have performed that test with our algorithm. For a target parameter set we chose one of our experimental data fitting results and slightly modified the parameter values (Tab. 4.1). With these parameters we generated the target data and ran the optimization two times. The surrogate data fitting results were very good (Figs. 4.3, 4.5) but not perfect and the underlying parameter combinations are different from the target values and are different from each other (Tab. 4.1). This shows that our fitting strategy is not able to find the global minimum in the total-error space. We think that this due to our choice of distance functions and the small population size. With a small population size the evolutionary algorithm will not be able to explore the whole error space and will get stuck in local minima. We think that our population size of N = 500 is too small, but due to our limited computational resources we could not test whether a larger population size would improve the results. Moreover the four distance functions appear not to emphasize the important distances between the final optimal solutions well enough. We have tried to include a fifth distance function optimizing the model with respect to the interspike interval trace, but could not test whether it significantly improved our fitting results as we would have needed an even larger population for this. We also observed that best initial solution (Fig. 4.2) is better than we had expected, which shows that with the free parameters chosen and the constraints defining a “good” model , it appears to be easy to obtain realistic spiking behaviour. It seems that the number of 18 free parameters is simply too large and that hence the resulting error space contains too many local minima. A way to reduce the number of free parameters per optimization run is to use a “parameter-peeling” approach by fitting the model to multiple traces one after the other where several ion channels were blocked (Keren et al., 2009; Roth & Bahl, 2009). Together with a MOO-strategy this should be an interesting approach to uniquely constrain an entire neuron model in one optimization run. 79 CHAPTER 5. DISCUSSION However if a combination with the pinching-strategy is desired this would require more distance functions and larger population sizes which surpasses our current computational resources and we are not sure whether this will be feasible experimentally. In any case it might not make sense to put effort in creating good strategies that are able to fit surrogate data perfectly and find the target parameter sets again, as it is not guaranteed that these methods directly transfer to experimental target data (Druckmann et al., 2008). Thus it might be a better way to further develop the algorithms that are already able to fit experimental data, but still do not find unique solutions for a given set of experimental recordings. 5.3.3. Experimental Data Optimization We were able to fit our model to experimental data. Only a very small number of studies were successful doing so (for example Druckmann et al., 2007; Keren et al., 2009; Prinz et al., 2003). Those studies however either used a single-compartment model or only focused on a specific region of the neuron. Furthermore none of these studies have presented how the optimized model generalizes to other current injections that were not used for the optimization. We instead have presented a method that is able to constrain an entire neuron model to several experimental data traces simultaneously (Figs. 4.7, 4.9, 4.10) and have shown that our optimized model predicts the spike frequency (Fig. 4.11) and even the detailed spiketrain for other current injections (Fig. 4.12). Optimal Parameters The final parameter results we have obtained after three evolutions are different, but all three solutions present good fits for the experimental data. Comparing the values in Tab. 4.2 we see interesting tendencies for some of the values. It would be fascinating to further investigate the suggested parameter correlations which could also reflect homeostatic adjustments in real neurons. The following hypotheses are however speculative and more optimization runs have to be performed to create a convincing statistical analysis: global Epas : It can be seen that the estimated global leak reversal potential is more negative than the resting potential ≈ -71 mV for all three results. This is expected as we know that HCN-channels push the membrane potential towards their reversal potential (Eh ≈ -47 mV) at rest. In order to remain at the resting potential a compensating leak conductance with a reversal potential below the resting potential is needed that balances the depolarizing h-current with a hyperpolarizing leak current. axosomatic : The value for the specific membrane resistance for the second result is rm estimated to be about 50 % smaller than for the other solutions and thus the leak conductance in this model is larger. This will reduce the excitability of the neuron which can be iseg hillock compensated by increasing the values of gbarsoma and gbarNat . Nap , gbarNat 80 5.3. FITTING RESULTS spinefactor: The estimates for this value were low in all trials and even below 1 for Trial 2. In other modelling studies spinefactors around 2 are used (Holmes, 1989). Our result might reflect a failure we have made in choosing the right proportions between the segments in our model, but it could also suggest lower spine densities in pyramidal neurons than previously assumed. gbarsoma Nat : This value appears to by correlated with decayNat . The sodium channel density in the apical dendrite needs to be adjusted properly to produce BAPs that lead to the right amount of dendritic source current to optimize the AP repolarization. If high soma values for gbarsoma Nat are found the decay needs to be fast (Trial 1, Trial 3), while if gbarNat is estimated to be low the decay must be slower (Trial 2). iseg gbarhillock Nat , gbarNat : These values are high for all solutions. This was not due to the search boundaries. We tried the optimization with the lower search boundary set to zero and also obtained high values. There has been a long lasting debate regarding the correct sodium channel density in the hillock and axon initial segment. Modelling studies suggested high sodium channel densities (Mainen et al., 1995) in order to initiate APs in the axon while experimental patch clamp recordings measured small sodium currents in axonal patches (Colbert & Pan, 2002). Only recently other experimental studies have agreed with high densities and suggested a ratio between the axon initial segment and somatic sodium channel density of around 50 (Kole et al., 2008). Consistent with these findings our results are 16, 42 and 20. iseg vshift2Nat : All three trials find that the voltage dependence of activation and inactivation of the sodium channel in the axon initial segment should be shifted to more negative potentials. Based on experimental recordings it was suggested that such a shift can explain the initiation of APs in the axon initial segment without assuming high sodium channel densities (Colbert & Pan, 2002). Our findings (-7.9 mV, -7.5 mV, -5.8 mV) are very close to the value found in the experiments (-7 mV) which gives more theoretical evidence that AP initiation in the axon initial segment is due to this axonal modification. 5.3.4. AP Initiation The optimal solution after all optimization trials showed AP initiation in the axon initial segment. If the AP was initiated in the soma the AP onset had a very distinct shape from the experimentally determined one and hence the underlying parameter set was associated with a large onset-error value and less considered during evolution. 5.3.5. Effects of Pinching With the help of our optimized model, we can now quantitatively explain the effects of pinching (Fig. 4.1): 81 CHAPTER 5. DISCUSSION Input Resistance The somatic input resistance increases, hence depolarization and firing rate increase in response to the same stimuli. This is due to the fact that the neuron loses a large amount of its accessible leaky surface area during pinching and therefore the same input current can charge the membrane more effectively. Resting Potential The resting potential is shifted to a more hyperpolarized level (≈ -1 mV). HCN-channels are open at rest and introduce a conductance that pushes the membrane potential towards their reversal potential (Eh ≈ -47 mV). As the density of these channels is highest in the distal regions of the apical dendritic tree, a large amount of HCN-channels is not accessible after pinching and therefore the resting potential is lowered. Voltage Threshold for AP Initiation The voltage threshold for spike initiation is shifted to a more hyperpolarized level (≈ -2 mV in the experiment, but only ≈ -0.5 mV in our model (not shown)). AP initiation occurs in the axon initial segment. In the intact neuron a significant amount of the activated axonal sodium current can escape into the dendrites and charge the dendritic capacitor, which acts as a current sink. Therefore more sodium channels need to be activated in order that the positive feedback of depolarization and sodium activation can initiate a spike. After pinching less sodium current can escape into the dendrites and less sodium channels are needed to initiate a sufficient local depolarization and hence the voltage threshold for AP initiation is reduced. We could not further minimize the threshold-difference between the experiment and our model. It was possible to obtain better results only if the sodium channel distribution was adjusted such that AP initiation occurred in the soma. We explain this discrepancy with the fact that the axon initial segment is too far away from the capacitive load of the apical dendrite. This might be solvable by modification of the axonal geometry and membrane parameters or it might help to modify the intracellular resistivity. We have tried several of these ideas, but none of them led to satisfying results. AHP Differences The AHP is increased by about 10 mV after pinching. The dendritic tree contains sodium and potassium channels as well. When a somatic AP occurs it can actively backpropagate into the dendritic tree (Fig. 4.14) which leads to a massive dendritic depolarization. Thus dendritic current flows back into the soma after a somatic spike. After pinching this depolarizing dendritic current is not available any more and hence the AHP is stronger (Fig. 4.15). 82 5.4. MODEL EVALUATION AP Height The total spike height (measured from the onset of an AP to its peak) is increased after pinching. This is expected as the depolarizing somatic currents cannot escape into the dendritic tree but instead lead to an enhanced somatic depolarization during an AP (Fig. 4.15). 5.4. Model Evaluation 5.4.1. Resting Potential We have studied the value of the resting potential in different locations along the apical dendrite (Fig. 4.13) and found that the resting potential is shifted in the distal part the neuron by about 4 mV. This was explained by the higher HCN-channel density in the tuft and the hyperpolarizing effect of the proximal potassium conductances. Experimental measurements (Stuart et al., 1997, Fig. 2B) and recent modelling studies (Keren et al., 2009) have also shown this modulation of the resting potential along the apical dendrite but suggested a higher change (≈ 9 mV). We think that this discrepancy is due to the missing HCN-channels along the apical dendrite in our model (see above). 5.4.2. Rapid AP Onset There has been a vigorous debate regarding the origin of the fast AP onset also called “the kink” in pyramidal neurons. This kink cannot be explained with standard HodgkinHuxley single-compartment models. Based on theoretical studies Naundorf et al. (2006) suggested a sodium channel cooperativity mechanism as a possible solution although there are no direct experimental findings supporting this idea. On the other hand it was argued that Hodgkin-Huxley models can explain this kink if the sudden influx of axonal lateral current is considered (McCormick et al., 2007). A combined experimental and modelling study was published (Yu et al., 2008) supporting the latter idea. However it was claimed that the amount of lateral current is insufficient to explain the rapid onset if the axonal morphology and membrane parameters are realistic.2 We have created a realistic somatic and axonal morphology (sec. 3.1). We use standard Hodgkin-Huxley models for the sodium current and our optimization algorithm has found densities and kinetics that are consistent with recent experimental findings (see above). The AP is initiated in the axon initial segment and travels antidromically through the soma into the dendrites (Fig. 4.14a). With these biologically realistic results we obtain very good fits for the experimentally determined AP kink (Fig. 4.7a). Wo do therefore 2 SfN-Conference 2008, Washington DC, USA: Baranauskas et al. “Why action potentials in the soma have sharp onset?” 83 CHAPTER 5. DISCUSSION support the lateral-current hypothesis and think that we have presented the first quantitative modelling study to support that idea. 5.4.3. AP Backpropagation Our reduced model shows BAPs after all three optimization trials, although we have not directly asked for this in the distance functions. We have analyzed BAPs only for Trial 1 in detail but the other solutions led to similar results. The peak amplitude of BAPs in our model decays sigmoidally with distance from the soma (Fig. 4.14b). This result is different from that of Keren et al. (2009, Fig. 10A) and does not match the experimental finding that the peak decays exponentially (Stuart et al., 1997, Fig. 2A). We find that the half-width of proximal BAPs increases linearly with distance (Fig. 4.14c). A linear increase with similar slope was also found experimentally (Stuart et al., 1997, Fig. 2C). The half-width slope in the tuft region of our model was higher but the experimental measurements are not precise enough in that region to judge this result. The underlying conductances during BAPs predicted by us (Fig. 4.14d,e,f) were similar to those predicted by Keren et al. (2009, Fig. 10B,C,D). We think that the mismatch for the peak amplitude is due to the missing HCN-channel conductance in the apical dendrite of our model. If we use an increasing HCN-channel density the leak conductance does also increase and therefore we expect the local sodium currents to have less depolarizing effects. As described above we have already suggested how apical HCN-channels could be introduced in our modelling study. 5.4.4. Currents Shaping the Somatic AP Waveform We have used our model to determine the main currents involved during a somatic AP (Fig. 4.15). We can see that the somatic AP onset is due to a sudden current influx from the axon that leads to the “kink” (see above). Then with a delay somatic sodium channels open and further enhance the depolarization. These two separate inward currents lead to the typical biphasic AP onset as seen in phase-plots of voltage traces of pyramidal neurons (This study, Fig. 1.5; Yu et al., 2008, Fig. 1C). The repolarization of the AP is shaped by a complex dialogue of sodium, potassium and lateral currents from the axon and the apical dendrite. It is interesting to see that with the described set of conductances we were able to reproduce the experimental repolarization well although we did not model further conductances. It was shown, for example, that the interplay of calcium channels and calcium-dependent potassium channels has strong effects on the AHP in CA1 pyramidal neurons (Golomb et al., 2007). Thus it seems that our ion channel composition can compensate for the missing conductances. 84 5.5. OUTLOOK 5.5. Outlook This thesis led to a model of a layer 5 pyramidal neuron that precisely reproduces experimental data, generalizes to other types of input, and that we have used to explain several biophysical mechanisms relevant to the function of pyramidal neurons. But significant differences compared to the experiment remain, for example in the resting potential and the shape of BAPs. One of the next steps will therefore be a modification of the HCNchannel distribution to get closer to the experimental findings. We will further improve the fitting strategy by increasing the number of individuals and by including additional distance functions. With these improvements we will then perform more optimization runs on the experimental data in order to create a rigorous statistical analysis of the homeostatic adjustments we have hypothesized. For our purposes we will need to acquire larger computational resources. We hope that this will finally lead to an accurate and efficient model of a layer 5 pyramidal neuron that will be beneficial for large-scale simulations of the cortex. 85 References Abbott, L. (1999). Lapicque’s introduction of the integrate-and-fire model neuron (1907). Brain Research Bulletin, 50(5-6), 303–304. Achard, P. & De Schutter, E. (2006). Complex parameter landscape for a complex neuron model. PLoS Comput Biol, 2(7), e94. Amini, B., Clark, J. W., & Canavier, C. C. (1999). Calcium dynamics underlying pacemaker-like and burst firing oscillations in midbrain dopaminergic neurons: a computational study. Journal of Neurophysiology, 82(5), 2249–61. Ariav, G., Polsky, A., & Schiller, J. (2003). Submillisecond precision of the input-output transformation function mediated by fast sodium dendritic spikes in basal dendrites of ca1 pyramidal neurons. Journal of Neuroscience, 23(21), 7750–7758. Baranauskas, G. & Martina, M. (2006). Sodium currents activate without a hodgkin-andhuxley-type delay in central mammalian neurons. J Neurosci, 26(2), 671–84. Barry, P. H. (1994). Jpcalc, a software package for calculating liquid junction potential corrections in patch-clamp, intracellular, epithelial and bilayer measurements and for correcting junction potential measurements. J Neurosci Methods, 51(1), 107–16. Barry, P. H. & Diamond, J. M. (1970). Junction potentials, electrode standard potentials, and other problems in interpreting electrical properties of membranes. Journal of Membrane Biology, 3(1), 93–122. Bekkers, J. M. & Häusser, M. (2007). Targeted dendrotomy reveals active and passive contributions of the dendritic tree to synaptic integration and neuronal output. Proc Natl Acad Sci USA, 104(27), 11447–52. Benda, J. & Herz, A. V. M. (2003). A universal model for spike-frequency adaptation. Neural Computation, 15(11), 2523–64. Berger, T., Larkum, M. E., & Lüscher, H. R. (2001). High i(h) channel density in the distal apical dendrite of layer v pyramidal cells increases bidirectional attenuation of epsps. Journal of Neurophysiology, 85(2), 855–68. 87 REFERENCES Bernstein, J. (1906). Untersuchungen zur Thermodynamik der bioelektrischen Ströme. Pflügers Archiv European Journal of Physiology, 112(9), 439–521. Borst, A. & Haag, J. (1996). The intrinsic electrophysiological characteristics of fly lobula plate tangential cells: I. passive membrane properties. J Comput Neurosci, 3(4), 313– 36. Bower, J. M. & Beeman, D. (1998). The book of GENESIS (2nd ed.): exploring realistic neural models with the GEneral NEural SImulation System. New York, NY, USA: Springer-Verlag New York, Inc. Bush, P. C. & Sejnowski, T. J. (1993). Reduced compartmental models of neocortical pyramidal cells. J Neurosci Methods, 46(2), 159–66. Carnevale, N. T. & Hines, M. L. (2006). The NEURON Book. New York, NY, USA: Cambridge University Press. Catterall, W. A. (1995). Structure and function of voltage-gated ion channels. Annu Rev Biochem, 64, 493–531. Cauller, L. J. & Connors, B. W. (1994). Synaptic physiology of horizontal afferents to layer i in slices of rat si neocortex. J Neurosci, 14(2), 751–62. Colbert, C. M. & Pan, E. (2002). Ion channel properties underlying axonal action potential initiation in pyramidal neurons. Nat Neurosci, 5(6), 533–8. Crivellato, E. & Ribatti, D. (2007). Soul, mind, brain: Greek philosophy and the birth of neuroscience. Brain Res Bull, 71(4), 327–36. Dalcin, L., Paz, R., Storti, M., & D’Elia, J. (2008). Mpi for python: Performance improvements and mpi-2 extensions. Journal of Parallel and Distributed Computing, 68(5), 655–662. De Schutter, E. & Bower, J. M. (1994). An active membrane model of the cerebellar purkinje cell. i. simulation of current clamps in slice. Journal of Neurophysiology, 71(1), 375–400. Deb, K. (2001). Multi-Objective Optimization Using Evolutionary Algorithms. Wiley. Deb, K., Pratap, A., Agarwal, S., & Meyarivan, T. (2002). A fast and elitist multiobjective genetic algorithm: Nsga-ii. Ieee T Evolut Comput, 6(2), 182–197. Destexhe, A. (2001). Simplified models of neocortical pyramidal cells preserving somatodendritic voltage attenuation. Neurocomputing, 38, 167–173. 88 REFERENCES Druckmann, S., Banitt, Y., Gidon, A., Schürmann, F., Markram, H., & Segev, I. (2007). A novel multiple objective optimization framework for constraining conductance-based neuron models by experimental data. Frontiers in neuroscience, 1(1), 7–18. Druckmann, S., Berger, T. K., Hill, S., Schürmann, F., Markram, H., & Segev, I. (2008). Evaluating automated parameter constraining procedures of neuron models by experimental and surrogate data. Biol Cybern, 99(4-5), 371–9. Eichner, H., Klug, T., & Borst, A. (2009). Neural simulations on multi-core architectures. Frontiers in Neuroinformatics, 3, 21. Fitzhugh, R. (1961). Impulses and physiological states in theoretical models of nerve membrane. Biophys J, 1(6), 445–466. Gentet, L., Stuart, G., & Clements, J. (2000). Direct measurement of specific membrane capacitance in neurons. Biophys J, 79(1), 314–320. Golding, N. L. & Spruston, N. (1998). Dendritic sodium spikes are variable triggers of axonal action potentials in hippocampal ca1 pyramidal neurons. Neuron, 21(5), 1189– 200. Golomb, D., Donner, K., Shacham, L., Shlosberg, D., Amitai, Y., & Hansel, D. (2007). Mechanisms of firing patterns in fast-spiking cortical interneurons. PLoS Comput Biol, 3(8), e156. Gurkiewicz, M. & Korngreen, A. (2007). A numerical approach to ion channel modelling using whole-cell voltage-clamp recordings and a genetic algorithm. PLoS Comput Biol, 3(8), e169. Hamill, O. P., Huguenard, J. R., & Prince, D. A. (1991). Patch-clamp studies of voltagegated currents in identified neurons of the rat cerebral cortex. Cereb Cortex, 1(1), 48–61. Hines, M. L., Davison, A. P., & Muller, E. (2009). Neuron and python. Frontiers in Neuroinformatics, 3, 1. Hines, M. L., Eichner, H., & Schürmann, F. (2008). Neuron splitting in compute-bound parallel network simulations enables runtime scaling with twice as many processors. J Comput Neurosci, 25(1), 203–210. Hodgkin, A. L. & Huxley, A. F. (1952a). Currents carried by sodium and potassium ions through the membrane of the giant axon of loligo. The Journal of Physiology, 116(4), 449–72. 89 REFERENCES Hodgkin, A. L. & Huxley, A. F. (1952b). A quantitative description of membrane current and its application to conduction and excitation in nerve. The Journal of Physiology, 117(4), 500–44. Holmes, W. R. (1989). The role of dendritic diameters in maximizing the effectiveness of synaptic inputs. Brain Res, 478(1), 127–37. Huguenard, J. R., Hamill, O. P., & Prince, D. A. (1988). Developmental changes in na+ conductances in rat neocortical neurons: appearance of a slowly inactivating component. Journal of Neurophysiology, 59(3), 778–95. Izhikevich, E. M. & Edelman, G. M. (2008). Large-scale model of mammalian thalamocortical systems. Proc Natl Acad Sci USA, 105(9), 3593–8. Johnston, D., Magee, J., Colbert, C., & Christie, B. (1996). Active properties of neuronal dendrites. Annu. Rev. Neurosci., 19, 165–186. Katz, Y., Menon, V., Nicholson, D. A., Geinisman, Y., Kath, W. L., & Spruston, N. (2009). Synapse distribution suggests a two-stage model of dendritic integration in ca1 pyramidal neurons. Neuron, 63(2), 171–7. Keren, N., Bar-Yehuda, D., & Korngreen, A. (2009). Experimentally guided modelling of dendritic excitability in rat neocortical pyramidal neurones. The Journal of Physiology, 587(Pt 7), 1413–37. Keren, N., Peled, N., & Korngreen, A. (2005). Constraining compartmental models using multiple voltage recordings and genetic algorithms. Journal of Neurophysiology, 94(6), 3730–42. Koch, C. (2004). Biophysics of Computation: Information Processing in Single Neurons (Computational Neuroscience). Oxford University Press, USA. Kole, M. & Stuart, G. (2008). Is action potential threshold lowest in the axon? Neurosci. Nat Kole, M. H. P., Hallermann, S., & Stuart, G. J. (2006). Single ih channels in pyramidal neuron dendrites: properties, distribution, and impact on action potential output. J Neurosci, 26(6), 1677–87. Kole, M. H. P., Ilschner, S. U., Kampa, B. M., Williams, S. R., Ruben, P. C., & Stuart, G. J. (2008). Action potential generation requires a high sodium channel density in the axon initial segment. Nat Neurosci, 11(2), 178–86. Kole, M. H. P., Letzkus, J. J., & Stuart, G. J. (2007). Axon initial segment kv1 channels control axonal action potential waveform and synaptic efficacy. Neuron, 55(4), 633–47. 90 REFERENCES Korngreen, A. & Sakmann, B. (2000). Voltage-gated k+ channels in layer 5 neocortical pyramidal neurones from young rats: subtypes and gradients. The Journal of Physiology, 525 Pt 3, 621–39. Lapicque, L. (2007). Quantitative investigations of electrical nerve excitation treated as polarization. 1907. Biol Cybern, 97(5-6), 341–9. Larkum, M. E., Senn, W., & Lüscher, H.-R. (2004). Top-down dendritic input increases the gain of layer 5 pyramidal neurons. Cereb Cortex, 14(10), 1059–70. Larkum, M. E., Zhu, J. J., & Sakmann, B. (2001). Dendritic mechanisms underlying the coupling of the dendritic with the axonal action potential initiation zone of adult rat layer 5 pyramidal neurons. The Journal of Physiology, 533(Pt 2), 447–66. LeMasson, G. & Maex, R. (2001). Introduction to equation solving and parameter fitting. In E. De Schutter (Ed.), Computational Neuroscience: Realistic Modeling for Experimentalists (pp. 1–21). CRC Press. Lindsay, K. A., Rosenberg, J. R., & Tucker, G. (2003). Analytical and numerical construction of equivalent cables. Mathematical biosciences, 184(2), 137–64. Loew, L. M. (1996). Potentiometric dyes: Imaging electrical activity of cell membranes. Pure Appl Chem, 68(7), 1405–1409. London, M. & Hausser, M. (2005). Dendritic computation. Annu. Rev. Neurosci., 28(1), 503–532. Losavio, B. E., Liang, Y., Santamaría-Pang, A., Kakadiaris, I. A., Colbert, C. M., & Saggau, P. (2008). Live neuron morphology automatically reconstructed from multiphoton and confocal imaging data. Journal of Neurophysiology, 100(4), 2422–9. Mainen, Z. F., Joerges, J., Huguenard, J. R., & Sejnowski, T. J. (1995). A model of spike initiation in neocortical pyramidal neurons. Neuron, 15(6), 1427–39. Markram, H. (2006). The Blue Brain Project. Nat Rev Neurosci, 7(2), 153–60. McCormick, D. A., Shu, Y., & Yu, Y. (2007). Neurophysiology: Hodgkin and huxley model–still standing? Nature, 445(7123), E1–2; discussion E2–3. Mel, B. W., Ruderman, D. L., & Archie, K. A. (1998). Translation-invariant orientation tuning in visual "complex" cells could derive from intradendritic computations. J Neurosci, 18(11), 4325–34. Morris, C. & Lecar, H. (1981). Voltage oscillations in the barnacle giant muscle fiber. Biophys J, 35(1), 193–213. 91 REFERENCES Nagumo, J., Arimoto, S., & Yoshizawa, S. (1962). An active pulse transmission line simulating nerve axon. Proceedings of the IRE, 50(10), 2061–2070. Naundorf, B., Wolf, F., & Volgushev, M. (2006). Unique features of action potential initiation in cortical neurons. Nature, 440(7087), 1060–3. Nevian, T., Larkum, M. E., Polsky, A., & Schiller, J. (2007). Properties of basal dendrites of layer 5 pyramidal neurons: a direct patch-clamp recording study. Nat Neurosci, 10(2), 206–214. Pearce, J. M. (2001). Emil heinrich du bois-reymond (1818-96). J Neurol Neurosurg Psychiatr, 71(5), 620. Piccolino, M. (1997). Luigi galvani and animal electricity: two centuries after the foundation of electrophysiology. Trends Neurosci, 20(10), 443–8. Pinsky, P. F. & Rinzel, J. (1994). Intrinsic and network rhythmogenesis in a reduced traub model for ca3 neurons. J Comput Neurosci, 1(1-2), 39–60. Poirazi, P., Brannon, T., & Mel, B. W. (2003). Pyramidal neuron as two-layer neural network. Neuron, 37(6), 989–99. Polsky, A., Mel, B. W., & Schiller, J. (2004). Computational subunits in thin dendrites of pyramidal cells. Nat Neurosci, 7(6), 621–627. Prinz, A. A., Billimoria, C. P., & Marder, E. (2003). Alternative to hand-tuning conductance-based models: construction and analysis of databases of model neurons. Journal of Neurophysiology, 90(6), 3998–4015. Prinz, A. A., Bucher, D., & Marder, E. (2004). Similar network activity from disparate circuit parameters. Nat Neurosci, 7(12), 1345–52. Rall, W. (1962). Theory of physiological properties of dendrites. Ann N Y Acad Sci, 96, 1071–92. Ray, S. & Bhalla, U. S. (2008). Pymoose: Interoperable scripting in python for moose. Frontiers in Neuroinformatics, 2, 6. Roth, A. & Bahl, A. (2009). Divide et impera: optimizing compartmental models of neurons step by step. The Journal of Physiology, 587(Pt 7), 1369–70. Sakmann, B. & Neher, E. (1984). Patch clamp techniques for studying ionic channels in excitable membranes. Annu. Rev. Physiol., 46, 455–72. 92 REFERENCES Schiller, J., Schiller, Y., Stuart, G., & Sakmann, B. (1997). Calcium action potentials restricted to distal apical dendrites of rat neocortical pyramidal neurons. The Journal of Physiology. Segev, I. (1994). The Theoretical Foundations of Dendritic Function: The Collected Papers of Wilfrid Rall with Commentaries (Computational Neuroscience). The MIT Press. Segev, I. & London, M. (2000). Untangling dendrites with quantitative models. Science, 290(5492), 744–50. Spruston, N. (2008). Pyramidal neurons: dendritic structure and synaptic integration. Nat Rev Neurosci, 9(3), 206–21. Stuart, G., Schiller, J., & Sakmann, B. (1997). Action potential initiation and propagation in rat neocortical pyramidal neurons. The Journal of Physiology, 505 ( Pt 3), 617–32. Stuart, G. & Spruston, N. (1998). Determinants of voltage attenuation in neocortical pyramidal neuron dendrites. J Neurosci, 18(10), 3501–10. Stuart, G. J. & Sakmann, B. (1994). Active propagation of somatic action potentials into neocortical pyramidal cell dendrites. Nature, 367(6458), 69–72. Traub, R. D., Buhl, E. H., Gloveli, T., & Whittington, M. A. (2003). Fast rhythmic bursting can be induced in layer 2/3 cortical neurons by enhancing persistent na+ conductance or by blocking bk channels. Journal of Neurophysiology, 89(2), 909–21. Traub, R. D., Contreras, D., Cunningham, M. O., Murray, H., LeBeau, F. E. N., Roopun, A., Bibbig, A., Wilent, W. B., Higley, M. J., & Whittington, M. A. (2005). Singlecolumn thalamocortical network model exhibiting gamma oscillations, sleep spindles, and epileptogenic bursts. Journal of Neurophysiology, 93(4), 2194–232. Van Geit, W., De Schutter, E., & Achard, P. (2008). Automated neuron model optimization techniques: a review. Biol Cybern, 99(4-5), 241–51. Vanier, M. C. & Bower, J. M. (1999). A comparative survey of automated parametersearch methods for compartmental neural models. J Comput Neurosci, 7(2), 149–71. Weaver, C. M. & Wearne, S. L. (2006). The role of action potential shape and parameter constraints in optimization of compartment models. Neurocomputing, 69(10-12), 1053–1057. Weaver, C. M. & Wearne, S. L. (2008). Neuronal firing sensitivity to morphologic and active membrane parameters. PLoS Comput Biol, 4(1), e11. 93 REFERENCES Wheeler, B. C. & Smith, S. R. (1988). High-resolution alignment of action potential waveforms using cubic spline interpolation. Journal of biomedical engineering, 10(1), 47–53. Winograd, M., Destexhe, A., & Sanchez-Vives, M. V. (2008). Hyperpolarizationactivated graded persistent activity in the prefrontal cortex. Proc Natl Acad Sci USA, 105(20), 7298–303. Yu, Y., Shu, Y., & McCormick, D. A. (2008). Cortical action potential backpropagation explains spike threshold variability and rapid-onset kinetics. J Neurosci, 28(29), 7260– 72. Zhu, J. J. (2000). Maturation of layer 5 neocortical pyramidal neurons: amplifying salient layer 1 and layer 4 inputs by ca2+ action potentials in adult rat tuft dendrites. The Journal of Physiology, 526 Pt 3, 571–87. 94 Eigenständigkeitserklärung Hiermit versichere ich, dass ich die vorliegende Diplomarbeit selbständig verfasst und keine anderen als die angegebenen Quellen und Hilfsmittel verwendet habe. München, den . . . . . . . . . . . . Unterschrift: . . . . . . . . . . . . . . . . . .