Review Article

David S. Warner, M.D., Editor

Clinical for Anesthesiologists Part I: Background and Basic Signatures

Patrick L. Purdon, Ph.D., Aaron Sampson, B.S., Kara J. Pavone, B.S., Emery N. Brown, M.D., Ph.D. Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

ABSTRACT

The widely used electroencephalogram-based indices for depth-of- assume that the same index value defines the same level of unconsciousness for all . In contrast, we show that different anesthetics act at different molecular targets and neural circuits to produce distinct brain states that are readily visible in the electroencephalogram. We present a two-part review to educate anesthesiologists on use of the unprocessed electroencephalogram and its spectro- gram to track the brain states of patients receiving anesthesia care. Here in part I, we review the biophysics of the electro- encephalogram and the neurophysiology of the electroencephalogram signatures of three intravenous anesthetics: propofol, dexmedetomidine, and ketamine, and four inhaled anesthetics: sevoflurane, isoflurane, desflurane, and . Later in part II, we discuss patient management using these electroencephalogram signatures. Use of these electroencephalogram signatures suggests a neurophysiologically based paradigm for brain state monitoring of patients receiving anesthesia care. (­ 2015; 123:937-60)

The Electroencephalogram and Brain patients over time as a three-dimensional plot (power by fre- Monitoring under General Anesthesia quency vs. time).11 Fleming and Smith12 devised the density- modulated or density spectral array, the two-dimensional plot Almost 80 yr ago, Gibbs et al. demonstrated that system- of the spectrogram, for the same purpose.13 Levy14 later sug- atic changes occur in the electroencephalogram and patient gested using multiple electroencephalogram features to track arousal level with increasing doses of ether or . effects. Despite further documentation of system- They stated that “a practical application of these observations atic relations among anesthetic doses, electroencephalogram might be the use of electroencephalogram as a measure of the 4,15–20 1 patterns, and patient arousal levels, use of the unpro- depth of anesthesia.” Several subsequent studies reported on cessed electroencephalogram and the spectrogram to monitor the relation between electroencephalogram activity and the the states of the brain under general anesthesia and 2–6 7 behavioral states of general anesthesia. Faulconer showed never became a standard practice in anesthesiology. in 1949 that a regular progression of the electroencephalo- Instead, since the 1990s, depth of anesthesia has been gram patterns correlated with the concentration of ether in tracked using indices computed from the electroencephalo- arterial blood. Bart et al. and Findeiss et al. used the spec- gram and displayed on brain monitoring devices.21–25 The trum—the decomposition of the electroencephalogram signal indices have been developed by recording simultaneously into the power in its frequency components—to show that the the electroencephalogram and the behavioral responses to electroencephalogram was organized into distinct oscillations various anesthetic agents in patient cohorts.26 Some of the at particular frequencies under general anesthesia.8,9 Bickford indices have been derived by using regression methods to et al.10 introduced the compressed spectral array or spectrogram relate selected electroencephalogram features to the behav- to display the electroencephalogram activity of anesthetized ioral responses.26–29 One index has been constructed by

This article is featured in “This Month in Anesthesiology,” page 1A. Drs. Purdon and Brown have summarized some of this work on the Web site www.anesthesiaEEG.com and have given several seminars on this topic in multiple venues during the last 2 yr. Submitted for publication April 8, 2013. Accepted for publication May 18, 2015. From the Department of Anesthesia, Critical Care, and Pain , Massachusetts General Hospital, Boston, Massachusetts, and Department of Anesthesia, Harvard Medical School, Boston, Mas- sachusetts (P.L.P.); Department of Anesthesia, Critical Care, and Pain Medicine, Massachusetts General Hospital, Boston, Massachusetts (A.S., K.J.P.); and Department of Anesthesia, Critical Care, and Pain Medicine, Massachusetts General Hospital, Boston, Massachusetts; Department of Anesthesia, Harvard Medical School, Boston, Massachusetts; Institute for Medical Engineering and Science and Harvard-Massachusetts Institute of Technology, Health Sciences and Technology Program; and Department of Brain and Cognitive Sciences, Massachusetts Institute of Technology, Cambridge, Massachusetts (E.N.B.). Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. All Rights Reserved. Anesthesiology 2015; 123:937–60

Anesthesiology, V 123 • No 4 937 October 2015 Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. Electroencephalography for Anesthesiologists

using classifier methods to derive a continuum of arousal lev- signatures observable in the unprocessed electroencephalo- els ranging from awake to profound unconsciousness from gram and the spectrogram. The new concept of defining the visually categorized electroencephalogram recordings.30,31 anesthetic state by using drug-specific electroencephalogram Another index has been constructed by relating the entropy signatures that relate to molecular and neural circuit mecha- of the electroencephalogram signal—its degree of disorder— nisms of anesthetic action would allow anesthesiologists to to the behavioral responses of the patients.32,33 The indices make more detailed and accurate assessments than those are computed from the electroencephalogram in near real based on electroencephalogram-based indices. The potential time and displayed on the depth-of-anesthesia monitor benefits of using the unprocessed electroencephalogram to as values scaled from 0 to 100, with low values indicating monitor anesthetic states have been recently stated.48,49 Cur- greater depth of anesthesia. The algorithms used in many rent brain monitors display the unprocessed electroencepha-

of the current depth-of-anesthesia monitors to compute the logram and the spectrogram.22,31,50 Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021 indices are proprietary. To define anesthetic states in terms of drug-specific elec- Although the electroencephalogram-based indices have troencephalogram signatures that relate to molecular and been in use for approximately 20 yr, there are several reasons neural circuit mechanisms of anesthetic action, we synthesize why they are not part of standard anesthesiology practice. different sources and levels of information: (1) formal behav- First, use of electroencephalogram-based indices does not ioral testing along with either simultaneous human electro- ensure that awareness under general anesthesia can be pre- encephalogram recordings or human intracranial recordings vented.34,35 Second, these indices, which have been devel- during anesthetic administration; (2) clinical observations oped from adult patient cohorts, are less reliable in pediatric of behavior along with simultaneous electroencephalogram 36,37 populations. Third, because the indices do not relate recordings during anesthetic administration; (3) the neuro- directly to the neurophysiology of how a specific anesthetic physiology and the molecular pharmacology of how the anes- exerts its effects in the brain, they cannot give an accurate thetics act at specific molecular targets in specific circuits; picture of the brain’s responses to the drugs. Finally, the indi- (4) the neurophysiology of altered states of arousal such as ces assume that the same index value reflects the same level non-rapid eye movement , (medically induced, of unconsciousness for all anesthetics. This assumption is pathologic, or hypothermia induced), hallucinations, and based on the observation that several anesthetics, both intra- paradoxical excitation; (5) time–frequency analyses of high- venous and inhaled agents, eventually induce slowing in the density electroencephalogram recordings; and (6) mathemat- 1,4,22 electroencephalogram oscillations at higher doses. The ical modeling of anesthetic actions in neural circuits. slower oscillations are assumed to indicate a more profound We present this new education paradigm in two parts. state of general anesthesia. Two anesthetics whose electroen- Here, in part I, we review the basic neurophysiology of the cephalogram responses frequently lead clinicians to doubt electroencephalogram and the neurophysiology and the elec- 38,39 40–42 index readings are ketamine and nitrous oxide. troencephalogram signatures of three intravenous anesthet- These agents are commonly associated with faster electro- ics: propofol, dexmedetomidine, and ketamine, and four encephalogram oscillations that tend to increase the value of inhaled anesthetics: sevoflurane, isoflurane, desflurane, and the indices at clinically accepted doses. Higher index values nitrous oxide. We explain, where possible, how the anesthet- cause concern as to whether the patients are unconscious. At ics act at specific receptors in specific neural circuits to pro- the other extreme, dexmedetomidine can produce profound 43 duce the observed electroencephalogram signatures. In part slow electroencephalogram oscillations and low index val- II, we discuss how knowledge of the different electroenceph- ues consistent with the patient being profoundly uncon- alogram signatures may be used in patient management. scious. However, the patient can be easily aroused from what 43,44 is a state of sedation rather than unconsciousness. The Electroencephalogram: A Window into These ambiguities in using electroencephalogram-based the Brain’s Oscillatory States indices to define brain states under general anesthesia and sedation arise because different anesthetics act at different Coordinated action potentials, or spikes, transmitted and molecular targets and neural circuits to create different states received by neurons, are one of the fundamental mechanisms of altered arousal45,46 and, as we show, different electroen- through which information is exchanged in the brain and cephalogram signatures.43,47 The signatures are readily visible central nervous system (fig. 1A).51,52 Neuronal spiking activ- as oscillations in the unprocessed electroencephalogram and ity generates extracellular electrical potentials,53 composed its spectrogram. We relate these oscillations to the actions of primarily of postsynaptic potentials and neuronal membrane the anesthetics at specific molecular targets in specific neural hyperpolarization (fig. 1A).53,54 These extracellular potentials circuits. are often referred to as local field potentials. Populations of Therefore, we propose a new approach to brain monitoring neurons often show oscillatory spiking and oscillatory local of patients receiving general anesthesia or sedation: train anes- field potentials that are thought to play a primary role in thesiologists to recognize and interpret anesthetic-induced coordinating and modulating communication within and brain states defined by drug-specific neurophysiological among neural circuits.52 Local field potentials produced in

Anesthesiology 2015; 123:937-60 938 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. EDUCATION Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

Fig. 1. The neurophysiological origins of the electroencephalogram. (A) Neuronal spiking activity–induced oscillatory extracel- lular electrical currents and potentials are two of the ways that information is transmitted, modulated, and controlled in the central nervous system. (B) The geometry of the neurons in the cortex favors the production of large extracellular currents and potentials. (C) The electroencephalogram recorded on the scalp is a continuous measure of the electrical potentials produced in the cortex. (D) Because the cortex (orange region) is highly interconnected with subcortical regions, such as the thalamus (yellow region), and the major arousal centers in the basal forebrain, hypothalamus, midbrain, and pons, profound changes in neural activity in these areas can result in major changes in the scalp electroencephalogram. A is reproduced, with permission, from Hughes and Crunelli: Thalamic mechanisms of electroencephalogram alpha rhythms and their pathological implications. Neuroscientist, 2005; 11:357–72. B is reproduced, with permission, from Rampil: A primer for electroencephalogram signal pro- cessing in anesthesia. Anesthesiology 1998; 89:980–1002.

the cortex can be measured at the scalp as the electroenceph- the scalp because the electric field decreases in strength as the alogram (fig. 1B). square of the distance from its source.56 However, because cor- The organization of the pyramidal neurons in the cortex tical and subcortical structures are richly interconnected, scalp favors the production of large local field potentials because the electroencephalogram patterns reflect the states of both cortical dendrites of the pyramidal neurons run parallel with each other and subcortical structures.57 Thus, the electroencephalogram and perpendicular to the cortical surface (fig. 1C). This geom- provides a window into the brain’s oscillatory states. etry creates a biophysical transmitting antenna that generates A growing body of evidence suggests that anesthet- large extracellular currents whose potentials can be measured ics induce oscillations that alter or disrupt the oscillations through the skull and scalp as the electroencephalogram.53,55,56 produced by the brain during normal information process- Subcortical regions, such as the thalamus (fig. 1D), produce ing.19,20,57–63 These anesthesia-induced oscillations are read- much smaller potentials that are more difficult to detect at ily visible in the electroencephalogram.

Anesthesiology 2015; 123:937-60 939 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. Electroencephalography for Anesthesiologists

Data Analysis spectral analysis by computing the spectrum (fig. 3C) and 64,65,69 To show the unprocessed electroencephalogram recordings the spectrogram (figs. 3D and 3E). and their spectrograms computed with the same methods For a given segment of electroencephalogram data, the on comparable displays, we have taken examples from cases spectrum (fig. 3C) provides a decomposition of the segment of patients receiving general anesthesia and sedation at our into its frequency components usually computed by Fourier 64,65 institution. These data were recorded following a protocol methods. The advantage of the spectrum is that it shows approved by the Massachusetts General Hospital Human the frequency decomposition of the electroencephalogram Research Committee to construct a database of electroen- segment for all of the frequencies in a given range (fig. 3C) cephalogram of recordings of patients receiving general anes- by plotting frequency on the x-axis and power on the y-axis. thesia and sedation. Electroencephalograms were recorded Power is commonly represented in decibels, defined as 10 using the Sedline monitor (Masimo Corporation, USA). times the log base 10 of the squared amplitude of a given Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021 The Sedline electrode array records approximately at posi- electroencephalogram frequency component. Electroenceph- tions Fp1, Fp2, F7, and F8, with the reference approximately alogram power can differ by orders of magnitude across fre- 1 cm above Fpz and the ground at Fpz. We required imped- quencies. Taking logarithms makes it easier to visualize on the ances less than 5kΩ in each channel. Our analyses use Fp1 same scale frequencies whose powers differ by orders of mag- nitude. The spectrum of a given data segment is thus a plot of as results were identical for Fp1 and Fp2. We also include 2 60 (amplitude) ) by frequency. analyses of human intracranial recordings and high-density power (10 log10 electroencephalogram recordings20 from previous studies. The frequency bands in the spectrum are named following The spectrograms were computed using the multitaper a generally accepted convention (table 1). Changes in power in method64,65 from the unprocessed electroencephalogram sig- these bands can be used to track the changes in the brain’s anes- nals recorded at a sampling frequency of 250 Hz. Individual thetic states. In the signal shown in figure 3A, the low-frequency spectra were computed in 3-s windows with 0.5-s overlap oscillation has a period of approximately 1 cycle per second or between adjacent windows. Multitaper spectral estimates 1 Hz (table 1, slow oscillation), whereas the period of the faster have near optimal statistical properties64,65 that substan- oscillation is at approximately 10 Hz (table 1, alpha oscillation). tially improve the clarity of spectral features. At present, no The spectrum (fig. 3C) also shows that this signal has power in current electroencephalogram monitor displays multitaper the delta range (1 to 4 Hz) and little to no power beyond 12 Hz. spectra or spectrograms. We present multitaper spectra in In addition to these conventional frequency bands, two this review to show the electroencephalogram signatures of other spectral features are commonly reported in electroen- different anesthetic drugs with the greatest clarity. cephalogram analyses in anesthesiology: the median frequency (fig. 3C, lower white curve) and the spectral edge frequency Time Domain and Spectral Measures of the (fig. 3C, upper white curve). The median frequency is the frequency that divides the power in the spectrum in half70,71 Brain’s and Anesthetic States (fig. 3C), whereas the spectral edge frequency is the frequency Many of the changes that occur in the brain with changes in below which 95% of the spectral power is located.71 In other anesthetic states can be readily observed in unprocessed elec- words, in the frequency range we use in our analyses of 0.1 to troencephalogram recordings (fig. 2). Different behavioral and 30 Hz, half of the power in the spectrum lies below the median neurophysiological states induced by anesthetics are associated and 95% of the power lies below the spectral edge frequency. In with different electroencephalogram waveforms. For example, figure 3C, the median frequency and spectral edge frequency figure 2 shows the electroencephalogram of the same patient in are 3.4 and 15.9 Hz, respectively, where the frequency range is different states of propofol-induced sedation and unconscious- 0.1 to 30 Hz. The median frequency and the spectral edge are ness.20,65 These include the awake state (fig. 2A), paradoxical displayed on commercial monitors and are useful clinically for excitation (fig. 2B), a sedative state (fig. 2C), the slow and alpha tracking whether spectrogram power is shifting to lower (lower oscillation anesthetic state (fig. 2D), the slow oscillation anes- median frequency and spectral edge) or higher (higher median thetic state (fig. 2E), burst suppression (fig. 2F), and the isoelec- frequency and spectral edge) frequencies. As we discuss, the tric state (fig. 2G). Visualization and analysis of the unprocessed interpretations of these shifts are anesthetic dependent. electroencephalogram is a form of time domain analysis.64,65 The spectrum displays the power content by frequency for This approach is commonly used in sleep medicine and sleep only a single segment of electroencephalogram data. Use of the research to define sleep states66 and also in epileptology to char- spectrum in electroencephalogram analyses during anesthesia acterize seizure states.67 care requires computing it on successive data segments. Suc- Reading the frequencies and amplitudes from the unpro- cessive computation of the spectrum across contiguous, often cessed electroencephalogram in real time in the operating overlapping, segments of data is termed the spectrogram64,65 room is challenging. If the frequencies of the oscillatory (fig. 3D and 3E). The spectrogram makes it possible to dis- components were known, then it would be possible to design play how the oscillations in the electroencephalogram change specific filters to extract these components (fig. 3, A and B). in time, with changes in the dosing of the anesthetics and/or The more practical and informative solution is to conduct a the intensity of arousal-provoking stimuli. The spectrogram is a

Anesthesiology 2015; 123:937-60 940 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. EDUCATION Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

Fig. 2. Unprocessed electroencephalogram signatures of propofol-induced sedation and unconsciousness. (A) Awake eyes open electroencephalogram pattern. (B) Paradoxical excitation. (C) Alpha and beta oscillations commonly observed during propofol-induced sedation (fig. 5). (D) Slow-delta and alpha oscillations commonly seen during unconsciousness. (E) Slow os- cillations commonly observed during unconsciousness at induction with propofol (fig. 6) and sedation with dexmedetomidine (fig. 11) and with nitrous oxide (fig. 13). (F) Burst suppression, a state of profound anesthetic-induced brain inactivation com- monly occurring in elderly patients,68 anesthetic-induced coma, and profound hypothermia (fig. 6, B and D). (G) Isoelectric electroencephalogram pattern commonly observed in anesthetic-induced coma and profound hypothermia. With the exception of the isoelectric state, the amplitudes of the electroencephalogram signatures of the anesthetized states are larger than the amplitudes of the electroencephalogram in the awake state by a factor of 5 to 20. All electroencephalogram recordings are from the same subject. Reproduced, with permission, from Brown et al. Chapter 50 in Miller’s Anesthesia, 8th edition, 2014.

three-dimensional structure (fig. 3D). However, it is plotted in is termed the density spectral array (fig. 3E),12,13 whereas the two dimensions by placing time on the x-axis, frequency on the three-dimensional plot of the spectrogram is termed the com- y-axis, and power through color coding on the z-axis (fig. 3E). As pressed spectral array (fig. 3D).10,22 We display the spectrogram discussed earlier, this two-dimensional plot of the spectrogram as a density spectral array and refer to it as the spectrogram.

Anesthesiology 2015; 123:937-60 941 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. Electroencephalography for Anesthesiologists Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

Fig. 3. Construction of the spectrogram. (A) A 10-s electroencephalogram (EEG) trace recorded under propofol-induced unconsciousness. (B) The electroencephalogram trace in A filtered into its two principal oscillations: the blue curve, an alpha (8 to 12 Hz) oscillation, and the green curve, a slow (0.1 to 1 Hz) oscillation. (C) The spectrum provides a decomposition of the electroencephalogram in A into power by frequency for all of the frequencies in a specified range. The range here is 0.1 to 30 Hz. Power at a given frequency is defined in decibels as the 10 times the log base 10 of the squared amplitude. The green horizontal line underscores the slow-delta frequency band and the blue horizontal line underscores the alpha frequency band used to compute the filtered signals in B. The median frequency, 3.4 Hz (dashed vertical line), is the frequency that divides the power in the spectrum in half. The spectral edge frequency, 15.9 Hz (solid vertical line), is the frequency such that 95% of the power in the spectrum lies below this value. (D) The three-dimensional (3D) spectrogram (compressed spectral array) displays the successive spectra computed on a 32-min electroencephalogram recording from a patient anesthetized with propofol.

Anesthesiology 2015; 123:937-60 942 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. EDUCATION Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

Fig. 4. Neurophysiological mechanisms of propofol’s actions in the brain. Propofol enhances γ-aminobutyric acid receptor type

A (GABAA)-mediated inhibition in the cortex, thalamus, and brainstem. Shown are three major sites of action: postsynaptic con- nections between inhibitory interneurons and excitatory pyramidal neurons in the cortex; the GABAergic neurons in the thalamic reticular nucleus (TRN) of the thalamus; and postsynaptic connections between GABAergic and galanergic (Gal) projections from the preoptic area (POA) of the hypothalamus and the monoaminergic nuclei, which are the tuberomammillary nucleus (TMN) that releases histamine (His), the locus ceruleus (LC) that releases norepinephrine (NE), the dorsal raphe (DR) that releases serotonin (5HT); the ventral periacqueductal gray (vPAG) that releases dopamine (DA); and the cholinergic nuclei that are the basal forebrain (BF), pedunculopontine tegmental (PPT) nucleus, and the lateral dorsal tegmental (LDT) nucleus that release acetylcholine (ACh). Also shown is the lateral hypothalamus (LH) that releases orexin. Adapted, with permission, from Brown, Purdon, and Van Dort: General anesthesia and altered states of arousal: A systems neuroscience analysis. Annu Rev Neurosci 2011; 34:601–28. Adapta- tions are themselves works protected by copyright. In order to publish this adaptation, authorization has been obtained both from the owner of the copyright of the original work and from the owner of copyright of the translation or adaptation.

Spectral analyses make it easier to visualize frequency con- Neurophysiology and Clinical tent, especially oscillations, and to detect subtle changes in fre- Electrophysiology of Selected Intravenous quency structure. Nevertheless, it is important to know both Anesthetics the time domain and spectral representations of a given behav- We review the neuropharmacology and clinical electro- ioral or neurophysiological state induced by an anesthetic. We physiology of propofol, dexmedetomidine, and ketamine. present both in our discussions in the following sections for For each anesthetic, we discuss the putative mechanism commonly used intravenous and inhaled anesthetics. through which its actions at specific molecular targets in specific neural circuits produce the electroencephalogram Fig. 3. (Continued ). Each spectrum is computed on a 3-s inter- signatures and the behavioral changes associated with its val and adjacent spectra have 0.5 s of overlap. The black curve anesthetic state. at minute 24 is the spectrum in C. (E) The spectrogram in D plot- ted in two dimensions (density spectral array). The black vertical Table 1. Spectral Frequency Bands curve is the spectrum in D. The lower white curve is the time course of the median frequency and the upper white curve is the Frequency Range time course of the spectral edge frequency. A–E were adapted, Name (Hertz, Cycles per Second) with permission, from Purdon and Brown, Clinical Electroen- Slow < 1 cephalography for the Anesthesiologist (2014), from the Partners Delta 1–4 69 Healthcare Office of Continuing Professional Development. Theta 5–8 Adaptations are themselves works protected by copyright. In or- Alpha 9–12 der to publish this adaptation, authorization has been obtained Beta 13–25 both from the owner of the copyright of the original work and Gamma 26–80 from the owner of copyright of the translation or adaptation.

Anesthesiology 2015; 123:937-60 943 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. Electroencephalography for Anesthesiologists Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021 Fig. 5. Spectrogram and the time domain signature of propofol-induced sedation. (A) Spectrogram shows slow-delta oscillations (0.1 to 4 Hz) and alpha-beta (8 to 22 Hz) oscillations in a volunteer subject receiving a propofol infusion to achieve and maintain a target effect-site concentration of 2 μg/ml, starting at time 0.20 The subject was responding correctly to the verbal but not to click train auditory stimuli delivered every 4-s for the entire 16 min, suggesting that she was becoming sedated.20 The lower and upper white curves are the median and the spectral edge frequencies, respectively. (B) Ten-second electroencephalogram trace recorded at minute 6 of the spectrogram in A.

Molecular and Neural Circuit Mechanisms of Propofol oscillations are larger than those of the gamma oscillations seen Propofol, the most widely administered anesthetic agent, is in the awake electroencephalogram (fig. 2A). When general used as an induction agent for sedation and maintenance of anesthesia has been maintained by a propofol infusion, a simi- general anesthesia. Low-dose propofol is frequently adminis- lar beta oscillation pattern is visible in patients after extuba- tered either as small boluses or by infusion for and tion as they lie quietly before transfer to the postanesthesia care diagnostic procedures that require only sedation. unit. A beta oscillation is also seen during paradoxical excita- The molecular mechanism of propofol has been well char- tion (fig. 2B), the state of euphoria or dysphoria with move- acterized. Propofol binds postsynaptically to γ-aminobutyric ments, that can occur when patients are sedated. The state is

acid type A (GABAA) receptors where it induces an inward termed paradoxical because a dose of propofol intended to chloride current that hyperpolarizes the postsynaptic neurons sedate results in excitation. Two mechanisms have been pro- thus leading to inhibition.72,73 Because the drug is lipid soluble posed to explain propofol-induced paradoxical excitation. One

and GABAergic inhibitory interneurons are widely distrib- involves GABAA1-mediated inhibition of inhibitory inputs uted throughout the cortex, thalamus, brainstem, and spinal from the globus pallidus to the thalamus leading to increased cord, propofol induces changes in arousal through its actions excitatory inputs from the thalamus to the cortex.46 This mech- at multiple sites (fig. 4). In the cortex, propofol induces inhi- anism is also the one through which the sedative zolpidem is bition by enhancing GABA-mediated inhibition of pyramidal postulated to induce arousal in minimally conscious patients.76 neurons.73 Propofol decreases excitatory inputs from the thala- The second mechanism, established in simulation studies, pos- mus to the cortex by enhancing GABAergic inhibition at the tulates that low-dose propofol induces transient blockade of thalamic reticular nucleus, a network that provides important slow potassium currents in cortical neurons.76 inhibitory control of thalamic output to the cortex. Because the thalamus and cortex are highly interconnected, the inhibi- The Electroencephalogram Signatures of Propofol Are tory effects of propofol lead not to inactivation of these circuits Slow-delta Oscillations on Induction but rather to oscillations in the beta (figs. 2C and 5) and alpha The electroencephalogram patterns observed during propo- (figs. 2D, 6 and 7) ranges. Propofol also enhances inhibition in fol general anesthesia depend critically on several factors, the the brainstem at the GABAergic projections from the preoptic most important of which is the rate of drug administration. area of the hypothalamus to the cholinergic, monoaminergic, When propofol is administered as a bolus for induction of and orexinergic arousal centers (fig. 4). Decreasing excitatory general anesthesia, the electroencephalogram changes within inputs from the thalamus and the brainstem to the cortex 10 to 30 s from an awake pattern with high-frequency, low- enhances hyperpolarization of cortical pyramidal neurons, an amplitude gamma and beta oscillations (fig. 2A) to patterns effect that favors the appearance of slow and delta oscillations of high-amplitude slow and delta oscillations (fig. 2E). The on the electroencephalogram (figs. 5–7).20,60,75 slow-delta and delta oscillations appear in the spectrogram as increased power between 0.1 to 5 Hz (fig. 6, A and B, The Electroencephalogram Signatures of Propofol between minutes 0 and 5) and in the time domain as high- Sedation and Paradoxical Excitation Are Beta-gamma amplitude oscillations (fig. 6C, minute 5.5, and fig. 6D, Oscillations minute 7.1). The electroencephalogram change is dramatic The electroencephalogram patterns seen during sedation are as the slow-delta oscillation amplitudes can be 5 to 20 times organized, regular beta-gamma oscillations (figs. 2C and 5) larger than the amplitudes of the gamma and beta oscilla- and slow-delta oscillations (fig. 5). The amplitudes of these tions seen in the electroencephalogram of an awake patient.43

Anesthesiology 2015; 123:937-60 944 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. EDUCATION Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

Fig. 6. Spectrogram and time domain electroencephalogram signatures of two patients receiving propofol for induction and maintenance of unconsciousness. (A) High slow-delta power after the 200-mg propofol bolus at minute 3 (green arrow) is evi- dent between minutes 3 and 5. The electroencephalogram transitions to robust slow-delta and alpha oscillations maintained by a propofol infusion at 100 μg kg−1 min−1. The lower and upper white curves are the median and the spectral edge frequencies, respectively. (B) After bolus doses of propofol (green arrows), the patient’s electroencephalogram transitions between three dif- ferent states: slow oscillations (minutes 5 to 8) after the 100-mg propofol bolus at minute 3; burst suppression (minutes 8 to 17) after two additional 50-mg propofol boluses; and slow-delta and alpha oscillations from minutes 17 to 25. Beginning at minute 24, the alpha band power decreases and broadens to the beta band. The slow-delta oscillation power decreases after minute 24. The dissipation of the slow-delta and alpha oscillation power as the patient emerges gives the appearance of a zipper open- ing. (C) Ten-second electroencephalogram traces recorded at minute 5.5 (slow-delta oscillations) and minute 24 (slow-delta and alpha oscillations) of the spectrogram in A. (D) Ten-second electroencephalogram traces showing slow oscillations at minute 7.1, burst suppression at minute 11.5, and slow-delta and alpha oscillations at minute 17 for the spectrogram in B. A–D were adapted, with permission, from Purdon and Brown, Clinical Electroencephalography for the Anesthesiologist (2014), from the Partners Healthcare Office of Continuing Professional Development.69 Adaptations are themselves works protected by copy- right. In order to publish this adaptation, authorization has been obtained both from the owner of the copyright of the original work and from the owner of copyright of the translation or adaptation.

The appearance of slow and delta oscillations (figs. 2 oscillations on the electroencephalogram with loss of con- and 6, C and D) within seconds of administering propo- sciousness (LOC).20,60,75 Loss of the oculocephalic reflex is fol for induction of general anesthesia coincides with loss of consistent with the anesthetic acting at cranial nerve nuclei responsiveness, loss of the oculocephalic reflex, apnea, and III, IV, and VI in the midbrain and pons.46,74 Apnea is most atonia.46,74 These electroencephalogram signatures and the likely due to the drug’s inhibition of the ventral and dorsal clinical signs are consistent with a rapid action of the anes- respiratory centers in the medulla and pons,76 whereas the thetic in the brainstem. After bolus administration, propofol brainstem component of atonia is most likely due to inhibi- quickly reaches the GABAergic inhibitory synapses emanat- tion of the pontine and medullary reticular nuclei.46 ing from the preoptic area of the hypothalamus onto the Administration of an additional propofol bolus, either major arousal centers in the brainstem and hypothalamus before or after intubation, can result in either enhance- (fig. 4). Action of the anesthetic at these synapses inhibits ment of the slow oscillation or the conversion of the slow the excitatory arousal inputs from the brainstem, favoring oscillation into burst suppression (fig. 6B, minutes 7 to hyperpolarization of the cortex, the appearance of slow-delta 14, and fig. 6D, minute 11.5). Burst suppression is a state

Anesthesiology 2015; 123:937-60 945 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. Electroencephalography for Anesthesiologists Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

Fig. 7. Spatiotemporal characterization of electroencephalogram alpha and slow oscillations observed during induction of and recovery from propofol-induced unconsciousness. (A) In the volunteer subject lying awake with eyes closed, spatially coherent alpha oscillations are observed over the occipital area. The alpha oscillations shift to the front of the head with loss of conscious- ness (LOC) where they intensify and become spatially coherent during unconsciousness. The alpha oscillations dissipate anteri- orly and return to the occipital area during return of (ROC) where they reintensify and are spatially coherent in the eyes-closed awake state. (B) During consciousness, there is broadband communication between the thalamus and the frontal cortex with beta and gamma activity in the electroencephalogram. Modeling studies suggest that during propofol-induced unconsciousness the spatially coherent alpha oscillations are highly structured rhythms in thalamocortical circuits.57 (C) Slow oscillations recorded 30 s after bolus induction of general anesthesia with propofol from grid electrodes implanted in a patient with epilepsy. The slow oscillations at nearby electrodes (red and green dots) are in phase (red and green traces), whereas the slow oscillation recorded at an electrode 2 cm away (blue dot) is out of phase (blue trace) with those at the other two locations. Neurons spike only (histograms) in a limited time window governed by the phase of the local slow oscillations. These slow oscil- lations are a marker of intracortical fragmentation with propofol as communication through spiking activity is restricted to local areas. The spatially coherent alpha oscillations and the disruption of neural spiking activity associated with the slow oscillations are likely to be two of the mechanisms through which propofol induces unconsciousness. A is adapted, with permission, from Purdon et al: Electroencephalogram signatures of loss and recovery of consciousness from propofol. Proc Natl Acad Sci U S A 2013; 110:E1142–51; and C is adapted, with permission, from Lewis et al. Rapid fragmentation of neuronal networks at the onset of propofol-induced unconsciousness. Proc Natl Acad Sci U S A 2012; 109:E3377–86. Adaptations are themselves works protected by copyright. In order to publish this adaptation, authorization has been obtained both from the owner of the copyright of the original work and from the owner of copyright of the translation or adaptation.

of unconsciousness and profound brain inactivation in as vertical lines in the spectrogram (fig. 6B, minutes 7 to which the electroencephalogram shows periods of electrical 14). When propofol is administered as an induction bolus, activity alternating with periods of isoelectricity or electri- patients, particularly elderly patients, can enter burst sup- cal silence. In the spectrogram, burst suppression appears pression within seconds.68

Anesthesiology 2015; 123:937-60 946 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. EDUCATION

As the effects of the bolus doses of propofol recede, the incoherent slow oscillations represent an impediment the electroencephalogram evolves into slow-delta oscilla- to normal intracortical communications.20,57,60,80 Together, tion and alpha oscillation patterns (fig. 6A, minute 7 and these two mechanisms likely contribute to the patient being fig. 6B, minute 15). The time of transition from the slow unconscious when the slow and alpha oscillations are vis- oscillations or burst suppression into a combined slow ible in the spectrogram of patients receiving propofol. These oscillation and alpha oscillation patterns depends on how highly coherent alpha oscillations, incoherent slow oscilla- profound the effect of the bolus dose was. Even if no addi- tions, and anteriorization most certainly contribute to the tional propofol is administered after the first bolus, it can loss of frontal-parietal effective connectivity80–82 that is also take several minutes for the transition to occur. The patient associated with propofol-induced LOC. shown in figure 6A received a second bolus dose of 50 mg

at minute 5. An infusion of propofol was started at a rate Electroencephalogram Signatures of Emergence from Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021 of 100 μg kg−1 min−1. Her slow-delta oscillations did not Propofol General Anesthesia evolve into the slow-delta and alpha oscillations until min- When the propofol infusion is discontinued, and the patient ute 7. In contrast, the patient shown in figure 6B stayed is allowed to emerge, the slow and alpha oscillations dis- in a state of burst suppression from minutes 7 to 14 after sipate and are gradually replaced by higher-frequency beta her two bolus doses of propofol. After the second bolus and gamma oscillations that have lower amplitudes (fig. 6A, dose, she received no additional propofol or other anesthet- minutes 25 to 30, and fig. 6B, minutes 23 to 27).20 In the ics and transitioned into slow-delta and alpha oscillations spectrogram, this gradual shift to higher-frequency beta and (fig. 6B, minutes 15 to 20, and fig. 6D, minute 17). If after gamma oscillations appears as a “zipper opening” pattern the bolus doses, propofol is administered as an infusion to (fig. 6A, minutes 25 to 30, and fig. 6B, minutes 23 to 27). maintain general anesthesia, the patient’s electroencephalo- The decrease in the amplitude is shown by the shift in the gram will continue to show the slow-delta and alpha oscil- spectrum from red to yellow. In the unprocessed electroen- lation patterns in the surgical plane (fig. 6A, minutes 9 to cephalogram, this change is marked by a gradual increase 26, and fig. 6C, minute 24). in frequency and decrease in amplitude of the oscillations. Simultaneously, there is dissipation of power in the slow and Slow-delta and Alpha Oscillations Are Markers of delta bands, which appears in the time domain as a flatten- Propofol-induced Unconsciousness ing of the unprocessed electroencephalogram. There is also Most brain function monitors used in anesthesiology record reversal of anteriorization with emergence (fig. 7A, return of only a few channels of the electroencephalogram from the consciousness and awake emergence) with loss of the coher- front of the head. An awake patient with eyes closed and a ent frontal alpha oscillations and return of the coherent full set of scalp electroencephalogram electrodes will show occipital alpha oscillations.20 The return of high-frequency the well-known eyes-closed alpha oscillations in the occipi- power in the electroencephalogram is consistent with return tal areas (fig. 7A, awake baseline).79 Concomitant with the of normal cortical activity and indirectly with return of nor- transition to LOC and the appearance of the slow and alpha mal thalamic and brainstem activity. Brainstem function oscillations is the phenomenon of anteriorization, in which returns in an approximate caudal (medulla and lower pons) the power in the alpha and beta bands of the electroencepha- to rostral (upper pons and midbrain) manner.46,74 Breathing logram shifts from the occipital area to the front of the head and gagging are controlled in the medulla and lower pons, (fig. 7A, LOC and unconsciousness).15,19,20 whereas the corneal and oculocephalic reflexes are controlled Although LOC due to propofol has a strong brainstem in the upper pons and the midbrain.46,74 Return of brain- effect, maintenance of unconsciousness involves the brain- stem, thalamic, and cortical activity is necessary to restore stem and other brain centers. Studies of propofol-induced the awake state. unconsciousness using a high-density (64-lead) electro- encephalogram have shown that the alpha oscillations are Burst Suppression and Medically Induced Coma highly coherent across the front of the head during uncon- When delivered in a sufficiently high dose, several anesthet- sciousness (fig. 7A, unconsciousness).19,20 An explanation for ics, including propofol, the barbiturates, and the inhaled this coherence is that propofol could be inducing an alpha ether drugs, induce burst suppression (figs. 2F and 6, B and oscillation within circuits linking the thalamus and the fron- D, minute 11.5, fig. 8, A and B).83–86 Burst suppression is tal cortex (fig. 7B).57 In contrast, the slow oscillations are induced by hypothermia for surgeries requiring total circula- not coherent.19,20 Studies of patients with intracranial elec- tory arrest87 and by administering anesthetics in the intensive trodes also show that the slow oscillations induced by bolus care unit for cerebral protection to treat intracranial hyper- administration of propofol are not coherent across the cortex tension or to treat .86,88–90 This latter state is (fig. 7C) and serve as a marker of phase-limited spiking activ- termed a medically induced coma. If the patient is in burst ity in the cortex (fig. 7C).60 We postulate that the highly orga- suppression, then, as the dose of the anesthetic is increased, nized coherent alpha oscillations most likely prevent normal the length of the suppression periods between the bursts communications between the thalamus and cortex, whereas increases. The dose can be increased to the point at which

Anesthesiology 2015; 123:937-60 947 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. Electroencephalography for Anesthesiologists Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

Fig. 8. Characterization of burst suppression. (A) The spectrogram in figure 6B from minutes 4 to 20. Burst suppression in the spectrogram shows as periods of blue (isoelectric activity) interspersed with periods of red-yellow (slow-delta and alpha oscilla- tions). The horizontal red line shows the principal period of burst suppression. (B) Unprocessed electroencephalogram record- ings corresponding to the spectrogram in A. The horizontal red lines at ±5 μV are the thresholds that separate burst events (amplitude ≥5 μV) from suppression events (amplitude < 5 μV). (C) The burst suppression probability provides an estimate of the instantaneous probability of the electroencephalogram being suppressed.98 Although it is apparent in the spectrogram and in the unprocessed electroencephalogram that the period of strong burst suppression extends from minutes 8 to 16, the burst suppression probability analysis shows that it does not completely subside until minute 17.

the electroencephalogram is isoelectric (fig. 2G). Within the measure of the instantaneous probability of the brain being bursts, the electroencephalogram can, in some cases, main- in a state of suppression that can be reliably computed using tain the brain dynamics that were present before the start state-space methods and used to track burst suppression of burst suppression.85,91,92 For example, if the patient is in in real time and to implement control systems for medical a state of slow and alpha oscillations before the burst sup- coma.99,100 A burst suppression probability of 0.5 means a pression, then the same pattern is present within the bursts 0.5 probability of being suppressed, whereas a burst suppres- (figs. 6B and 8A). In addition to deep general anesthesia, sion probability of 0.75 means a 0.75 probability of being burst suppression is also observed in other conditions of pro- suppressed. found brain inactivation including coma93 and in children with significantly compromised brain development.94 The Ketamine observation that multiple different mechanisms induce burst suppression are consistent with a recently proposed neuro- Neural Circuit Mechanisms of Ketamine nal metabolic mechanism of burst suppression.85 Transient Ketamine, an anesthetic adjunct and an , acts increases and decreases in extracellular calcium, leading to primarily by binding to N-methyl-d-aspartate (NMDA) synaptic disfacilitation, could also play a role in determining receptors in the brain and spinal cord.45 Ketamine is a suppression duration.84 channel blocker so that to be effective, the channels have The burst suppression ratio or suppression ratio is a time to be open.101 Because in general, the channels on inhibi- domain measure used to track quantitatively the level of tory interneurons are more active than those on pyramidal burst suppression. The burst suppression ratio is a number neurons, ketamine at low-to-moderate doses has its primary between 0 and 1, which measures the fraction of time in effect on inhibitory interneurons (fig. 9A).102,103 By block- a given time interval that the electroencephalogram is sup- ing inputs to inhibitory interneurons, ketamine allows pressed.95,96 The burst suppression ratio is displayed on some downstream excitatory neurons to become disinhibited or brain function monitors and is one of the measures used more active.45,46 This is why cerebral metabolism increases in electroencephalogram-based indices to assess depth of with low doses of ketamine. Hallucinations, dissociative anesthesia.97,98 A refinement of the burst suppression ratio, states, euphoria, and dysphoria are common with low-dose termed the burst suppression probability, has been recently ketamine because brain regions, such as the cortex, hippo- developed (fig. 8C).100 The burst suppression probability is a campus, and the amygdala, continue to communicate but

Anesthesiology 2015; 123:937-60 948 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. EDUCATION Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

Fig. 9. Neurophysiology and electroencephalogram signatures of ketamine. (A) At low doses, ketamine blocks preferentially the actions of glutamate N-methyl-d-aspartate receptors on γ-aminobutyric acid (GABA)ergic inhibitory interneurons in the cortex and subcortical sites such as the thalamus, hippocampus, and the limbic system. The antinociceptive effect of ketamine is due in part to its blockade of glutamate release from peripheral afferent (PAF) neurons in the dorsal root ganglia (DRG) at their syn- apses on to projection neurons (PNs) in the spinal cord. (B) Spectrogram showing the beta-gamma oscillations in the electroen- cephalogram of a 61-yr-old woman who received ketamine administered in 30 mg and 20 mg doses (green arrows) for a vacuum dressing change. Blocking the inhibitory action of the interneurons in cortical and subcortical circuits helps explain why ket- amine produces beta oscillations as its electroencephalogram signature. (C) Ten-second electroencephalogram trace recorded at minute 5 from the spectrogram in B. A is reproduced, with permission, from Brown, Purdon, and Van Dort: General anesthesia and altered states of arousal: A systems neuroscience analysis. Annu Rev Neurosci. 2011;324:601–28. B and C were adapted from Purdon and Brown, Clinical Electroencephalography for the Anesthesiologist (2014), with permission, from the Partners Healthcare Office of Continuing Professional Development.69 Adaptations are themselves works protected by copyright. In order to publish this adaptation, authorization has been obtained both from the owner of the copyright of the original work and from the owner of copyright of the translation or adaptation.

with less modulation and control by the inhibitory interneu- in the prefrontal cortex due in part to increased glutamate rons. Information is processed without proper coordination activity at non-NMDA glutamate receptors.104 Analgesia is in space and time.45,46 The hallucinatory effects are likely due in part to the action of ketamine on glutamate NMDA enhanced by disruption of dopaminergic neurotransmission receptors at the dorsal root ganglia, the first synapse of the

Anesthesiology 2015; 123:937-60 949 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. Electroencephalography for Anesthesiologists

pain pathway in the spinal cord where glutamate is the pri- of these inhibitory pathways from the preoptic area to the mary neurotransmitter (fig. 9A).101 As the dose of ketamine arousal centers. Activation of inhibitory inputs from the pre- is increased, the NMDA receptors on the excitatory gluta- optic area is postulated to be an essential component of how matergic neurons are also blocked and consciousness is lost. nonrapid eye movement sleep is initiated.113,114 Sedation by dexmedetomidine is further enhanced due to blockage The Electroencephalogram Signatures of Ketamine of presynaptic release of norepinephrine, leading to loss of Sedation Are Beta and Gamma Oscillations excitatory inputs from the locus ceruleus to the basal fore- Given the preference of ketamine for NMDA receptors on brain, intralaminar nucleus of the thalamus, and cortex115 inhibitory interneurons, whose inhibition results in increased and decreased thalamocortical connectivity.116 cerebral metabolic rate, cerebral blood flow, and hallucina- 105–107 tions, it is no surprise that ketamine is associated with The Electroencephalogram Signatures of Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021 an active electroencephalogram pattern. When ketamine is Dexmedetomidine Are Slow-delta Oscillations and administered alone in low dose, the electroencephalogram Spindles shows fast oscillations (fig. 9, B and C) in the high beta, low The relation between the actions of dexmedetomidine in the pre- gamma range between 25 and 32 Hz. The beta-gamma oscil- optic area and the initiation of non-rapid eye movement sleep is lation did not start until 2 min after the initial ketamine dose. important to appreciate because it helps explain the similarities Compared with propofol and dexmedetomidine (discussed in the electroencephalogram patterns between this anesthetic in the section The Electroencephalogram Signatures of Dex- and those observed in nonrapid eye movement sleep. Dexme- medetomidine Are Slow-delta Oscillations and Spindles), detomidine administered as a low-dose infusion induces a level the ketamine slow oscillation is less regular (fig. 9C). The of sedation in which the patient responds to minimal auditory 61-yr old whose spectrogram is shown in figure 9B required or tactile stimulation. The electroencephalogram shows a com- sedation and analgesia for change of a vacuum dressing over bination of slow-delta oscillations with spindles, which are 9 to a panniculectomy site. At minute 0, she received a 50-mg 15 Hz oscillations that occur in bursts lasting 1 to 2 s (figs. 10B bolus of ketamine in two doses of 30 and 20 mg, 2 min apart. and 11, A and B).43,44 In the frequency domain, the dexme- The dressing change started at minute 10. Five minutes after detomidine spindles appear as streaks in the high alpha and low administering the 20-mg ketamine dose, the patient con- beta bands between 9 to 15 Hz (fig. 11A). The spindles occur tinued to breathe spontaneously and was unresponsive to in a similar frequency range as the alpha oscillations seen with verbal commands and the nociceptive stimulation from the propofol but have much less power than the alpha oscillations procedure. Although the beta and gamma oscillations lasted (figs. 6, A and B).43 Because the color scales on the spectrograms for only 27 min, the sedative effect of being unresponsive to in figures 6, A and B, and 11A are the same, the plots provide an verbal commands persisted for several minutes after the beta- informative comparison of the amplitudes of the propofol alpha gamma oscillations had disappeared. oscillations and the dexmedetomine spindles. The dexmedeto- midine spindles closely resemble the spindles that define stage II Dexmedetomidine nonrapid eye movement sleep.43,44 Dexmedetomidine is used as a sedative in the The slow-delta oscillations are apparent in the spectro- and as a sedative and anesthetic adjunct in the operating room. gram as power from 0 to 4 Hz (fig. 11A). The 44-yr-old, Compared with propofol, patients are easily arousable when 59-kg female patient, whose spectrogram and time domain sedated with dexmedetomidine, with minimal to no respiratory traces are shown in figures 11, A and B, received a 1 μg/ depression. Unlike propofol and the , dexme- kg loading infusion of dexmedetomidine over 10 min, fol- detomidine cannot also be used as a hypnotic agent. lowed by a maintenance dexmedetomidine infusion of 0.65 μg kg−1 h−1—a rate in the intermediate to high end of the Neural Circuit Mechanisms of Dexmedetomidine sedative range—for the creation of a left forearm arteriove- Dexmedetomidine alters arousal primarily through its nous fistula for . During the , the patient was

actions on presynaptic α2 adrenergic receptors on neurons sedated, meaning that she responded to verbal queries from projecting from the locus ceruleus. Binding of dexmedetomi- the anesthesiologist and moved in response to nociceptive

dine to the α2 receptors hyperpolarizes locus coeruleus neu- stimulation from the surgery. rons decreasing norepinephrine release.108–110 The behavioral When the rate of the dexmedetomidine infusion is effects of dexmedetomidine are consistent with this proposed increased, spindles disappear and the amplitude of the slow- mechanism of action.111 Hyperpolarization of locus ceruleus delta oscillations increases (fig. 11, C and D). This electro- neurons results in loss of inhibitory inputs to the preoptic encephalogram pattern of slow-delta oscillations closely area of the hypothalamus (fig. 10A). The preoptic area sends resembles nonrapid eye movement sleep stage III or slow-wave GABAergic and galanergic inhibitory projections to the sleep.66 The slow-delta oscillations appear again as intense major arousals centers in the midbrain, pons, and hypothal- power in the slow oscillation band (fig. 11C). The power in the amus (fig. 10A).45,46,112 Hence, loss of the inhibitory inputs slow oscillation band for the higher dose of dexmedetomidine from the locus ceruleus results in sedation due to activation (fig. 11C) is considerably stronger than the slow oscillations

Anesthesiology 2015; 123:937-60 950 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. EDUCATION Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

Fig. 10. Neurophysiology of dexmedetomidine (dex). (A) Dexmedetomidine acts presynaptically to block the release of norepi- nephrine (NE) from neurons projecting from the locus coeruleus (LC) to the basal forebrain (BF), the preoptic area (POA) of the ­hypothalamus, and the intralaminar nucleus (ILN) of the thalamus and the cortex. Blocking the release of NE in the POA leads to activation of its inhibitory γ-aminobutyric acid (GABA)ergic and galanergic (Gal) projections to dorsal raphe (DR) that releases serotonin (5HT), the tuberomammillary nucleus (TMN) that releases histamine (His), the LC, the ventral periaqueductal gray (PAG) that releases dopamine, the lateral dorsal tegmental (LDT) nucleus, and the pedunculopontine tegmental (PPT) nucleus that release acetylcholine (ACh). These actions lead to decreased arousal by inhibition of the arousal centers. (B) Ten-second electroencephalogram segment showing spindles, intermittent 9 to 15 Hz oscillations (underlined in red), characteristic of dex- medetomidine sedation. (C) The spindles are most likely produced by intermittent oscillations between the cortex (orange region) and thalamus (light green region). DRG = dorsal root ganglia; PAF = peripheral afferent; PN = projection neuron. A is adapted, with permission, from Brown, Purdon, and Van Dort: General anesthesia and altered states of arousal: A systems neuroscience analysis. Annu Rev Neurosci 2011; 34:601–28. Adaptations are themselves works protected by copyright. In order to publish this adaptation, authorization has been obtained both from the owner of the copyright of the original work and from the owner of copyright of the translation or adaptation. DRG = dorsal root ganglia; PAF = peripheral afferent; PN = projection neuron.

for the lower dose of dexmedetomine (fig. 11A). The spectro- unresponsive to verbal queries from the anesthesiologist and gram is that of a 48-yr-old, 65-kg female patient who received but moved in response to changes in the level of nociceptive dexmedetomidine as a loading infusion of 1 μg/kg over 10 min stimulation from the procedure. and a maintenance infusion of 0.85 μg kg−1 h−1 for placement Propofol frontal alpha oscillations (fig. 7A), dexmedeto- of a left forearm arteriovenous fistula. The intense slow-delta midine spindles45 (fig. 11, A and B), and sleep spindles117 oscillation persisted for the duration of the procedure. Dur- are all thought to be generated by thalamocortical loop ing the surgery, the patient was sedated, meaning that she was mechanisms (figs. 7B and 10B). Although the propofol

Anesthesiology 2015; 123:937-60 951 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. Electroencephalography for Anesthesiologists Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

Fig. 11. Spectrograms and time domain electroencephalogram signatures of dexmedetomidine-induced sedation. (A) Spectro- gram of the electroencephalogram of a 59-kg patient receiving a 0.65 μg kg−1 h−1 dexmedetomidine infusion to maintain sedation. The spectrogram shows spindles (9 to 15 Hz oscillations) and slow-delta oscillations. (B) Ten-second electroencephalogram trace recorded at minute 60 from the spectrogram in A emphasizing spindles (red underlines). (C) Spectrogram of the electroencepha- logram of a 65-kg patient receiving a 0.85 μg kg−1 h−1 dexmedetomidine infusion to maintain sedation. (D) Ten-second electro- encephalogram trace recorded at minute 40 from the spectrogram in C showing the slow-delta oscillations. A–D were adapted, with permission, from Purdon and Brown, Clinical Electroencephalography for the Anesthesiologist (2014), from the Partners Healthcare Office of Continuing Professional Development.69 Adaptations are themselves works protected by copyright. In order to publish this adaptation, authorization has been obtained both from the owner of the copyright of the original work and from the owner of copyright of the translation or adaptation.

frontal alpha oscillations are highly coherent and continuous brief interruptions in neuronal activity.118 Similarly, both in time,19,20,57 the sleep and dexmedetomidine spindles are slow oscillations and spindles under dexmedetomidine are brief and episodic.43 Sleep spindles are a marker for nonrapid smaller than their counterparts under propofol, which may eye movement stage II sleep, which is a state of unconscious- reflect a lower level of disruption in neuronal activity under ness that is less profound than nonrapid eye movement stage dexmedetomidine compared with propofol.43 Consequently, III or slow-wave sleep.66 Dexmedetomidine spindles are con- cortical and thalamocortical activities are likely to be more sistent with a light state of sedation.43 profoundly inhibited under propofol-induced unconscious- Propofol-induced slow oscillations likely result from ness compared with either slow-wave sleep or dexmedeto- decreased excitatory inputs to the cortex due to propofol’s midine-induced sedation. This difference in cortical and GABAA-mediated inhibition arousal centers in the midbrain, thalamocortical activity suggests why patients can be aroused pons, and hypothalamus (fig. 4). Dexmedetomine-induced from sleep and dexmedetomidine-induced sedation but not slow oscillations likely result from decreased excitatory from propofol-induced unconsciousness.43,60 inputs to the cortex due to dexmedetomidine’s disinhibi- tion of the inhibitory circuits emanating from the preoptic Clinical Electrophysiology of the Inhaled area of the hypothalamus to the arousal centers along with Anesthetics decreased adrenergically-mediated excitatory inputs to the basal forebrain, the intralaminar nucleus of the thalamus, Neurophysiological Mechanisms of Inhaled of Anesthetic and to the cortex directly (fig. 10A). Propofol-induced slow Action oscillations gate brief periods of neuronal activity (fig. 7B).60 The principal inhaled anesthetics are the ether derivatives In contrast, sleep slow-wave oscillations are associated with sevoflurane, isoflurane, and desflurane. The inhaled agents are

Anesthesiology 2015; 123:937-60 952 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. EDUCATION

total anesthetics in that they can maintain all of the required time, the slow and delta oscillations dissipate. The loss of the behavioral and physiological characteristics of general anes- alpha and slow-delta oscillation power again appears in the thesia without adjunct. In practice, the inhaled anesthetics spectrogram as a zipper opening pattern. are rarely used alone but more commonly in conjunction Isoflurane (fig. 12, E and F) and desflurane (fig. 12, G and with intravenous drugs, nitrous oxide, and muscle relaxants H) have similar patterns to sevoflurane (fig. 12, C and D). At to provide balanced general anesthesia, which maximizes the sub-MAC concentrations, they also show strong alpha and desired effects while minimizing side effects. The inhaled slow-delta oscillations. When the concentration of isoflurane agents are known to create their behavioral and physiologi- or desflurane is increased to MAC levels and above, a theta cal effects through binding at multiple targets in the brain oscillation fills in between the delta and alpha bands (fig.

and central nervous system including binding to GABAA 12E, minute 30, and fig. 12G, minute 28). As in the case

receptors and enhancing GABAergic inhibition; blockade of of sevoflurane, on emergence from isoflurane and oxygen or Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021 two-pore potassium channels and ­hyperpolarizing-activated desflurane and oxygen anesthesia, there is loss of the theta cyclic nucleotide-gated channels; and blocking glutamate oscillation power, followed by dissipation of alpha and slow- release by binding to NMDA receptors.72 Studies of minimal delta oscillation power, and reappearance of the power in alveolar concentration (MAC) have shown that the muscle the beta and gamma bands (fig. 12E, minutes 78 to 80, and relaxant effects of the inhaled anesthetics are the result of fig. 12G, minutes 85 to 90). These observations suggest that direct action at one or more of these receptors and channels by analogy with sevoflurane, enhanced GABAergic inhibi- in the spinal cord.119 This mechanism is also consistent with tion is likely a primary but not the only mechanism through the observation that spinal cord motor-evoked potentials are which isoflurane and desflurane induce their anesthetic difficult to record in patients receiving anesthetic concentra- states. This relation between MAC and theta oscillations is tions of an ether anesthetic.120 useful clinically. The appearance of theta oscillations indi- Gibbs et al.1 showed that the electroencephalogram of cates a more profound state of unconsciousness and immo- patients receiving ether show slow-delta oscillations and bility for an inhaled ether anesthetic. alpha oscillations at surgical levels of general anesthesia. As with propofol, burst suppression can be induced by When sevoflurane is administered at sub-MAC concentra- administering a sufficiently high dose of any one of the tions to achieve surgical levels of general anesthesia, the inhaled ether anesthetic drugs.82,85 electroencephalogram shows slow-delta oscillations and coherent alpha oscillations similar to those of propofol.47 High-amplitude Slow-delta Oscillations Mark the This observation suggests that for sevoflurane, as for propo- Transition to High-flow Nitrous Oxide fol, enhanced GABAergic inhibition may be its dominant Nitrous oxide has been used as a sedative and as an anes- mechanism of action.47 However, unlike propofol, sevoflu- thetic since the late 1800s.3,121 Today, it is most often used rane also shows a small coherent theta oscillation,47 which as an anesthetic adjunct because, unlike the ether anesthet- may reflect one of its non-GABAergic mechanisms. ics, nitrous oxide is not sufficiently potent by itself to pro- duce general anesthesia. Attempts to administer sufficient The Electroencephalogram Signatures of Sevoflurane, doses of only nitrous oxide to produce unconsciousness pre- Isoflurane, and Desflurane Are Alpha, Slow-delta, and dictably result in nausea and vomiting.41 Unlike the ether Theta Oscillations anesthetics, administering nitrous oxide with oxygen is not At sub-MAC concentrations, sevoflurane shows strong commonly thought to produce slow and alpha oscillations. alpha and slow-delta oscillations (fig. 12, A–D) that closely Instead, nitrous oxide is associated with prominent beta and resemble those of propofol (fig. 6). As the concentration of gamma40 oscillations and, possibly, with a relative decrease in sevoflurane is increased to MAC levels and above, a strong power in the slow and delta oscillation bands.41 theta oscillation appears creating a distinctive pattern of A common practice at our institution is to use an inhaled evenly distributed power from the slow oscillation range up ether anesthetic with oxygen for maintenance of general anes- through the alpha range (fig. 12, A and C). This approxi- thesia and to switch to a high concentration of nitrous oxide mately equal power from the slow oscillation range through with oxygen with high total flow rates near the end of the case to the alpha range (fig. 12, A and C) is typical during main- to hasten the patient’s emergence. In this situation, a distinc- tenance with sevoflurane at or above MAC. The theta oscil- tive electroencephalogram pattern appears with the transi- lation power appears to fill in between the slow-delta and tion to nitrous oxide (illustrated in fig. 13). In anticipation alpha oscillation power. As the concentration of sevoflurane of emergence, a 34-yr-old patient was maintained on 0.5% is decreased, the theta oscillations dissipate first. This is evi- isoflurane and 58% oxygen during the last minutes of a lapa- dent in figure 12A as the theta oscillations dissipate when the roscopic cholecystectomy. At minute 82, the anesthetic was concentration of sevoflurane is lowered. On emergence, as in switched to 0.2% isoflurane, 75% nitrous oxide, and 24% the case of propofol (fig. 6A, minute 27, and fig. 6B, minute oxygen. With the switch, the total flow rate was increased 25), the alpha oscillations transition to lower amplitude beta from 3 to 7 l/min. Two minutes after the switch, the power and gamma oscillations (fig. 12A, minute 180). At the same in the slow delta and alpha bands begins to decline (fig.

Anesthesiology 2015; 123:937-60 953 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. Electroencephalography for Anesthesiologists Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

Fig. 12. Spectrograms and time domain electroencephalogram signatures of sevoflurane, isoflurane, and desflurane at surgi- cal levels of unconsciousness. The inspired concentration of the anesthetics is the blue trace in the upper part of each panel. Green arrows below each panel are propofol bolus doses. (A) At sub-minimal alveolar concentrations (MACs) (minutes 40 to 60), the spectrogram of sevoflurane resembles that of propofol (fig. 6, A and B). As the concentration of sevoflurane is increased (minutes 100 to 120), theta (5 to 7Hz) oscillations appear. The theta oscillations dissipate when the sevoflurane concentration (blue curve) is decreased. (B) Ten-second electroencephalogram trace of sevoflurane recorded at minute 40 of the spectro- gram in A. (C) The spectrogram of sevoflurane shows constant alpha, slow, delta and theta oscillations at a constant con- centration of 3%. (D) Ten-second electroencephalogram trace of sevoflurane recorded at minute 30 of the spectrogram in C. (E) At sub-MAC concentrations (minutes 16 to 26), the spectrogram of isoflurane resembles that of propofol (fig. 6, A and B) and sub-MAC sevoflurane (A). Theta oscillations strengthen as the isoflurane concentration increases toward MAC. (F) Ten-second electroencephalogram trace of isoflurane recorded at minute 40 of the spectrogram in E. (G) At the sub-MAC concentrations shown here, the spectrogram of desflurane resembles propofol with very low theta oscillation power. (H) Ten-second electroen- cephalogram trace of isoflurane recorded at minute 40 of the spectrogram in G. A, C, E, and G were adapted, with permission, from Purdon and Brown, Clinical Electroencephalography for the Anesthesiologist (2014), from the Partners Healthcare Office of Continuing Professional Development.69 Adaptations are themselves works protected by copyright. In order to publish this adaptation, authorization has been obtained both from the owner of the copyright of the original work and from the owner of copyright of the translation or adaptation.

13A, minute 84). Four minutes after the switch, a profound 5 Hz for approximately 4 min (fig. 13, minutes 86 to 90, slow-delta oscillation appears in the electroencephalogram blue area in the spectrogram) before they evolve to the beta- (fig. 13A, minute 86, and fig. 13B, minute 86). These slow- gamma oscillations that are more commonly associated with delta oscillations dominate with near loss of all power above nitrous oxide (fig. 13A, minute 90, and fig. 13B, minute 90).

Anesthesiology 2015; 123:937-60 954 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. EDUCATION Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

Fig. 13. Slow-delta and beta-gamma oscillations associated with nitrous oxide. (A) Prior to emergence, a patient was main- tained on 0.5% isoflurane and 58% oxygen. At minute 82, the composition of the anesthetic gases was changed to 0.2% isoflurane (blue curve) in 75% nitrous oxide (green curve) and 24% oxygen. The total gas flow was increased from 3 to 7 l/min. The alpha, theta, and slow oscillation power decreased from minutes 83 to 85. At minute 86, the power in the theta to beta bands decreased considerably (blue area) as the slow-delta oscillation power increased. At minute 89, the slow-delta oscillation power decreased and the beta-gamma oscillations appeared at minute 90. The flow rates and anesthetic concentrations were maintained constant between minutes 82 and 91. Isoflurane was turned off at minute 91. (B) Ten-second electroencephalogram traces of the slow-delta oscillation at minute 86.7 and the beta-gamma oscillations at minute 90.8. A and B were adapted, with permission, from Purdon and Brown, Clinical Electroencephalography for the Anesthesiologist (2014), from the Partners Health- care Office of Continuing Professional Development.69 Adaptations are themselves works protected by copyright. In order to publish this adaptation, authorization has been obtained both from the owner of the copyright of the original work and from the owner of copyright of the translation or adaptation.

The appearance of the beta-gamma oscillations and the loss of median pontine reticular formation to the basal forebrain the slow-delta oscillations take place from minutes 90 to 94. and to the thalamus.124,125 The beta-gamma oscillations The patient was extubated at minute 110 (not shown). may have a mechanism similar to that of ketamine (fig. 9B) The appearance of slow-delta oscillations we commonly whose increased beta-gamma activity depends critically on observe in the switch from one of the ether anesthetics, halo- NMDA-mediated inactivation of inhibitory interneurons.45 thane or propofol, to a high concentration (> 70%) nitrous oxide has been previously reported.122–124 The slow-delta Discussion 122–124 oscillations tend to be transient, and the beta-gamma Anesthesiologists administer selected combinations of drugs oscillations usually reported with nitrous oxide40,41 appear as to create the anesthetic state best suited for the patient and the slow-delta oscillations dissipate. These slow-delta oscil- the given surgical or diagnostic procedure. A growing body of lations are noticeably different from the slow waves seen evidence suggests that the behavioral effects of the anesthet- during propofol induction (fig. 6) and during deep dexme- ics are due to neural oscillations induced by their actions at detomidine sedation (fig. 11D) in that they are accompanied specific molecular targets in specific neural circuits. Anesthe- by a substantial reduction in spectral power at all frequencies sia-induced oscillations are 5 to 20 times larger than normal above 10 Hz (fig. 13A, minutes 86 to 90). Although the brain oscillations, likely disrupt normal brain communication mechanism of this slow oscillation is unknown, one possible (fig. 1) and are readily visible in the unprocessed electroenceph- explanation is the blockade of NMDA receptor–mediated alogram (fig. 2) and its spectrogram (fig. 3). Therefore, for three excitatory inputs from the parabrachial nucleus and the widely used intravenous anesthetics—propofol, ketamine, and

Anesthesiology 2015; 123:937-60 955 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. Electroencephalography for Anesthesiologists Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021

Fig. 14. Different anesthetics (propofol, sevoflurane, ketamine, and dexmedetomidine), different electroencephalogram sig- natures, and different molecular and neural circuit mechanisms. (A) Anesthetic-specific differences in the electroencephalo- gram are difficult to discern in unprocessed electroencephalogram waveforms. (B) In the spectrogram, it is clear that different anesthetics produce different electroencephalogram signatures. The dynamics the electroencephalogram signatures can be related to the molecular targets and the neural circuits at which the anesthetics act to create altered states of arousal. Propofol and sevoflurane enhanceγ -aminobutyric acid (GABA)ergic inhibition, sevoflurane binds at GABA receptors and other molecu- lar targets, ketamine blocks N-methyl-d-aspartate (NMDA) glutamate receptors, and dexmedetomidine is a presynaptic alpha adrenergic agonist. A and B were adapted, with permission, from Purdon and Brown, Clinical Electroencephalography for the Anesthesiologist (2014), from the Partners Healthcare Office of Continuing Professional Development.69 Adaptations are them- selves works protected by copyright. In order to publish this adaptation, authorization has been obtained both from the owner of the copyright of the original work and from the owner of copyright of the translation or adaptation.

dexmedetomidine—we related the electroencephalogram sig- to reduce or not use certain anesthetics, for example, not to use natures to the molecular targets and the neural circuits in the ether anesthetics during spinal cord monitoring, to facilitate brain at which these drugs most likely act (figs. 4–11). For the the detection of clinically significant changes in unprocessed inhaled ether-derived anesthetics such as sevoflurane, isoflu- electroencephalogram patterns.120, 130 Hence, the neurologist’s rane, and desflurane, we observed that, with the exception of paradigm does not require understanding the electroencepha- the theta oscillations that appear around 1 MAC and beyond, logram signatures of the anesthetics. their electroencephalogram patterns during maintenance and Our paradigm synthesizes research from the last 80 yr. emergence closely resemble those seen in propofol (fig. 12). Gibbs et al.1 showed that under general anesthesia, electro- Nitrous oxide is known to be associated with increased beta encephalogram patterns changed with level of unconscious- and gamma oscillations and likely decreased slow-delta oscilla- ness. Several investigators showed that different anesthetics tions. However, we demonstrated that nitrous oxide also pro- have different electroencephalogram patterns and that these duces profound slow-delta oscillations during the transition patterns were visible in the spectra of the electroencepha- from an inhaled ether anesthetic (fig. 13). logram.10,12–14,22 In 1959, Martin et al.6 wrote, “The big In contrast to brain state monitoring based on the electro- question, of course, remains unanswered; namely, do the encephalogram-derived indices, which assumes that the same electroencephalographic patterns defined in the several clas- index value defines for any anesthetic the same level of uncon- sifications presented constitute a valid measure of depth of sciousness, the unprocessed electroencephalogram and the anesthesia. We believe that this question unanswerable until spectrogram define a broader range of brain states. Therefore, a precise definition of ‘depth of anesthesia’ is made.” Relating under our paradigm, their use should enable a more nuanced the electroencephalogram patterns to the mechanisms was not characterization of brain states guided by mechanistic insights. possible at the time because lipid solubility was the primary Our paradigm for brain state monitoring differs also from that theory of anesthetic mechanisms. The concept that anesthet- used by neurologists and clinical neurophysiologists. These ics bind to specific molecular targets was not promulgated clinicians commonly use unprocessed electroencephalogram until 1984,131 and the detailed neural circuit descriptions patterns to identify seizures in the intensive care unit,67,126 to of how anesthetic actions at specific molecular targets could characterize evoked potentials,127 to detect ischemia during lead to altered states of arousal were not reported until more neurophysiological monitoring in the operating room,128 and than 25 yr later.45,46 In recent years, more has been learned to characterize sleep stages.129 Their paradigm relies on identify- about the biophysics of the electroencephalogram53 and about ing patterns in the unprocessed electroencephalogram to diag- how altered states of arousal induced by anesthetics relate to nose seizures or ischemia irrespective of the anesthetic. To do electroencephalogram activity.19,20,43,47,57,60 Hence, it is now so, clinical neurophysiologists frequently ask anesthesiologists possible to link electroencephalogram signatures to anesthetic

Anesthesiology 2015; 123:937-60 956 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. EDUCATION

state and the actions of the drugs at specific molecular targets 3. Faulconer A, Pender JW, Bickford RG: The influence of par- tial pressure of nitrous oxide on the depth of anesthesia and and in specific neural circuits (fig. 14). the electro-encephalogram in man. Anesthesiology 1949; Today, the unprocessed electroencephalogram and the spec- 10:601–9 trogram are displayed on several brain monitors.31,50,132 Accurate 4. Kiersey DK, Bickford RG, Faulconer A Jr: Electro- spectral analyses are required to track accurately the anesthetic encephalographic patterns produced by thiopental sodium during surgical operations; description and classification. Br effects. For this reason, we computed our spectrograms using J Anaesth 1951; 23:141–52 multitaper spectral methods. For a given data length, multita- 5. Galla SJ, Rocco AG, Vandam LD: Evaluation of the traditional per methods have been shown to be the optimal nonparametric signs and stages of anesthesia: An electroencephalographic spectral techniques in the sense of giving the spectral estimates and clinical study. Anesthesiology 1958; 19:328–38 64,65,133 6. Martin JT, Faulconer A Jr, Bickford RG: Electroencephalography with the highest resolution and the lowest variance. As in anesthesiology. Anesthesiology 1959; 20:359–76 a result, the multitaper methods make it easier to identify the 7. Faulconer A Jr: Correlation of concentrations of ether in arte- Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021 spectral features of anesthetic-specific signatures. rial blood with electro-encephalographic patterns occurring For the information we have covered in part I to be useful during ether-oxygen and during nitrous oxide, oxygen and ether anesthesia of human surgical patients. Anesthesiology in the management of patients receiving general anesthesia 1952; 13:361–9 and sedation, it is important to describe how the electroen- 8. Bart AJ, Homi J, Linde HW: Changes in power spectra of cephalogram patterns change as different drugs are combined electroencephalograms during anesthesia with fluroxene, because this is the more common scenario in anesthesiol- and ethrane. Anesth Analg 1971; 50:53–63 9. Findeiss JC, Kien GA, Huse KO, Linde HW: Power spec- ogy practice. The effects of combinations of anesthetics on tral density of the electroencephalogram during halothane the electroencephalogram will be the topic of part II. An and cyclopropane anesthesia in man. Anesth Analg 1969; animated version of portions of parts I and II are available 48:1018–23 at www.AnesthesiaEEG.com. In part III, we will review the 10. Bickford RG, Fleming N, Billinger T: Compression of EEG data. Trans Am Neurol Assoc 1971; 96:118–22 neurological examination for anesthesiologists. 11. Myers RR, Stockard JJ, Fleming NI, France CJ, Bickford RG: The use of on-line telephonic computer analysis of the E.E.G. Acknowledgments in anaesthesia. Br J Anaesth 1973; 45:664–70 12. Fleming RA, Smith NT: An inexpensive device for analyzing Supported by grants DP1-OD003646 (to Dr. Brown), and monitoring the electroencephalogram. Anesthesiology DP2-OD006454 (to Dr. Purdon), and TR01-GM104948 (to 1979; 50:456–60 Dr. Brown) from the National Institutes of Health, Bethesda, 13. Levy WJ, Shapiro HM, Maruchak G, Meathe E: Automated Maryland, and funds from the Department of Anesthesia, EEG processing for intraoperative monitoring: A comparison Critical Care, and Pain Medicine, Massachusetts General of techniques. Anesthesiology 1980; 53:223–36 Hospital, Boston, Massachusetts. 14. Levy WJ: Intraoperative EEG patterns: Implications for EEG monitoring. Anesthesiology 1984; 60:430–4 15. Tinker JH, Sharbrough FW, Michenfelder JD: Anterior shift Competing Interests of the dominant EEG rhythm during anesthesia in the Java Masimo, Irvine, California, has signed an agreement with Mas- monkey: Correlation with anesthetic potency. Anesthesiology sachusetts General Hospital, Boston, Massachusetts, to license 1977; 46:252–9 the signal processing algorithms developed by Drs. Brown 16. John ER, Prichep LS, Kox W, Valdés-Sosa P, Bosch-Bayard and Purdon for analysis of the electroencephalogram to track J, Aubert E, Tom M, di Michele F, Gugino LD, diMichele F: the brain states of patients receiving general anesthesia and Invariant reversible QEEG effects of anesthetics. Conscious sedation for incorporation into their brain function monitors. Cogn 2001; 10:165–83 The other authors declare no competing interests. 17. Gugino LD, Chabot RJ, Prichep LS, John ER, Formanek V, Aglio LS: Quantitative EEG changes associated with loss and return of consciousness in healthy adult volunteers anaesthetized Correspondence with propofol or sevoflurane. Br J Anaesth 2001; 87:421–8 Address correspondence to Dr. Brown or Dr. Purdon: 18. Feshchenko VA, Veselis RA, Reinsel RA: Propofol-induced alpha rhythm. Neuropsychobiology 2004; 50:257–66 ­Department of Anesthesia, Critical Care, and Pain ­Medicine, Massachusetts General Hospital, 55 Fruit Street, GRB-444, 19. Cimenser A, Purdon PL, Pierce ET, Walsh JL, Salazar-Gomez AF, Harrell PG, Tavares-Stoeckel C, Habeeb K, Brown EN: ­Boston, Massachusetts 02114. [email protected]; Tracking brain states under general anesthesia by using [email protected]. Information on purchasing global coherence analysis. Proc Natl Acad Sci U S A 2011; reprints may be found at www.anesthesiology.org or on the 108:8832–7 masthead page at the ­beginning of this issue. Anesthesiology’s 20. Purdon PL, Pierce ET, Mukamel EA, Prerau MJ, Walsh articles are made freely­ accessible to all readers, for personal JL, Wong KF, Salazar-Gomez AF, Harrell PG, Sampson AL, use only, 6 months from the cover date of the issue. Cimenser A, Ching S, Kopell NJ, Tavares-Stoeckel C, Habeeb K, Merhar R, Brown EN: Electroencephalogram signatures of loss and recovery of consciousness from propofol. Proc Natl References Acad Sci U S A 2013; 110:E1142–51 1. Gibbs FA, Gibbs LE, Lennox WG: Effects on the electroen- 21. Schneider G, Gelb AW, Schmeller B, Tschakert R, Kochs E: cephalogram of certain drugs which influence nervous activ- Detection of awareness in surgical patients with EEG-based ity. Arch Intern Med 1937; 60:154–66 indices— and patient state index. Br J 2. Faulconer A, Bickford RG: Electroencephalography in anes- Anaesth 2003; 91:329–35 thesiology, American Lecture Series. Springfield, Thomas, 22. Rampil IJ: A primer for EEG signal processing in anesthesia. 1960, pp 28–50 Anesthesiology 1998; 89:980–1002

Anesthesiology 2015; 123:937-60 957 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. Electroencephalography for Anesthesiologists

23. Palanca BJ, Mashour GA, Avidan MS: Processed electroen- 42. Foster BL, Liley DT: Effects of nitrous oxide sedation on rest- cephalogram in depth of anesthesia monitoring. Curr Opin ing electroencephalogram topography. Clin Neurophysiol Anaesthesiol 2009; 22:553–9 2013; 124:417–23 24. Bol CJJG, Danhof M, Stanski DR, Mandema JW: Pharmacokinetic- 43. Akeju O, Pavone KJ, Westover MB, Vazquez R, Prerau MJ, pharmacodynamic characterization of the cardiovascular, hyp- Harrell PG, Hartnack KE, Rhee J, Sampson AL, Habeeb K, notic, EEG and ventilatory responses to dexmedetomidine in Gao L, Lei G, Pierce ET, Walsh JL, Brown EN, Purdon PL: the rat. J Pharmacol Exp Ther 1997; 283:1051–8 A comparison of propofol- and dexmedetomidine-induced 25. Barnard JP, Bennett C, Voss LJ, Sleigh JW: Can anaesthe- electroencephalogram dynamics using spectral and coher- tists be taught to interpret the effects of ence analysis. Anesthesiology 2014; 121:978–89 on the electroencephalogram? Comparison of performance 44. Huupponen E, Maksimow A, Lapinlampi P, Särkelä M, with the BIS and spectral entropy. Br J Anaesth 2007; Saastamoinen A, Snapir A, Scheinin H, Scheinin M, Meriläinen 99:532–7 P, Himanen SL, Jääskeläinen S: Electroencephalogram spin- 26. Kearse LA Jr, Manberg P, Chamoun N, deBros F, Zaslavsky dle activity during dexmedetomidine sedation and physi- A: Bispectral analysis of the electroencephalogram cor- ological sleep. Acta Anaesthesiol Scand 2008; 52:289–94 Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021 relates with patient movement to skin incision during 45. Brown EN, Purdon PL, Van Dort CJ: General anesthesia and propofol/nitrous oxide anesthesia. Anesthesiology 1994; altered states of arousal: A systems neuroscience analysis. 81:1365–70 Annu Rev Neurosci 2011; 34:601–28 27. Glass PS, Bloom M, Kearse L, Rosow C, Sebel P, Manberg P: 46. Brown EN, Lydic R, Schiff ND: General anesthesia, sleep, and Bispectral analysis measures sedation and effects coma. N Engl J Med 2010; 363:2638–50 of propofol, , isoflurane, and alfentanil in healthy 47. Akeju O, Westover MB, Pavone KJ, Sampson AL, Hartnack volunteers. Anesthesiology 1997; 86:836–47 KE, Brown EN, Purdon PL: Effects of sevoflurane and propo- 28. Prichep LS, Gugino LD, John ER, Chabot RJ, Howard B, fol on frontal electroencephalogram power and coherence. Merkin H, Tom ML, Wolter S, Rausch L, Kox WJ: The Patient Anesthesiology 2014; 121:990–8 State Index as an indicator of the level of hypnosis under 48. Kertai MD, Whitlock EL, Avidan MS: Brain monitoring with general anaesthesia. Br J Anaesth 2004; 92:393–9 electroencephalography and the electroencephalogram- 29. Drover D, Ortega HR: Patient state index. Best Pract Res Clin derived bispectral index during . Anesth Anaesthesiol 2006; 20:121–8 Analg 2012; 114:533–46 30. Schultz B, Kreuer S, Wilhelm W, Grouven U, Schultz A: [The 49. Bennett C, Voss LJ, Barnard JP, Sleigh JW: Practical use of the Narcotrend monitor. Development and interpretation algo- raw electroencephalogram waveform during general anes- rithms]. Anaesthesist 2003; 52:1143–8 thesia: The art and science. Anesth Analg 2009; 109:539–50 31. Kreuer S, Wilhelm W: The Narcotrend monitor. Best Pract 50. Drover DR, Schmiesing C, Buchin AF, Ortega HR, Tanner Res Clin Anaesthesiol 2006; 20:111–9 JW, Atkins JH, Macario A: Titration of sevoflurane in elderly 32. Viertiö-Oja H, Maja V, Särkelä M, Talja P, Tenkanen N, patients: Blinded, randomized clinical trial, in non-cardiac Tolvanen-Laakso H, Paloheimo M, Vakkuri A, Yli-Hankala surgery after beta-adrenergic blockade. J Clin Monit Comput A, Meriläinen P: Description of the Entropy algorithm as 2011; 25:175–81 applied in the Datex-Ohmeda S/5 Entropy Module. Acta 51. Kandel ER, Schwartz JH, Jessell TM, Siegelbaum SA, Hudspeth Anaesthesiol Scand 2004; 48:154–61 AJ: Nerve cells and behavior, Principles of Neural Science, 5th 33. Jäntti V, Alahuhta S: Spectral entropy—What has it to do edition. New York, McGraw-Hill, 2013, pp 19–35 with anaesthesia, and the EEG? Br J Anaesth 2004; 93:150–1; 52. Buzsáki G: Synchronization by oscillation, Rhythms of the author reply 151–2 Brain. Oxford; New York, Oxford University Press, 2006, pp 34. Avidan MS, Zhang L, Burnside BA, Finkel KJ, Searleman AC, 136–174 Selvidge JA, Saager L, Turner MS, Rao S, Bottros M, Hantler C, 53. Buzsáki G, Anastassiou CA, Koch C: The origin of extracellu- Jacobsohn E, Evers AS: Anesthesia awareness and the bispec- lar fields and currents—EEG, ECoG, LFP and spikes. Nat Rev tral index. N Engl J Med 2008; 358:1097–108 Neurosci 2012; 13:407–20 35. Avidan MS, Jacobsohn E, Glick D, Burnside BA, Zhang L, 54. hughes SW, Crunelli V: Thalamic mechanisms of EEG alpha Villafranca A, Karl L, Kamal S, Torres B, O’Connor M, Evers rhythms and their pathological implications. Neuroscientist AS, Gradwohl S, Lin N, Palanca BJ, Mashour GA, Group 2005; 11:357–72 B-RR: Prevention of intraoperative awareness in a high-risk 55. Purves D, Augustine GJ, Fitzpatrick D, Hall WC, LaMantis surgical population. N Engl J Med 2011; 365:591–600 AS, McNamara JO, White LE: Neuroscience, 4th edition. 36. Samarkandi AH: The bispectral index system in — Sunderland, Sinauer Associates, Inc., 2008, pp 715–17 Is it related to the end-tidal concentration of inhalation anes- 56. hamalainen M, Hari R, Ilmoniemi RJ, Knuutila J, Lounasmaa thetics? Middle East J Anaesthesiol 2006; 18:769–78 OV: Magnetoencephalography—Theory, instrumentation, 37. Tirel O, Wodey E, Harris R, Bansard JY, Ecoffey C, Senhadji and applications to noninvasive studies of the working L: Variation of bispectral index under TIVA with propofol in human brain. Reviews of Modern Physics 1993; 65:413–497 a paediatric population. Br J Anaesth 2008; 100:82–7 57. Ching S, Cimenser A, Purdon PL, Brown EN, Kopell NJ: 38. hayashi K, Tsuda N, Sawa T, Hagihira S: Ketamine increases Thalamocortical model for a propofol-induced alpha-rhythm the frequency of electroencephalographic bicoherence associated with loss of consciousness. Proc Natl Acad Sci U S peak on the alpha spindle area induced with propofol. Br J A 2010; 107:22665–70 Anaesth 2007; 99:389–95 58. Supp GG, Siegel M, Hipp JF, Engel AK: Cortical hypersyn- 39. Tsuda N, Hayashi K, Hagihira S, Sawa T: Ketamine, an NMDA- chrony predicts breakdown of sensory processing during antagonist, increases the oscillatory frequencies of alpha- loss of consciousness. Curr Biol 2011; 21:1988–93 peaks on the electroencephalographic power spectrum. Acta 59. Chauvette S, Crochet S, Volgushev M, Timofeev I: Properties Anaesthesiol Scand 2007; 51:472–81 of slow oscillation during slow-wave sleep and anesthesia in 40. yamamura T, Fukuda M, Takeya H, Goto Y, Furukawa K: Fast cats. J Neurosci 2011; 31:14998–5008 oscillatory EEG activity induced by analgesic concentrations 60. Lewis LD, Weiner VS, Mukamel EA, Donoghue JA, Eskandar of nitrous oxide in man. Anesth Analg 1981; 60:283–8 EN, Madsen JR, Anderson WS, Hochberg LR, Cash SS, Brown 41. Foster BL, Liley DT: Nitrous oxide paradoxically modulates EN, Purdon PL: Rapid fragmentation of neuronal networks slow electroencephalogram oscillations: Implications for at the onset of propofol-induced unconsciousness. Proc Natl anesthesia monitoring. Anesth Analg 2011; 113:758–65 Acad Sci U S A 2012; 109:E3377–86

Anesthesiology 2015; 123:937-60 958 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. EDUCATION

61. Li D, Voss LJ, Sleigh JW, Li X: Effects of volatile anesthetic 80. Lee U, Ku S, Noh G, Baek S, Choi B, Mashour GA: Disruption agents on cerebral cortical synchronization in sheep. of frontal-parietal communication by ketamine, propofol, Anesthesiology 2013; 119:81–8 and sevoflurane. nestA hesiology 2013; 118:1264–75 62. Wang K, Steyn-Ross ML, Steyn-Ross DA, Wilson MT, Sleigh 81. Imas OA, Ropella KM, Ward BD, Wood JD, Hudetz AG: JW: EEG slow-wave coherence changes in propofol-induced Volatile anesthetics disrupt frontal-posterior recurrent infor- general anesthesia: Experiment and theory. Front Syst mation transfer at gamma frequencies in rat. Neurosci Lett Neurosci 2014; 8:215 2005; 387:145–50 63. Vizuete JA, Pillay S, Ropella KM, Hudetz AG: Graded defrag- 82. Jordan D, Ilg R, Riedl V, Schorer A, Grimberg S, Neufang S, mentation of cortical neuronal firing during recovery of con- Omerovic A, Berger S, Untergehrer G, Preibisch C, Schulz E, sciousness in rats. Neuroscience 2014; 275:340–51 Schuster T, Schröter M, Spoormaker V, Zimmer C, Hemmer 64. Mitra PP, Bokil H: Nonparametric Quadratic Estimates, B, Wohlschläger A, Kochs EF, Schneider G: Simultaneous Observed Brain Dynamics. Edited by Mirta PP, Bokil H. New electroencephalographic and functional magnetic reso- York, Oxford University Press, 2008, pp 194–196 nance imaging indicate impaired cortical top-down process-

ing in association with anesthetic-induced unconsciousness. Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021 65. Babadi B, Brown EN: A review of multitaper spectral analy- Anesthesiology 2013; 119:1031–42 sis. IEEE Trans Biomed Eng 2014; 61:1555–64 83. Rampil IJ, Lockhart SH, Eger EI II, Yasuda N, Weiskopf RB, 66. McCarley RW: Neurobiology of REM and NREM sleep. Sleep Cahalan MK: The electroencephalographic effects of desflu- Med 2007; 8:302–30 rane in humans. Anesthesiology 1991; 74:434–9 67. Niedermeyer E: Abnormal EEG patterns: Epileptic and par- 84. Amzica F: Basic physiology of burst-suppression. Epilepsia oxysmal, Electroencephalography: Basic Principles, Clinical 2009; 50(suppl 12):38–9 Applications, and Related Fields, 5th edition. Edited by Niedermeyer E, da Silva FL. Philadelphia, Lippincott Williams 85. Ching S, Purdon PL, Vijayan S, Kopell NJ, Brown EN: & Wilkins, 2005, pp 255–80 A neurophysiological-metabolic model for burst suppres- sion. Proc Natl Acad Sci U S A 2012; 109:3095–100 68. Besch G, Liu N, Samain E, Pericard C, Boichut N, Mercier M, Chazot T, Pili-Floury S: Occurrence of and risk factors for 86. Mirsattari SM, Sharpe MD, Young GB: Treatment of refrac- electroencephalogram burst suppression during propofol- tory status epilepticus with agents remifentanil anaesthesia. Br J Anaesth 2011; 107:749–56 isoflurane and desflurane. Arch Neurol 2004; 61:1254–9 69. Purdon PL, Brown EN: Clinical Electroencephalography for 87. Stecker MM, Cheung AT, Pochettino A, Kent GP, Patterson T, Anesthesiologists, Partners Healthcare Office of Continuing Weiss SJ, Bavaria JE: Deep hypothermic circulatory arrest: Professional Development, 2014. Available at: www. I. Effects of cooling on electroencephalogram and evoked AnesthesiaEEG.com potentials. Ann Thorac Surg 2001; 71:14–21 70. Schwilden H, Schüttler J, Stoeckel H: Closed-loop feedback 88. Doyle PW, Matta BF: Burst suppression or isoelectric control of anesthesia by quantitative EEG anal- encephalogram for cerebral protection: Evidence from met- ysis in humans. Anesthesiology 1987; 67:341–7 abolic suppression studies. Br J Anaesth 1999; 83:580–4 71. Drummond JC, Brann CA, Perkins DE, Wolfe DE: A com- 89. Rossetti AO, Reichhart MD, Schaller MD, Despland PA, parison of median frequency, spectral edge frequency, a fre- Bogousslavsky J: Propofol treatment of refractory sta- quency band power ratio, total power, and dominance shift in tus epilepticus: A study of 31 episodes. Epilepsia 2004; the determination of depth of anesthesia. Acta Anaesthesiol 45:757–63 Scand 1991; 35:693–9 90. Hunter G, Young GB: Status epilepticus: A review, with 72. Hemmings HC Jr, Akabas MH, Goldstein PA, Trudell JR, emphasis on refractory cases. Can J Neurol Sci 2012; Orser BA, Harrison NL: Emerging molecular mechanisms 39:157–69 of general anesthetic action. Trends Pharmacol Sci 2005; 91. Lewis LD, Ching S, Weiner VS, Peterfreund RA, Eskandar 26:503–10 EN, Cash SS, Brown EN, Purdon PL: Local cortical dynamics 73. Bai D, Pennefather PS, MacDonald JF, Orser BA: The of burst suppression in the anaesthetized brain. Brain 2013; general anesthetic propofol slows deactivation and 136(Pt 9):2727–37 desensitization of GABA(A) receptors. J Neurosci 1999; 92. Kroeger D, Amzica F: Hypersensitivity of the anesthesia- 19:10635–46 induced comatose brain. J Neurosci 2007; 27:10597–607 74. Brown EN, Solt K, Purdon PL, Akeju O: Monitoring brain 93. Koenig MA, Kaplan PW, Thakor NV: Clinical neurophysi- states during general anesthesia and sedation, Miller's ologic monitoring and brain injury from cardiac arrest. Anesthesia, 8th edition. Edited by Miller RD, Eriksson LI, Neurol Clin 2006; 24:89–106 Fleisher LA, Wiener-Kronish JP, Cohen NH, Young WL, eds. 94. Yamatogi Y, Ohtahara S: Early-infantile epileptic enceph- Philadelphia, Elsevier; 2014, pp 1524–1540 alopathy with suppression-bursts, Ohtahara syndrome; 75. Devor M, Zalkind V: Reversible analgesia, atonia, and loss its overview referring to our 16 cases. Brain Dev 2002; of consciousness on bilateral intracerebral microinjection of 24:13–23 pentobarbital. Pain 2001; 94:101–12 95. Rampil I: Consciousness, awareness, and the clinician. Can 76. Williams ST, Conte MM, Goldfine AM, Noirhomme Q, J Anesth 2003; 50:R1–5 Gosseries O, Thonnard M, Beattie B, Hersh J, Katz DI, Victor 96. Rampil IJ, Weiskopf RB, Brown JG, Eger EI II, Johnson BH, JD, Laureys S, Schiff ND: Common resting brain dynam- Holmes MA, Donegan JH: I653 and isoflurane produce sim- ics indicate a possible mechanism underlying zolpidem ilar dose-related changes in the electroencephalogram of response in severe brain injury. Elife 2013; 2:e01157 pigs. Anesthesiology 1988; 69:298–302 77. McCarthy MM, Brown EN, Kopell N: Potential network mech- 97. Bruhn J, Bouillon TW, Shafer SL: Bispectral index (BIS) and anisms mediating electroencephalographic beta rhythm burst suppression: Revealing a part of the BIS algorithm. changes during propofol-induced paradoxical excitation. J Clin Monit Comput 2000; 16:593–6 J Neurosci 2008; 28:13488–504 98. Voss L, Sleigh J: Monitoring consciousness: The current sta- 78. Feldman JL, Del Negro CA: Looking for inspiration: New tus of EEG-based depth of anaesthesia monitors. Best Pract perspectives on respiratory rhythm. Nat Rev Neurosci 2006; Res Clin Anaesthesiol 2007; 21:313–25 7:232–42 99 Chemali J, Ching S, Purdon PL, Solt K, Brown EN: Burst 79. Redlich FC, Callahan A, Mendelson RH: Electroencephalographic suppression probability algorithms: State-space methods changes after eye opening and visual stimulation. Yale J Biol for tracking EEG burst suppression. J Neural Eng 2013; Med 1946; 18:367–76 10:1741–2560

Anesthesiology 2015; 123:937-60 959 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited. Electroencephalography for Anesthesiologists

100. Shanechi MM, Chemali JJ, Liberman M, Solt K, Brown EN: connectivity is a neural correlate of dexmedetomidine- A brain-machine interface for control of medically-induced induced unconsciousness. Elife 2014; 4:e04499 coma. PLoS Comput Biol 2013; 9: e1003284 117. Destexhe A, Contreras D, Steriade M: Mechanisms underly- 101. Sinner B, Graf BM: Ketamine. Handb Exp Pharmacol 2008; ing the synchronizing action of corticothalamic feedback 182:313–33 through inhibition of thalamic relay cells. J Neurophysiol 102. Olney JW, Farber NB: Glutamate receptor dysfunction and 1998; 79:999–1016 schizophrenia. Arch Gen Psychiatry 1995; 52:998–1007 118. Nir Y, Staba RJ, Andrillon T, Vyazovskiy VV, Cirelli C, Fried 103. Seamans J: Losing inhibition with ketamine. Nat Chem Biol I, Tononi G: Regional slow waves and spindles in human 2008; 4:91–3 sleep. Neuron 2011; 70:153–69 104. Moghaddam B, Adams B, Verma A, Daly D: Activation of 119. Antognini JF, Schwartz K: Exaggerated anesthetic glutamatergic neurotransmission by ketamine: A novel step requirements in the preferentially anesthetized brain. in the pathway from NMDA receptor blockade to dopami- Anesthesiology 1993; 79:1244–9 nergic and cognitive disruptions associated with the pre- 120. Lyon R, Feiner J, Lieberman JA: Progressive suppression frontal cortex. J Neurosci 1997; 17:2921–7 of motor evoked potentials during general anesthesia: The Downloaded from http://pubs.asahq.org/anesthesiology/article-pdf/123/4/937/373943/20151000_0-00034.pdf by guest on 28 September 2021 105. Cavazzuti M, Porro CA, Biral GP, Benassi C, Barbieri GC: phenomenon of “anesthetic fade.” J Neurosurg Anesthesiol Ketamine effects on local cerebral blood flow and metabo- 2005; 17:13–9 lism in the rat. J Cereb Blood Flow Metab 1987; 7:806–11 121. Andrews E: The oxygen mixture, a new anesthetic combina- 106. Strebel S, Kaufmann M, Maître L, Schaefer HG: Effects of ket- tion. Chic Med Exam 1868; 9:656–61 amine on cerebral blood flow velocity in humans. Influence 122. Avramov MN, Shingu K, Mori K: Progressive changes in of pretreatment with midazolam or esmolol. Anaesthesia electroencephalographic responses to nitrous oxide in 1995; 50:223–8 humans: A possible acute drug tolerance. Anesth Analg 107. Vollenweider FX, Leenders KL, Oye I, Hell D, Angst J: 1990; 70:369–74 Differential psychopathology and patterns of cerebral glu- 123. hagihira S, Takashina M, Mori T, Mashimo T: The impact of cose utilisation produced by (S)- and ®-ketamine in healthy nitrous oxide on electroencephalographic bicoherence dur- volunteers using positron emission tomography (PET). Eur ing isoflurane anesthesia. Anesth Analg 2012; 115:572–7 Neuropsychopharmacol 1997; 7:25–38 124. Pavone KJ, Akeju O, Sampson AL, Ling K, Purdon PL, Brown 108. Jorm CM, Stamford JA: Actions of the hypnotic anaesthetic, EN: Nitrous oxide-induced slow and delta oscillations. Clin dexmedetomidine, on noradrenaline release and cell firing Neurophysiol 2015; 10.1016/j.clinph.2015.06.001. in rat locus coeruleus slices. Br J Anaesth 1993; 71:447–9 125. Jevtović-Todorović V, Todorović SM, Mennerick S, Powell S, 109. Nacif-Coelho C, Correa-Sales C, Chang LL, Maze M: Dikranian K, Benshoff N, Zorumski CF, Olney JW: Nitrous Perturbation of ion channel conductance alters the hyp- oxide (laughing gas) is an NMDA antagonist, neuroprotec- notic response to the α2-adrenergic agonist dexmedetomi- tant and neurotoxin. Nat Med 1998; 4:460–3 dine in the locus coeruleus of the rat. Anesthesiology 1994; 126. Moura LM, Shafi MM, Ng M, Pati S, Cash SS, Cole AJ, Hoch DB, 81:1527–34 Rosenthal ES, Westover MB: Spectrogram screening of adult 110. Nelson LE, Lu J, Guo T, Saper CB, Franks NP, Maze M: The EEGs is sensitive and efficient. 2014; 83:56–64 α2-adrenoceptor agonist dexmedetomidine converges on 127. Hansen H, Classen J: EEG and evoked potentials in neu- an endogenous sleep-promoting pathway to exert its seda- roanesthesia, intraoperative neurological monitoring, tive effects. Anesthesiology 2003; 98:428–36 and neurointensive care, Electroencephalography: Basic 111. Nelson LE, Guo TZ, Lu J, Saper CB, Franks NP, Maze M: The Principles, Clinical Applications, and Related Fields, 5th sedative component of anesthesia is mediated by GABA(A) edition. Edited by Niedermeyer E, da Silva FL. Philadelphia, receptors in an endogenous sleep pathway. Nat Neurosci Lippincott Williams & Wilkins, 2005, pp 1137–64 2002; 5:979–84 128. Blume WT, Sharbrough FW: EEG monitoring during carotid end- 112. Saper CB, Scammell TE, Lu J: Hypothalamic regulation of arterectomy and open heart surgery, Electroencephalography: sleep and circadian rhythms. Nature 2005; 437:1257–63 Basic Principles, Clinical Applications, and Related Fields, 5th 113. Sherin JE, Elmquist JK, Torrealba F, Saper CB: Innervation edition. Edited by Niedermeyer E, da Silva FL. Philadelphia, of histaminergic tuberomammillary neurons by GABAergic Lippincott Williams & Wilkins, 2005, pp 815–28 and galaninergic neurons in the ventrolateral preoptic 129. Prerau MJ, Hartnack KE, Obregon-Henao G, Sampson A, nucleus of the rat. J Neurosci 1998; 18:4705–21 Merlino M, Gannon K, Bianchi MT, Ellenbogen JM, Purdon 114. Morairty S, Rainnie D, McCarley R, Greene R: Disinhibition of PL: Tracking the sleep onset process: An empirical model of ventrolateral preoptic area sleep-active neurons by adenos- behavioral and physiological dynamics. PLoS Comput Biol ine: A new mechanism for sleep promotion. Neuroscience 2014; 10:978–89 2004; 123:451–7 130. Sloan TB: Anesthetic effects on electrophysiologic record- 115. España RA, Berridge CW: Organization of noradrener- ings. J Clin Neurophysiol 1998; 15:217–26 gic efferents to arousal-related basal forebrain structures. 131. Franks NP, Lieb WR: Do general anaesthetics act by competi- J Comp Neurol 2006; 496:668–83 tive binding to specific receptors? Nature 1984; 310:599–601 116. Akeju O, Loggia ML, Catana C, Pavone KJ, Vazquez R, 132. Kelley SD: Monitoring Consciousness: Using the Bispectral Rhee J, Contreras Ramirez V, Chonde DB, Izquierdo-Garcia Index, 2nd edition. Boulder, Covidien, 2010, pp 6 D, Arabasz G, Hsu S, Habeeb K, Hooker JM, Napadow V, 133. Bronez T: On the performance advantage of multitaper spec- Brown EN, Purdon PL: Disruption of thalamic functional tral analysis. IEEE Trans Signal Process 2002; 40:2941–46

Anesthesiology 2015; 123:937-60 960 Purdon et al. Copyright © 2015, the American Society of Anesthesiologists, Inc. Wolters Kluwer Health, Inc. Unauthorized reproduction of this article is prohibited.