Shock-induced termination of cardiac arrhythmias

Similar documents
Shock-induced termination of cardiac arrhythmias

Omar Sami. Muhammad Abid. Muhammad khatatbeh

The Electrocardiogram

Cardiac Cycle. Each heartbeat is called a cardiac cycle. First the two atria contract at the same time.

PART I. Disorders of the Heart Rhythm: Basic Principles

Systems Biology Across Scales: A Personal View XXVII. Waves in Biology: Cardiac Arrhythmia. Sitabhra Sinha IMSc Chennai

Full file at

Modeling Chaos in the Heart

ECG. Prepared by: Dr.Fatima Daoud Reference: Guyton and Hall Textbook of Medical Physiology,12 th edition Chapters: 11,12,13

BIPN100 F15 Human Physiology I (Kristan) Problem set #5 p. 1

Interpreting Electrocardiograms (ECG) Physiology Name: Per:

Division Ave. High School AP Biology. cell body. signal direction

Comparison of Different ECG Signals on MATLAB

The Nervous System AP Biology

Where are the normal pacemaker and the backup pacemakers of the heart located?

Cardiac muscle is different from other types of muscle in that cardiac muscle

Introduction. Circulation

Lab #3: Electrocardiogram (ECG / EKG)

Nervous System. Nervous system cells. Transmission of a signal 2/27/2015. Neuron

The Normal Electrocardiogram

Nerve. (2) Duration of the stimulus A certain period can give response. The Strength - Duration Curve

Arrhythmias. 1. beat too slowly (sinus bradycardia). Like in heart block

A Dynamic model of Pulmonary Vein Electrophysiology. Harry Green 2 nd year Ph.D. student University of Exeter

Ameen Alsaras. Ameen Alsaras. Mohd.Khatatbeh

The Function of an ECG in Diagnosing Heart Conditions. A useful guide to the function of the heart s electrical system for patients receiving an ECG

Neurophysiology scripts. Slide 2

Anesthesia Assistants Review Course

Lab 2. The Intrinsic Cardiac Conduction System. 1/23/2016 MDufilho 1

Electrocardiography Biomedical Engineering Kaj-Åge Henneberg

Cardiovascular system

Lab Activity 24 EKG. Portland Community College BI 232

THE CARDIOVASCULAR SYSTEM. Heart 2

Physiology sheet #2. The heart composed of 3 layers that line its lumen and cover it from out side, these layers are :

DEFIBRILLATORS ATRIAL AND VENTRICULAR FIBRILLATION

Chapter 12: Cardiovascular Physiology System Overview

EKG Abnormalities. Adapted from:

Rhythmical Excitation of the Heart

WHAT S THAT RHYTHM I AM HEARING? GUIDE TO AUSCULTATION OF ARRHYTHMIAS IN HORSES

Chapter 3 Biological measurement 3.1 Nerve conduction

DEFIBRILLATORS. Prof. Yasser Mostafa Kadah

Department of medical physiology 7 th week and 8 th week

Chapter 26. Media Directory. Dysrhythmias. Diagnosis/Treatment of Dysrhythmias. Frequency in Population Difficult to Predict

Introduction to Neurobiology

FIBER TYPES - oxidative metabolism is the main form here - ATPase activity is relatively low

Management strategies for atrial fibrillation Thursday, 20 October :27

V. TACHYCARDIAS Rapid rhythm abnormalities

Arrhythmias. Simple-dysfunction cause abnormalities in impulse formation and conduction in the myocardium.

UNDERSTANDING YOUR ECG: A REVIEW

Engineering involved in Cardiac Arrest Management Sahar Sorek

Principles and Applications of Electrical Circuits and Signal Theories in EP

UNDERSTANDING ELECTROPHYSIOLOGY STUDIES

آالء العجرمي أسامة الخضر. Faisal Muhammad

Arrhythmias. Pulmonary Artery

Cardiac Telemetry Self Study: Part One Cardiovascular Review 2017 THINGS TO REMEMBER

Step by step approach to EKG rhythm interpretation:

Cardiac physiology. b. myocardium -- cardiac muscle and fibrous skeleton of heart

Quantitative Electrophysiology

CARDIOVASCULAR SYSTEM

CARDIOCOMP Electrocardiography Primer

ECG Interpretation Cat Williams, DVM DACVIM (Cardiology)

Newer pacemakers also can monitor your blood temperature, breathing, and other factors and adjust your heart rate to changes in your activity.

Cardiac Properties MCQ

Objectives of the Heart

OBJECTIVE. 1. Define defibrillation. 2. Describe Need and history of defibrillation. 3. Describe the principle and mechanism of defibrillation.

ANATOMY AND PHYSIOLOGY OF NEURONS. AP Biology Chapter 48

EKG Sensor Product Number: ENEKG189

CASE 10. What would the ST segment of this ECG look like? On which leads would you see this ST segment change? What does the T wave represent?

Thursday, January 22, Nerve impulse

Medical Cyber-Physical Systems

Collin County Community College. ! BIOL Anatomy & Physiology! WEEK 5. The Heart

Understanding the 12-lead ECG, part II

Chapter 20 (2) The Heart

Axon Nerve impulse. Axoplasm Receptor. Axomembrane Stimuli. Schwann cell Effector. Myelin Cell body

Catheter Ablation. Patient Education

3/26/15 HTEC 91. EKG Sign-in Book. The Cardiac Cycle. Parts of the ECG. Waves. Waves. Review of protocol Review of placement of chest leads (V1, V2)

A Cellular Automata Model For Dynamics and Control of Cardiac Arrhythmias DANNY GALLENBERGER (CARROLL UNIVERSITY) XIAOPENG ZHAO (UTK) KWAI WONG (UTK)

Marah karablieh. Osama khader. Muhammad khatatbeh. 0 P a g e

Sample. Analyzing the Heart with EKG. Computer

Chapter 16: Arrhythmias and Conduction Disturbances

Paramedic Rounds. Tachyarrhythmia's. Sean Sutton Dallas Wood

Cardiac arrhythmias. Janusz Witowski. Department of Pathophysiology Poznan University of Medical Sciences. J. Witowski

Chapter 44 Neurons and Nervous Systems

Electrocardiography I Laboratory

Farah Khreisat. Raghad Abu Jebbeh. Faisal Mohammad. 1 P a g e

Enhanced Cellular Automation Model For Detecting Heart Fibrillation

a lecture series by SWESEMJR

Chapter 45 Neurons and Nervous Systems

EE 791 Lecture 2 Jan 19, 2015

This presentation will deal with the basics of ECG description as well as the physiological basics of

Cellular Bioelectricity

Branches of the Nervous System

Outline. Electrical Activity of the Human Heart. What is the Heart? The Heart as a Pump. Anatomy of the Heart. The Hard Work

Paroxysmal Supraventricular Tachycardia PSVT.

STRUCTURAL ELEMENTS OF THE NERVOUS SYSTEM

current, and acting like

Analyzing the Heart with EKG

Local Anesthetics. Xiaoping Du Room E417 MSB Department of Pharmacology Phone (312) ;

Circulatory system of mammals

Transcription:

Shock-induced termination of cardiac arrhythmias Group members: Baltazar Chavez-Diaz, Chen Jiang, Sarah Schwenck, Weide Wang, and Jinglei Zhang Cardiac arrhythmias, also known as irregular heartbeat, occur when the electrical activity of the heart is irregular, disrupting the flow of blood to the heart (1). Arrhythmias can result in either the heart beating too quickly as in ventricular tachycardia, or too slowly, as in bradycardia (2). While many arrhythmias are not life-threatening, they are still a leading cause of morbidity and mortality. In fact, over 300,000 individuals die suddenly yearly in the United States, and in most cases it is thought that arrhythmias are the cause (2). Furthermore, the Center for Disease Control estimated that 2.66 million people would have atrial fibrillation in 2010, indicating that they are a world-wide problem (2). Now, the pumping of the heart is controlled through electrical impulses. Specialized cells within the right atrium known as the sinoatrial node fire spontaneously about 70 times per minute (2). These firings lead to the coordinated contraction of the atria and then, after a slight delay, the ventricles. As a result, in order to treat cardiac arrhythmias, the irregular electrical activity must be removed and the regular activity recovered. Defibrillation is commonly used to terminate cardiac arrhythmias such as ventricular tachycardia (3). Defibrillation consists of delivering a controlled amount of electrical energy to the heart using a device called a defibrillator. This suppresses the chaotic cardiac action potentials caused by the arrhythmia and allows the normal rhythm to be reestablished (3). There are three main protocols commonly used in commercial defibrillators, which are based on monophasic and both symmetric and asymmetric biphasic shocks. The monophasic protocol consists of only one phase while the biphasic consists two phases, where the polarity of the electrodes is switched during the second phase. Biphasic shocks can be either symmetrical or asymmetrical, and the asymmetrical shocks can have either a shorter first or second phase. Biphasic shocks are thought to be more efficient than monophasic shocks based upon empirical evidence found by Ideker s group in the late 1980s (3). However, it is not yet fully understood why biphasic shocks are more efficient. As a result, in our project, we will quantitatively compare the efficiency of the three different protocols. Our model uses the simplest geometry that can maintain a propagating action potential- a one-dimensional ring of cardiac tissue (see Figure 1) (3). Under normal circumstances, an action potential is initialized and travels around the ring. In our case, the shock will be modeled as the application of currents through electrodes on either side of the ring. These electrodes can either inject or subtract electrical charges during the shocks. A defibrillatory shock will be defined as successful if all wave propagation is removed within one second of the end of the shock (3). If any wave activity is still present, the shock will be characterized as unsuccessful.

Figure 1. Schematic of the ring of cardiac tissue. The two areas represent the locations of the two electrodes. In order to fully understand our model, we now need to switch from a more macroscopic standpoint to a microscopic view. By expanding upon the Beeler-Reuter equations, which describe the electrical activity of cardiac myocytes, the membrane potential can by calculated by solving : ( ) where V m is the membrane potential (3,4). This equation becomes important when considering how the membrane potential changes when the shock of defibrillation is applied, as this allows for an understanding of action potential waves which can determine the success or failure of defibrillation. Now, membrane potential refers to the difference in electric potential between the inside and the outside of cells. In humans, there is a much higher proportion of sodium (Na + ) outside the cell than in and much more potassium (K + ) inside the cell than out (5). As a result, sodium moves into the cell and potassium out via ion channels according to their concentration gradients. In addition, membrane potentials are usually held stable at around -70 to -80 millivolts (5). However, when this potential rises past a certain threshold, an action potential can occur. Action potentials occur in several types of cells, including neurons, muscle cells and endocrine cells. During action potentials, the membrane potential of the cells changes as the cell is first depolarized by the opening of voltage-gated Na + channels and then repolarized by the opening of voltage-gated K + channels (see Figure 2) (5). Following an action potential, there is a refractory period where the potential falls below the resting potential. This prevents the action potential from travelling the opposite direction and continues the flow of information through the cells. Figure 2. A typical action potential showing first the depolarization and then repolarization stages.

Referring back to our equation, this helps explain the first of the three terms. In this term, C m represents the capacitance of the myocyte membrane, or its ability to store charge, I BR is the membrane current, I ep is the current associated with electroporation, and I fu accounts for the possible stimulation of the tissue through an anode break. Looking first at I BR, we can split this term into four parts: In this case, I K and I X both represent potassium currents outward, I Na represents the sodium current in, and I S represents the calcium current inward. I K and I Na represent the currents generated during an action potential by the opening of voltage-gated sodium and potassium channels and I X represents the current caused by the presence of leaky potassium channels in the cells which are consistently letting small amounts of potassium out of the cell. Next, I ep accounts for the electroporation phenomenon. This phenomenon occurs when a shock is applied and results in the creation of pores in the membrane of the cell. These pores allow for ion flow which prevents the membrane potential from reaching an unrealistic level. Moving on to the next two terms in the equation, represents the extra-cellular potential and denotes intra-cellular diffusion of the electrical potential. The second term, ( ), describes the diffusion of the potential from the inside of the cell out. By taking the divergence of this term, we can understand how the diffusion is changing with time. Similarly, the third term,, describes the diffusion from the outside of the cell into the cell over time. Now, within our model, we will be considering seven main parameters that can influence the success or failure of defibrillation. The first is the shock waveform, or the three basic protocols of defibrillation described above. The second, shock duration, refers to how long the shock is applied. In the simulations done by Bragard et al., this was kept constant. Next, shock energy refers to how strong the shock is. The strength of the shock with be varied from E= 1 V/cm to E= 10V/cm. Next, shock timing refers to the location of the action potential wave front and back at the time of the shock. The size of the action potential is also important in determining defibrillation outcome as it is constantly varying as the dynamics are quasi-periodic. The heterogeneity of the cardiac tissue also plays a role as heterogeneities create more polarization sites within the tissue which makes excitation easier. Finally, the system size, or the size of the ring, plays a role. The system size in our model will be set to 6.7 cm, as was done by Bragard et al. In addition, our spatial discretization is 0.025cm. Also, during the first ten milliseconds of the shock, the time step with be set to 0.001 milliseconds. At every time following, it will be set to 0.01 milliseconds. Simulations were done by Bragard et al. to determine which of the three protocols was most efficient. We plan to mimic the simulations to both replicate their results and then test how the data changes when either the shock duration is changed or when the asymmetric biphasic shock has a shorter first phase. The simulations conducted were split into three main parts. First, the research team used an artificial neural network to classify four different types of successful defibrillations (see Figure 3). The first type, direct block, occurs when the front of the

action potential is near the anode when current is applied and is then directly blocked. This is shown by the dark block in the center of the color plot (a) followed by a vertical strip showing lower membrane potential, indicating a successful defibrillation. The second type, direct annihilation, occurs when two action potentials potentiate around the ring in opposite directions and annihilate each other. This is shown in figure 3(b) where the colors come to a peak and then die out. Delayed block occurs when the action potential generated propagates around the ring until it reaches a refractory region, or an area of lower membrane potential, and then dies out (see Figure 3(c)). This mechanism tends to last longer and the action potential can travel around the ring multiple times before dying out. The fourth and final mechanism is called direct activation. This is seen when the entire ring is excited to a higher membrane potential, as is shown in the bright red in the lower right graph of figure 3. Because the whole ring is activated, the action potential can no longer travel about the ring, rendering the defibrillation successful when the ring relaxes again once more. Figure 3. Color plots of the four mechanisms. (a) is direct block, (b) is annihilation, (c) delayed block, and (d) direct activation. Colors represent membrane potential with lighter colors indicating depolarization and darker color indicating repolarization. Time is on the x-axis and position on the ring on the y-axis. The proportions of the different mechanisms as well as the number of successes were determined for energy levels ranging from 1 to 10 V/cm. At lower energy levels (E= 1 V/cm), the monophasic protocol was most efficient with 27.50% of defibrillations being successful (3). In addition, the direct block mechanism was seen more at lower energy levels and only during monophasic runs. At higher energy levels (E= 7-10 V/cm), greater than 90 % of all defibrillations were successful for each of the three protocols (3). In addition, direct activation was the most predominant mechanism seen at higher energies. Finally, at energies ranging from 2 to 5 V/cm, the asymmetric biphasic protocol was found to be most efficient. Next, the efficiency of the shock was quantified using a Dose-Response curve: ( ) β 0 + β 1 E, where p is the probability of success, β 0 and β 1 are parameters fit to the curve, and E is the applied energy. Using this curve, it was determined that in order to have 90% of defibrillations be successful, the energy applied during the asymmetric biphasic shock should

range between 5.83 and 5.85 V/cm (3). This again indicated that the asymmetric biphasic protocol was more efficient than the other two, which required energy levels of 6.77-6.80 and 6.02-6.04 for monophasic and symmetric biphasic respectively (3). Finally, two parameters, (action potential duration) and b (action potential location on the ring), were varied and the probability of successful defibrillation mapped out on color histograms. The wave back of the action potential was used rather than the wave front to record the results as previous studies have shown that perturbations along the ring are more sensitive in the wave back. This was confirmed again by Bragard et al. as along the y-axis, constant b values yield similar results, indicating its relevance (3). At low energy (E = 1V/cm), monophasic and asymmetric biphasic protocols have similar efficiencies (see Figure 4). In addition, the direct activation mechanism was not present at low energies and so was not included. Because direct activation works by exciting a large portion of tissue through multiple electrodes, or in the case of our model, by exciting the whole ring, it was not seen at low energies. Also, the direct block mechanism was only relevant in the monophasic protocol at low energies. Furthermore, because the regions of high probability of success are localized, if b and are known when the shock is applied, the probability of defibrillation can be predicted with high accuracy at low energy. Figure 4. Color histograms showing the probability of successful defibrillation. The top row represents monophasic shock, the middle symmetric biphasic, and the bottom asymmetric biphasic. The first column shows the direct block mechanism, the second annihilation, the third delayed block, and the fourth is an overlay of the first three columns.

When the energy was slightly increased (E = 3V/cm), the monophasic protocol became less efficient than both biphasic protocols. In addition, the results were not as clear cut as they were for the lower energy. Rather than just yielding results of either 100% or 0% defibrillation, intermediate values such as 50% defibrillation as seen. This indicates that we cannot predict the probability of defibrillation by only knowing b and for higher energies. Next, when the energy was increased to 5V/cm, the direct block mechanism was no longer seen. However, the direct activation mechanism predominated and had the highest success rate of defibrillation throughout the three protocols. In addition, at lower values, the success of defibrillation was increased. This is expected because at lower values, there is more tissue to excite resulting in increased propagation which assists defibrillation, resulting in more successes. Finally, when the energy was increased to 7 V/cm, direct activation was dominant for biphasic shocks, and the asymmetric biphasic protocol was the most efficient (see Figure 5). Also, successful defibrillations were uniformly spread throughout the variations of the two parameters in both biphasic protocols. This is expected as both asymmetric and symmetric biphasic protocols render both electrodes equivalent meaning that the probability of a successful defibrillation should be evenly dispersed. Also, for the monophasic protocol, there was less chance of a successful defibrillation for any of the mechanisms when b is located at the cathode. This results from the wave back at shock initiation being too close to the hyperpolarized region, which causes less refractory and thus shock failure. Figure 5. Color histograms showing the probability of successful defibrillation. The top row represents monophasic shock, the middle symmetric biphasic, and the bottom asymmetric biphasic. The first column shows the annihilation mechanism, the second delayed block, the third direct activation, and the fourth is an overlay of the first three columns.

In conclusion, the energy levels used throughout the experiment by Bragard et al. are values that could be used in a commercial defibrillator. Monophasic defibrillators, which are no longer commonly used, produce shocks vary around 5000 Volts. Biphasic defibrillators range from 200-2200 Volts with most common biphasic defibrillators using 1700 Volts. This corroborates the conclusion that biphasic shocks are more efficient than monophasic, as less energy is needed to successfully defibrillate. This is important to consider as lower energy shocks cause less damage to tissue (3). In order to expand upon the work done by Bragard et al., we plan to simulate how the rates of success of defibrillation change when the shock duration is shorten, as shorter shocks cause less damage to the tissue than longer shocks. In addition, we plan to test the differing efficiency of two asymmetric biphasic shocks. The first asymmetric shock will have a longer first phase and a shorter second phase as was tested in Bragard et al.. The second will have a shorter first phase and longer second phase. The simulations discussed above will be replicated and expanded on as we change these two other parameters.

References: 1. Center for Disease Control. Atrial Fibrillation Fact Sheet. (2010). 2. Keating, M. T. and Sanguinetti, M. C. Molecular and cellular mechanisms of cardiac arrhythmias. Cell. 104 (2001) 569-580. 3. Bragard, J. et al. Shock-induced termination of reentrant cardiac arrhythmias: Comparing monophasic and biphasic shock protocols. Chaos: An Interdisciplinary Journal of Nonlinear Science. 23 (2013). 4. Beeler, G.W. and Reutter, H. Reconstruction of the action potential of ventricular myocardial fibers. Journal of Physiology. 268 (1977) 177-210. 5. Barnett, M.W. and Larkman, P.M. The action potential. Practical Neurology 7 (2007) 192-197.