Technical







IQCB Blueprint Coverage


This unit addresses III. Technical (3 Hours). The competent clinical neurophysiologist must acquire knowledge of electronics and instrumentation related to the EEG and ERPs.




This unit covers:
A. Topographical Representation of the EEG
B. Electrodes and Acquisition Systems
C. Instrumentation (Acquisition and Review Parameters/Settings)
D. Montages
E. Electrical/Clinical Safety

A. Topographical representation of the EEG


Please click on the podcast icon below to hear a full-length lecture over Section A.



Quantitative electroencephalography (qEEG) involves the mathematical analysis of the EEG signals to extract quantitative information about the brain's electrical activity. One crucial aspect of qEEG is the topographical representation, which provides a spatial map of EEG activity across the scalp. Topographical mapping is essential for visualizing the distribution and intensity of electrical potentials, facilitating the identification of regional brain activity patterns and abnormalities.

The raw EEG signals undergo preprocessing steps, including filtering, artifact removal, and segmentation into epochs. This preprocessing ensures that the signals are clean and ready for further analysis. The signals are then subjected to spectral analysis, usually via Fourier Transform, to decompose them into their constituent frequency bands (e.g., delta, theta, alpha, beta, gamma). Each frequency band is associated with different brain states and functions.

The processed data from each electrode are used to create a spatial representation of brain activity. This involves interpolating the values between electrodes to generate a continuous map. Topographical maps can be color-coded to represent the intensity of electrical activity at different scalp locations. These maps visually display how specific frequency bands or other EEG metrics (e.g., power, coherence) are distributed across the scalp.

Topographical EEG maps are used in clinical settings to identify and localize abnormal brain activity associated with various neurological conditions, such as epilepsy, brain injuries, and psychiatric disorders. In research, these maps help study brain function and connectivity, investigate neural mechanisms underlying cognitive processes, and assess interventions' effects.

The benefits of topographical mapping are enhanced visualization, brain activity localization, and diagnostic accuracy.

Enhanced visualization: Topographical representation provides an intuitive and detailed visualization of brain activity, making it easier for clinicians and researchers to interpret complex EEG data.

Brain function localization: Topographical EEG can help localize specific brain functions and detect abnormalities in particular regions by mapping the spatial distribution of electrical activity.

Improved diagnostic accuracy: In clinical practice, topographical maps enhance the accuracy of diagnosing neurological conditions by revealing subtle abnormalities that may not be apparent in raw EEG traces.



The Development of Z-Score Training

To meet the need for z-score training, Robert Thatcher, developer of the NeuroGuide database, in cooperation with Thomas Collura of BrainMaster Technologies, developed Live Z-Score Training (LZT). Beginning in about 1996 (Thatcher et al., 2019), they started with a simple 1- or 2-channel approach that quickly developed into a 4-channel system for training clients by providing real-time information about how closely their EEG matched a database of age-matched typical controls. There were approximately 72 variables available within the Thatcher database, including power, relative power, phase and coherence values, and more for the standard EEG frequencies.

Caption: Z-PLUS LZT Live Z-Score Training with Z-Bars and Z-Maps Display

The client’s EEG values were compared to these values, and differences were displayed in standard deviations. The goal was to train toward z-zero standard deviations.



Caption: Tom Collura

Subsequent advances resulted in 19-channel, real-time surface z-score training and then 19-channel LORETA z-score training, using the LORETA source localization method to train in 3 dimensions. Collura and his colleagues ultimately shifted to the E. Roy John database, BrainDX. Then, they shifted again to a database known as qEEGPro, which was developed using client EEG recordings cleaned of clinical EEG patterns identified by a client questionnaire (qEEG.pro/database/). This approach is based upon questionable assumptions, and many qEEG researchers consulted about this expressed skepticism regarding this database approach (personal communication, John Anderson, 2019-2021).


Caption: qEEGPro 19-Channel sLORETA Z-Score Training

Robert Thatcher and his team continued to work with the Neuroguide database and have continued to develop the features of 19-channel z-score training, most recently with the development of swLORETA, a more precise and accurate iteration of the LORETA source localization method.

Caption: NeuroNavigator swLORETA

This has resulted in a large base of clinicians using this approach. More than 50 publications have presented evidence of the efficacy of z-score neurofeedback training, suggesting widespread acceptance within the scientific literature (https://www.appliedneuroscience.com/PDFs/Z_Score_NFB_Publications.pdf).


Clinical Database Development

Concurrently with the development of the normative database came another approach to protocol development. Skilled and experienced practitioners who also trained new neurofeedback clinicians began to develop what came to be known as a clinical database.

This approach was based on the clinician’s personal experience with often hundreds of clients and their perception of the similarities and differences in the findings associated with those clients. These individuals also consulted the published electroencephalography and neurofeedback research and identified correlations between that information and their experiences.

This led to several semi-automated approaches to assessment and training.

In his article in NeuroConnections (Swingle, 2014), Paul Swingle stated, “For Clinicians, the most accurate databases are clearly clinical.” He describes several logical problems with using normative databases, such as the issue of using a group of supposedly "normal" individuals as a reference standard with which to compare a client with clinical symptoms.



Caption: Paul Swingle

Swingle questioned the fundamental premise of symptom-free individuals selected for a normative database having no underlying abnormalities. He offers the example of heritable disorders such as migraine, schizophrenia, and others, which may have verifiable neurophysiological components that have not manifested due to a lack of triggering factors. In other words, genetic characteristics or predispositions that have not been expressed in the form of identifiable symptoms.

This would lead to a database of individuals who pass all the screening tests yet have notable EEG findings that will become components of the normal standard. Because of this, Swingle suggests that such comparisons are not useful for accurate client assessment and training.

Others also have called into question the concept of the normal individual and pointed out that a symptom-free individual may still have abnormal findings on a clinical EEG assessment strictly based on visual inspection of the EEG by a skilled electroencephalographer (Johnstone & Gunkelman, 2008). They also stated that a database comparison identifies differences from average, not from optimal. The term, normal, is difficult to define, particularly regarding a highly variable measure like the EEG.

Proponents of clinical databases claim exceptional accuracy when used with clinical populations because they were developed from experience with similar individuals experiencing similar symptoms, causal factors, symptom progression, and treatment histories from the general medical and psychiatric communities and psychotherapeutic clinicians.

Because of this reported accuracy and the clinical database’s close alignment with client symptomology, they are often more descriptive and, some say, user-friendly. This is one of the concerns regarding the use of a normative database.

A normative database analysis may produce 200-300 pages of tables, graphs, and topographic brain maps like the ones below. This impressive array of information can be quite overwhelming and difficult to interpret. Questions arise regarding which components of the analysis are important, which relate to the client’s symptoms, and which are useful metrics for planning client training sessions.

Caption: Image from NeuroGuide Normative Database, Applied Neuroscience.


Caption: Image from LORETA Progress Report by Phil Jones.


Caption: Image from NeuroGuide Normative Database, Applied Neuroscience.


Image from Low Resolution Electromagnetic Tomography (LORETA), Roberto Pascual-Marqui.


Caption: Image from NeuroGuide Normative Database, Applied Neuroscience.


On the other hand, a clinical database is often more descriptive. Some only provide graphs and tables, but most also suggest relationships between the findings and probable client symptoms. Swingle’s Clinical Q provides a data analysis that shows differences from expected data and offers "probes" for client symptoms or behaviors that may be consistent with these findings.

Caption: Clinical Q example report.


Peter Van Deusen

One of the earliest clinical databases was created by Peter Van Deusen (Ribas et al., 2016), who initially studied with Joel Lubar. He aimed to identify client symptoms and behavioral issues without utilizing diagnostic categories so that clients could be trained using neurofeedback approaches no matter what the issue.

The culmination of his years of work with clients, training practitioners, and studying the EEG was an approach known as The Learning Curve (TLC). TLC organized client symptoms and findings into six broad categories and identified training approaches to train these clients.



Caption: Peter Van Deusen.


Richard Soutar

Another approach to this method, and arguably the most comprehensive, is the New Mind Maps developed by Richard Soutar. This system has a variety of levels of analysis up to 19 channels and a detailed report that covers most of the typical metrics and provides narrative content and protocol recommendations for standard amplitude training and z-score training. It offers advice about additional interventions such as audio-visual entrainment (AVE).



Caption: Richard Soutar


Caption: New Mind Maps, Richard Soutar – http:/www.newmindmaps.com.


Caption: New Mind Maps, Richard Soutar – http:/www.newmindmaps.com.


John Demos

The Jewel Clinical Database, developed by John Demos, author of Getting Started with Neurofeedback (2nd ed.), is another clinical database for ages 7-19 and adults that produces surface and sLORETA representations (maps, graphs, etc.) and training recommendations based on client checklists.

Caption: The Jewel Clinical Datavase, John Demos.




Caption: John Demos.


John Anderson

John Anderson developed an assessment for the Nexus/BioTrace system known as the NewQ.



Caption: John Anderson

This system is like the Clinical Q of Paul Swingle and uses 6 locations, recorded sequentially, two locations at a time. There are four age ranges: 6-11, 12-15, 16-20, and 21+ (adult).

Caption: NewQ, John Anderson.

The following is the disclaimer from the NewQ assessment. These cautions are worth keeping in mind when utilizing such an approach to assessment.
Disclaimer: This assessment tool is based upon a general understanding of the EEG literature plus the author's clinical experience. It should not be viewed as a statistically validated or rigorously referenced instrument and constitutes a set of clinical observations. It is not a diagnostic instrument, and results must be evaluated based on the client's presenting concerns. The clinical judgment of the practitioner must remain primary in any assessment process.

The clinical database approach is valuable in providing practitioners with a perspective on client assessment based on the author’s knowledge and experience. It can be a helpful shortcut for beginning practitioners and experienced clinicians as they benefit from different clinical perspectives.

However, because the sensor locations are often recorded sequentially (except for Soutar’s 19-channel version of New Mind Maps), the assessments cannot calculate such important metrics as phase and coherence, except for sensor locations recorded simultaneously. Also, the perspective of the designer/author is quite subjective and, while reflecting knowledge and experience, may also reflect unconscious biases and assumptions that are not rigorously grounded in published research.

Ideally, there will eventually be a blending of the normative database approach with the clinical assessment to the point where a truly accurate and effective "expert" system can be designed to develop best practices for neurofeedback training.


A Guide to Interpreting qEEG Topographic Maps

Reading topographic maps of the EEG may seem straightforward and relatively simple. Z-score maps highlight areas that deviate from typical values compared to normative databases adjusted for age and sometimes gender and handedness. When an area on the map shows excessive activity in a particular EEG frequency, targeted sensor placement and effective client training can help normalize these levels. Conversely, reduced activity in an area may result in efforts to enhance it.

However, the reality is more complex. EEG recordings often exhibit significant artifacts from multiple sources, such as environmental interference (e.g., 50- or 60-Hz electrical noise) and physiological factors like eye blinks, movements, heartbeats, and muscle contractions. We must clean EEG data to ensure its integrity. This requires distinguishing between genuine EEG activity and transient phenomena such as drowsiness, sleep, or normal variants that do not signify pathology. While important for an accurate EEG report, these EEG features should not factor into statistical analyses.

Creating topographic EEG maps should be considered one of the last stages of a clinical assessment, not its primary focus.

Therefore, adhering to a careful progression from data collection to comprehensive analysis is essential instead of relying on automated artifact rejection algorithms and immediately generating maps.

Focusing neurofeedback training on areas associated with problematic symptoms is important for effective intervention. Simply targeting any abnormality detected in the EEG may not address the underlying issues and could lead to unintended consequences.

Atypical EEG findings can arise from various factors, including exceptional skills, compensatory changes due to illness or injury, developmental differences, or unique characteristics that do not necessarily indicate pathology. Therefore, careful clinical assessment and interpretation are necessary to determine whether observed deviations require correction.

By focusing on symptom-based training associated with understanding the clinical picture, clinicians can ensure that neurofeedback protocols address specific concerns and optimize outcomes for individuals undergoing training.

The following is an example of an EEG recording of a 76-year-old male with complaints of "brain fog," memory problems, lack of energy, slow cognitive processing, and difficulty sleeping.


Eyes-Closed Linked Ears (ECLE) Montage

This is a 50-uV scale, 10-sec display. Note the persistent ECG artifact in multiple channels, most clearly seen in reference channels (red outline at the bottom). The peak alpha frequency is approximately 8 Hz, the amplitude is 12-25 uV at the parietal sensors, and there is a small electrode pop in the F4 sensor (blue outline in this and subsequent montages).






Eyes-Closed Longitudinal Bipolar (ECLBP) Montage

This is a 50-uV scale, 10-second display. The alpha frequency is 8 Hz, and the amplitude is 10-30 uV in parietal-occipital derivations. Note that the rhythmic activity (frontal alpha) seen in the linked ears montage is not present in prefrontal–frontal derivations in this bipolar montage, indicating it was the result of reference contamination in the previous montage.





Eyes-Closed Average Reference (ECAVE) Montage

This is a 50-uV scale, 10-second display. The alpha frequency is 8-9 Hz, and the amplitude is 5-15 uV at the occipital and 10-18 uV at the parietal sensors. Note the delta activity at the parietal sensors (green outline) and EMG artifact at the occipital sensors.






Eyes-Closed Laplacian (ECLP) Montage

This is a 400-microampere (uA) scale, 10-second display. The alpha frequency is 8-9 Hz, with the highest current density in parietal sensors. EMG artifact continues in occipital sensors, and delta is more pronounced in the parietal area. Electrode pop in F4 sensor. The lack of prefrontal and frontal alpha activity suggests reference contamination in the linked ears montage above.






Eyes-Closed Linked Ears (ECLE) Montage

A linked ears montage in NeuroGuide with FFT absolute power spectral display at the top right with a line indicating the highest amplitude at 9 Hz is at the P4 electrode.




The same image showing the maximum power at 8 Hz is at the C4 electrode.





The same linked ears montage image shows the peak activity at 7.5 Hz, which is generally 1-3 SD greater than typical values at multiple locations (see z-score indicators on the left side of tracings next to electrode location labels).





Eyes-Closed Longitudinal Bipolar (ECLBP) Montage

This spectral display shows a longitudinal bipolar montage indicating the maximum z-scores at 2.5 Hz.






Eyes-closed longitudinal bipolar montage showing standard deviations at 8.5 Hz.






Eyes-Closed Average Reference (ECAVE) Montage

Average reference montage showing deviations at 2.5 Hz.







Eyes-Closed Laplacian (ECLP) Montage

Laplacian montage showing current source density (CSD) z-scores at 3 Hz.






Eyes-Closed Linked Ears (ECLE) Montage

These absolute power topographic maps represent 1 minute and 30 seconds of a recording from an eyes-closed linked ears montage. Each small head map represents a virtual view of the top of the head with the nose at the top. The absolute power (microvolts squared) values correspond to the colored scale below each map. Red represents the greatest value, and blue is the lowest value for each 1 Hz frequency bin. The P4 electrode shows a power value of 36 in the 9 Hz frequency bin. Each bin has its own scale.


Z-Score Absolute Power

1-Hz frequency bin maps showing maximum to minimum deviations compared to the NeuroGuide normative database for a linked ears montage. The scale is from -3 (blue) to +3 (red) standard deviations (SD). The excess activity at 1 Hz is likely due to the ECG artifact noted earlier. The heart beats at about 1 beat per second, which equals 1 Hz. Excess activity is seen at 7-9 Hz.

Frequency band maps (like delta, theta, alpha, and beta) show the entire frequency band and lack the resolution of the individual 1-Hz frequency bin maps. The delta map generally shows excess delta activity, which misidentifies the ECG artifact that was seen at 1 Hz in the frequency bin maps and the visual inspection of the EEG.



These are relative power topographic maps representing 1 minute and 30 seconds of data from an eyes-closed linked ears montage. They show relative power values (percentages) that compare the value in each 1 Hz bin to the broadband EEG (0.5 – 30 Hz). This information is displayed as a colored scale below each map, with red representing the highest percentage and blue representing the lowest percentage for each 1-Hz frequency bin. Note that 9 Hz contains 33 percent of the total EEG power at the P4 electrode location. Each head map has its own scale.


These are z-score relative power 1-Hz frequency bin maps showing maximum to minimum deviations compared to the NeuroGuide normative database using a linked ears montage. The scale is from -3 (blue) to +3 (red) standard deviations (SD). Note that this page represents relative power, showing the relative value of each 1 Hz bin compared to the EEG as a whole. This can result in areas showing incorrect abnormal z-score values at some frequencies because other frequencies are abnormal in the opposite direction. For example, frontal, central, and parietal activity at 8 Hz is excessive in the image below, causing apparent deficient activity in the same areas at multiple frequencies because 8 Hz takes up too much of the percentage "pie."

The previous examples show the progression of EEG analysis from viewing the recorded EEG through spectral analysis to topographic maps representing absolute and relative power and z-score maps showing deviations from expected values when client results are compared to a normative database.

The raw tracings and the spectral displays show examples from multiple montages (sensor comparisons) and help to highlight that what is seen is highly dependent upon which comparisons are used. So far, The topographic map examples have all used the linked ears montage and only showed one perspective on the EEG results. Now, we will look at the same data presented as topographic maps using different montages.



Eyes-Closed - 3 Montages

Eyes-closed average montage absolute power topographic map for 1-20 Hz, eyes-closed linked ears absolute power topographic map for 1-20 Hz, and eyes-closed Laplacian absolute power topographic map for 1-20 Hz.


Eyes-closed average z-score absolute power topographic map for 1-20 Hz, eyes-closed linked ears z-score absolute power topographic map for 1-20 Hz, and eyes-closed Laplacian montage z-score absolute power topographic map for 1-20 Hz. Notice the differences between the average, Laplacian, and linked ears reference maps. Which one should we follow when completing our assessment?



The average reference and Laplacian montages show fairly similar results and correspond to some of the other indicators and may generally represent the client’s presenting issues of brain fog, memory issues, and slow cognitive processing.

Several posts and Neurofeedback Tutor have addressed the issue of montage selection. Nunez and Srinivasan's (2006) Electrical Fields of the Brain (2nd ed.) is an excellent resource for more in-depth information.

For this example, we are confronted with significant differences between the linked ears results and average reference and Laplacian montage results, when there is excess activity in the 2-6 Hz range of delta and theta.

The Laplacian montage uses an average of the current flow from electrodes immediately surrounding the electrode of interest as a localized average reference using current rather than voltage as its metric. This has been described as more accurate in identifying local activity while minimizing general effects such as those from medication, drowsiness, and others. This can help highlight local abnormalities, which can be lost or masked by other montages.

The average reference montage uses an average of all scalp electrodes to serve as the reference for each electrode, thus eliminating the ear or mastoid reference. This helps remove contributions from these reference electrodes, common to all electrode pairings when using the linked ears or linked mastoid reference montage.

Each of these montage choices has benefits and limitations. The benefits have been mentioned above, but what are the limitations? The average reference montage, like all montages, is subject to the differential amplifier's common mode rejection phenomenon. Sources (electrical activity such as EEG, ECG, EMG, EOG, EMF) that are the same in frequency and, to a lesser extent, in amplitude are rejected, while sources that are different are retained. If multiple sources (electrode locations) of delta activity contribute to the average, and this average is then compared to a location that does not show delta activity, then there is a difference between the signals. That difference is retained and displayed in the topographic maps and, of course, in the EEG tracings, resulting in apparently abnormal delta activity where it does not exist. The same is true of the linked ears/mastoid reference and, to a lesser extent, of the Laplacian montage.

Therefore, we look for agreement among multiple montages, being particularly attentive to the various bipolar montages that allow revealing comparisons when viewing the EEG tracings.

In the present example, though earlier we mentioned that the delta activity was confined to the 1 Hz effect from the ECG (heartbeat) artifact, we can see from the Laplacian and average montages that there is substantial agreement on a broader frequency distribution of delta/theta activity that exceeds statistical significance. Thus, the recommendation is to begin training by focusing on these excesses.










Glossary



bins: frequency ranges, which may be as narrow as 1 Hz, into which the EEG power spectrum is divided.

Eyes-Closed Average Reference (ECAVE) montage: a method of EEG electrode montage where the average of all EEG channels is subtracted from each individual channel's EEG signal. This technique is often used to remove common noise sources and enhance the signal-to-noise ratio, particularly during eyes-closed conditions when the brain is relatively more relaxed.

Eyes-Closed Laplacian (ECLP) montage:
a spatial filtering technique in qEEG where the Laplacian transform is applied to the EEG signal recorded during eyes-closed conditions. This transform emphasizes local changes in EEG activity while reducing the influence of distant sources, thus enhancing spatial resolution.

Eyes-Closed Linked Ears (ECLE) montage: referencing the EEG signal recorded during eyes-closed conditions to an electrode placed on each earlobe, with the two ear electrodes linked together. This montage is used to minimize common noise sources and provide a stable reference for EEG analysis.

Eyes-Closed Longitudinal Bipolar (ECLBP) montage: a bipolar referencing method in qEEG where each EEG channel is computed as the difference between neighboring electrodes. This montage is often applied during eyes-closed conditions to enhance the detection of local variations in EEG activity.

montage: EEG recording configuration that groups electrodes (combines derivations) to monitor EEG activity.

z-score absolute power:
a statistical measure used in qEEG analysis to quantify the deviation of absolute power values within specific frequency bands from a normative database. It indicates how many standard deviations a particular absolute power measurement is from the mean of the reference population, providing a measure of relative EEG activity levels across different frequency bands.


Test Yourself


Customers enrolled on the ClassMarker platform should click on its logo to take 10-question tests over this unit (no exam password).


REVIEW FLASHCARDS ON QUIZLET


Click on the Quizlet logo to review our chapter flashcards.




Visit the BioSource Software Website


BioSource Software offers Physiological Psychology, which satisfies BCIA's Physiological Psychology requirement, and Neurofeedback100, which provides extensive multiple-choice testing over the Biofeedback Blueprint.



Assignment


Now that you have completed this unit, explain the rationale for z-score training using a normative database.

References


Ayers, M. E. (1977). EEG Neurofeedback to bring individuals out of Level Two Coma [Abstract]. Biofeedback and Self-Regulation, 20(3), 304–305.

Berger, H (1929). Über das elektroenkephalogramm des menschen. Arch Psychiatr Nervenkr 87(1), 527–570.

Budzynski, T. H. (1977). Tuning in on the twilight zone. Psychology Today, 11, 38–44.

Budzynski, T. H. (1996). Brain brightening: Can neurofeedback improve cognitive process? Biofeedback, 24, 14–17.

Byers, A. P. (1998). Neurotherapy reference library (2nd ed.). Applied Psychophysiology and Biofeedback.

Evans, J. R., Dellinger, M. B., & Russell, H. L. (Eds.) (2020). Neurofeedback: The first fifty years. Academic Press.

Fehmi, L. G., & Robbins, J. (2008). Sweet surrender-Discovering the benefits of synchronous alpha brainwaves (pp. 231-254). Measuring the immeasurable: The scientific case for spirituality. Sounds True, Inc.

Fein, G., Galin, D., Johnstone, J., Yingling, C. D., Marcus, M., & Kiersch, M. E. (1983).EEG power spectra in normal and dyslexic children. I. Reliability during passive conditions. Electroencephalogr Clin Neurophysiol, 55(4), 399-405. https://doi.org/10.1016/0013-4694(83)90127-x

Hardt, J. V., & Kamiya, J. (1978). Anxiety change through electroencephalographic alpha feedback seen only in high anxiety subjects. Science, 201(4350).79-81. https://doi.org/10.1126/science.663641 PMID: 663641.

John, E. R., Ahn, H., & Prichep, L., Trepetin, M., Brown, D., & Kaye, H. (1980). Developmental equations for the electroencephalogram. Science, 210(4475), 1255-1258. https://doi.org/10.1126/science.7434026

Johnstone, J. & Gunkelman, J. (2003). Use of databases in QEEG evaluation. Journal of Neurotherapy, 7, 3-4, 31-52. https://doi.org/10.1300/J184v07n03_02

Kamiya, J. (1968). Conscious control of brain waves. Psychology Today, 1, 56-60.

Legarda, S. B., McMahon, D., Othmer, S., & Othmer S. (2011). Clinical neurofeedback: Case studies, proposed mechanism, and implications for pediatric neurology practice. J Child Neurol., 26(8),1045-1051. https://doi.org/10.1177/0883073811405052

Leong, S. L., Vanneste, S., Lim, J., Smith, M., Manning, P., & De Ridder, D. (2018). A randomised, double-blind, placebo-controlled parallel trial of closed-loop infraslow brain training in food addiction. Sci Rep 8(11659). https://doi.org/10.1038/s41598-018-30181-7

Lubar, J. F., & Bahler, W. W. (1976). Behavioral management of epileptic seizures following EEG biofeedback training of the sensorimotor rhythm. Biofeedback Self Regul, 1(1), 77-104. https://doi.org/10.1007/BF00998692. PMID: 825150.

Lubar, J. F., Shabsin, H. S, Natelson, S. E., Holdson, S. E., Holder, G. S., Whittsett, S. F., Pamplin, W. E., & Krulikowski, D. I. (1981). EEG operant conditioning in intractable epileptics. Archives of Neurology, 38, 700–704.

Lubar, J. F., & Shouse, M. N. (1976). EEG and behavioral changes in a hyperkinetic child concurrent with training of the sensorimotor rhythm (SMR): A preliminary report. Biofeedback Self Regul,1(3), 293-306. https://doi.org/10.1007/BF01001170. PMID: 990355.

Marzbani, H., Marateb, H. R., & Mansourian, M. (2016). Neurofeedback: A comprehensive review on system design, methodology and clinical applications. Basic and clinical neuroscience, 7(2), 143–158.

Monastra, V. J., Lubar, J. F., Linden, M., VanDeusen, P., Green, G., Wing, W., Phillips, A., & Fenger, T. N. (1999). Assessing attention deficit hyperactivity disorder via quantitative electroencephalography: An initial validation study. Neuropsychology, 13(3), 424–433. https://doi.org/10.1037/0894-4105.13.3.424

Nuwer, M. R., & Coutin-Churchman, P. (2014). Brain mapping and quantitative electroencephalogram. Encyclopedia of the Neurological Sciences (2nd ed.). Elsevier.

Oken, B. S., & Chiappa, K. H. (1988). Short-term variability in EEG frequency analysis. Electroencephalogr Clin Neurophysiol, 69, 191–198. https://doi.org/10.1016/0013-4694(88)90128-9

Peniston, E. G., & Kulkosky, P. J. (1989). Alpha-theta brain wave training and beta-endorphin levels in alcoholics. Alcoholism, Clinical and Experimental Research, 13, 271 – 279. https://doi.org/10.1111/J.1530-0277.1989.TB00325.X

Peniston, E. G., & Kulkosky, P. J. (1990). Alcoholic personality and alpha-theta brain wave training. Medical Psychotherapy, 3, 37–55.

Putman, J. A., Othmer, S. F., Othmer, S., & Pollock, V. E. (2005). TOVA results following inter-hemispheric bipolar EEG training. Journal of Neurotherapy: Investigations in Neuromodulation, Neurofeedback and Applied Neuroscience, 9(1), 37-52. https://doi.org/10.1300/J184v09n01_04

Ribas, V. R., Ribas, R., & Martins, H. (2016). The learning curve in neurofeedback of Peter Van Deusen: A review article. Dementia & Neuropsychologia, 10(2), 98–103. https://doi.org/10.1590/S1980-5764-2016DN1002005

Rogala, J., Jurewicz, K., Paluch, K., Kublik, E., Cetnarski, R., & Wróbel, A. (2016). The do's and don'ts of neurofeedback training: A review of the controlled studies using healthy adults. Frontiers in Human Neuroscience, 10, 301. https://doi.org/10.3389/fnhum.2016.00301

Shouse, M. N., & Lubar, J. F. (1979). Operant conditioning of EEG rhythms and ritalin in the treatment of hyperkinesis. Biofeedback and Self-Regulation, 4, 299–312. https://doi.org/10.1007/BF00998960

Sittenfeld, P., Budzynski, T., & Stoyva, J. (1976). Differential shaping of EEG theta rhythms. Biofeedback and Self-Regulation, 1, 31–45. https://doi.org/10.1007/BF00998689

Smith, M. L., Collura, T. F., Ferrara, J., & de Vries, J. (2014). Infra-slow fluctuation training in clinical practice: A technical history. NeuroRegulation, 1(2), 187–207. https://doi.org/10.15540/nr.1.2.187

Smith, M. L., Leiderman, L., & de Vries, J. (2017). Infra-slow fluctuation (ISF) for autism spectrum disorders. In T. F. Collura & J. A. Frederick (Eds.), Handbook of clinical QEEG and neurotherapy. Routledge Taylor and Francis Group.

Sterman, M. B. (1976). Effects of brain surgery and EEG operant conditioning on seizure latency following monomethylhydrazine in the cat. Exp. Neurol., 50, 757-765. https://doi.org/10.1016/0014-4886(76)90041-8

Sterman, M. B., & Friar, L. (1972). Suppression of seizures in an epileptic following sensorimotor EEG feedback training. Electroencephalogr Clin Neurophysiol,33(1), 89-95. https://doi.org/10.1016/0013-4694(72)90028-4. PMID: 4113278.

Sterman, M. B., Lopresti, R. W., & Fairchild, M. D. (1969). Electroencephalographic and behavioral studies of monomethylhydrazine toxicity in the cat. AMRL-TR-69-3, Aerospace Medical Research Laboratory, Air Force Systems Command, Wright-Patterson Air Force Base, Ohio.

Sterman, M. B., LoPresti, R. W., & Fairchild, M. D. (2010). Electroencephalographic and behavioral studies of monomethyl hydrazine toxicity in the cat. Journal of Neurotherapy, 14(4), 293-300, https://doi.org/10.1080/10874208.2010.523367

Strehl, U. (2014). What learning theories can teach us in designing neurofeedback treatments. Frontiers in Human Neuroscience, 8(894).

Swingle, P. (2014). Clinical versus normative databases: Case studies of clinical Q assessments. NeuroConnections.

Thatcher, R. W. (1998). Normative EEG databases and EEG biofeedback. Journal of Neurotherapy, 2(4), 8-39. https://doi.org/10.1300/J184v02n04_02

Thatcher, R. W., Lubar, J. F., & Koberda, J. L. (2019). Z-Score EEG biofeedback: Past, present, and future. Biofeedback, 47(4), 89–103. https://doi.org/10.5298/1081-5937-47.4.04


B. Electrodes and Acquisition Systems


Please click on the podcast icon below to hear a full-length lecture over Section B.



Electrodes detect biological signals. They are also transducers since they convert energy from one form to another. Four types of EEG electrodes are shown below: gold cup, gold flat, silver cup, and silver/silver-chloride ring.





 





Cap systems like the NeXus EEG cap share a connector containing a pin or other type of connector for each electrode. These connectors plug all electrodes into the amplifier at the same time.




Consider how EEG electrodes work. In response to chemical and electrical synaptic messages, the dendrites of cortical pyramidal neurons develop excitatory postsynaptic potential (EPSPs) and inhibitory postsynaptic potentials (IPSPs) that travel about 10-12 centimeters as a current of ions through the cortex, blood vessels, glial cells, interstitial fluid, meninges, and skull to electrodes located on the scalp. This process is called volume conduction.


Electrodes transform this current of ions into a current of electrons that flows through the cable into an electroencephalograph’s input jack.





The EEG signal is attenuated during volume conduction. The volume-conducted signal that reaches EEG electrodes is measured in microvolts or millionths of a volt.

How do EEG electrodes work? When an EEG electrode is filled with a conductive gel or paste, the electrode metal donates ions to the electrolyte. In turn, the electrolyte contributes ions to the metal surface. Electrodes create a DC voltage between the electrode metal and the electroconductive gel or paste. Signal conduction succeeds as long as electrode and electrolyte ions are freely exchanged.

Recording Problems

Conduction breaks down during polarization when chemical reactions produce separate positive and negative charge regions where the electrode and gel make contact. DC flows across the connection between an electrode and the scalp. The current carries positive ions to the more negative region of this junction and negative ions to the more positive region. This build-up of ions polarizes the electrode to favor current flow in one direction and resists flow in the other.

When an electrode is polarized, ion exchange is reduced, and impedance increases, weakening the signal reaching the electroencephalograph. This problem can result from routine clinical use. Electrode manufacturers control this problem using silver/silver-chloride or gold electrodes that resist polarization.

Bias potentials are a second potential recording problem. They result from the exchange of metal ions donated by the electrodes and electrolytes in the absence of a biological current. Bias potentials can be prevented using electrodes with intact surfaces and identical materials (e.g., all gold or silver).


Recording the EEG with Three Leads

We record scalp electrical activity using three recording electrodes: active, reference, and ground. We place the active electrode over a scalp site, which is an EEG voltage source. We can locate the reference electrode over the scalp or neutral, but not electrically inactive, sites like an earlobe or the mastoid bone. Finally, we can place the ground electrode on an earlobe, mastoid bone, or scalp (Demos, 2019). The ground electrode is grounded to the amplifier.

Active and reference sensors are identical in construction and are each balanced inputs. They are interchangeable! However, some technologies require that you designate a sensor as a reference. For example, linked ears reference.

In the graphic below, the active (+) is red, the reference (-) is black, and the ground electrode (Gnd/Ref) is white.






The voltages of the active and reference inputs are based on the ground.






EEG Apparatus

An electroencephalograph consists of the following stages: differential amplifier, gain amplifier, analog-to-digital converter, digital and FFT filters, and optical isolator.






Signal Amplification

The biological signals monitored in biofeedback are very weak. The EEG signal, for example, is measured in microvolts (millionths of a volt). These signals must first be amplified over several stages to isolate the signal we are interested in and then drive displays. Stereo amplifiers perform the same tasks when they boost audio signals above the noise floor to levels that can power loudspeakers.

Amplifiers share the properties of input sensitivity and gain. Input sensitivity is the maximum voltage level an amplifier can accept without producing clipping and distortion. The graphic below shows the same EEG signal with different sensitivity. The top tracing shows greater sensitivity than the bottom tracing, as evidenced by its substantially greater voltage swings.





Gain is an amplifier's ability to increase the magnitude of an input signal to create a higher output voltage. Gain is the ratio of output/input, which is different for AC and DC systems. An amplifier that produces a 1-mV output from a 1-μV input has a gain of 1,000.


Differential Amplifiers

The EEG signal is first boosted by a differential amplifier and then by a gain amplifier. A differential amplifier, also called a balanced amplifier, amplifies the difference between the two inputs: the active (input 1) and reference (input 2). In the diagram below, the triangle represents the amplifier, and the black circles represent the input voltages. Graphic © Hand Robot/Shutterstock.com.




A differential amplifier combines two (or more) identical single-ended amplifiers with balanced inputs. The inputs are referenced to a common ground to compare the resulting signals. The amplifiers are 180o out of phase, so signals that differ in frequency, amplitude, and phase are amplified. Only signal components that differ between two inputs are retained and amplified as output. Signals that are out of phase or possess different amplitudes are "seen" by the common-mode rejection process as different and are retained.

The graphic below was redrawn from John Demos' BCIA-recommended Getting Started with EEG Neurofeedback (2nd ed.). A differential amplifier rejects the common voltage (e.g., 3 feet) and outputs the voltage difference (e.g., 4 feet). A single-ended amplifier outputs the entire voltage (e.g., 7 feet, EEG artifact, and signal value).





Frequency is the number of cycles per second (Hz). Frequency graphic © Bany's beautiful art/Shutterstock.com.



Amplitude is the signal voltage or power measured in microvolts or picowatts. Amplitude graphic © petrroudny43/Shutterstock.com.





Phase is the similarity in timing of the waves at two locations. Note in the plot on the right. The two signals are 180o out of phase so that the top signal peaks when the bottom signal reaches its trough. Phase graphic © petrroudny43/Shutterstock.com.




The recording below shows 4 tracings. The first one shows the Fp1 electrode referenced to linked ears and has an event circled in red. The second tracing shows the O2 electrode also referenced to linked ears and has a distinct EEG event circled in red. When the two electrodes, Fp1 and O2, are referenced (compared) to each other (the third tracing), the differences are retained, showing an example of common mode rejection (CMR). The fourth tracing shows the linked ears compared to each other, resulting in complete rejection of the identical signals. Typically, CMR occurs within the amplifier, although the resulting signals may be displayed in differing comparisons (montages) within the software, resulting in a similar function.




How does this reduce artifact? When there is no EEG activity, identical noise signals reach each amplifier. The differential amplifier subtracts these signals, canceling out the artifact. The output of a perfect differential amplifier would be 0.


The Challenges of Recording Infra-Slow EEG Activity

An AC amplifier has severe limitations when recording infra-slow (0-1 Hz) EEG activity. AC amplifiers exacerbate artifact effects. Client movement, eye movement, sweat, and transient field artifacts produce significant voltage changes. Long time constants over 80 s are recommended to integrate artifact-induced voltages over 2-4-min periods. However, persistent artifacts like eye movement will consistently degrade the signal-to-noise ratio of client feedback.

Infra-slow recording requires DC-coupled amplifiers with a large dynamic range produced by 24-bit A/D converters to prevent saturation by slow drifts in baseline voltage. Standard EEG electrodes made of gold, steel, or tin are unacceptable because they suffer capacitance or energy storage, blocking lower frequencies. Silver/silver-chloride electrodes are ideal because they are reversible and do not polarize.

The clinician must distinguish slow artifacts from infra-slow signals. Eccrine sweat glands produce standing millivolt-range potentials. While these can be eliminated by partial skin puncturing, this practice risks infection transmission. Clinicians can identify eye blink and eye movement artifacts by their characteristic location. Body tilt, cough and strain, hyperventilation, and tongue movements produce high amplitude diffuse very slow potentials (Miller et al., 2007).


Common-Mode Rejection

A differential amplifier’s separation of signal from artifacts is measured by the common-mode rejection ratio (CMRR). Since these amplifiers cancel out noise imperfectly, signal and noise will be boosted. The CMRR specification compares the degree to which a differential amplifier boosts signal (differential gain) and artifact (common-mode gain). CMRR = differential gain/common-mode gain.

CMRR should be measured at 50/60Hz where the strongest artifacts, like power line (50/60Hz) noise, are found. The smallest acceptable ratio is 100 dB (100,000:1), which means that the signal is boosted 100,000 times more than competing noise. State-of-the-art equipment exceeds a 180-dB ratio. Lower ratios could result in unacceptable contamination of biological signals.

The graphic shows common-mode rejection when the common signal is in and out of phase.





You can take nine steps to maximize common-mode rejection:

(1) ensure that skin-electrode impedances are balanced within 1-3 Kohm. If both actives receive identical noise signals, the imbalance will make the signals look different and prevent complete subtraction of noise.

(2) active electrodes should be equidistant from the artifact source.

(3) active, reference, and ground sensors should be the same distance from each other.

(4) when using two or more channels, the ground and each active should be the same distance apart.

(5) ensure that there is a good ground connection. A deficient ground connection can make different voltages appear identical, defeating common-mode rejection.

(6) identify artifact sources. You can use a portable electroencephalograph or electromyograph like a Geiger counter. Move the unit around the room with EEG sensors connected but held in your hand. Artifact sources should produce the largest display values.

(7) remove the artifact sources you find. For example, fluorescent lights can be replaced with fixtures that produce less 50/60Hz noise.

(8) remove unused sensor cables from the encoder so they do not function as antennas for 50/60Hz artifacts.

(9) position the electroencephalograph and electrode cable to reduce artifact reception. Use the location and angle that yield the lowest readings when not attached to a patient (Thompson & Thompson, 2015).


The Effect of Electrode Location on Common Mode Rejection

Brain activity is more similar when electrodes are close together and less similar when they are farther apart. A differential amplifier may reject actual EEG voltages detected by adjacent electrodes. The sensors were placed at the same anatomical location (Fp1-Fp1) for maximum cancellation, as shown by the flat line in the recording below.





Differential Input Impedance

An amplifier’s differential input impedance further reduces the effect of unequal impedances. As EEG signals enter the amplifier, they are dropped across a network of resistors, presenting a differential input impedance in the Gohm (billion ohms) range. State-of-the-art instruments now exceed 10 Gohms. The differential input impedance must be at least 100 times skin-electrode impedance so that 99% or more of the signal can reach the electroencephalograph.

Why is this important? Stronger signals help an amplifier differentiate EEG activity from noise, producing more accurate feedback.


Sampling the EEG Signal

An analog-to-digital (A/D) converter samples the EEG signal at a fixed sampling interval. The sampling rate is the number of measurements taken within a given period. The sampling rate must be high enough to represent the measured signal accurately.

According to the Nyquist-Shannon sampling theorem, an A/D converter's sampling rate should be at least twice the highest frequency component you intend to sample.

The American Clinical Neurophysiology Society (ACNS) guidelines recommend a minimum sampling rate of at least three times the high-frequency filter setting for digitization. This means at least 100 samples per second (sps) for a 35-Hz high-pass filter and at least 200 sps for a 70-Hz high-pass filter (Halford et al., 2016).

A sampling rate of 128 sps is acceptable for visually inspecting the EEG. A rate of 256 sps is typical, and rates from 500-1000 sps are preferred. The graphic below shows the same EEG signal sampled at 32 and 256 sps. The vertical scale (signal amplitude) is identical for both rates.

Sampling at too slow rates results in aliasing, where an analog signal seems to have a lower frequency than it does. "Phantom" slow activity results from too few samples per second. An 11-Hz signal is sampled at 12 and 200 sps. The 12-sps rate produces an aliasing signal shown in black. Graphic redrawn by minaanandag on Fiverr.com.






Resolution Depends on Bit Depth

An A/D converter's resolution is limited by the smallest signal amplitude it can sample. A bit number is the number of voltage levels an A/D converter can discern. ACNS (Halford et al., 2016) recommends a 16-bit resolution, which can discriminate among 65,536 voltage levels and achieve 0.05-μV resolution. Lower A/D converter resolutions overemphasize small voltage increases.


Signal Properties

EEG signals may be described by their frequency and amplitude. A/D conversion utilizes digital filters to break the EEG into its component frequencies. Prism graphic © kmls/Shutterstock.com.




The movie below is a 19-channel BioTrace+ /NeXus-32 display of EEG activity from 1-64 Hz activity broken into component delta, theta, alpha, and beta frequency bands by digital filters © John S. Anderson.




Recall that frequency is the number of cycles completed each second (Hz). The longer the wavelength, the slower the frequency. The delta, theta, alpha, and beta bands can be defined by wave frequency, wave shape or morphology, and context. EEG graphic © Jeniffer Fontan/Shutterstock.com.





The next graphic illustrates the inverse relationship between wavelength and frequency. The time scale on the horizontal axis is in milliseconds. The amplitude scales are different for the upper (-10 to 10 μV) and lower (-50 to +50 μV) tracings.





The graphic below shows a 9.5-Hz alpha wave. There are 9.5 peaks during a second.






Also, recall that amplitude is signal voltage or power measured in microvolts or picowatts. The alpha wave below has a 20-μV amplitude.





The EEG signal is sent to an integrator to measure signal amplitude in microvolts (μV) or picowatts. Integrators use four methods to calculate the voltage. The peak-to-peak method provides the largest estimate, equivalent to the energy contained between the positive and negative maximum values of the original AC waveform, which is 2 times the peak value. The peak voltage is 0.5 of the peak-to-peak value. The root mean square (RMS) voltage is 0.707 of the peak value and 20% higher than the average voltage. The average voltage is 0.637 of the peak value. Amplitude graphic © Pepermpron/Shutterstock.com.




Conversion among these methods is straightforward. If the peak-to-peak voltage is 20 μV, peak voltage is 10 μV, root mean square voltage is 7.07 μV, and average voltage is 6.37 μV.





Glossary


100 db: dB stands for decibel, a unit used to measure the intensity or level of sound, electrical signal, or other physical quantities. A decibel is a logarithmic unit that expresses the ratio of two values of a physical quantity. 100 dB represents a voltage level that is 100 times greater than the reference level,

active electrode: the electrode that is placed over a site that is a known EEG generator like Cz.

amplitude: signal strength measured in microvolts or picowatts.

analog-to-digital converter (ADC): an electronic device that converts continuous signals to discrete digital values.

bias potential: spurious voltage produced by the exchange of metal ions donated by the electrodes and electrolytes in the absence of a biological current.

bit number: the number of voltage levels that an A/D converter can discern. A resolution of 16 bits means that the converter can discriminate among 65,536 voltage levels.

common-mode rejection ratio (CMRR): the degree by which a differential amplifier boosts signal (differential gain) and artifact (common-mode gain).

differential amplifier (balanced amplifier): a device that boosts the difference between two inputs: the active (input 1) and reference (input 2).

differential input impedance:
the opposition to an AC signal entering a differential amplifier as it is dropped across a resistor network.

digital filter: device that mathematically removes unwanted or extracts valuable aspects of a sampled, discrete-time signal.

electrode: a specialized conductor that converts biological signals like the EEG into currents of electrons.

frequency (Hz): the number of complete cycles that an AC signal completes in a second, usually expressed in hertz.

gain: an amplifier's ability to increase the magnitude of an input signal to create a higher output voltage; the ratio of output/input voltages.

ground electrode: a sensor placed on an earlobe, mastoid bone, or the scalp that is grounded to the amplifier.

input sensitivity: the maximum voltage level an amplifier can accept without producing clipping and distortion.

microvolt (μV): the unit of amplitude (signal strength) that is one-millionth of a volt.

Nyquist-Shannon sampling theorem: the perfect reconstruction of the analog signal requires sampling at two times its highest frequency. A signal whose highest frequency is 1000 Hz should be sampled

peak voltage: 0.5 of the peak-to-peak voltage.

peak-to-peak voltage: the voltage contained between the positive and negative maximum values of the original AC waveform.

phase: the degree to which the peaks and valleys of two waveforms coincide.

phase distortion: the displacement of the EEG waveform in time.

picowatt: billionths of a watt.

polarization: chemical reactions produce separate regions of positive and negative charge where an electrode and electrolyte make contact, reducing ion exchange.

power (W): the rate at which energy is transferred, which is proportional to the product of current and voltage. Power is measured in watts.

quantitative EEG (qEEG): digitized statistical brain mapping using at least a 19-channel montage to measure EEG amplitude within specific frequency bins.

reference electrode: the electrode placed over a less-electrically active site like the mastoid bone behind the ear.

resolution: degree of detail in a biofeedback display (0.1 μV) or the number of voltage levels an A/D converter can discriminate (16 bits or discrimination among 65,536 voltage levels).

sampling rate: the number of samples of a signal that are taken per second to represent the continuous signal digitally. It is measured in hertz (Hz) and is often denoted as samples per second (sps) or kilohertz (kHz).

volume conduction: the movement of biological signals through interstitial fluid.

volt (V): unit of electrical potential difference (electromotive force) that moves electrons in a circuit.

voltage (E): the amount of electrical potential difference (electromotive force).

voltohmmeter: a device that uses a DC signal to measure resistance in an electric circuit, such as between active and reference electrodes.

watt (W): a unit of power used to express signal strength in the qEEG.


TEST YOURSELF ON CLASSMARKER


Click on the ClassMarker logo below to take a 10-question exam over this entire unit.


REVIEW FLASHCARDS ON QUIZLET


Click on the Quizlet logo to review our chapter flashcards.


Visit the BioSource Software Website


BioSource Software offers Physiological Psychology, which satisfies BCIA's Physiological Psychology requirement, and Neurofeedback100, which provides extensive multiple-choice testing over the Biofeedback Blueprint.



Assignment


Now that you have completed this module, explain why bit-depth and sampling rate matter.

References


Andreassi, J. L. (2000). Psychophysiology: Human behavior and physiological response. Lawrence Erlbaum and Associates, Inc.

Basmajian, J. V. (Ed.). (1989). Biofeedback: Principles and practice for clinicians. Williams & Wilkins.

Cacioppo, J. T., & Tassinary, L. G. (Eds.). (1990). Principles of psychophysiology. Cambridge University Press.

Collura, T. F. (2014). Technical foundations of neurofeedback. Taylor & Francis.

Demos, J. N. (2005). Getting started with neurofeedback. W. W. Norton & Company.

Fisch, B. J. (1999). Fisch and Spehlmann's EEG primer (3rd ed.). Elsevier.

Floyd, T. L. (1987). Electronics fundamentals: Circuits, devices, and applications. Merrill Publishing Company.

Grant, A. (2015). Four elements earn permanent seats on the periodic table. Science News.

Halford, J. J., Sabau, D., Drislane, F. W., Tsuchida, T. N., & Sinha, S. R. (2016). American Clinical Society Guideline 4: Recording clinical EEG on digital media. Journal of Clinical Neurophysiology, 33(4), 317-319. https://doi.org/10.1080/21646821.2016.1245563

Hughes, J. R. (1994). EEG in clinical practice (2nd ed.). Butterworth-Heinemann.

Kubala, T. (2009). Electricity 1: Devices, circuits, and materials (9th ed.). Cengage Learning.

Libenson, M. H. (2010). Practical approach to electroencephalography. Saunders Elsevier.

Miller, J. W., Kim, W. S., Homes, M. D., & Vanhatalo, S. (2007). Ictal localization by source analysis of infraslow activity in DC-coupled scalp EEG recordings. NeuroImage, 35(2), 583-597. https://doi.org/10.1016/j.neuroimage.2006.12.018

Montgomery, D. (2004). Introduction to biofeedback. Module 3: Psychophysiological recording. Association for Applied Psychophysiology and Biofeedback.

Nilsson, J. W., & Riedel, S. A. (2008). Electric circuits (8th ed.). Pearson Prentice-Hall.

Peek, C. J. (2016). A primer of traditional biofeedback instrumentation. In M. S. Schwartz, & F. Andrasik (Eds.). (2016). Biofeedback: A practitioner's guide (4th ed.). The Guilford Press.

Stern, R. M., Ray, W. J., & Quigley, K. S. (2001). Psychophysiological recording (2nd ed.). Oxford University Press.

Thompson, M., & Thompson, L. (2015). The biofeedback book: An introduction to basic concepts in applied psychophysiology (2nd ed.). Association for Applied Psychophysiology and Biofeedback.

Wadman, W. J., & Lopes da Silva, F. H. (2011). In D. L. Schomer & F. H. Lopes da Silva (Eds.). Niedermeyer's electroencephalography: Basic principles, clinical applications, and related fields (6th ed.). Lippincott Williams & Wilkins.


C. Instrumentation (Acquisition and Review Parameters/Settings)


Quantitative Electroencephalography (qEEG) is a powerful tool in neuropsychology and neuroscience for analyzing the brain's electrical activity. Ensuring accurate and reliable data collection involves several critical parameters and settings.

Please click on the podcast icon below to hear a full-length lecture over Section C.

First, electrode placement and the number of electrodes are fundamental. Standardized placement using the International 10-20 or 10-10 system is essential for consistent and reproducible electrode locations across subjects and sessions (Klem et al., 1999). Typically, a high-density cap with 19, 32, 64, or more electrodes captures detailed spatial information (Nuwer, 1997).

Second, the sampling rate is another crucial parameter. The minimum sampling rate should be at least 256 Hz, but higher rates, such as 512 Hz or 1024 Hz, are preferred for capturing fast neural dynamics and avoiding aliasing (Miller et al., 2009).

Third, data acquisition settings also play a vital role. Filter settings should include a high-pass filter set around 0.5-1 Hz and a low-pass filter set around 70-100 Hz to remove artifacts and noise outside the frequency range of interest (Luck, 2014). A 50/60 Hz notch filter should also be applied to eliminate power line noise, depending on the regional electrical supply frequency.

Fourth, maintaining low impedance levels is essential for signal quality. Electrode impedances should be kept below 5 kΩ to ensure good signal quality and reduce artifacts, with lower impedance values preferable (Ferree et al., 2001).

Fifth, artifact management is another critical aspect of qEEG. Implementing real-time artifact rejection algorithms can help exclude segments contaminated by eye movements, muscle activity, or electrical noise. Post-processing methods, such as Independent Component Analysis (ICA), are useful for artifact correction (Jung et al., 2000).

Sixth, the choice of reference electrode can significantly affect the results. Common references include linked earlobes, mastoids, or an average reference, and this choice should be consistent across subjects and sessions (Yao, 2001).

Seventh, the recording environment should be controlled to minimize environmental noise and ensure subject comfort. This includes performing recordings in a quiet, electrically shielded room with controlled temperature and lighting.

Finally, proper subject preparation is essential for accurate qEEG data. Subjects should avoid caffeine, heavy meals, and intense physical activity before the recording session. Ensuring they are well-rested and relaxed can help reduce movement artifacts.

These settings and parameters are critical for obtaining high-quality qEEG data, which can provide valuable insights into brain function and dysfunction.

The movie below shows parameters and settings when processing a 19-channel EEG recording in NeuroGuide © John S. Anderson.





EEG Filters Define the Signal

EEG filters select signals of interest and minimize artifacts. In this section, we will review high-pass, low-pass, bandpass, and notch filters. In the graphic below, the range of frequencies passed through a filter is called the passband, and the range that is sharply attenuated is called the stopband. Filter graphic redrawn by minaanandag at Fiverr.com.





A high-pass filter only passes frequencies higher than a set value (e.g., 1 Hz). A low-pass filter only passes frequencies lower than a specified value (e.g., 40 Hz). A bandpass filter passes frequencies between the set values, the "band" of the filter (e.g., 1-40 Hz). Graphic redrawn by minaanandag at Fiverr.com.




10-Hz Low-Pass Filter

The movie below shows the output of a 10-Hz low-pass filter with a vertical scale of 0-50 μV © John S. Anderson.




20-Hz Low-Pass Filter

The movie below shows the output of a 20-Hz low-pass filter with a vertical scale of 0-50 μV © John S. Anderson.





30-Hz Low-Pass Filter

The movie below shows the output of a 30-Hz low-pass filter with a vertical scale of 0-50 μV © John S. Anderson.





40-Hz Low-Pass Filter

The movie below shows the output of a 40-Hz low-pass filter with a vertical scale of 0-50 μV © John S. Anderson.





10-Hz High-Pass Filter

The movie below shows the output of a 10-Hz high-pass filter with a vertical scale of 0-50 μV © John S. Anderson.




20-Hz High-Pass Filter

The movie below shows the output of a 20-Hz high-pass filter with a vertical scale of 0-50 μV © John S. Anderson.





30-Hz High-Pass Filter

The movie below shows the output of a 30-Hz high-pass filter with a vertical scale of 0-50 μV © John S. Anderson.




Bandpass Filters


1-40-Hz Bandpass Filter

The movie below shows the output of a 1-40-Hz bandpass filter with a vertical scale of 0-50 μV © John S. Anderson.






8-12-Hz Bandpass Filter

The movie below shows the output of an 8-12-Hz bandpass filter with a vertical scale of 0-50 μV © John S. Anderson.





The movie below shows the output of three bandpass filters for delta, theta, and alpha © John S. Anderson.




The movie below, generously provided by John S. Anderson, shows a "raw" or "wave" display of oscillating electrical information using a positive/negative scale with 0.0 in the middle with the voltage displayed as peak-to-peak μV.





The movie © John S. Anderson shows the same alpha waveform plotted along two scales. The top display plots the waveform on a scale that ranges from -20 to +20 μV. The bottom "amplitude" display plots the voltage on a scale that ranges from 0- 50 μV where all values are positive.






The movie © John S. Anderson shows the conversion of the complex EEG signal into its spectral components.







The movie © John S. Anderson shows the spectrum magnitude (average amplitude over a given time) in the top display and power (μV2) in the bottom display.





The movie © John S. Anderson shows the same alpha activity displayed in terms of amplitude (positive voltages), power or amplitude2 (picowatts/resistance), and percent power (signal power as a percentage of total EEG power from 0-100%).






Notch Filter

A notch filter suppresses a narrow band of frequencies produced by line current (e.g., 50/60Hz artifact). The stopband is the range of frequencies attenuated by a notch filter. Use notch filters as a last resort. Stop band graphic © Pepermpron/Shutterstock.com.



The narrated video below © John S. Anderson displays the same 21-channel recording viewed using different montages with a 60-Hz notch filter on and off.





Digital Filters

Digital filters, like a digital signal processing (DSP) chip, use digital processors to exclude unwanted frequencies. First, an analog-to-digital converter (ADC) samples and digitizes the analog signal, representing signal voltages as binary numbers. Second, a DSP chip performs calculations on the binary numbers. Third, a digital-to-analog converter (DAC) may transform the sampled, digitally-filtered signal back to analog form.

Three main methods of digital filtering are Fast Fourier Transformation (FFT), finite impulse response (FIR), and infinite impulse response (IIR).

FFT filters convert the EEG signal into a set of sine waves that vary in frequency, amplitude, and phase.

FIR filters have a finite-duration impulse response and calculate a moving weighted average of digitally-sampled voltages.

IIR filters have an infinite impulse response and employ feedback to calculate a moving weighted average of digitally-sampled voltages.

FFT, FIR, and IIR methods enjoy four advantages over analog filters. First, a clinician can retrospectively adjust the filter settings as they review the EEG record since digital filters are programmable. Second, digital filters can be designed to minimize phase distortion (displacement of the EEG waveform in time). Third, digital filters are stable over time and across a range of temperatures. Fourth, digital filters accurately process low-frequency signals. Graphic © Fouad A. Saad/Shutterstock.com shows the digital reconstruction of an analog waveform.





Since these three digital filtering methods can yield different statistical values, they cannot be used interchangeably. Only compare FFT statistics with themselves and not FIR or IIR statistics (Thompson & Thompson, 2016).

Below is a BioGraph ® Infiniti EEG three-dimensional FFT display. Frequency is displayed on the X-axis, amplitude on the Y-axis, and time on the Z-axis.





Glossary


amplitude: signal strength measured in microvolts or picowatts.

bandpass filter: the filter that passes frequencies between the set values, the "band" of the filter (e.g., 1-40 Hz).

digital filter: device that mathematically removes unwanted or extracts valuable aspects of a sampled, discrete-time signal.

finite impulse response (FIR) filter: filter with a finite-duration impulse response.

impedance test: the automated or manual measurement of skin-electrode impedance.

infinite impulse response (IIR) filter: a filter with an infinite impulse response and employ feedback as they calculate a moving weighted average of digitally-sampled voltages.

input sensitivity: the maximum voltage level an amplifier can accept without producing clipping and distortion.

low-pass filter: a filter that only passes frequencies lower than a set value (e.g., 40 Hz).

microvolt (μV): the unit of amplitude (signal strength) that is one-millionth of a volt

notch filter: a filter that suppresses a narrow band of frequencies, such as those produced by line current at 50/60Hz.

passband: the range of frequencies that is passed through a filter.

picowatt: billionths of a watt.

power (W): the rate at which energy is transferred, which is proportional to the product of current and voltage. Power is measured in watts.

stopband: the range of frequencies that is sharply attenuated by a filter.

watt (W): a unit of power used to express signal strength in the qEEG.


TEST YOURSELF ON CLASSMARKER


Click on the ClassMarker logo below to take a 10-question exam over this entire unit.


REVIEW FLASHCARDS ON QUIZLET


Click on the Quizlet logo to review our chapter flashcards.


Visit the BioSource Software Website


BioSource Software offers Physiological Psychology, which satisfies BCIA's Physiological Psychology requirement, and Neurofeedback100, which provides extensive multiple-choice testing over the Biofeedback Blueprint.



Assignment


Now that you have completed this module, in which units does your data acqusition system measure signal power?

References


Andreassi, J. L. (2000). Psychophysiology: Human behavior and physiological response. Lawrence Erlbaum and Associates, Inc.

Basmajian, J. V. (Ed.). (1989). Biofeedback: Principles and practice for clinicians. Williams & Wilkins.

Cacioppo, J. T., & Tassinary, L. G. (Eds.). (1990). Principles of psychophysiology. Cambridge University Press.

Collura, T. F. (2014). Technical foundations of neurofeedback. Taylor & Francis.

Demos, J. N. (2005). Getting started with neurofeedback. W. W. Norton & Company.

Fisch, B. J. (1999). Fisch and Spehlmann's EEG primer (3rd ed.). Elsevier.

Floyd, T. L. (1987). Electronics fundamentals: Circuits, devices, and applications. Merrill Publishing Company.

Grant, A. (2015). Four elements earn permanent seats on the periodic table. Science News.

Halford, J. J., Sabau, D., Drislane, F. W., Tsuchida, T. N., & Sinha, S. R. (2016). American Clinical Society Guideline 4: Recording clinical EEG on digital media. Journal of Clinical Neurophysiology, 33(4), 317-319. https://doi.org/10.1080/21646821.2016.1245563

Hughes, J. R. (1994). EEG in clinical practice (2nd ed.). Butterworth-Heinemann.

Kubala, T. (2009). Electricity 1: Devices, circuits, and materials (9th ed.). Cengage Learning.

Libenson, M. H. (2010). Practical approach to electroencephalography. Saunders Elsevier.

Miller, J. W., Kim, W. S., Homes, M. D., & Vanhatalo, S. (2007). Ictal localization by source analysis of infraslow activity in DC-coupled scalp EEG recordings. NeuroImage, 35(2), 583-597. https://doi.org/10.1016/j.neuroimage.2006.12.018

Montgomery, D. (2004). Introduction to biofeedback. Module 3: Psychophysiological recording. Association for Applied Psychophysiology and Biofeedback.

Nilsson, J. W., & Riedel, S. A. (2008). Electric circuits (8th ed.). Pearson Prentice-Hall.

Peek, C. J. (2016). A primer of traditional biofeedback instrumentation. In M. S. Schwartz, & F. Andrasik (Eds.). (2016). Biofeedback: A practitioner's guide (4th ed.). The Guilford Press.

Stern, R. M., Ray, W. J., & Quigley, K. S. (2001). Psychophysiological recording (2nd ed.). Oxford University Press.

Thompson, M., & Thompson, L. (2015). The biofeedback book: An introduction to basic concepts in applied psychophysiology (2nd ed.). Association for Applied Psychophysiology and Biofeedback.

Wadman, W. J., & Lopes da Silva, F. H. (2011). In D. L. Schomer & F. H. Lopes da Silva (Eds.). Niedermeyer's electroencephalography: Basic principles, clinical applications, and related fields (6th ed.). Lippincott Williams & Wilkins.


D. Montages


Peer-reviewed evidence suggests that more EEG channels provide a more accurate assessment and achieve superior clinical or performance outcomes than fewer channels (Lau et al., 2012).

Please click on the podcast icon below to hear a full-length lecture over Section D.


Although some practitioners conduct assessments and training with a single channel, assessments and training methods that use more than one channel have become more available. For example, the cost of a full 19-channel EEG assessment has decreased substantially. Such an assessment can provide EEG amplitude data from all sites in the 10-20 system and calculate metrics such as coherence and phase that give information on how well the 10-20 sites communicate. These data from multi-channel methods are beneficial for complex symptom profiles like those associated with Autism Spectrum Disorders, epilepsy, and traumatic brain injury (Thompson & Thompson, 2016). Graphic © Chaikom/Shutterstock.com.




A channel is an EEG amplifier output resulting from scalp electrical activity from three electrode/sensor connections to the scalp. These sensors are known as active, reference, and ground electrodes, though they are more appropriately called positive +, negative - and reference. They are placed on the head in the following manner: an active or positive electrode is placed over a known EEG generator like Cz. A reference or negative electrode may be located on the scalp, earlobe, or mastoid. A ground/reference electrode may also be placed on an earlobe or mastoid (Thompson & Thompson, 2015).

Active and reference sensors are identical balanced inputs that are interchangeable. However, some neurofeedback data acquisition systems require the designation of a specific sensor as a "reference," as in a linked-ears reference.

The graphic below was redrawn from John Demos' BCIA-recommended Getting Started with EEG Neurofeedback (2nd ed.). The ear references are connected as a common reference for the four active electrodes (F7, T3, T4, and T5).





A derivation is the assignment of two electrodes to an amplifier's inputs 1 and 2. For example, Fp1 to O2 means that Fp1 is placed in input 1 and O2 in input 2.

A montage groups electrodes together (combines derivations) to record EEG activity (Thomas, 2007).

All montages compare EEG activity between one or more pairs of electrode sites.

Modern amplifiers record all input sensors in reference to a common sensor - often Cz - and all montage (sensor comparison) changes are performed in the software. Amplifiers no longer require manual switching of electrodes between inputs.

The narrated video below © John S. Anderson displays the same 21-channel recording viewed using different montages with a 60-Hz notch filter on and off.




Montage Options and Their Consequences


Referential (Monopolar) Montage

A referential (monopolar) montage places one active electrode (A) on the scalp and a "neutral" reference (R) and ground (G) on the ear or mastoid. Graphic adapted by minaanandag on fiverr.com.




A referential montage assumes that the EEG activity seen on the computer screen represents the active (+) site because the reference (-) site is assumed to be neutral (i.e., producing no EEG activity) and because of the subtraction of signals produced by noise and artifacts that are common to both active and reference sites (common-mode rejection).

The graphics below were redrawn from John Demos' BCIA-recommended Getting Started with EEG Neurofeedback (2nd ed.). In the right diagram, the active electrode "sees" 7 microvolts while the reference "sees" 0 microvolts for a total voltage of 7 microvolts.






In the photograph below, the blue cable would be used for the active electrode, the yellow cable with an ear clip for reference, and the black cable with an ear clip for the ground.



However, this montage is vulnerable to artifacts from the contraction of facial muscles (Demos, 2019).  The ear reference is also known to produce reference contamination, where EEG signals from this electrode are contributed or added to other electrodes via the mechanism of the differential amplifier, where anything different between the "active" and "reference" sensors is retained. This commonly results in alpha activity produced by posterior alpha sources close to the ear.

The graphic below was redrawn from John Demos' BCIA-recommended Getting Started with EEG Neurofeedback (2nd ed.). A differential amplifier rejects the common voltage (e.g., 3 feet) and outputs the voltage difference (e.g., 4 feet). A single-ended amplifier outputs the entire voltage (e.g., 7 feet, EEG artifact, and signal value).





Sequential (Bipolar) Montage

A sequential (bipolar) montage presents a sequence of comparisons of positive (+) and negative (-) electrodes (often called ‘active’ and ‘reference’) that are attached to sites on the scalp. Therefore, the reference electrode is considered to be a second active electrode. The ground (G) electrode is attached to the scalp, to an earlobe, or over the mastoid process. Graphic adapted by minaanandag on fiverr.com.




The sequential (bipolar) montage detects the difference in EEG between the positive and the negative electrodes (active and reference), as the referential montage does. Still, now the signal for the channel represents the difference between the two sources of EEG activity.

The graphics below were redrawn from from John Demos' BCIA-recommended Getting Started with EEG Neurofeedback (2nd ed.). In the diagram on the right, the active "sees" 7 microvolts while the reference "sees" 3 microvolts. A differential amplifier subtracts these voltages, leaving 4 microvolts.



In cases when 19 channels are used, this montage is usually presented with electrode pairs shown in sequence. Note that only the black cable for the ground has an ear clip, as shown in the photograph below. 



When used as a single channel, this montage does not detect localized EEG activity well because it shows only the difference between the A and R signals. However, when used as part of a 19-channel assessment, it localizes EEG events related to epilepsy. This montage can also reduce artifacts when the A and R electrodes are relatively close.

A sequential montage is frequently used in neurofeedback and trains the difference between EEG activity at the A and R electrodes. However, when neurofeedback training produces a change, it remains uncertain whether it is because of a change in EEG at the A, R, or both electrodes.

The EEG record appears more similar when sensors are closer together and less similar when they are farther apart (Fp1-O2). When electrodes are spaced close together (Fp1-Fp2), this montage may reject actual EEG activity (Fp1-Fp1). Graphic © John S. Anderson.





Montages for 21 Recording Electrodes



Please click on the podcast icon below to hear a full-length lecture over Section D.


All montages compare EEG activity between one or more pairs of electrode sites.

The choice of montage does not alter the raw cortical electrical activity itself, but rather filters, enhances, or diminishes aspects of it based on the spatial relationship between electrodes and the dipolar sources within the brain.

A montage defines how each EEG channel is constructed by determining which electrodes are compared to each other. Broadly, montages fall into two categories: bipolar and referential.

In a bipolar montage, each channel represents the voltage difference between two adjacent or anatomically aligned electrodes. This configuration emphasizes local voltage gradients and is especially useful for identifying focal abnormalities, such as epileptiform discharges, through phase reversals—a key diagnostic feature. The longitudinal bipolar montage, also known as the "double banana," arranges channels along the anterior-posterior axis, while the transverse montage organizes them across the coronal plane from left to right. Both are classic examples of bipolar configurations and are commonly employed in clinical EEG for their effectiveness in spatial localization and waveform morphology analysis.

In contrast, a referential montage displays the voltage at each active electrode relative to a common reference point. This reference may be a single electrode, a pair of electrodes, or a mathematically computed value such as the average of all electrodes. Referential montages are particularly suited for assessing global brain activity, hemispheric asymmetries, and background rhythms. Examples include the average reference montage, where each electrode is referenced to the mean of all electrodes; the Cz reference montage, which uses the vertex (Cz) as the fixed reference; and the linked ears montage, which references all electrodes to the average of the earlobes (A1 and A2). The Laplacian montage is a specialized form of referential montage in which each electrode is compared to a weighted average of its immediate neighbors, enhancing the spatial resolution of focal activity.

This section will examine the structure, clinical utility, and interpretive implications of each of these montages. By understanding the fundamental distinction between bipolar and referential configurations—and the strengths and limitations of specific montages within these categories—clinicians can select the most appropriate montage for a given diagnostic question. Furthermore, the ability to apply multiple montages to the same EEG data set, known as re-montaging, enables a more comprehensive and nuanced analysis, ultimately enhancing the accuracy of neurophysiological interpretation in both routine and complex clinical contexts.

Longitudinal Bipolar Montage

The longitudinal bipolar (double banana) montage, is one of the most widely employed configurations in clinical electroencephalography. It consists of a series of bipolar derivations arranged in parallel, linking electrodes from the frontal to the occipital poles along the midline and lateral chains of the scalp. For example, one chain may include the derivations Fp1–F3, F3–C3, C3–P3, and P3–O1, forming a continuous pathway along the left hemisphere. A mirror set is applied on the right: Fp2–F4, F4–C4, C4–P4, and P4–O2. These pairings create an electrical map that is optimal for detecting activity propagating along the anterior-posterior axis of the cerebral cortex.








Strengths

The primary strength of this montage lies in its capacity to enhance phase reversals, which are the inversion points of waveform polarity along the bipolar chain and serve as a crucial diagnostic cue for localizing the maximum field of a focal discharge. When a sharp wave or spike occurs in a cortical region sampled between two electrodes, the voltage gradient between them becomes pronounced, resulting in an easily identifiable phase reversal that pinpoints the site of maximal activity. This makes the montage particularly adept at detecting focal epileptiform activity, especially when it is organized along the sagittal plane (Niedermeyer & da Silva, 2005). The montage is also intuitive and standardized, which facilitates inter-reader reliability across clinical institutions.

Limitations

However, the longitudinal bipolar montage is not without limitations. Because each channel is composed of a bipolar derivation, activity that is spatially generalized across adjacent electrode pairs—such as in generalized epileptic discharges or diffuse slowing—may be attenuated or even cancelled out, particularly if it appears in phase across a chain. Additionally, activity that is vertically or obliquely oriented relative to the anteroposterior axis may be poorly resolved due to the montage’s directional bias. Furthermore, the montage may be less sensitive to activity with a deep cortical origin, such as from mesial temporal structures, which may not produce robust potentials at the scalp surface.


Transverse Montage

The transverse (coronal bipolar) montage is an alternative bipolar configuration in which electrodes are linked across the coronal plane of the scalp, from lateral to medial or from one hemisphere to the other. For example, a typical chain might include F7–F3–Fz–F4–F8 across the frontal row, followed by a second chain of T3–C3–Cz–C4–T4 at the central level, and posteriorly T5–P3–Pz–P4–T6, extending through the parietal and temporal regions. These linkages are designed to visualize activity spreading horizontally across the hemispheres or in a left-right configuration.






Strengths

The transverse montage is especially advantageous in situations where lateralization is of clinical concern. It is often used as a complementary montage to the longitudinal bipolar in epilepsy workups, as it can delineate whether a waveform or discharge is restricted to one hemisphere, crosses the midline, or is more pronounced on a particular side. By emphasizing horizontal field distribution, the transverse montage improves detection of temporal lobe discharges, which often extend laterally and may appear less prominent in a longitudinal montage. Moreover, the transverse montage is adept at highlighting hemispheric asymmetries in background rhythm and interictal epileptiform activity, offering a different spatial perspective than traditional sagittal derivations.


Limitations

Nonetheless, this montage also presents specific challenges. Its less common use makes it less familiar to many general neurologists, requiring a higher level of interpretive skill and spatial visualization. The absence of anterior-posterior connections can obscure phase reversals along that axis, making it less effective for localizing discharges that originate in midline or parasagittal regions. In addition, the reliance on lateral electrodes introduces greater sensitivity to artifact from facial and temporal muscle activity, which can be a confounding factor in certain recordings.


Average Reference Montage

The average reference montage is conceptually different from bipolar montages in that it is constructed by referencing each active electrode to the mean potential of all scalp electrodes. Mathematically, the signal from each electrode is recalculated as the difference between that electrode and the averaged signal from the full array. The underlying assumption is that the sum of all scalp-recorded potentials approximates zero, creating a neutral reference. In practice, this configuration is implemented in modern digital EEG systems using real-time computational averaging across electrodes, excluding those contaminated by artifact or poor contact. The image below was adapted from Lopez et al. (2017). In most cases, the midline electrodes are also included in these calculations.






Strengths

The primary advantage of the average reference montage is its reference-free nature, which provides a theoretical spatial neutrality. This can be particularly useful in visualizing diffuse, low-amplitude cerebral activity, including generalized spike-wave discharges, slow wave abnormalities, and subtle background fluctuations. Because each electrode is displayed with respect to the same computed reference, spatial comparisons across channels are more direct, making this montage highly valuable in quantitative EEG (qEEG) and source localization studies (Fisch, 1999).


Limitations

However, the montage is also highly susceptible to contamination by a single noisy electrode. If one electrode has a high amplitude artifact (e.g., due to muscle or movement), this can affect the computed average, thereby distorting every channel in the montage. Moreover, the assumption that the average voltage equals zero is only valid when electrodes are evenly distributed and the cortical activity is spatially balanced—conditions rarely met in practice. In the presence of significant focal abnormalities, the average reference can over- or under-estimate the true field of the activity and may even introduce artifactual inversions or misleading asymmetries (Loddenkemper et al., 2014).


Cz Reference Montage

The Cz reference montage is a unipolar configuration in which every scalp electrode is referenced to a single, fixed electrode located at the vertex of the scalp—Cz, the central midline location. This configuration is straightforward and provides consistent spatial orientation for each channel, allowing the reader to compare the amplitude and waveform morphology at different sites relative to a central standard. It is commonly used in event-related potential (ERP) studies, sleep scoring, and pediatric EEG, where ease of interpretation and temporal clarity are prioritized. The image below was adapted from Lopez et al. (2017). In most cases, the midline electrodes are also included in these calculations.




Strengths

One of the key benefits of the Cz reference montage is its ability to reveal lateralized discharges. Because Cz is equidistant from the hemispheres, discharges arising from left or right hemispheres produce strong voltage differences and clear waveforms. The fixed reference also facilitates temporal alignment of discharges, which is useful in assessing spike timing, propagation patterns, and hemispheric synchrony.


Limitations

Nonetheless, Cz is a problematic reference point for activity that originates near the vertex itself, such as discharges from the supplementary motor area, parasagittal cortex, or midline regions. Since the reference is physically close to the source, the recorded potential difference may be minimal or even absent, leading to false negatives. Furthermore, Cz may itself pick up cortical signals, which undermines the assumption of a neutral reference, particularly during periods of midline activation. This limitation becomes particularly relevant when interpreting generalized or midline spike-wave complexes or high-frequency sleep spindles that originate near central sites (Niedermeyer & da Silva, 2005).



Linked Ears Reference (A1-A2 reference or LE) Montage

The linked ears reference montage is a referential EEG configuration in which all scalp electrodes are referenced to a common average of the left and right earlobe electrodes (A1 and A2). These electrodes are presumed to be electrically inactive relative to cerebral sources, providing a stable baseline for assessing cortical activity. In practice, the average of A1 and A2 serves as a reference for each channel, offering a consistent point of comparison across the scalp. This montage has historically been widely used in both routine clinical EEG and intraoperative monitoring due to its simplicity and relatively noise-free reference point (Niedermeyer & da Silva, 2005).





Strengths

The primary strength of the linked ears montage lies in its relative ease of interpretation and its ability to provide clear visualization of regional asymmetries. Because all electrodes are referenced to a common point, focal abnormalities such as interictal epileptiform discharges are often well-demarcated, especially when lateralized. It is also less susceptible to widespread muscle artifacts that may affect central or midline reference points, allowing for improved identification of background rhythms (Nuwer, 1997). Additionally, its long-standing use in clinical practice has made it familiar to many neurologists and EEG technologists, contributing to reproducibility and interpretative consistency.


Limitations

However, the montage is not without limitations. The assumption that the earlobes are electrically inactive is not always valid; in some cases, especially in infants or during high-amplitude discharges, the ears may pick up cortical activity, introducing spurious signals into the reference (Gloor, 1985). Furthermore, asymmetry in earlobe impedance or local artifact can lead to lateralized distortions that mimic or obscure true cerebral asymmetries. The linked ears reference also tends to exaggerate posterior alpha activity and may underrepresent midline or deep sources due to its relatively lateral reference position. As such, while useful in many scenarios, this montage should be interpreted in conjunction with other configurations to ensure accurate localization and characterization of EEG findings.



Laplacian Montage

The Laplacian montage is a type of spatially weighted referential EEG montage designed to enhance the localization of cortical activity by emphasizing signals originating directly beneath each electrode while attenuating those from more distant sources. In this configuration, each electrode is referenced to a weighted average of the surrounding electrodes, thereby approximating the second spatial derivative of the potential field. This mathematical approach enhances local voltage differences and minimizes the effects of broadly distributed or volume-conducted signals. As a result, the Laplacian montage provides improved spatial resolution and is particularly useful in research applications and advanced clinical settings where precise source localization is required (Nunez & Srinivasan, 2006).







Strengths

The Laplacian montage offers significant advantages in localizing focal cortical activity with high spatial precision. Unlike referential montages that rely on ear or mastoid electrodes, the Laplacian configuration excludes these distant references, thereby reducing their influence and minimizing the risk of introducing artifact from electrically active or asymmetrically placed reference sites. Instead, it calculates a spatially weighted average of the surrounding electrodes to serve as the reference for each site, emphasizing the electrical potential directly beneath the electrode of interest. This approach suppresses volume-conducted and broadly distributed signals, allowing for clearer isolation of local cortical activity.

As a result, the Laplacian montage enhances the visibility of subtle, focal EEG features that may be obscured in standard referential or bipolar configurations. It is particularly effective for detecting focal epileptiform discharges, localized slowing, and sensorimotor rhythms, especially when applied in high-density EEG systems where spatial sampling is sufficient to support accurate Laplacian calculations (McFarland et al., 1997). Furthermore, by minimizing global signals and common-mode noise, it offers greater resistance to artifacts from eye movements, muscle activity, and non-neuronal sources such as the diffuse effects of sedative medications or metabolic encephalopathy. These characteristics make the Laplacian montage especially useful in advanced clinical evaluations, functional brain mapping, and neurophysiological research where precise localization is critical. While not typically used as a primary montage in standard EEG interpretation, it serves as a valuable adjunctive tool for refining diagnostic accuracy in complex or ambiguous cases.


Limitations

However, the Laplacian montage also presents notable limitations. Its effectiveness is highly dependent Despite its advantages, the Laplacian montage has important limitations that affect its clinical application. Its accuracy depends on high-density, evenly spaced electrode arrays, making it less reliable in standard low-density systems like the 10–20 montage. A key concern is the edge effect—electrodes at the scalp periphery, such as Fp1, Fp2, F7, F8, O1, and O2, lack surrounding electrodes on all sides, reducing the precision of spatial averaging in these regions. Inadequate electrode coverage can distort rather than enhance localization of cortical signals. Additionally, in some instances, the montage may appear to introduce or redistribute EEG activity, though this is difficult to verify due to the influence of the reference system and signal processing. Furthermore, the spatial filtering inherent to the Laplacian method attenuates slow or widespread activity, limiting its sensitivity to generalized abnormalities such as diffuse slowing or generalized spike-wave discharges (Gordon & Rzempoluck, 2004; Srinivasan et al., 1996). Therefore, while highly effective for focal analysis, the Laplacian montage is best used as a complementary method alongside conventional configurations in clinical EEG interpretation.


Montage Selection Strategy

Obviously, our goal when viewing the EEG is to identify areas that deviate from typical behavior, correlate those differences with client symptoms, and design a training protocol or group of protocols to address those differences.

The longitudinal bipolar and transverse montages are most effective in demonstrating the key electrographic features of focal abnormalities—namely, clear localization, polarity, and sharp morphology—making them essential in identifying epileptiform discharges and delineating their spatial distribution. The average reference montage, while less sensitive to focal events, contributes significantly to the assessment of hemispheric symmetry and background rhythm integrity, particularly in diffuse or generalized processes. The Cz reference montage provides a reliable lateralized perspective and is especially useful for differentiating left versus right hemispheric activity; however, it lacks optimal resolution for detecting midline sources due to its placement directly over the central vertex.

The linked ears reference montage, commonly used in routine clinical EEG, offers consistency in visualizing lateralized abnormalities and is less susceptible to generalized artifact. However, its reliance on ear electrodes as a common reference introduces potential for asymmetry and contamination, especially in pathological conditions or with poor electrode impedance. In contrast, the Laplacian montage offers superior spatial resolution for identifying focal cortical sources by emphasizing local field potentials and suppressing distant or volume-conducted activity. It is particularly advantageous in high-density EEG systems and functional mapping but requires careful interpretation due to limitations in edge accuracy and reduced sensitivity to generalized abnormalities.

Understanding the comparative strengths and limitations of each montage—bipolar for sharpness and polarity, average reference for symmetry and background, Cz for lateralization, linked ears for routine focal analysis, and Laplacian for high-resolution localization—enables a more nuanced and precise interpretation of EEG findings. This integrative approach enhances diagnostic accuracy and supports informed clinical decision-making in epilepsy evaluation and broader neurophysiological assessment.


Re-Montaging

In clinical electroencephalography, the montage used to display EEG data fundamentally shapes how brain activity is visualized and interpreted. Each montage, whether longitudinal bipolar, transverse, average reference, or Cz reference, has distinct strengths and limitations. However, an essential feature of modern EEG analysis is that clinicians are not limited to a single montage when evaluating an EEG recording. Because digital EEG systems preserve the raw voltage potential at each electrode, clinicians can reconfigure how these signals are displayed by applying different montages to the same epoch of data. This process, known as re-montaging, allows for a more complete and nuanced understanding of the recorded brain activity.

Re-montaging enables clinicians to re-express the same underlying electrical signals in different spatial contexts. A particular waveform or discharge may appear more prominent, localized, or sharply contoured in one montage and more diffuse or less distinct in another. For example, a spike that is clearly localized with a phase reversal in a longitudinal bipolar montage may appear attenuated or blended with surrounding activity in an average reference montage. Conversely, generalized discharges may be more clearly appreciated in referential montages, such as the average or Cz reference, whereas bipolar montages may obscure them due to cancellation effects.

Clinicians re-montage not only to improve visualization of specific events but also to clarify ambiguous findings. A sharply contoured waveform seen in a bipolar montage can be cross-checked in a referential montage to determine whether it reflects true cortical activity or a muscle artifact. Similarly, rhythmic slowing that appears focal in one configuration may prove to be part of a more diffuse process when re-displayed using another montage. This comparative approach helps avoid misinterpretation, particularly in cases where benign variants or technical artifacts could otherwise be mistaken for pathologic activity.

Re-montaging also supports more accurate localization. Some montages are aligned along specific anatomical axes and are therefore more sensitive to discharges propagating in particular directions. The longitudinal bipolar montage, for instance, follows the anterior-posterior axis and is well-suited for identifying phase reversals in that orientation. However, it may not capture lateral propagation across hemispheres as effectively as the transverse montage. By re-montaging into both configurations, a clinician can more precisely triangulate the source of epileptiform discharges or assess whether a waveform is confined to one hemisphere or crosses the midline.

Moreover, the ability to re-montage is indispensable in certain clinical scenarios. When evaluating patients for focal epilepsy, the need to localize seizure onset to a particular lobe or hemisphere is paramount. Re-montaging can reveal subtle asymmetries or focal features that might not be visible in the original display. In cases of suspected encephalopathy, re-montaging into an average reference montage may better demonstrate diffuse slowing or triphasic waves. In sleep EEGs, where vertex sharp waves and sleep spindles may be crucial, the choice of reference becomes especially important, as certain references may suppress or distort these midline phenomena.

Digital EEG systems have made re-montaging straightforward and immediate. Rather than being constrained by the montage selected during acquisition, clinicians can explore different configurations at any point during interpretation. This flexibility is endorsed by the American Clinical Neurophysiology Society, which recommends the use of both bipolar and referential montages and emphasizes the need for clarity, simplicity, and interpretability in montage design. Their guidelines further encourage the use of at least 16 recording channels and the full complement of 10–20 system electrodes to ensure comprehensive spatial sampling and re-montage capability.

Ultimately, the purpose of EEG is to accurately identify and characterize brain activity, whether normal or abnormal, diffuse or focal, transient or rhythmic. No single montage can fully reveal all aspects of this activity. By using multiple montages to re-express the same epoch of EEG, clinicians can compensate for the spatial and referential limitations of individual configurations. This approach allows for the confirmation of findings across montages, facilitates more precise localization, and strengthens diagnostic confidence. The ability to re-montage is not just a technical convenience, it is a critical component of rigorous EEG interpretation and a powerful tool in clinical neurophysiology.

When we identify an EEG feature representing slowed alpha, high amplitude frontal theta activity, excess fast activity, lack of a typical alpha response, or any other finding in one montage, we want to verify and validate that finding using additional montages.

Following the average reference montage, the Laplacian montage can be useful for further zeroing in on the areas of interest.

Finally, the linked ears montage must be consulted if only to identify areas of likely reference contamination that may affect subsequent topographic z-score maps, phase, coherence and network analyses, and other downstream evaluations.


Use Consistent Settings

One important consideration for viewing the EEG is to use consistent settings. The EEG is often viewed using a 50 μV y-scale (vertical axis) and a 30 mm per second x-axis (horizontal) equivalent "chart speed" display.

In the early days of EEG, tracings were drawn by pens suspended over moving chart paper. A typical speed for adult EEG was 30 mm per second and for pediatric EEG was 15 mm per second. Now that almost all EEG is digital, similar equivalent displays show the EEG in this format. This is so the waves appear consistently the same every time they are viewed and can also be compared to reference sources. If different visual display settings are used, then it is difficult to identify wave patterns and abnormally high or low voltage values by visual inspection.

Most modern EEG software uses display time settings indicated in seconds rather than chart speed equivalents. In the NeuroGuide database, the setting that appears most similar to 30 mm per second is 10 seconds per page. Also, in this program, the y-scale setting adjusts automatically depending on the highest voltage present in the recording. It doesn’t allow for a constant setting, though the desired setting can be made manually each time the montage is changed.











Glossary



50/60 Hz: external artifacts transmitted by nearby electrical sources.

active electrode: an electrode placed over a site that is a known EEG generator like Cz.

amplitude: the strength of the EEG signal measured in microvolt or picowatts.

artifact: false signals like 50/60Hz noise produced by line current.

asynchronous waves: neurons depolarize and hyperpolarize independently.

average reference montage: an EEG referencing technique where each electrode is referenced to the arithmetic mean of all scalp electrode potentials; used to highlight focal activity and reduce common signals.

bipolar montage: a montage in which each EEG channel represents the voltage difference between two adjacent scalp electrodes.

channel: an EEG amplifier output that is the result of scalp electrical activity from three electrode/sensor connections to the scalp.

Cz: the central midline electrode in the 10–20 system, often used as a common reference point in unipolar montages.

derivation: the assignment of two electrodes to an amplifier's inputs 1 and 2.

desynchrony: pools of neurons fire independently due to stimulation of specific sensory pathways up to the midbrain and high-frequency stimulation of the reticular formation and nonspecific thalamic projection nuclei.

differential amplifier (balanced amplifier): a device that boosts the difference between two inputs: the active (input 1) and reference (input 2).

dipolar sources: brain-generated electrical fields that produce opposing voltages detectable at the scalp; critical in EEG localization.

epoch: a time segment of EEG recording, usually lasting a few seconds, used to examine waveform characteristics within a specific interval.

focal activity: EEG signals that originate from a localized area of the brain, as opposed to generalized activity.

frequency (Hz): the number of complete cycles that an AC signal completes in a second, usually expressed in hertz.

ground electrode: a sensor placed on an earlobe, mastoid bone, or the scalp that is grounded to the amplifier.

hertz (Hz): unit of frequency measured in cycles per second.

inion: the bony prominence on the back of the skull.

International 10-10 system: a modified combinatorial system for electrode placement that expands the 10-20 system to 75 electrode sites to increase EEG spatial resolution and improve detection of localized evoked potentials.

International 10-20 system: a standardized procedure for placing 21 recording and one ground electrode on adults.

Laplacian montage: a referential EEG configuration in which each electrode is referenced to a weighted average of its immediately surrounding electrodes. This enhances local cortical activity while suppressing distant or volume-conducted signals, improving spatial resolution for source localization.

linked ears reference montage: a referential EEG configuration in which all scalp electrodes are referenced to the average of the left (A1) and right (A2) earlobe electrodes. This setup assumes the earlobes are electrically neutral and provides a common reference point for assessing cortical activity.

localization: the process of determining the origin of EEG activity within the brain, often aided by montage selection and waveform characteristics.

longitudinal bipolar montage: a bipolar montage where electrodes are connected front-to-back along the anterior-posterior axis; also known as the "double banana" montage.

mastoid bone: the bony prominence behind the ear.

microvolt (μV): the unit of amplitude (signal strength) that is one-millionth of a volt.

monopolar recording: a recording method that uses one active and one reference electrode.

montage: a grouping of electrodes (combining derivations) to record EEG activity.

nasion: the depression at the bridge of the nose.

phase reversal: a change in polarity between adjacent electrodes in a bipolar montage, indicating the likely location of maximal voltage and aiding in source localization.

polarity: the direction of waveform deflection (positive or negative) in EEG, relevant for determining the source and direction of electrical activity.

polarization: chemical reactions produce separate regions of positive and negative charge where an electrode and electrolyte make contact, reducing ion exchange.

posterior dominant rhythm (PDR): the highest-amplitude frequency detected at the posterior scalp when eyes are closed.

preauricular point: the slight depression located in front of the ear and above the earlobe.

Quantitative EEG (qEEG): digitized statistical brain mapping using at least a 19-channel montage to measure EEG amplitude within specific frequency bins.

reference electrode: an electrode placed on the scalp, earlobe, or mastoid.

referential (monopolar) montage: the placement of one active electrode (A) on the scalp and a neutral reference (R) and ground (G) on the ear or mastoid.

re-montaging: the process of re-displaying EEG data using a different montage configuration, allowing clinicians to view the same recorded electrical activity from alternative spatial perspectives to enhance interpretation, localization, and artifact differentiation.

response stereotypy: a person’s unique response pattern to stressors of identical intensity.

sequential (bipolar) montage: placement of active (A) and reference (R) sensors on active scalp sites and the ground (G) to an earlobe or mastoid.

spatial resolution: the ability of an EEG montage to distinguish electrical activity arising from different areas of the brain.

synchrony: the coordinated firing of pools of neurons due to pacemakers and mutual coordination.

tragus: the flap at the opening of the ear.

transverse montage: a bipolar EEG montage that links electrodes horizontally across the head, providing sensitivity to lateralized and horizontal propagation of activity.

vertex (Cz): the intersection of imaginary lines drawn from the nasion to inion and between the two preauricular points in the International 10-10 and 10-20 systems.


TEST YOURSELF ON CLASSMARKER


Click on the ClassMarker logo below to take a 10-question exam over this entire unit.


REVIEW FLASHCARDS ON QUIZLET


Click on the Quizlet logo to review our chapter flashcards.


Visit the BioSource Software Website


BioSource Software offers Physiological Psychology, which satisfies BCIA's Physiological Psychology requirement, and Neurofeedback100, which provides extensive multiple-choice testing over the Biofeedback Blueprint.



Assignment


Now that you have completed this module, summarize the strengths and limitations of a sequential montage.

References


Andreassi, J. L. (2000). Psychophysiology: Human behavior and physiological response. Lawrence Erlbaum and Associates, Inc.

Acharya, J. N., Hani, A. J., Thirumala, P. D., & Tsuchida, T. N. (2016). American Clinical Neurophysiology Society Guideline 3: A Proposal for Standard Montages to Be Used in Clinical EEG. Journal of Clinical Neurophysiology: Official Publication of the American Electroencephalographic Society, 33(4), 312–316. https://doi.org/10.1097/WNP.0000000000000317

Basmajian, J. V. (Ed.). (1989). Biofeedback: Principles and practice for clinicians. Williams & Wilkins.

Breedlove, S. M., & Watson, N. V. (2023). Behavioral neuroscience (10th ed.). Sinauer Associates, Inc.

Cacioppo, J. T., & Tassinary, L. G. (Eds.). (1990). Principles of psychophysiology. Cambridge University Press.

Carvalhaes, C., & Acacio de Barros, J. (2014). The surface Laplacian technique in EEG: Theory and methods. International Journal of Psychophysiology, 97(3), 174–188.

Collura, T. F. (2014). Technical foundations of neurofeedback. Taylor & Francis.

Demos, J. N. (2019). Getting started with neurofeedback (2nd ed.). W. W. Norton & Company.

Ferree, T. C., Luu, P., Russell, G. S., & Tucker, D. M. (2001). Scalp electrode impedance, infection risk, and EEG data quality. Clinical Neurophysiology, 112(3), 536-544. https://doi.org/10.1016/s1388-2457(00)00533-2

Fisch, B. J. (1999). Fisch and Spehlmann's EEG primer (3rd ed.). Elsevier.

Floyd, T. L. (1987). Electronics fundamentals: Circuits, devices, and applications. Merrill Publishing Company.

Garces, A., Laciar, E,, Patiño, H., & Valentinuzzi, M. (2007). Artifact removal from EEG signals using adaptive filters in cascade. Journal of Physics: Conference Series, 90(1), 012081. https://10.1088/1742-6596/90/1/012081

Gloor, P. (1985). EEG analysis and the localization of epileptic foci. In D. Daly & T. Pedley (Eds.), Current practice of clinical electroencephalography (pp. 139–174). Raven Press.Gordon, E., & Rzempoluck, E. (2004). Clinical EEG and neuroscience, 35(2), 57–68.

Grant, A. (2015). Four elements earn permanent seats on the periodic table. Science News.

Halford, J. J., Sabau, D., Drislane, F. W., Tsuchida, T. N., & Sinha, S. R. (2016). American Clinical Society Guideline 4: Recording clinical EEG on digital media. Journal of Clinical Neurophysiology, 33(4), 317-319. https://doi.org/10.1080/21646821.2016.1245563

Hugdahl, K. (1995). Psychophysiology: The mind-body perspective. Harvard University Press.

Hughes, J. R. (1994). EEG in clinical practice (2nd ed.). Butterworth-Heinemann.

Jung, T. P., Makeig, S., Humphries, C., Lee, T. W., McKeown, M. J., Iragui, V., & Sejnowski, T. J. (2000). Removing electroencephalographic artifacts by blind source separation. Psychophysiology, 37(2), 163-178. PMID: 10731767

Khazan, I. Z. (2019). Biofeedback and mindfulness in everyday life: Practical solutions for improving your health and performance. W. W. Norton & Company.

Klass, D. W. (2008). The continuing challenge of artifacts in the EEG. EEG artifacts. American Society of Electroneurodiagnostic Technologists, Inc. https://doi.org/10.1080/00029238.1995.11080524

Klem, G. H., Lüders, H. O., Jasper, H. H., & Elger, C. (1999). The ten-twenty electrode system of the International Federation. Electroencephalography and Clinical Neurophysiology, 52(3), 3-6. PMID: 10590970

Kubala, T. (2009). Electricity 1: Devices, circuits, and materials (9th ed.). Cengage Learning.

Lau, T. M., Gwin, J. T., & Ferris, D. P. (2012). How many electrodes are really needed for EEG-based mobile brain imaging? Journal of Behavioral and Brain Science, 2(3), 387-393. https://doi.org/10.4236/jbbs.2012.23044

Lebby, P. C. (2013). Brain imaging: A guide for clinicians. Oxford University Press.

Libenson, M. H. (2010). Practical approach to electroencephalography. Saunders Elsevier.

Lin, F., Witzel, T., Hamalainen, M. S., Dale, A. M., Belliveau, J. W., & Stufflebeam, S. M. (2004). Spectral spatiotemporal imaging of cortical oscillations and interactions in the human brain. NeuroImage, 2(3), 582-595. https://dx.doi.org/10.1016%2Fj.neuroimage.2004.04.027

López, S., Gross, A., Yang, S., Golmohammadi, M., Obeid, I., & Picone, J. (2016). An analysis of two common reference points for EEGs. IEEE Signal Processing in Medicine and Biology Symposium, 2016, https://doi.org/10.1109/SPMB.2016.7846854 Luck, S. J. (2014). An introduction to the event-related potential technique. MIT press.

McFarland, D. J., McCane, L. M., David, S. V., & Wolpaw, J. R. (1997). Spatial filter selection for EEG-based communication. Electroencephalography and Clinical Neurophysiology, 103(3), 386–394. https://doi.org/10.1016/S0013-4694(97)00022-2

Min, B.-K. (2007). The top-down function of prestimulus EEG alpha activity. Dissertation.

Miller, K. J., Sorensen, L. B., Ojemann, J. G., & den Nijs, M. (2009). Power-law scaling in the brain surface electric potential. PLoS Computational Biology, 5(12), e1000609. https://doi.org/10.1371/journal.pcbi.1000609

Montgomery, D. (2004). Introduction to biofeedback. Module 3: Psychophysiological recording. Association for Applied Psychophysiology and Biofeedback.

Niedermeyer, E., & da Silva, F. L. (2005). Electroencephalography: Basic principles, clinical applications, and related fields (5th ed.). Lippincott Williams & Wilkins.

Nilsson, J. W., & Riedel, S. A. (2008). Electric circuits (8th ed.). Pearson Prentice-Hall.

Nuwer, M. R. (1997). Assessment of digital EEG, quantitative EEG, and EEG brain mapping: Report of the American Academy of Neurology and the American Clinical Neurophysiology Society. Neurology, 49(1), 277-292. https://doi.org/10.1212/wnl.49.1.277

Peek, C. J. (2016). A primer of traditional biofeedback instrumentation. In M. S. Schwartz, & F. Andrasik (Eds.). (2016). Biofeedback: A practitioner's guide (4th ed.). The Guilford Press.

Picton, T. W., & Hillyard, S. A. (1972). Cephalic skin potentials in electroencephalography. Encephalogr Clin Neurophysiol, 33, 419-424. https://doi.org/10.1016/0013-4694(72)90122-8

Pfister, H., Kaynig, V., Botha, C. P., Bruckner, S., Dercksen, V., & Hege, H.-C. (2012). Visualization in connectomics. Mathematics and Visualization, 37. https://doi.org/10.1007/978-1-4471-6497-5_21

Stern, R. M., Ray, W. J., & Quigley, K. S. (2001). Psychophysiological recording (2nd ed.). Oxford University Press.

Thomas, C. (2007). What is a montage? EEG instrumentation. American Society of Electroneurodiagnostic Technologists, Inc.

Thompson, M., & Thompson, L. (2015). The biofeedback book: An introduction to basic concepts in applied psychophysiology (2nd ed.). Association for Applied Psychophysiology and Biofeedback.

Yao, D. (2001). A method to standardize a reference of scalp EEG recordings to a point at infinity. Physiological Measurement, 22(4), 693. https://doi.org/10.1088/0967-3334/22/4/305
 

E. Electrical/Clinical Safety


Electricity makes most biofeedback applications possible. Biological signals like skeletal muscle and cortical voltages are streams of charged atoms or molecules called ions.

Please click on the podcast icon below to hear a full-length lecture over Section E.



The hardware that monitors these signals is powered by batteries or wall outlets that supply currents of electrons. Graphic © Vladimir Popovic/iStockphoto.com.






Without a basic understanding of electricity and the circuits used in biofeedback instruments, we might mistakenly accept readings produced by equipment misuse or breakdown. "Garbage in, garbage out."

BCIA Blueprint Coverage


This unit addresses III. Instrumentation and Electronics - A. Essential Terms and Concepts.
 



This unit covers Basic Terms and Metrics, EEG Recording, and Safety Precautions.

Basic Terms and Metrics


Building Blocks of Matter

The matter comprising our universe occupies space and possesses mass. Matter can assume solid, liquid, gaseous, and plasma states. Graphic © magnetix/Shutterstock.com.








Atoms are basic units of matter consisting of a central nucleus that contains protons, neutrons, and orbiting electrons.

Atom


The positively charged nucleus contains most of an atom's mass in the form of positively charged protons and uncharged neutrons. Each proton carries a positive charge equal and opposite to the electron's negative charge. Negatively charged electrons rotate around the nucleus at varying distances and participate in chemical reactions. The number of electrons equals the number of protons in an atom, balancing the electrical charge of the nucleus. In other words, an atom’s net charge is zero, and an atom is said to be neutral. Graphic © Designua/Shutterstock.com depicts a carbon atom.


>





Elements contain identical atoms and cannot be reduced by common chemical reactions. Of the 118 elements confirmed to date, calcium (Ca), carbon (C), hydrogen (H), nitrogen (N), oxygen (O), and phosphorous (P) are the most important to human life (Grant, 2015). Calcium (Ca), chloride (Cl), potassium (K), and sodium (Na) are critical to generating physiological potentials like the EEG. Elements are neutrally charged since their atoms contain equal protons and electrons. Graphic © Designua/ Shutterstock.com.






How does a carbon atom differ from a sodium atom? The difference lies in the number of protons located in the nucleus. A carbon atom has 6 protons, while a sodium atom has 11. The total number of protons determines the atomic number. The number of protons and neutrons approximates the atomic weight.

Ions are atoms or molecules charged by the gain or loss of electrons. The biological potentials produced by cortical neurons (EEG), eccrine sweat glands (EDA), and skeletal muscles (SEMG) are currents of ions. The ions most responsible for these signals are chloride (Cl-), potassium (K+), and sodium (Na+).

terms



Electric Current

Charge (Q) indicates the imbalance between positively and negatively charged particles in a given place or between two locations. Charge is measured in coulombs. Lightning graphic © New Africa/shutterstock.com.





Current (I) is the movement of electrons through a conductor. Current flows because atoms and molecules contain two types of electrical charge: positive and negative. Opposite charges attract while identical charges repel each other. When there is a difference in the overall charge of atoms between two points—for example, between two ends of a wire—negatively charged electrons will flow toward the positively charged end of the wire, creating an electric current.


Listen to a mini-lecture on Current © BioSource Software LLC. Graphic © Designua/Shutterstock.com.







Electrons are also affected by the materials in their path. Conductors like copper allow electron movement, while the insulators enclosing the wires oppose their movement.




Listen to a mini-lecture on Conductors © BioSource Software LLC. Graphic © demarcomedia/Shutterstock.com.



Conductors and Insulators

Biological signals like the EEG travel through interstitial fluid, which is the fluid between the cells in our bodies. Signals like the EEG bump through body fluids as a current of colliding ions (not electrons) until they reach the skin. This process, called volume conduction, allows us to eavesdrop on the scalp's cortical potentials instead of inserting electrodes inside the brain.

Electrodes are specialized conductors that convert biological signals like the EEG into currents of electrons. Surface EEG electrodes function like an antenna to detect the EEG signals produced by macrocolumns of cortical neurons. Currents of ions, atoms with positive or negative charges, volume conduct to the scalp (like an FM radio broadcast), and electrodes convert this signal to a current of electrons (Stern et al., 2001). Graphic © Zyabich/Shutterstock.com.





Insulation from body fat, connective tissue, and the epidermis (outer skin layer) interferes with ion current flow and can significantly reduce surface EMG readings. Like the rubber covering the wiring of a muscle electrode, insulators block the flow of electric currents. In natural and fabricated insulators, a large number of electrons in their final energy level produces a cohesiveness that resists electron loss due to collision. The best insulators, like rubber, possess the maximum number of outer-level electrons (Nilsson & Riedel, 2008).


Measuring Current

Measuring current means we learn how much "x" has passed by a point over a fixed period. The "amount" of electric current is measured in amperes (A). You have used 1 ampere of current when 1 coulomb (6.24 x 1018 or 6 billion billion electrons) has passed a point in 1 second (Kubala, 2009).

Electric current graphic © Designua/shutterstock.com.






DC and AC

Electricity travels as either a direct current (DC) or alternating current (AC). Direct current (DC) is the flow of electricity in one direction—from negative to positive. A difference in electrical potential pressures electrons to move. The negative end of a wire repels electrons (e-) while the positive end attracts them. Biological signals representing peripheral blood flow (blood volume pulse and skin temperature), respiration, and skin electrical activity are all DC signals.

When we plot DC signals against time, they never completely reverse direction over a second. The electroencephalogram (EEG) contains both DC (slow cortical potentials) and AC (slow cortical potentials and delta through 40-Hz) waveforms. BioGraph ® Infiniti blood volume pulse (BVP) display.





In the space of a second, an alternating current (AC) regularly reverses direction 50 or 60 times. The frequency of an alternating current is the number of cycles completed per second or hertz (Hz). Electrical potentials detected from the cerebral cortex (EEG), heart (ECG), and skeletal muscles (SEMG) all contain AC waveforms (Kubala, 2009). Check out the YouTube video AC and DC Differences.

BioGraph ® Infiniti 60-Hz artifact display. The software uses an auto-scale feature to keep the fluctuating signal on the screen.





The movie below is a single-channel BioTrace+ /NeXus-32 display of EEG activity from 1-64 Hz activity broken into component delta, theta, alpha, and beta frequency bands by digital filters © John S. Anderson.




Electromotive Force (EMF)

What forces electrons to move through a circuit? Electrons flow when there is a difference in electrical potential or charge.

A flashlight works because its battery contains negative and positive poles. These two regions of opposite charge produce an electrical potential difference called the electromotive force (EMF) that drives the current ahead. The electrical potential difference can be considered the "strength" of the current. A battery's negative pole repels electrons (e-) while its positive pole attracts them, resulting in current flow. If the battery's two poles had identical charges, electrons would stay put instead. No potential difference, no current, and no light (Nilsson & Riedel, 2008).

Note the negative and positive poles of the circuit that cause the current to move through the wire below. Electromagnetic field graphic © tersetki/Shutterstock.com.





Electromagnetic Fields Carry Energy Instead of Electrons

The classic model of electrons traveling through conductors in two directions is an explanatory fiction. Electrons don't travel from a battery to a light bulb or power plant to your microwave. As the Veritasium video shows, electromagnetic fields (shown above) travel and carry energy in one direction. This is true for sunlight, powerlines, and neurons. Watch the YouTube video, The Big Misconception About Electricity.


Voltage

A battery's pressure on electrons flowing through a flashlight is measured in volts (E). A typical flashlight battery is rated at 1.5 volts. One volt is the potential difference required to make 1 coulomb (6.24 x 1018 electrons) perform 1 joule of work. Voltage indexes signal power (Nilsson & Riedel, 2008).

When monitoring biological signals, you will record signals ranging from microvolts or μV (millionths of a volt) to millivolts or mV (thousandths of a volt). EEG and SEMG amplitudes are measured in microvolts (μV), usually less than 100 μV.

In neurofeedback, clinicians and researchers increasingly express the quantitative EEG (qEEG) signal strength, digitized statistical brain mapping using at least a 19-channel montage to measure EEG amplitude within specific frequency bins in picowatts (trillionths of a watt).


Watts

An electric current’s overall power depends on the amount of current flowing through a circuit (measured in amperes) and the electric potential driving it (measured in volts). Electric power is measured in watts (W). One watt is equal to one ampere moving at one volt. Multiplying amperes by volts produces the number of watts. For example, an appliance that uses 10 amperes and runs on 115 volts consumes 1150 watts of power (Kubala, 2009). Below are 21- and 32-channel Mitsar amplifier systems featured on the NovaTech EEG website.






Resistance

The electrons moving through a conductor encounter opposition which reduces current flow. This phenomenon is called resistance (R) in DC circuits and impedance (Z) in AC circuits and is measured in ohms (Ω). Resistance depends on the number of electrons in an atom's outermost energy level.




Listen to a mini-lecture on Resistance and Conductance © BioSource Software LLC. Graphic © Peter Hermes Furian/Shutterstock.com. Electrons are red, protons are green, and neutrons are gray.




Increasing the number of electrons in this level binds these electrons more tightly. This cohesiveness reduces the loss of electrons due to collisions with free electrons.

Resistance is a practical concern in biofeedback. Biological signals compete with stronger false signals for a biofeedback instrument's attention. Clinicians clean, abrade, and apply conductive gel to their clients' skin when monitoring the brain (EEG) and skeletal muscles (SEMG). These precautions improve signal reception since dead skin, oil, and dirt block biological potentials from reaching electrodes.







Dry electrodes like BrainMaster's Freedom 20R do not require time-consuming skin preparation and the application of conductive paste.




Skin resistance is also a biological signal, in its own right, that reflects emotional and cognitive processes. Clinicians measure skin resistance level (SRL) by running an AC or DC across the inner surface of the fingers or palm. SRL is expressed in Kohms of resistance per cm2. Typical values range from 0-500 Kohms/cm2. Lower values reflect more intense sweat gland activity since moisture reduces resistance.


Conductance

Resistance and conductance are mirror images of each other. Resistance is the reciprocal of conductance. Where resistance measures the opposition free electrons encounter, conductance (G) indexes how easily they travel through a conductor like copper or silver. The graphic depicts resistors in a computer circuit © Kovakchuk Oleksandr/Shutterstock.com.





Resistance is expressed in ohms (Ω). Conductance is now measured in Siemens and was previously measured in mhos (mho is ohm spelled backward). Skin conductance is one index of eccrine sweat gland activity.


Ohm's Law

Ohm’s law states that the “amount” of current (I) flowing through a conductor is equal to the voltage (E) (the “push”) divided by the resistance (R). These values are measured in amperes, volts, and ohms, respectively (Nilsson & Riedel, 2008).

Please click on the podcast icon below to hear a full-length lecture over Section E.


Ohm’s law can be used to find any value in a DC circuit: Voltage (E) = current (I) x resistance (R). Graphic © Emre Terim/Shutterstock.com.

For example, using actual units, 10 volts = 2 amperes x 5 ohms. Check out the YouTube video MAKE Presents: Ohms Law.

Graphic © VectorMine/Shutterstock.com. In the diagram below, voltage supplies the "push" while resistance opposes current movement.





Ohm's law is helpful because it describes the relationship between voltage, current, and resistance. We can use this law to show two ways to detect adequate voltages. Ohm's law graphic © Ergun_Pinar/Shutterstock.com.



First, if voltage (E) = current (I) x resistance (R), then we can increase the voltage by increasing current or resistance. Hardware designers use this relationship to increase the voltage reaching an electroencephalograph. When EEG voltages (current) enter an electroencephalograph's amplifier, they are dropped across a network of resistors (resistance). This large differential input impedance increases the EEG voltage seen by an electroencephalograph, which helps separate EEG voltages from artifacts.

Second, we can restate Ohm's law from the standpoint of current. If current (I) = voltage (E) / resistance (R), then we can increase current by increasing voltage or reducing resistance. This relationship is the reason clinicians prepare the skin when monitoring the EEG. Skin abrasion and application of conductive gel/paste minimize resistance. This increases the current reaching EEG electrodes, which helps an electroencephalograph distinguish EEG activity from artifacts.


Impedance

In AC circuits, current periodically reverses direction. This introduces frequency, the number of cycles completed each second. Frequency is measured in hertz (Hz). When an AC travels through a circuit at a given frequency, it encounters a complex form of opposition called impedance (Z), measured in ohms (Ω). Impedance reduces current flow between electrodes and the brain surface.



Clinicians perform an impedance test to determine whether they have correctly cleaned and abraded the skin and applied electrodes with sufficient gel or paste (Andreassi, 2007). Excessive impedance means that a weak biological signal must compete at a disadvantage with false electrical signals like power line artifacts. This could contaminate the EEG signal so severely that the electroencephalograph displays power line fluctuations instead of cortical activity.

We measure skin-electrode impedance by passing an AC through pairs of electrodes. An impedance test can be manually performed with a separate impedance meter. Graphic from the bio-medical.com website.






Software integrated with a data acquisition system and sensors may also perform an impedance test.





After the practitioner has positioned all electrodes, they should check their impedances or offsets using methods appropriate for their equipment. Electrodes that show excessive values can be reapplied after removing them to prepare the electrode site, if necessary.

Unless skin-electrode impedance is low (under 5 KΩ for research and 20 KΩ for training) and balanced (under 1-3 KΩ ), diverse artifacts like 50/60 Hz and movement can contaminate the EEG signal, as seen in the P3 and Pz electrodes. Graphic © eegatlas-online.com.





When the skin-electrode impedance at two sites is unequal, the resulting signals will appear to have different amplitudes when they reach the amplifier, regardless of the actual values. Unbalanced impedance will also increase DC offset values due to the battery effect. The amplifier will boost the resulting inaccurate input, which will be displayed to your client.

When a clinician fails to ensure low and balanced impedances at the start or during a training session, feedback regarding signal amplitude within specific frequency bands will be inaccurate. The wrong thresholds may be selected.

Michael and Lynda Thompson provided an example of an impedance problem that developed during a session because a hyperactive child scratched his ears, resulting in high and imbalanced impedances. Following corrective action that restored acceptable impedance values, high-beta activity (24-32 Hz) declined from 10-15 to 4 μV, gamma activity (45-58 Hz) declined below 2 μV, and SMR and beta activity returned to previous session values (Thompson & Thompson, 2015, p. 66).


DC Offset

DC offset is a voltage that results from combinations of factors, including electrode and gel/paste materials, interactions with skin, environment (humidity and temperature), and sweat gland activity due to stress level. The DC offset value should be consistent across all sensors and less than 25,000 μV, ideally below 10,000 μV. DC offset graphic © John S. Anderson.






Ohm's Law for AC Circuits

We can extend Ohm's law to AC circuits by substituting impedance (z) for resistance and using lowercase letters for voltage and current. The revised expression is voltage = current x impedance (e = i x z). This means that voltage is the product of a current flowing across an impedance. In actual units, 50 volts = 10 amperes x 5 ohms.


Open and Closed Circuits

Broken electrode cables significantly cause equipment malfunction since they prevent electron movement. Clinicians perform a continuity test to check if a cable is damaged. An impedance meter sends an AC signal down the cable to measure opposition to current flow. If there is a break, there is no continuity, and the circuit is described as open. Impedance will be infinite since current cannot flow across space.


A blown fuse illustrates an open circuit.

A filament in a fuse melts to create an open circuit when the current exceeds safe values. Graphic © AlexLMX/Shutterstock.com.




If the cable is free of breaks (continuous), the circuit is described as closed instead. Impedance will approach 0 Kohms since the current can easily travel through the circuit. Graphic © imagedb.com/ Shutterstock.com. The top diagram depicts an open circuit (light bulb off), while the bottom diagram shows a closed circuit (light bulb on).



Behavioral tests, also called tracking tests, check whether the circuit is closed and evaluate the performance of the entire data acquisition system.

For example, when monitoring EEG activity, a clinician can test the performance of the entire signal chain (EEG sensor, differential amplifier, gain amplifier, cable, encoder, and computer) by asking a client to close and then open the eyes. If the computer display mirrors these actions, the behavioral test has been passed, confirming no breaks in the cable.


Short Circuit

A short circuit results when an unintended connection is made between two points of a circuit. Graphic © Designua/Shutterstock.com.







The new path has lower resistance than the original circuit and should measure close to 0 Kohms on an impedance meter. The reduced resistance draws electrons through the short and may increase current flow to levels that can melt circuitry and injure clients (Nilsson & Riedel, 2008). Short circuit graphic © torook/Shutterstock.com.



Visualize a bare wire inside an electroencephalograph touching its metal case. The AC powering this equipment could leak through the metal case and injure anyone touching this surface.

Safety Precautions


Like computer-based data acquisition systems, line-powered equipment can expose clients and practitioners to shock hazards. Both should avoid contact with metal surfaces, and water spills should be immediately cleaned. Graphic © DenisNata/Shutterstock.com.







Exposure to Current Can Injure and Cause Death

A 1-second exposure to a current exceeding 5 mA can injure. An 18-mA current can affect breathing. A 50-mA current can cause fatal ventricular fibrillation in which the heart chambers cannot pump blood (Peek, 2016). Graphic © AlfaMD/Shutterstock.com.





Biomedical engineers prevent shock hazards through ground fault interrupt circuits, optical isolation, fiber optic connections, and telemetry. Graphic © ESB Professional/Shutterstock.com.






Ground Fault Interrupt Circuit

A ground fault interrupt circuit is designed into some power outlets to shut down power when a short circuit occurs. This protective circuit monitors current leakage. When harmful leakage is detected (> 5 mA), it triggers a circuit breaker that shuts off power to the equipment, protecting the client, therapist, and hardware.

Montgomery (2004) recommended plugging the entire biofeedback system into the same power outlet to create a common ground so that current leakage in any of your equipment will trigger the ground fault interrupt circuit.






Optical Isolation

Optical isolation protects a client from hardware receiving AC power. An optical isolator (opto-isolator) converts a biological signal into a beam of light using an LED source, the light crosses a dialectic barrier (insulation) located in the center (open circuit), and a phototransistor reconverts the light into an electrical signal.






Fiber Optic Connections

Fiber optic connections, thin, flexible cables that transmit digital signals as pulses of light, transmit photons between the electrodes and data acquisition system. This prevents current from leaking from a computer to a client since electrons cannot travel through fiber optic cables. This approach also reduces contamination by electrical artifacts like power line noise.





Telemetry

Telemetry can wirelessly transmit physiological data from a battery-powered encoder unit to a computer many meters away. This technology protects clients from shock since current surges cannot travel across a Bluetooth connection (Montgomery, 2004). MindMedia's NeXus-10 featured below communicates wirelessly with a computer for data acquisition.



Glossary



alternating current (AC): an electric current that periodically reverses its direction.

ampere (A): the unit of electrical current or the flow rate of electrons through a conductor. One volt dropped across one ohm of resistance produces a current flow of one ampere.

atom: the basic unit of matter consisting of a central nucleus that contains protons and neutrons and orbiting electrons.

atomic number: the number of protons in the nucleus of an atom that defines an element.

atomic weight: the approximate number of protons and neutrons in the nucleus of an atom.

average voltage: 0.637 of the peak voltage.

charge (Q): the imbalance between the number of positively and negatively charged particles in a given place or between two locations.

closed circuit: a complete path that allows electrons to travel from the power source, through the conductor and resistance, and back to the source.

common-mode rejection ratio (CMRR): the degree by which a differential amplifier boosts signal (differential gain) and artifact (common-mode gain).

conductance (G): the ability of a material like copper or silver to carry an electric current. Conductance is measured in siemens (formerly mhos).

conductor: a material that readily allows electron movement, like a copper wire.

continuity test: a procedure to ensure that a circuit is closed. For example, a cable is not broken.

coulomb: approximately 6.24 x 1018 or 6 billion billion electrons.

current (I): the movement of electrons through a conductor measured in amperes (A).

DC offset: the voltage that results from combinations of factors, including electrode and gel/paste materials, interactions with skin, environment (humidity and temperature), and sweat gland activity due to stress level.

differential amplifier (balanced amplifier): a device that boosts the difference between two inputs: the active (input 1) and reference (input 2).

differential input impedance:
the opposition to an AC signal entering a differential amplifier as it is dropped across a resistor network.

direct current (DC):
an electric current that flows in only one direction, as in a flashlight.

dry electrodes: electrodes that do not require the use of conductive gel or paste to establish electrical contact with the scalp. Instead, they utilize specialized materials or designs to interface directly with the skin, potentially reducing setup time and increasing user comfort compared to conventional wet electrodes.

electromotive force (EMF): a difference in electrical potential that "pushes" electrons to move in a circuit.

electron: a negatively-charged particle that rotates around the nucleus at varying distances and participates in chemical reactions.

elements: substances that contain identical atoms and cannot be reduced by common chemical reactions.

energy level: one of an electron's possible orbits around a nucleus at a constant distance.

frequency (Hz): the number of complete cycles that an AC signal completes in a second, usually expressed in hertz.

ground fault interrupt circuit: a protective device that opens a circuit—shutting down power—when current leakage exceeds 5 mA.

hertz (Hz): the unit of frequency measured in cycles per second.

high-pass filter: a filter that only passes frequencies higher than a set value (e.g., 1 Hz).

impedance (Z): complex opposition to an AC signal measured in Kohms.

impedance meter: device that uses an AC signal to measure impedance in an electric circuit, such as between active and reference electrodes.

impedance test: the automated or manual measurement of skin-electrode impedance.

insulator: material that resists the flow of electricity, like glass and rubber.

mho: the unit of conductance replaced by the siemen.

microsiemen (μS): the unit of conductance that is one-millionth of a siemen.

microvolt (μV): the unit of amplitude (signal strength) that is one-millionth of a volt.

milliampere (mA): unit of electrical current that is one-thousandth of an ampere.

millivolt (mV): unit of amplitude (signal strength) that is one-thousandth of a volt.

nucleus: central mass of an atom that contains protons and neutrons.

ohm (Ω): the unit of impedance or resistance.

Ohm's law: voltage (E) = current (I) X resistance (R). The “amount” of current (I) flowing through a conductor is equal to the voltage (E) or “push” divided by the resistance (R).

open circuit: an incomplete path that prevents electron movement from the power source, through the conductor, and back to the source. For example, a broken sensor cable.

optical isolation: a device that converts a biological signal into a beam of light, the light crosses a gap (open circuit), and a photoreceptor reconverts the light into an electrical signal.

picowatt: billionths of a watt.

polarization: chemical reactions produce separate regions of positive and negative charge where an electrode and electrolyte make contact, reducing ion exchange.

power (W): the rate at which energy is transferred, which is proportional to the product of current and voltage. Power is measured in watts.

proton: positively charged subatomic particle found in the nucleus of an atom.

resistance (R): the opposition to a DC signal by a resistor measured in ohms.

short circuit: a lower-resistance electrical circuit created by the unintended contact between components that accidentally diverts the current.

skin conductance level (SCL): a tonic measurement of how easily an AC or DC passes through the skin, expressed in microsiemens.

skin resistance level (SRL): a tonic (resting) measurement of the opposition to an AC or DC as it passes through the skin, expressed in Kohms.

telemetry:
remote monitoring and transmission of information. An encoder measures physiological activity and transmits these data to a computer for analysis.

ventricular fibrillation: a medical emergency in which the lower heart chambers contract in a rapid and unsynchronized fashion and cannot pump blood.

volt (V): unit of electrical potential difference (electromotive force) that moves electrons in a circuit.

voltage (E): the amount of electrical potential difference (electromotive force).

watt (W): a unit of power used to express signal strength in the qEEG.


TEST YOURSELF ON CLASSMARKER


Click on the ClassMarker logo below to take a 10-question exam over this entire unit.


REVIEW FLASHCARDS ON QUIZLET


Click on the Quizlet logo to review our chapter flashcards.


Visit the BioSource Software Website


BioSource Software offers Physiological Psychology, which satisfies BCIA's Physiological Psychology requirement, and Neurofeedback100, which provides extensive multiple-choice testing over the Biofeedback Blueprint.



Assignment


Now that you have completed this module, explain why low-and-balanced skin-electrode impedances are important in neurofeedback training. Describe the precautions you take to achieve acceptable impedance values. How do you measure impedance with your neurofeedback system?


References


Andreassi, J. L. (2000). Psychophysiology: Human behavior and physiological response. Lawrence Erlbaum and Associates, Inc.

Basmajian, J. V. (Ed.). (1989). Biofeedback: Principles and practice for clinicians. Williams & Wilkins.

Cacioppo, J. T., & Tassinary, L. G. (Eds.). (1990). Principles of psychophysiology. Cambridge University Press.

Collura, T. F. (2014). Technical foundations of neurofeedback. Taylor & Francis.

Demos, J. N. (2005). Getting started with neurofeedback. W. W. Norton & Company.

Fisch, B. J. (1999). Fisch and Spehlmann's EEG primer (3rd ed.). Elsevier.

Floyd, T. L. (1987). Electronics fundamentals: Circuits, devices, and applications. Merrill Publishing Company.

Grant, A. (2015). Four elements earn permanent seats on the periodic table. Science News.

Halford, J. J., Sabau, D., Drislane, F. W., Tsuchida, T. N., & Sinha, S. R. (2016). American Clinical Society Guideline 4: Recording clinical EEG on digital media. Journal of Clinical Neurophysiology, 33(4), 317-319. https://doi.org/10.1080/21646821.2016.1245563

Hughes, J. R. (1994). EEG in clinical practice (2nd ed.). Butterworth-Heinemann.

Kubala, T. (2009). Electricity 1: Devices, circuits, and materials (9th ed.). Cengage Learning.

Libenson, M. H. (2010). Practical approach to electroencephalography. Saunders Elsevier.

Miller, J. W., Kim, W. S., Homes, M. D., & Vanhatalo, S. (2007). Ictal localization by source analysis of infraslow activity in DC-coupled scalp EEG recordings. NeuroImage, 35(2), 583-597. https://doi.org/10.1016/j.neuroimage.2006.12.018

Montgomery, D. (2004). Introduction to biofeedback. Module 3: Psychophysiological recording. Association for Applied Psychophysiology and Biofeedback.

Nilsson, J. W., & Riedel, S. A. (2008). Electric circuits (8th ed.). Pearson Prentice-Hall.

Peek, C. J. (2016). A primer of traditional biofeedback instrumentation. In M. S. Schwartz, & F. Andrasik (Eds.). (2016). Biofeedback: A practitioner's guide (4th ed.). The Guilford Press.

Stern, R. M., Ray, W. J., & Quigley, K. S. (2001). Psychophysiological recording (2nd ed.). Oxford University Press.

Thompson, M., & Thompson, L. (2015). The biofeedback book: An introduction to basic concepts in applied psychophysiology (2nd ed.). Association for Applied Psychophysiology and Biofeedback.

Wadman, W. J., & Lopes da Silva, F. H. (2011). In D. L. Schomer & F. H. Lopes da Silva (Eds.). Niedermeyer's electroencephalography: Basic principles, clinical applications, and related fields (6th ed.). Lippincott Williams & Wilkins.