Top Banner
Editors: Sadock, Benjamin J.; Sadock, Virginia A. Title: Kaplan & Sadock's Comprehensive Textbook of Psychiatry, 8th Edition Copyright ©2005 Lippincott Williams & Wilkins > Table of Contents > Volume I > 1 - Neural Sciences > 1.14: Applied Electrophysiology 1.14: Applied Electrophysiology Nashaat N. Boutros M.D. Frederick A. Struve Ph.D. Part of "1 - Neural Sciences" HISTORY AND OVERVIEW Over a span of more than 70 years, the field of human scalp electroencephalography (EEG) has expanded from its beginnings as a highly controversial phenomenon to its maturity as an investigative and clinical method with wide applications in medicine and neuroscience. Originally tied to neurology and psychiatry, EEG methods have enjoyed expanded use in the study of central nervous system (CNS) effects of a variety of metabolic, endocrinological, toxic, pharmacological, and traumatic events. Recent decades have witnessed the development and refinement of topographic quantitative EEG (QEEG) methods applied to clinical and research problems, and the present era promises the technology to simultaneously record multichannel EEG and functional imaging during magnetic resonance imaging (MRI) scanning. Furthermore, the basic field of EEG has given birth to the emergence of significant sister fields of polysomnography and magnetoencephalography. Origins The lengthy transition from laboratory experiment to eventual acceptance of EEG was plagued by intense controversy. P.172 Despite the continuing accumulation of experimental evidence of brain-derived electrical potentials, beginning with Richard Caton's discovery in 1874 of spontaneous electrical activity recorded from the exposed cortex of cats, rabbits, and monkeys, the notion that electrical potentials emanated from the brain was rejected for nearly 50 years by leading authorities. Caton's work was replicated by Vasili Danilevsky's 1877 report that electrical
28

Translate

May 08, 2017

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Translate

Editors: Sadock, Benjamin J.; Sadock, Virginia A.Title: Kaplan & Sadock's Comprehensive Textbook of Psychiatry, 8th EditionCopyright ©2005 Lippincott Williams & Wilkins> Table of Contents > Volume I > 1 - Neural Sciences > 1.14: Applied Electrophysiology1.14: Applied ElectrophysiologyNashaat N. Boutros M.D.Frederick A. Struve Ph.D.Part of "1 - Neural Sciences"HISTORY AND OVERVIEWOver a span of more than 70 years, the field of human scalp electroencephalography (EEG) has expanded from its beginnings as a highly controversial phenomenon to its maturity as an investigative and clinical method with wide applications in medicine and neuroscience. Originally tied to neurology and psychiatry, EEG methods have enjoyed expanded use in the study of central nervous system (CNS) effects of a variety of metabolic, endocrinological, toxic, pharmacological, and traumatic events. Recent decades have witnessed the development and refinement of topographic quantitative EEG (QEEG) methods applied to clinical and research problems, and the present era promises the technology to simultaneously record multichannel EEG and functional imaging during magnetic resonance imaging (MRI) scanning. Furthermore, the basic field of EEG has given birth to the emergence of significant sister fields of polysomnography and magnetoencephalography.OriginsThe lengthy transition from laboratory experiment to eventual acceptance of EEG was plagued by intense controversy. P.172

Despite the continuing accumulation of experimental evidence of brain-derived electrical potentials, beginning with Richard Caton's discovery in 1874 of spontaneous electrical activity recorded from the exposed cortex of cats, rabbits, and monkeys, the notion that electrical potentials emanated from the brain was rejected for nearly 50 years by leading authorities. Caton's work was replicated by Vasili Danilevsky's 1877 report that electrical oscillations recorded from the animal brain could be altered by strong sensory stimuli, and, later, in 1883, Fleischl von Marxow demonstrated that changes in brain electrical activity produced by sensory stimulation could be abolished by chloroform. Three further historical highlights of note include (1) the 1891 demonstration by Adolph Beck that the dog visual cortex produced large electrical potentials when the eyes were rhythmically illuminated (thus laying the experimental foundation for EEG photic driving), (2) Beck and Napoleon Cybulski's 1892 report that local injury to the cortex could alter the characteristics of recorded spontaneous electrical activity, and (3) Cybulski's 1914 report that brain-wave seizure discharges could be induced in the cortex by applying electrical stimulation to the cortex (thus presaging the use of EEG in epilepsy). Despite this stream of successful experimental work, the EEG phenomenon remained largely insecure.The focused perseverance of Hans Berger, a biologically oriented Professor of Psychiatry and Director of the Psychiatric Clinic in Jena, Germany, finally brought EEG to a position of acceptance and clinical usefulness. After years of unsuccessful attempts to record brain waves from humans (he was able to obtain recordings from animals), he finally succeeded in recording the human EEG in 1924, and, in 1929, he published the first in his classic series of 23 papers describing many aspects of the human EEG. Among his vast achievements, he demonstrated that

Page 2: Translate

brain electrical activity came from neurons and not blood vessels or connective tissue, that recordings from patients with brain tumors contained high-voltage slow waves (his recording technique did not permit localization), that waking alpha waves were blocked by eye opening, and that the characteristics of EEG activity change with age, sensory stimulation, state of consciousness, and physiochemical state of the body. He coined the word electroencephalogram. However, acceptance was still temporarily delayed when Lord Adrian, a Nobel laureate neurophysiologist, claimed that Berger's findings “were impossible.” Later, in 1934, Adrian publicly confirmed Berger's work, and the field of EEG was born.Epilepsy and Classical NeurologyDespite the fact that EEG originated in psychiatry, the strongest initial impetus for its use came from neurology, particularly the study of epilepsy. The years from 1934 to 1940 saw a marked proliferation of EEG studies focused on structural brain lesions and a variety of seizure disorders. In 1934, the team led by Fred Gibbs discovered the classic three-per-second spike-and-wave complex, which proved to be specific to petit mal absence attacks. Before the decade ended, they had described EEG patterns associated with grand mal and myoclonic seizures and a diffuse spike-and-wave pattern that was slower in frequency than petit mal (and given the confusing name of petit mal variant) and that was associated with grand mal seizures and a high incidence of mental retardation. They also introduced the term psychomotor seizures (now complex partial seizures [CPSs]) and described the EEG manifestations characterizing an ictal psychomotor attack. Later, in 1947 and 1948, they described the anterior temporal spike focus that became the interictal EEG correlate of this disorder. The other side of the neurological coin, structural brain lesions, was also advanced through landmark, new EEG discoveries during this early decade. In 1935, O. Foerster and H. Altenberger reported from Germany that focal slow waves in the EEG recording often appeared near brain tumors, and, later, Grey Walter made a major advance by demonstrating a technique for EEG localization of brain tumors. Under the leadership of Herbert Jasper in Montreal, direct cortical EEG recording began to be introduced during neurosurgery, and, by the close of the decade, in 1941, Denis Williams at Oxford began using EEG recordings to study and localize traumatic intracranial injuries received during World War II.PsychiatryStarting in approximately 1938, a flurry of continuing EEG investigations began to reveal an increase in the overall prevalence of minor abnormalities in almost all psychiatric populations as compared to healthy or nonpsychiatric controls, a finding that remains undisputed today. On the other hand, two major factors led to the rapid disillusionment of the field of psychiatry with EEG. The first was the lack of specificity of EEG abnormalities to known psychiatric syndromes. The second factor, alluded to previously, was the continuing discovery of EEG abnormalities correlating with epilepsy, tumors, encephalopathies, stroke syndromes, and coma. The fact that discoveries of significant EEG changes accompanying neurological problems were occurring while those EEG abnormalities associated with psychiatric symptomatology continued to be minimal (compared to those related to neurological disorders) and noncontributory to the diagnostic process led the field of clinical EEG (and later clinical neurophysiology) to become squarely a subspecialty of the field of neurology, with nearly a complete lack of interest in EEG among psychiatrists.The recent significant surge of interest in the neurobiology of psychiatric disorders, the emergence of the clinical field of neuropsychiatry, and the unprecedented advances in computerized analyses of EEG and other neurophysiological signals have resulted in a strong

Page 3: Translate

rekindling of interest in electrophysiology among psychiatrists. Less than a decade ago, John Hughes undertook the mammoth task of compiling a comprehensive outline of the broad area of EEG and psychiatry with 181 significant references selected for citation from before 1950 until 1994. When such compilations are inspected, the findings reveal that more than one-half of the EEG-psychiatry references appear after 1980, and one-third were written within the 5 years preceding Hughes' 1995 report. Continued inspection of the literature indicates that this trend has not abated.ELECTROENCEPHALOGRAPHYA given brain wave is no more than the transient difference in electrical potential (greatly amplified) between any two points on the scalp or between some electrode placed on the scalp and a reference electrode located elsewhere on the head (i.e., ear lobe or nose). In a simplistic sense, the EEG is nothing more than an extremely sensitive voltmeter, with the unit of measurement being the microvolt, or millionth of a volt. Typical EEG signals range from approximately 30 to 80 µV, but they can be as low as 10 µV in some tracings or as high as 150 or 200 µV in some high-voltage “spike” discharges. The difference in electrical potential measured between any two EEG electrodes fluctuates or oscillates rapidly, usually many times per second. It is this oscillation that produces the characteristic “squiggly line” that even many lay persons now recognize as the appearance of “brain waves.”The earliest EEG recordings involved only one pair of electrodes, or one channel of recording, and although this could detect certain normal and abnormal features, effective clinical application remained for the P.173

future. Soon, the breakthrough ability to record two channels of brain waves emerged, and it became possible to record activity simultaneously from homologous locations in each hemisphere. Before long, the rapid advances in recording technology allowed four- and eight-channel recordings to be made, and EEG became a viable clinical tool. Eventually, 10-, 12-, and 16-channel recording machines became the standard work horses of clinical and research EEG laboratories around the world. EEG equipment capable of simultaneous recording from 64 (or even many more) channels is available but is largely confined to special research applications. The ability to simultaneously record brain waves from many scalp locations is important, because it allows direct comparisons between homologous cortical regions, permits recording arrays to locate focal or regional abnormal features more clearly, and increases the ability to detect various artifacts (i.e., waveforms of nonbrain origin) that can contaminate the recording.Scalp EEG cannot detect the electrical activity generated by a single neuron or even by several neurons close to the scalp. Rather, the recorded EEG signals that are seen are the result of summated field potentials generated by excitatory postsynaptic potentials (EPSPs) and inhibitory postsynaptic potentials (IPSPs) in vertically oriented pyramidal cells of the cortex. An EPSP in a dendrite produces electrical negativity in the immediately surrounding area, but the electrical field becomes positive with increasing distance from the source. The reverse occurs with an IPSP, generating an electrical positivity nearby and a negative field at a distance. The summation of EPSPs and IPSPs is enhanced, because the neurons are tightly packed together and oriented vertically in parallel. In addition, large aggregates of these neurons may receive similar input, thus making it likely that they may respond in unison over time. Because of the manner in which the dominant intrinsic brain waves are generated, EEG is maximally sensitive to cortical neuronal activity and relatively insensitive to electrical potentials generated from subcortical

Page 4: Translate

regions. However, there are some minor exceptions, because subcortical neuronal events can sometimes influence cortical neuronal firing via afferent transmissions along subcortical-cortical tracts.Probably the first observation about brain waves, going back to the time of Caton, was that the recorded potentials oscillate and repeat in a rhythmic fashion. Indeed, the term intrinsic rhythms is often used for normal activity, and the term dysrhythmic is used for activity that might be abnormal. Within reasonable limits, the repetitive rhythmic nature of the EEG is stable across individuals and within individuals over time, barring the introduction of pathophysiologic events.Limitations of Scalp ElectroencephalographyEEG continues to be one of the few objective measures of brain function. However, appreciation of its strengths in clinical and research settings must also be tempered with recognition of its limitations.Because of the limitations of scalp EEG, a normal EEG can never constitute positive proof of absence of brain dysfunction. With several diseases with established brain pathophysiology, such as multiple sclerosis, deep subcortical neoplasm, some seizure disorders, and Parkinson's disease and other movement disorders, to name only a few, a substantial incidence of patients with normal EEGs may be encountered. Nonetheless, as is discussed later in the chapter, a normal EEG can often provide convincing evidence for excluding certain types of brain pathology that may present with behavioral or psychiatric symptoms.Brain Coverage and ImpedanceBecause the human brain is encased and protected in a bony skull, large areas of cortex are inaccessible to scalp EEG recording. Although approximately one-third of the outer convexity of the cortex may be within reach, much cortical area consists of mesial, inferior, and deep buried cortical tissue that is removed from the proximity of electrodes that are confined to external scalp placement. Electrical events generated in these areas may not be detected by scalp electrodes. Furthermore, substantial impedance to electrical conduction from skin, skull, dura, and brain tissue exists between the source of generated electrical potentials and the detecting electrode on the scalp. Weak electrical signals, even those close to the surface, may escape detection. It has been demonstrated that electrical potentials recorded from the cortical surface are much higher in voltage than potentials recorded simultaneously at the surface of the scalp and that depth electrode recordings often show activity that is attenuated and distorted or not visible at the scalp.Paroxysmal Discharges and Recording LengthMany types of EEG abnormality, particularly abnormalities of brain wave frequency, such as generalized or focal slowing, tend to be present from the beginning of the recording, and the recording length generally is not a limiting factor in their detection. For example, in several clinical situations, such as suspected delirium or suspected nonconvulsive status, a 10-minute wake EEG often provides the needed diagnostic information. However, other significant abnormalities, including focal and diffuse spike or spike-wave complexes and several controversial paroxysmal dysrhythmias, occur episodically against a background of more or less normal activity. In cases in which sporadic paroxysmal discharges occur frequently during a tracing, a limited recording length may not be problematic. However, paroxysmal abnormal discharges are often widely spaced, may occur only a few times in a long tracing, or may be confined to certain recording states, such as stage I or II sleep. In these cases, a short recording may fail to detect infrequent sporadic discharges and thus are falsely negative.Nonspecificity of Results

Page 5: Translate

There are only a limited number of ways in which brain electrical activity can respond to normal or pathological influences. Brain waves can only reflect change by becoming faster or slower in frequency or lower or higher in voltage, or perhaps some combination of these two responses. Thus, the same or similar abnormal EEG patterns can emerge from different etiological causes. For example, a neoplasm, subdural hematoma, brain abscess, cerebral vascular accident (CVA), closed head injury, or aneurysm may result in similar, although not always identical, focal EEG slowing. Generalized slowing is a common abnormal finding for which the etiological causes are legion and include cortical atrophy, drug-induced encephalopathy, electroconvulsive therapy (ECT), encephalitis, certain endocrine disorders (hypothyroidism and hypopituitarism), porphyria, head trauma, lead exposure, hypocalcemia, and Wernicke's encephalopathy, to name but a few. The nonspecificity of results or the failure to specifically denote etiology is a genuine limitation but one that is not as bleak as this paragraph suggests. More often than not, information from the patient's symptoms, clinical course and history, and other laboratory results identifies a probable cause for the EEG findings. Furthermore, EEGs are often ordered for specific reasons in cases in which a pathophysiological process is already suspected.RECORDINGMuch has been written about the complexities of EEG recording and interpretation and the corresponding high level of skill needed to obtain an adequate EEG. What may be insufficiently recognized is the fact that there are also clinical situations in which a greatly simplified EEG secured by a properly trained registered nurse or resident can have substantial diagnostic usefulness. There are some important EEG P.174

findings of particular relevance to emergency room settings and, possibly, even to some acute psychiatric admission or triage units that can be assessed in only 10 minutes by using only a 10- or 12-channel recording instrument by those with a minimum level of technical skill. Cases presenting with moderate to marked confusion and agitation, delirium, or possible nonconvulsive status may have diffuse EEG abnormalities that are more or less continuous in the tracing, once the recording is turned on. Such findings (if present) do not require sophisticated localization studies, and their presence, as well as their absence, is diagnostically relevant. Prompt access to an EEG laboratory may not always be possible, especially on weekends or evenings, and on-site screening may thus be helpful.Other than the circumscribed (yet potentially highly useful) screening EEGs described previously, recording the EEG does, in fact, require a considerable amount of skill and experience. It is not merely a technical act performed by a technician. The unfolding clinical EEG tracing is a constantly moving and shifting parade of complex waveforms recorded simultaneously from numerous scalp locations, and the EEG patterns differ dramatically during wakefulness, drowsiness, and various sleep levels. The appearance of the EEG also changes from one recording montage to another while a host of normal and abnormal EEG waveforms and contaminating artifacts must be identified in their obvious and subtle forms. In addition to the necessary skills of accurate electrode application and machine operation, the better technologists are also capable of sophisticated EEG interpretation. It may not be obvious how important this is. EEG abnormalities do not always emerge in clear-cut, textbook form but instead may be distorted and, hence, ambiguous. Interpretative ability is necessary to recognize probable abnormalities and then to arrange recording montages and states of patient alertness (wake or sleep) in ways that might enhance or bring out the patterns and allow a more definite

Page 6: Translate

interpretation by the electroencephalographer. A minimum of 1 year of full-time training, including didactic instruction and supervised, hands-on recording and interpretation experience, is necessary for an EEG technologist to achieve competence. Formal training schools for EEG technologists exist in many places, and the graduates can become registered EEG technologists by taking and passing a two-part written and practical examination.Electrode PlacementAs EEG emerged into the clinical arena, electrodes were simply placed on the scalp symmetrically by eye, using salient landmarks on the head as reference points, and not all laboratories used the same placement system. Eventually, in 1947, it was decided at an International EEG Congress held in London that some effort should be made to standardize the system of electrode placement, so that clinical and research findings would be more directly comparable across different laboratories. The challenge was taken up by Jasper, who developed the 10/20 International System of Electrode Placement, which has become standard worldwide since 1958. Without going into lengthy technical detail, the 10/20 system simply measures the distance between readily identifiable landmarks on the head and then locates electrode positions at 10 percent or 20 percent of that distance in an anterior-posterior or transverse direction (Fig. 1.14-1). Electrodes are then designated by an uppercase letter denoting the brain region beneath that electrode and a number, with odd numbers used for the left hemisphere and with even numbers signifying the right hemisphere (the subscript Z denotes midline electrodes). Thus, the O2 electrode is placed over the right occipital region, and the P3 lead is found over the left parietal area.

FIGURE 1.14-1 International 10/20 Electrode Placement System. (Courtesy of Grass, Astro-Med, Inc. Product Group.)Although most laboratories use 21 scalp electrodes for standard recordings, the 10/20 system provides for additional electrodes to provide greater coverage, if needed, and the American EEG Society has even developed a nomenclature for the designation of as many as 75 defined electrode locations (Fig. 1.14-2). However, it must be stressed that extremely large numbers of scalp electrodes, although no doubt impressive, are unnecessary for currently established clinical EEG applications. For currently accepted clinical indications,

Page 7: Translate

P.175

optimally useful EEG recordings can be achieved with only 21 or, at the most, 32 scalp electrodes. However, large electrode sensor arrays, including 125 or even 256 scalp leads (Fig. 1.14-3), may be needed for specialized research applications involving source analyses and three-dimensional dipole characterization in which it has been estimated that the limit at which additional unique information may be obtained is between 200 and 300 electrodes.

FIGURE 1.14-2 An expanded 75-electrode array developed by the Electrode Nomenclature Committee of the American Electroencephalography Society. The four electrode positions in black are given new names. Previous designations of T3 and T4 have been renamed (black electrodes) as T7 and T8. The T5 and T6 locations in the original placement system are now renamed (black electrodes) as P7 and P8. Such extensive placement systems are primarily used for special research studies and are only rarely used for clinical recordings.

Page 8: Translate

FIGURE 1.14-3 A 256-electrode-dense sensor array. Such large arrays are often used in research requiring optimal resolution for source analyses and three-dimensional dipole localization. (Courtesy of Electrical Geodesies, Inc.)Although tedious, the 10/20 placement system has several advantages. Because the placement system is based on rigorous measurements, electrode placement error, particularly asymmetrical placement of electrodes for homologous electrode pairs, is greatly minimized. The system also renders recordings entirely comparable between laboratories, as well as across serial tracings obtained from a single subject. Because percentages of distances between landmarks on the head are used for placement locations, scalp electrodes overlie the same cortical regions despite differences in head size. Furthermore, the relationships between electrodes placed on the scalp and underlying brain structures have been well established (Fig. 1.14-4) by using placements on cadavers (with holes drilled at the electrode sites to later identify the cortical area under the electrode), as well as recent studies using computed tomography (CT) scanning.

Page 9: Translate

FIGURE 1.14-4 A left-lateral diagram of the head showing the locations of the routine 10-20 electrodes (left-side electrode locations F7 and T3 and the new electrode placement [T1]) in relation to the temporal pole. (Modification of figure printed courtesy of Grass, Astro-Med, Inc. Product Group.)It has been suggested that, in cases of suspected temporal lobe abnormality unconfirmed by traditional electrodes, a closer examination of the temporal area should be attempted, because the anterior temporal lobe is not well covered by the standard 10/20 placement system. The F7 and F8 electrodes are over the posterior-inferior-frontal lobe and, hence, forward of the temporal pole, whereas the T3 and T4 electrodes are behind the anterior temporal region. Some laboratories now add new electrodes (T1 and T2) or simply relocate the F7 and F8 electrodes to this new position. The placement of the T1 and T2 anterior temporal electrodes is based on the distance from the lateral canthus of the eye to the external auditory canal, with electrodes placed at one-third of this total distance anterior to the auditory canal and 1 cm up from a line connecting these two landmarks (Fig. 1.14-4). However, the F7 and F8 electrodes may detect potentials spreading from the anterior temporal cortex, particularly if the voltages of the discharges are high.Scalp electrodes must be applied carefully. The skin under the electrode must be clean and completely free of oil or grease. A common practice is to rub the area with a slightly abrasive cleansing electrolyte material that also removes some of the superficial epidermis. When this is done, a metal disc electrode can be applied to the scalp by using a conducting electrode paste. Electrode impedance should be maintained at equal to or less than 3,000 Ω. The whole electrode

Page 10: Translate

application procedure should not be uncomfortable for the subject. In some laboratories, subdermal needle electrodes are used. Although they can reduce the electrode application time, they have the significant disadvantage of coming loose spontaneously or when the head is moved. They also involve a concern about transmission of disease, especially human immunodeficiency virus (HIV).Special ElectrodesNasopharyngeal (NP) electrodes can be inserted into the NP space through the nostrils and can be closer to the temporal lobe than scalp electrodes (Fig. 1.14-1; these leads are designated Pg1 and Pg2 in the 10/20 placement system). No actual penetration of tissue occurs. The NP lead is a long (as long as 15 cm for adults), curved S- or Z-shaped insulated wire with a silver ball (the electrode) on the tip, which is inserted in the nostril and then rotated laterally, so that P.176

the ball is in contact with the roof of the nasopharynx. With a cooperative patient and a skilled technologist, the procedure can be well tolerated. Although this lead is presumed to be better positioned to detect activity from the orbitofrontal cortex, temporal pole, and hippocampus, it has numerous disadvantages. Chief among these are a high propensity to produce pulse and respiration artifact and the fact that NP leads cannot be used when a deviated septum or nasal inflammatory process is present. They are also contraindicated with many psychiatric patients displaying behaviors, such as confusion, agitation, or belligerence, that could pull the leads out, possibly lacerating the nasal passage. Their use may also interfere with obtaining a sleep-activated EEG, and, not infrequently, otherwise cooperative patients simply refuse the procedure.Sphenoidal electrodes use a hollow needle through which a fine electrode that is insulated, except at the tip, is inserted between the zygoma and the sigmoid notch in the mandible, until it is in contact with the base of the skull lateral to the foramen ovale. This is an invasive procedure that must be done by a physician and requires a signed consent form. The yield of positive results from these specialized electrodes, over and above findings present in conventional scalp recordings, is still controversial. In general, the yield from NP leads has not been high, although, with sphenoidal leads, positive results as great as 40 percent have been reported from seizure patients who had no other specific changes in the waking or sleep EEG.Montage SelectionA common misconception is that the EEG records the voltage detected at each electrode site. Instead, each “squiggly line” on the EEG chart represents the shifting or oscillating difference in electrical potential between two electrodes. Thus, in a multichannel recording, the activity from each channel represents the shifting difference in microvoltage between two selected electrodes. When 10, 16, or even many more electrodes are placed on the head, the number of possible electrode pairs becomes large, and how these pairs are arranged among the recording channels can become complex. In EEG parlance, the way electrode pairs are arranged for a recording is called a montage, and, although many montages are possible, only a limited number have become popular and useful. Not uncommonly, several montages are used during a recording to sample the brain electrical activity.Montages are designed to facilitate the detection of EEG abnormalities in different brain regions and to facilitate comparisons between left and right hemisphere activity. There are general guidelines for how montages are to be set up. The most important rule is simplicity of the montages. Additional rules include the stipulation that odd numbers refer to the left side, whereas even numbers refer to right-side electrodes. Furthermore, left-side electrodes are routinely

Page 11: Translate

displayed on top of or before right-side electrodes. Similarly, anterior electrodes are displayed on top of or before more posteriorly placed electrodes.There are two main types of montages: referential and bipolar. With referential montages, all electrodes are referenced to a single common reference point that commonly consists of linked ears (the mastoid prominence can be used in place of the ear lobe), with variations being left or right ear reference alone, ipsilateral ear reference in which all electrodes in one hemisphere are referenced to the ear on that side, or a contralateral ear reference in which all electrodes in one hemisphere are referenced to the opposite-side ear. Referential montages are useful for judging the magnitude of the abnormality (in terms of how large a sharp wave or slow wave is in microvolts). Bipolar montages, on the other hand, are useful (and, indeed, are much more widely used than referential montages) for pinpointing the area of maximal abnormality or the exact source of an abnormal activity. In bipolar montages, electrodes are referenced from one scalp location to a nearby scalp location in chains of electrodes going across the head from front to back (Fig. 1.14-5) or from left to right (Fig. 1.14-6).

FIGURE 1.14-5 Example of an 18-channel bipolar montage with anterior to posterior linkages. The numbers between electrode locations designate recording channels. Thus, the number 6 means channel 6, which measures the difference in electrical potential between the F3 and C3 electrodes. (From Tyner FS. Fundamentals of EEG Technology: Basic Concepts and Methods. Vol 1. New York: Raven Press; 1985, with permission.)

Page 12: Translate

FIGURE 1.14-6 Example of a 16-channel transverse bipolar montage. (From Tyner FS. Fundamentals of EEG Technology: Basic Concepts and Methods. Vol 1. New York: Raven Press; 1985, with permission.)P.177

The majority of abnormal cerebral activities tend to appear at the surface as negative potentials. One can think of a given channel of EEG activity as being derived from two inputs. By convention, the first electrode of a pair constitutes input 1, whereas the second electrode provides input 2. Thus, in the electrode pair C3-P3, the first electrode C3 constitutes input 1. The direction of the pen deflection is based on whether input 1 (the first electrode in a pair) is, relatively speaking, “more negative” or “less negative” (i.e., relatively more “positive”) than the second electrode (input 2). If the first electrode in a recording pair (input 1) is closer to the source of a negative field and, hence, more “negative” than the second electrode (input 2), even though both electrodes may be within the field, there is an upward pen deflection. Conversely, if the first electrode of a pair is more distant from the source of the field than the second electrode and, hence, less negative than the second electrode (which is the same as saying that it is, relatively speaking, more “positive”), the pen deflection is downward. There is no denying that it takes some time to become accustomed to these polarity principles. However, they lead to important techniques for localizing certain abnormal features. As bipolar pairs of electrodes move in a longitudinal or transverse direction from one side of a strong and highly localized negative field to the other side, the pen deflection changes direction as the first electrode in a given pair (input 1) shifts from being relatively more negative to relatively more positive than the second electrode (input 2). This change of pen deflection is called a phase reversal and is a powerful

Page 13: Translate

method for localization of sharply focal abnormalities. By contrast, monopolar montages localize by identifying the electrode with the highest amplitude of the abnormality (Fig. 1.14-7).

FIGURE 1.14-7 Illustration of bipolar (phase-reversal) and monopolar (highest amplitude) localization of a focal negative spike discharge at the left anterior temporal (T1) electrode. See the text for an explanation.With rapidly advancing computer technology, it is possible today to record the EEG with one montage and to reconnect digitally the electrodes in any sequence desired. This has the advantage of being able to examine the entire EEG record in all possible configurations. One particular montage configuration deserves special mention, for it may be particularly useful in

Page 14: Translate

psychiatric EEG. This is a montage that combines referential and bipolar electrode arrangements. Following four bipolar connections from the frontal regions through the temporal areas and ending in the occipital region, a referential placement connects each posterior temporal region (T5 and T6) to the opposite ear. This arrangement allows activity of low amplitude to be highlighted by the referential electrodes for further examination via the bipolar electrode pairs. This montage is commonly referred to as the Queen Square montage (Fig. 1.14-8).

FIGURE 1.14-8 Diagram of the Queen Square montage. This is an 18-channel montage modified to include two referential leads to highlight temporal lobe activity.The appearance of EEG activity varies greatly from one recording montage to another. Large interelectrode distances often (but not always) yield higher voltages, whereas a close spacing between electrodes in a pair tends to reduce voltage, because when both electrodes overlie nearly the same portion of an electrical field the potential difference between them is small. Furthermore, specific EEG patterns visible in one montage may be distorted or even completely canceled out in another montage. Although some montages may permit a differentiation of activity between two or more brain regions, other montage choices may not do so. For example, EEG sleep patterns are well visualized and well differentiated in central and occipital regions when a common (monopolar) reference recording is made. However, differentiation between central and occipital sleep activity is no longer possible when bipolar anterior-posterior linkages are used (C3-O1 and C4-O2), and, with transverse bipolar links between homologous electrodes,

Page 15: Translate

the sleep patterns may not be visible at all (Fig. 1.14-9). The issue is not merely academic. Discharges of interest to the electroencephalographer, whether they be clinically abnormal or controversial, that are detectable in some recording montages may be completely or nearly undetectable, even though they are currently “firing” when a different montage is being used (Fig. 1.14-10).

FIGURE 1.14-9 Alteration of appearance of brain waves (sleep patterns) with change in recording montage. Note that the monopolar montage (top four channels) yields higher amplitudes and greater differentiation between central and occipital activity. Similar input to members of an electrode pair (C3-O1 and C4-O2) can reduce voltage in the bipolar derivation. Note the absence of differentiation between central and occipital activity in the bipolar derivation. Note the extreme cancellation of activity in the last two bipolar derivations.

Page 16: Translate

FIGURE 1.14-10 A: Fourteen-per-second and six-per-second positive spike discharges (a controversial pattern), independent left and right temporal-parietal-occipital area (monopolar montage). B: The top two channels show these discharges with the same monopolar montage as

Page 17: Translate

channels 3 and 4 in A, whereas lower channels show bipolar cancellation of the discharges, even though all electrodes in montage A are present. The female patient was 32 years of age with a closed head injury.SensitivityThe amplification used in EEG recording is adjustable and can be increased to visualize low-voltage signals or decreased to prevent recording pens from reaching their deflection limits and “squaring off,” thus distorting the shape of the top of the waveform. Although the accepted standard sensitivity across laboratories for most recording situations is 7 µV for each mm of pen deflection, the sensitivity may be altered, if necessary, to increase the clarity of the EEG information being obtained. For example, it may be necessary to sharply decrease the amplification to 10, 15, or even 20 µV per mm to visualize the complete waveform shape in certain high-voltage seizure discharges. Conversely, there are situations, such as recordings to document electrocerebral silence, in which it is important to maximize the ability to detect brain wave activity. In such situations, a high amplification of 1.0 or 0.5 µV per mm might be selected, along with the use of referential montages or bipolar runs with large interelectrode distances, to further enhance low-voltage registration. The EEG recording indicates the sensitivity setting at the beginning of the record and at any point in the recording at which the sensitivity was changed.Frequency Filter SettingsNearly all of the EEG activity that is analyzed for clinical or research purposes falls within the frequency range of 0.5 to 40.0 or 50.0 Hz. Conventional EEG recordings usually use a high-frequency filter setting of 70 Hz, which means that brain waves become progressively attenuated in amplitude the more that they increase above this filter setting. At the other end of the spectrum, most laboratories set the low-frequency filter at 1.0 Hz to reduce the registration of frequencies below this level. Unfortunately, scalp electrodes pick up a variety of electrical potentials of nonbrain origin, and many of these have frequencies within or close to the EEG frequency spectrum. Frequency filters may, to P.178

some degree, mitigate against the distorting effects of frequencies generated by nonbrain sources. However, filters must be used judiciously and with caution, because they can also filter out real brain waves that one wishes to see.Although the low-frequency filter can be adjusted downward to 0.3 Hz or even 0.1 Hz to capture slow waves, this is seldom done in routine recordings. More commonly, the low-frequency filter is moved upward to 5 Hz to eliminate unwanted slow potentials known to be artifact. Chief among these unwanted slow waves are those generated by electrical activity of the skin during sweating (galvanic skin response), and they can be of sufficiently high amplitude that they completely obliterate genuine EEG activity in the affected recording channels (usually bilateral frontal-anterior temporal areas). Increasing the low-frequency filter setting to 5 Hz totally eliminates this source of contamination in the recording (Fig. 1.14-11) but does so at the expense of attenuating any real generalized or focal slow activity that may also be present. It is much more common to adjust the high-frequency filter downward from 70 Hz to 35 Hz or even 15 Hz to eliminate or reduce unwanted muscle potential from the recording (Fig. 1.14-12). Again, the choice to do this involves a compromise, because such lowering of the high-frequency filter setting may make the accurate detection of certain fast spike discharges problematic.

Page 18: Translate

FIGURE 1.14-11 Effect of low-frequency filter setting on perspiration artifact (F7 and T3 electrodes) during sleep recording. Adjusting the low-frequency filter upward to 5 Hz completely eliminates the artifact and also eliminates the normal slow wave components of sleep but does not alter the 14-Hz sleep spindles.

FIGURE 1.14-12 Effect of adjusting the high-frequency filter setting on muscle potential artifact (generated by having the patient grind his teeth repeatedly). Muscle potential seen at the “normal” filter setting of 70 Hz is attenuated when the filter setting is lowered to 35 Hz and is completely removed when it is set at 15 Hz. Lowering the high-frequency filter introduces the risk of attenuating or removing (i.e., filtering out) abnormal spike discharges from the tracing.Special ActivationsOver the years, electroencephalographers have recognized that certain activating procedures tend to increase the probability that abnormal discharges, particularly spike or spike-wave seizure discharges, will occur. Some activating techniques remain standard in many laboratories, others are used only rarely for specific purposes, and still others introduced in the past have largely been abandoned, because they were not easy to use or involved risk.Medication ActivationAlthough the use of drugs to induce EEG changes enjoyed a certain vogue in the past, this type of activation is essentially no longer used in clinical work today. During the 1940s and 1950s, the convulsant drug, pentylenetetrazol (Metrazol), was sometimes used to activate seizure discharges in the EEG, but, if not used with caution, it could precipitate actual overt grand mal seizures during the EEG recording. The use of bemegride, another drug with convulsant properties, was also used, and, although it was presumed to be safer, it also was not without risk. Of particular relevance to psychiatry, Russell Monroe began using α-chloralose in the 1960s as a specific activator of EEG abnormalities in psychiatric patients. Although it was reported to be effective with psychiatric patients, it was said to be particularly effective in activating paroxysmal EEG discharges in a high proportion of patients with aggressive episodic dyscontrol syndromes.P.179

Page 19: Translate

HyperventilationStrenuous hyperventilation is one of the oldest, and still one of the most frequently used, activation procedures in clinical laboratories. While remaining reclined with the eyes closed, the patient is asked to overbreathe through their open mouth with deep breaths for 1 to 4 minutes, depending on the laboratory (3 minutes is common). The normal EEG response to hyperventilation (referred to in EEG parlance as a build-up) consists of an increase in generalized medium- to high-voltage synchronous slow waves in the delta range, which then quickly subside when overbreathing stops. Not everyone has a build-up response to hyperventilation, and children are far more likely to respond with diffuse EEG slowing than are adults. In terms of activating EEG abnormality, hyperventilation is especially effective in eliciting the classic diffuse three-per-second spike-and-wave complex of petit mal seizures when the pattern does not first appear in the standard wake tracing, and, to a lesser degree, it may activate other synchronous diffuse spike-wave patterns. Activation of focal seizure activity has been reported much less frequently. An interesting observation is that significantly low blood glucose levels have been associated with large, synchronous delta wave hyperventilation build-ups, and, because of this, a large delta wave build-up in an adult may signal the existence of covert, unsuspected pathological hypoglycemia. If this suspicion should present itself during a recording, a good idea would be to give a sugar drink to the patient and then to repeat hyperventilation later. If the glucose ingestion reduces or abolishes the large hyperventilation build-up, the suspicion of hypoglycemia is reinforced. In general, hyperventilation is one of the safest EEG activating procedures, and, for the majority of the population, it presents no physical risk. However, it may pose a risk for patients with cardiopulmonary disease or risk factors for cerebral vascular pathophysiology.Photic StimulationIn the earliest days of EEG, it was known that the frequency of normal EEG activity recorded from posterior scalp regions could be made (within narrow limits) to follow the frequency of a flickering light that was flashed slightly faster or slower than the intrinsic brain wave frequencies, a phenomenon that came to be referred to as photic-driving. When it also became known that photic-driving would sometimes cause paroxysmal discharges to occur in the EEG, photic stimulation (PS) emerged as a technique for eliciting EEG abnormalities. Although there is some variation between laboratories, PS generally involves placing an intense strobe light approximately 12 inches in front of the subject's closed eyes and flashing at frequencies that can range from 1 to 50 Hz, depending on how the procedure is carried out. Retinal damage does not occur, because each strobe flash, although intense, is extremely brief in duration. Some laboratories sample independent flash frequencies separately and randomly, although almost no one samples all frequencies between 1 and 50 Hz. Other laboratories use a zoom technique in which the flashes start at a low frequency, such as 1 Hz, and are then gradually and continuously increased to much higher flash frequencies. In some individuals, PS produces facial and eye muscle jerks, called a photomyoclonic response (PMR). More than 40 years ago,