Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-dfsvx Total loading time: 0 Render date: 2024-04-29T14:24:12.408Z Has data issue: false hasContentIssue false

13 - Psychophysiology Principles, Pointers, and Pitfalls

from SECTION FOUR - DATA ACQUISITION, REDUCTION, ANALYSIS, AND INTERPRETATION: CONSIDERATIONS AND CAVEATS

Published online by Cambridge University Press:  27 July 2009

Anita Miller
Affiliation:
Visiting Assistant Professor of Psychology Skidmore College
James Long
Affiliation:
Computer electrical and software engineer and the owner James Long Company
Louis A. Schmidt
Affiliation:
McMaster University, Ontario
Sidney J Segalowitz
Affiliation:
Brock University, Ontario
Get access

Summary

INTRODUCTION

Psychophysiology focuses on physiological processes associated with human sensory, motor, cognitive, emotional, and social functions. Developmental psychophysiology centers on the emergence of such processes in youngsters. Over the past several decades, technological advances have revolutionized the field. Marked progress has been made in psychology and neuroscience as well as in electrical engineering and applied mathematics. Advances in circuit boards and silicon chips have facilitated manufacturing of accurate, stable, and predictable devices for amplifying and filtering analog physiological signals, converting them to a digital format, and recording sizable datasets on personal computers. In addition, the computational efficiency and storage capacity of digital hardware have increased significantly, and software tools for signal processing have grown more sophisticated and widely available. Such technological advances have created a trend toward increased performance for a given price, and complete commercial laboratory systems have made human psychophysiology measures increasingly accessible to more investigators conducting basic and applied research.

As psychophysiology tools become widely available, needs increase for introductory tutorials for conducting psychophysiology assessments. Despite the ease of obtaining turn-key equipment and recording physiological data, fundamental challenges remain inherent to the work. For instance, most psychophysiological measures have multiple determinants. Bioelectric signals are often a composite of multiple physiological processes that co-occur or interact. For example, brain recordings contain ocular and muscle activity, and voluntary breathing influences heart rate variability. In addition, noise sources can mimic physiological processes, such as the AC power frequency overlapping with physiological activity and body movements confounding skin conductance responses.

Type
Chapter
Information
Developmental Psychophysiology
Theory, Systems, and Methods
, pp. 367 - 423
Publisher: Cambridge University Press
Print publication year: 2007

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Ahmed, S. S., Levinson, G. E., Schwartz, C. J., & Ettinger, P. O. (1972). Systolic time intervals as measures of the contractile state of the left ventricular myocardium in man. Circulation, 46, 559–571.CrossRefGoogle ScholarPubMed
Akselrod, S., Gordon, D., Ubel, F. A., Shannon, D. C., Barger, A. C., & Cohen, R. J. (1981). Power spectrum analysis of heart rate fluctuations: A quantitative probe of beat-to-beat cardiovascular control. Science, 213, 220–222.CrossRefGoogle ScholarPubMed
Alexander, G. E., & Moeller, J. R. (1994). Application of the scaled subprofile model to functional imaging in neuropsychiatric disorders: A principal component approach to modeling brain function and disease. Human Brain Mapping, 2, 79–94.CrossRefGoogle Scholar
Allen, J. J. B., Coan, J. A., & Nazarian, M. (2004). Issues and assumptions on the road from raw signals to metrics of frontal EEG asymmetry in emotion. Biological Psychology, 67, 183–218.CrossRefGoogle ScholarPubMed
Als, H., Duffy, F. H., McAnulty, G. B., Rivkin, M. J., Vajapeyam, S., Mulkern, R. V., Warfield, S. K., Huppi, P. S., Butler, S. C., Conneman, N., Fischer, C., & Eichenwald, E. C. (2004). Early experience alters brain function and structure. Pediatrics, 113, 846–857.CrossRefGoogle ScholarPubMed
American Electroencephalographic Society. (1994). Guideline thirteen: Guidelines for standard electrode position nomenclature. Journal of Clinical Neurophysiology, 11, 111–113.CrossRef
Artman, M., Mahoney, L., & Teitel, D. F. (2002). Neonatal cardiology. New York: McGraw-Hill.Google Scholar
Attuel, P., Leporho, M. A., Ruta, J., Lucet, V., Steinberg, C., Azancot, A., & Coumel, P. (1986). The evolution of the sinus heart rate and variability as a function of age from birth to 16 years. In Doyle, E. F. (Ed.), Pediatric cardiology: Proceeding of the Second World Congress 1985. New York: Springer-Verlag.CrossRefGoogle Scholar
Bakeman, R., & Gottman, J. M. (1997). Observing interaction: An introduction to sequential analysis (2nd ed.). New York: Cambridge University Press.CrossRefGoogle Scholar
Balaban, M. T. (1995). Affective influences on startle in five-month-old infants: Reactions to facial expressions of emotion. Child Development, 66, 28–36.CrossRefGoogle Scholar
Bar-Haim, Y., Marshall, P. J., & Fox, N. A. (2000). Developmental changes in heart period and high-frequency heart period variability from 4 months to 4 years of age. Developmental Psychobiology, 37, 44–56.3.0.CO;2-7>CrossRefGoogle ScholarPubMed
Basar, E. (1980). EEG brain dynamics: Relations between EEG and brain evoked potentials. Amsterdam, Netherlands: Elsevier.Google Scholar
Basar, E. (1998). Brain oscillations: Principles and approaches. Berlin, Germany: Elsevier.CrossRefGoogle Scholar
Belik, J., & Pelech, A. (1988). Thoracic electric bioimpedance measurement of cardiac output in the newborn infant. Journal of Pediatrics, 113, 890–895.CrossRefGoogle ScholarPubMed
Bell, M. A. (2002). Power changes in infant EEG frequency bands during a spatial working memory task. Psychophysiology, 39, 450–458.CrossRefGoogle ScholarPubMed
Bell, M. A. & Fox, N. A. (1992). The relations between frontal brain electrical activity and cognitive development during infancy. Child Development, 63, 1142–1163.CrossRefGoogle ScholarPubMed
Ben-Shakhar, G. (1985). Standardization within individuals: Simple method to neutralize individual differences in skin conductance. Psychophysiology, 22, 292–299.CrossRefGoogle ScholarPubMed
Berger, H. (1929). On the electroencephalogram of man. I. Archives Psychiatry Nervenkr, 87, 527–570.CrossRefGoogle Scholar
Berne, R. M., & Levy, M. N. (2001). Cardiovascular physiology (8th ed.). London: Mosby.Google Scholar
Berntson, G. G., Bigger, J. T., Eckberg, D. L., Grossman, P., Kaufmann, P. G., Malik, M., Nagaraja, H. N., Porges, S. W., Saul, J. P., Stone, P. H., & Molen, M. W. (1997). Heart rate variability: Origins, methods, and interpretive caveats. Psychophysiology, 34, 623–648.CrossRefGoogle ScholarPubMed
Berntson, G. G., & Cacioppo, J. T. (2002). Psychophysiology. In D'Haenen, H., Boer, J. A., & Willner, P. (Eds.), Biological psychiatry (Vol. 1, pp. 123–138). West Sussex, UK: John Wiley & Sons.CrossRefGoogle Scholar
Berntson, G. G., Cacioppo, J. T., Binkley, P. F., Uchino, B. N., Quigley, K. S., & Fieldstone, A. (1994). Autonomic cardiac control:III. Psychological stress and cardiac response in autonomic space as revealed by pharmacological blockades. Psychophysiology, 31, 599–608.CrossRefGoogle ScholarPubMed
Berntson, G. G., Lozano, D. L., Chen, Y. J., & Cacioppo, J. T. (2004). Where to Q in PEP. Psychophysiology, 41, 333–337.CrossRefGoogle Scholar
Berntson, G. G., Quigley, K. S., Jang, J. F., & Boysen, S. T. (1990). An approach to artifact identification: Applications to heart period data. Psychophysiology, 27, 586–598.CrossRefGoogle Scholar
Bertrand, O., Perrin, F., & Pernier, J. (1985). A theoretical justification of the average reference in topographic evoked potential studies. Electroencephalography and Clinical Neurophysiology, 62, 462–464.CrossRefGoogle ScholarPubMed
Blumenthal, T. D., Cuthbert, B. N., Filion, D., Hackley, S., Lipp, O. V., & Boxtel, A. (2005). Committee report: Guidelines for human startle eyeblink electromyographic studies. Psychophysiology, 42, 1–15.CrossRefGoogle ScholarPubMed
Boik, R. J. (1981). A priori tests in repeated measures designs: Effects of nonsphericity. Psychometrika, 46, 241–255.CrossRefGoogle Scholar
Boucsein, W. (1992). Electrodermal activity. New York: Plenum.CrossRefGoogle Scholar
Bradley, M. M. (2000). Emotion and motivation. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (Eds.), Handbook of psychophysiology (2nd ed., pp. 602–642). New York: Cambridge University Press.Google Scholar
Brunia, C. H. M. (1997). Gaiting in readiness. In Lang, P. J. & Simons, R. F.. Attention and orienting: Sensory and motivational processes (pp. 281–306). Mahwah, NJ: Lawrence Erlbaum Associates.Google Scholar
Bush, L. K., Hess, U., & Wolford, G. (1993). Transformations for within-subject designs: A Monte Carlo investigation. Psychological Bulletin, 101, 147–158.Google Scholar
Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (Eds.) (2000). Handbook of psychophysiology (2nd ed.), New York: Cambridge University Press.Google Scholar
Campbell, D. T., & Stanley, J. C. (1963). Experimental and quasi-experimental designs for research. Boston: Houghton Mifflin Company.Google Scholar
Chapman, L. J., & Chapman, J. P. (2001). Commentary on two articles concerning generalized and specific cognitive deficits. Journal of Abnormal Psychology, 110, 31–39.CrossRefGoogle ScholarPubMed
Chiarenza, G. A. (1998). Editorial: The richness of developmental psychophysiology. Journal of Psychophysiology, 12, 220–222.Google Scholar
Cohen, J. (1960). Coefficient of agreement for nominal scales. Educational and Psychological Measurement, 20, 37–46.CrossRefGoogle Scholar
Cohen, J. (1968). Weighted kappa: Nominal scale agreement with provision for scaled disagreement or partial credit. Psychological Bulletin, 70, 213–220.CrossRefGoogle ScholarPubMed
Coles, M. G. H., Donchin, E., & Porges, S. W. (1986). Psychophysiology: Systems, processes, and applications. New York: Guilford.Google Scholar
Cook, E. W., & Miller, G. A. (1992). Digital filtering: Background and tutorial for psychophysiologists. Psychophysiology, 29, 350–367.CrossRefGoogle ScholarPubMed
Cooley, J. W., & Tukey, J. W. (1965). An algorithm for the machine calculation of complex Fourier series. Mathematics of Computation, 19, 297–301.CrossRefGoogle Scholar
Crider, A. (1993) Electrodermal response lability-stability: Individual difference correlates. In Roy, J. C., Boucsein, W., Fowles, D. C., & Gruzelier, J. H. (Eds.), Progress in electrodermal research (pp. 173–186). New York: Plenum.CrossRefGoogle Scholar
Croft, R. J., & Barry, R. J. (2000). Removal of ocular artifact from the EEG: A review. Neurophysiologie Clinique, 30, 5–19.CrossRefGoogle ScholarPubMed
Cycowicz, Y. M., & Friedman, D. (1997). A developmental study of the effect of temporal order on the ERPs elicited by novel environmental sounds. Electroencephalography and Clinical Neurophysiology, 103, 304–318.CrossRefGoogle ScholarPubMed
Davidson, R. J., Jackson, D. C., & Larson, C. L. (2000). Human electroencephalography. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (Eds.), Handbook of psychophysiology (2nd ed., pp. 27–52). New York: Cambridge University Press.Google Scholar
Davidson, R. J. & Tomarken, A. J. (1989). Laterality and emotion: An electrophysiological approach. In Boller, F. & Grafman, J. (Eds.), Handbook of neuropsychology (Vol. 3, pp. 419–441). Amsterdam, Netherlands: Elsevier Science.Google Scholar
Dawson, M. E., Schell, A. M., & Filion, D. L. (2000). The electrodermal system. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (Eds.), Handbook of psychophysiology (2nd Ed., pp. 200–223). New York: Cambridge University Press.Google Scholar
Delorme, A., & Makeig, S. (2004). EEGLAB: An open source toolbox for analysis of single-trial EEG dynamics including independent component analysis. Journal of Neuroscience Methods, 15, 9–21.CrossRefGoogle Scholar
Dien, J. (1998). Issues in the application of the average reference: Review, critiques, and recommendations. Behavior Research Methods, Instruments, and Computers, 30, 34–43.CrossRefGoogle Scholar
Dimberg, U. (1990). Facial electromyography and emotional reactions. Psychophysiology, 27, 481–494.CrossRefGoogle ScholarPubMed
DiPietro, J. A., Caughy, M. O., Cusson, R., & Fox, N. A. (1994). Cardiorespiratory functioning of preterm infants: Stability and risk associations for measures of heart rate variability and oxygen saturation. Developmental Psychobiology, 27, 137–152.CrossRefGoogle ScholarPubMed
Donchin, E., & Heffley, E. (1978). Multivariate analysis of event-related potential data: A tutorial review. In Otto, D. (Ed.), Multidisciplinary perspectives in event-related brain potential research (EPA-600/9-77043, pp. 555–572). Washington, DC: U.S. Government Printing Office.Google Scholar
Donchin, E., & Herning, R. I. (1975). A simulation study of the efficacy of step-wise discriminant analysis in the detection and comparison of event-related potentials. Electroencephalography and Clinical Neurophysiology, 38, 51–68.CrossRefGoogle Scholar
Donchin, E., Ritter, W., & McCallum, C. (1978). Cognitive psychophysiology: The endogenous components of the ERP. In Callaway, E., Tueting, P., & Koslow, S. H. (Eds.), Event-related brain potentials in man (pp. 349–411). New York: Academic Press.Google Scholar
Duffy, F. H., Iyer, V. G., & Surwillo, W. W. (1989). Clinical electroencephalography and topographic brain mapping: Technology and practice. New York: Springer-Verlag.CrossRefGoogle Scholar
Dumermuth, G., & Molinari, L. (1987). Spectral analysis of the EEG: Some fundamentals revisited and some open problems. Neuropsychobiology, 17, 85–99.CrossRefGoogle ScholarPubMed
Eckberg, D. L. (1983). Human sinus arrhythmia as an index of vagal cardiac outflow. Journal of Applied Physiology, 54, 961–966.CrossRefGoogle ScholarPubMed
Edelberg, R. (1972). Electrical activity of the skin: Its measurement and uses in psychophysiology. In Greenfield, N. S. & Sternbach, R. A. (Eds.), Handbook of psychophysiology (pp. 367–418). New York: Holt.Google Scholar
Elbert, T., Lutzenberger, W., Rockstroh, B., & Birbaumer, N. (1985). Removal of ocular artifacts from the EEG – A biophysical approach to the EOG. Electroencephalography and Clinical Neurophysiology, 60, 455–463.CrossRefGoogle Scholar
El-Sheikh, M. (2005). Stability of respiratory sinus arrhythmia in children and young adolescents: A longitudinal examination. Developmental Psychobiology, 46, 66–74.CrossRefGoogle Scholar
Fabiani, M., Gratton, G., & Coles, M. G. H. (2000). Event-related brain potentials: Methods, theory, and applications. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (Eds.), Handbook of psychophysiology (2nd ed., pp. 53–84). New York: Cambridge University Press.Google Scholar
Fabiani, M., Gratton, G., Karis, D., & Donchin, E. (1987). The definition, identification, and reliability of measurement of the P300 component of the event-related brain potential. In Ackles, P. K., Jennings, J. R., & Coles, M. G. H. (Eds.), Advances in psychophysiology (Vol. 1, pp. 1–78). Greenwich, CT: JAI.Google Scholar
Falkenstein, M., Hohnsbein, J., Hoormann, J., & Blanke, L. (1990). Effects of errors in choice reaction tasks on the ERP under focused and divided attention. In Brunia, C. H. M., Gaillard, A. W. K., & Kok, A. (Eds.), Psychophysiological brain research (pp. 192–195). Tilburg, The Netherlands: Tilburg University Press.Google Scholar
Ferree, T. C., Luu, P., Russell, G. S., & Tucker, D. M. (2001). Scalp electrode impedance, infection risk, and EEG data quality. Clinical Neurophysiology, 112, 536–544.CrossRefGoogle ScholarPubMed
Fisch, B. J. (1991). Artifacts. In Fisch, B. J. (Ed.), Spehlmann's EEG Primer (2nd Ed., pp. 107–126). New York: Elsevier.Google Scholar
Fowles, D. C. (1986). The eccrine system and electrodermal activity. In Coles, M. G. H., Donchin, E., & Porges, S. W. (Eds.), Psychophysiology: Systems, processes, and applications (pp. 51–96). New York: Guilford.Google Scholar
Fowles, D. C., Christie, M. J., Edelberg, R., Grings, W. W., Lykken, D. T., & Venables, P. H. (1981). Publication recommendations for electrodermal measurements. Psychophysiology, 18, 232–239.CrossRefGoogle ScholarPubMed
Fox, N. A., & Calkins, S. D. (1993). Multiple-measure approaches to the study of infant emotion. In Lewis, M. & Haviland, J. M. (Eds.), Handbook of emotions (pp. 203–219). New York: Guilford.Google Scholar
Fox, N. A., & Porges, S. W. (1985). The relation between neonatal heart period patterns and developmental outcome. Child Development, 56, 28–37.CrossRefGoogle ScholarPubMed
Fox, N. A., Schmidt, L. A., & Henderson, H. A. (2000). Developmental psychophysiology: Conceptual and methodological perspectives. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (Eds.), Handbook of psychophysiology (2nd ed., pp. 665–686). New York: Cambridge University Press.Google Scholar
Fridlund, A. J., & Cacioppo, J. T. (1986). Guidelines for human electromyographic research. Psychophysiology, 23, 567–589.CrossRefGoogle ScholarPubMed
Friedman, D. (2000). Event-related brain potential investigations of memory and aging. Biological Psychology, 54, 175–206.CrossRefGoogle ScholarPubMed
Galles, S. J., Miller, A., Cohn, J. F., & Fox, N. A. (2002). Estimating parasympathetic control of heart rate variability: Two approaches to quantifying vagal tone. Psychophysiology, 39, S37.Google Scholar
Gasser, T., Bacher, P., & Mocks, J. (1982). Transformations towards the normal distribution of broad band spectral parameters of the EEG. Electroencephalography and Clinical Neurophysiology, 53, 119–124.CrossRefGoogle ScholarPubMed
Gevins, A. S. (1984). Analysis of the electromagnetic signals of the human brain: Milestones, obstacles, and goals. IEEE Transactions in Biomedical Engineering, 31, 833–850.CrossRefGoogle ScholarPubMed
Gevins, A., Le, J., Martin, N. K., Brickett, P., Desmond, J., & Reutter, B. (1994). High resolution EEG: 124–channel recording, spatial deblurring and MRI integration methods. Electroencephalography and Clinical Neurophysiology, 90, 337–358.CrossRefGoogle ScholarPubMed
Glaser, E. M., & Ruchkin, D. S. (1976). Principles of neurobiological signal analysis. New York: Academic Press.Google Scholar
Graham, F. K. (1979). Distinguishing among orienting, defense, and startle reflexes. In Kimmel, H. D., Olst, E. H., & Orlebeke, J. F. (Eds.), The orienting reflex in humans (pp. 137–167). Hillsdale, NJ: Lawrence Erlbaum Associates.Google Scholar
Graham, F. K. (1992). Attention: The heartbeat, the blink, and the brain. In Campbell, B. A., Hayne, H., & Richardson, R. (Eds.), Attention and information processing in infants and adults: Perspectives from human and animal research (pp. 3–29). Hillsdale, NJ: Lawrence Erlbaum Associates.Google Scholar
Graham, F. K., Anthony, B. J., & Zeigler, B. L. (1983). The orienting response and development processes. In Siddle, D. (Ed.), Orienting and habituation: Perspectives in human research (pp. 371–430). Sussex, UK: John Wiley & Sons.Google Scholar
Gratton, G., Coles, M. G. H., & Donchin, E. (1983). A new method for off-line removal of ocular artifact. Electroencephalography and Clinical Neurophysiology, 55, 468–484.CrossRefGoogle ScholarPubMed
Gratton, G., Kramer, A. F., Coles, M. G., & Donchin, E. (1989). Simulation studies of latency measures of components of the event-related brain potential. Psychophysiology, 26, 233–248.CrossRefGoogle ScholarPubMed
Greene, W. A., Turetsky, B., & Kohler, C. (2000). General laboratory safety. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (Eds.), Handbook of psychophysiology (2nd ed., pp. 951–977). New York: Cambridge University Press.Google Scholar
Grossman, P., Karemaker, J., & Wieling, W. (1991). Prediction of tonic parasympathetic cardiac control using respiratory sinus arrhythmia: The need for respiratory control. Psychophysiology, 28, 201–216.CrossRefGoogle ScholarPubMed
Grossman, P., Beek, J., & Wientjes, C. (1990). A comparison of three quantification methods for estimation of respiratory sinus arrhythmia. Psychophysiology, 27, 702–714.CrossRefGoogle ScholarPubMed
Gunnar, M. R. (2003). Integrating neuroscience and psychological approaches in the study of early experiences. Annals of the New York Academy of Science, 1008, 238–247.CrossRefGoogle Scholar
Harris, W. S., Schoenfeld, C. D., & Weissler, A. M. (1967). Effects of adrenergic receptor activation and blockade of the systolic pre-ejection period, heart rate, and arterial pressure in man. The Journal of Clinical Investigation, 46, 1704–1714.CrossRefGoogle Scholar
Harver, A., & Lorig, T. S. (2000). Respiration. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (Eds.), Handbook of psychophysiology (2nd ed., pp. 265–293). New York: Cambridge.Google Scholar
Hernes, K. G., Morkrid, L., Fremming, A., Odegarden, S., Martinsen, O. G., & Storm, H. (2002). Skin conductance changes during the first year of life in full-term infants. Pediatric Research, 52, 837–843.CrossRefGoogle ScholarPubMed
Hirsch, J. A., & Bishop, B. (1981). Respiratory sinus arrhythmia: How breathing patterns modulates heart rate. American Journal of Physiology, 241, 620–629.Google Scholar
Horowitz, P., & Hill, W. (1989). The art of electronics (2nd ed.). New York: Cambridge University Press.Google Scholar
Horst, R. L., & Donchin, E. (1980). Beyond averaging II: Single trial classification of exogenous event-related potentials using step-wise discriminant analysis. Electroencephalography and Clinical Neurophysiology, 48, 113–126.CrossRefGoogle Scholar
Howard, L., & Polich, J. (1985). P300 latency and memory span development. Developmental Psychology, 21, 283–289.CrossRefGoogle Scholar
Huettel, S. A., McKeown, M. J., Song, A. W., Hart, S., Spencer, D. D., Allison, T., & McCarthy, G. (2004). Linking hemodynamic and electrophysiological measures of brain activity: Evidence from functional MRI and intracranial field potentials. Cerebral Cortex, 14, 165–173.CrossRefGoogle ScholarPubMed
Hugdahl, K. (1995). Psychophysiology: The mind-body perspective. Cambridge, MA: Harvard University Press.Google Scholar
Jacobson, L., & Sapolsky, R. (1991). The role of the hippocampus in feedback regulation of the hypothalamic-pituitary-adrenocortical axis. Endocrinology Review, 12, 118–134.CrossRefGoogle ScholarPubMed
Jennings, J. R., Berg, W. K., Hutcheson, J. S., Obrist, P., Porges, S., & Turpin, G. (1981). Publication guidelines for heart rate studies in man. Psychophysiology, 18, 226–231.CrossRefGoogle ScholarPubMed
Jennings, J. R., Cohen, M. J., Ruchkin, D. S., & Fridlund, A. J. (1987). Editorial policy on analysis of variance with repeated measures. Psychophysiology, 24, 474–478.CrossRefGoogle Scholar
Jennings, J. R. & Molen, M. W. (2002). Cardiac timing and the central regulation of action. Psychology Research, 66, 337–349.CrossRefGoogle Scholar
John, E. R., Ahn, H., Prichep, L. S., Trepetin, M., Brown, D., & Kaye, H. (1980). Development equations for the electroencephalogram. Science, 210, 1255–1258.CrossRefGoogle ScholarPubMed
John, E. R., Prichep, L. S., & Easton, P. (1987). Normative data banks and neurometrics. Basic concepts, methods, and results of norm constructions. In Gevins, A. S. & Remond, A. (Eds.), Methods of analysis of brain electrical and magnetic signals (pp. 449–495). Amsterdam, Netherlands: Elsevier Science.Google Scholar
John, M. S., Brown, D. K., Muir, P. J., & Picton, T. W. (2004). Recording auditory steady-state responses in young infants. Ear and Hearing, 25, 539–553.CrossRefGoogle ScholarPubMed
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, 163–178.CrossRefGoogle ScholarPubMed
Kayser, J., & Tenke, C. E. (2003). Optimizing PCA methodology for ERP component identification and measurement: Theoretical rational and empirical evaluation. Clinical Neurophysiology, 114, 2307–2325.
Kendall, P. T., & McCreary, E. K. (1980). Muscles: Testing and function (3rd ed.). Baltimore: Williams & Wilkins.Google Scholar
Kirschbaum, C., & Hellhammer, D. H. (1989). Salivary cortisol in psychobiological research: An overview. Neuropsychobiology, 22, 150–169.CrossRefGoogle ScholarPubMed
Kubicek, W. G., Karnegis, J. N., Patterson, R. P., Witsoe, D. A., & Mattson, R. H. (1966). Development and evaluation of an impedance cardiograph system. Aerospace Medicine, 37, 1208–1212.Google Scholar
Kubicek, W. G., Patterson, R. P., & Witsoe, D. A. (1970). Impedance cardiography as a noninvasive method of monitoring cardiac function and other parameters of the cardiovascular system. Annals of the New York Academy of Science, 170, 724–732.CrossRefGoogle Scholar
Lagercrantz, H., Hanson, M., Evrard, P., & Rodeck, C. (Eds.) (2002). The newborn brain: Neuroscience and clinical applications. Cambridge, UK: Cambridge University Press.Google Scholar
Lamberts, R., Visser, K. R., & Zijlstra, W. G. (1984). Impedance cardiography. Assen, The Netherlands: Van Gorcum.Google Scholar
Lang, P. J., Bradley, M. M., & Cuthbert, B. N. (1990). Emotion, attention, and the startle reflex. Psychological Review, 97, 377–398.CrossRefGoogle ScholarPubMed
Lang, P. J., Bradley, M. M., & Cuthbert, B. N. (2005). International affective picture system (IAPS): Affective ratings of pictures and instruction manual. Technical Report A-6. University of Florida, Gainesville, FL.Google Scholar
Lins, O. G., Picton, T. W., Berg, P. & Scherg, M. (1993a). Ocular artifacts in EEG and event-related potentials: II. Scalp topography. Brain Topography, 6, 51–63.CrossRefGoogle Scholar
Lins, O. G., Picton, T. W., Berg, P., & Scherg, M. (1993b). Ocular artifacts in recording EEGs and event-related potentials: II. Source dipoles and source components. Brain Topography, 6, 65–78.CrossRefGoogle Scholar
Lovallo, W. R. &, Thomas, T. L. (2000). Stress hormones in psychophysiological research: Emotional, behavioral, and cognitive interpretations. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (Eds.), Handbook of psychophysiology (2nd ed., pp. 342–367). New York: Cambridge University Press.Google Scholar
Luck, S. J., & Hillyard, S. A. (1994). Electrophysiological correlates of feature analysis during visual search. Psychophysiology, 31, 291–308.CrossRefGoogle ScholarPubMed
Lykken, D. T., Rose, R. J., Luther, B., & Maley, M. (1966). Correcting psychophysiological measures for individual differences in range. Psychological Bulletin, 66, 481–484.CrossRefGoogle ScholarPubMed
Lykken, D. T., & Venables, P. H. (1971). Direct measurement of skin conductance: A proposal for standardization. Psychophysiology, 8, 656–672.CrossRefGoogle ScholarPubMed
MacGregor, J. (2000). Introduction to the anatomy and physiology of children. Oxford, UK: Routledge.Google Scholar
Makeig, S., Jung, T. P., Bell, A. J., Ghahremani, D., & Sejnowski, T. J. (1997). Blind separation of auditory event-related brain responses into independent components. Proceedings of the National Academy of Science, 94, 10979–10984.CrossRefGoogle ScholarPubMed
Markand, O. N. (1994). Brainstem auditory evoked potentials. Journal of Clinical Neurophysiology, 11, 319–342.CrossRefGoogle ScholarPubMed
Marshall, P. J., Bar-Haim, Y., & Fox, N. A. (2002). Development of the EEG from 5 months to 4 years of age. Clinical Neurophysiology, 113, 1199–1208.CrossRefGoogle ScholarPubMed
Martin, C. E., Shaver, J. A., Thompson, M. E., Reddy, P. S., & Leonard, J. J. (1971). Direct correlation of external systolic time intervals with internal indices of left ventricular function in man. Circulation, 44, 419–431.CrossRefGoogle ScholarPubMed
Matthews, K. A., Salomon, K., Kenyon, K., & Allen, M. T. (2002). Stability of children's and adolescents' hemodynamic responses to psychological challenge: A three-year longitudinal study of a multiethnic cohort of boys and girls. Psychophysiology, 39, 826–834.CrossRefGoogle ScholarPubMed
McCall, W. A. (1923). How to experiment in education. New York: Macmillan.Google Scholar
McCarthy, G. (1999). Event-related potentials and functional MRI: A comparison of localization in sensory, perceptual, and cognitive tasks. Electroencephalography and Clinical Neurophysiology Supplement, 49, 3–12.Google ScholarPubMed
McCubbin, J. A., Richardson, J. E., Langer, A. W., Kizer, J. S., & Obrist, P. A. (1983). Sympathetic neuronal function and left ventricular performance during behavioral stress in humans: The relationship between plasma catecholamines and systolic time intervals. Psychophysiology, 20, 102–110.CrossRefGoogle ScholarPubMed
McGrath, J. J., & O'Brien, W. H. (2001). Pediatric impedance cardiography: Temporal stability and intertask consistency. Psychophysiology, 38, 479–484.CrossRefGoogle ScholarPubMed
McIsaac, H., & Polich, J. (1992). Comparison of infant and adult P300 from auditory stimuli. Journal of Experimental Child Psychology, 53, 115–128.CrossRefGoogle ScholarPubMed
Miller, A., & Tomarken, A. J. (2001). Task-dependent changes in frontal brain asymmetry: Effects of incentive cues, outcome expectancies, and motor responses. Psychophysiology, 38, 500–511.CrossRefGoogle ScholarPubMed
Miller, G. A., Gratton, G., & Yee, C. M. (1988). Generalized implementation of an eye movement correction procedure. Psychophysiology, 25, 241–243.CrossRefGoogle Scholar
Möcks, J., & Verleger, R. (1991). Multivariate methods in biosignal analysis: Application of principal component analysis to event-related potentials. In Weitkunat, R. (Ed.), Digital biosignal processing (pp. 399–458). Amsterdam, Netherlands: Elsevier Science.Google Scholar
Mohapatra, S. N. (1981). Non-invasive cardiovascular monitoring by electrical impedance technique. London: Pittman Medical Ltd.Google Scholar
Näätänen, R. (2003). Mismatch negativity: Clinical research and possible applications. International Journal of Psychophysiology, 48, 179–188.CrossRefGoogle ScholarPubMed
Näätänen, R., & Picton, T. (1987). The N1 wave of the human electric and magnetic response to sound: A review and an analysis of the component structure. Psychophysiology, 24, 375–425.CrossRefGoogle ScholarPubMed
Näätänen, R., & Teder, W. (1991). Attention effects on the auditory event-related potential. Acta Otolaryngology Supplement, 491, 161–167.CrossRefGoogle ScholarPubMed
Nelson, C. A. (1994). Neural correlates of recognition memory in the first postnatal year. In Dawson, G. & Fischer, K. W. (Eds.), Human behavior and the developing brain (pp. 269–313). New York: Guilford.Google Scholar
Nelson, C. A., & Monk, C. S. (2001). The use of the event-related potentials in the study of cognitive development. In Nelson, C. A. & Luciana, M. (Eds.), Handbook of developmental cognitive neuroscience (pp. 125–136). Cambridge, MA: MIT Press.Google Scholar
Newlin, D. B., & Levenson, R. W. (1979). Pre-ejection period: Measuring beta-adrenergic influences upon the heart. Psychophysiology, 16, 546–553.CrossRefGoogle ScholarPubMed
Niedermeyer, E. (1993). Maturation of the EEG: Development of waking and sleep patterns. In Niedermeyer, E. & Silva, F. H. Lopes da (Eds.), Electroencephalography: Basic principles, clinical applications, and related fields (pp. 167–191). Baltimore, MD: Williams & Wilkins.Google Scholar
Nitschke, J. B., Miller, G. A., & Cook, E. W. (1998). Digital filtering in EEG/ERP analysis: Some technical and empirical comparisons. Behavior Research Methods, Instruments, and Computers, 30, 54–67.CrossRefGoogle Scholar
Nuñez, P. L. (1981). Electrical fields of the brain: The neurophysics of EEG. New York: Oxford University Press.Google Scholar
Obrist, P. A., Light, K. C., James, S. A., & Strogatz, D. S. (1987). Cardiovascular responses to stress: I. Measures of myocardial response and relationships to high resting systolic pressure and parental hypertension. Psychophysiology, 24, 65–78.CrossRefGoogle Scholar
Pascual-Marqui, R. D., Esslen, M., Kochi, K., & Lehmann, D. (2002). Functional imaging with low-resolution brain electromagnetic tomography (LORETA): A review. Methods, Findings, and Experiments in Clinical Pharmacology, 24 Supplement, 91–95.Google ScholarPubMed
Pascual-Marqui, R. D., Michel, C. M., & Lehmann, D. (1994). Low resolution electromagnetic tomography: A new method for localizing electrical activity in the brain. International Journal of Psychophysiology, 18, 49–65.CrossRefGoogle Scholar
Penney, B. C., Patwardhan, N. A., & Wheeler, H. B. (1985). Simplified electrode array for impedance cardiography. Medical and Biological Engineering and Computing, 23, 1–7.CrossRefGoogle ScholarPubMed
Pernier, J., Perrin, F., & Bertrand, O. (1988). Scalp current density fields: Concept and properties. Electroencephalography and Clinical Neurophysiology, 69, 385–389.CrossRefGoogle ScholarPubMed
Perrin, F., Bertrand, O., & Pernier, J. (1987). Scalp current density mapping: Value and estimation from potential data. IEEE Transactions in Biomedical Engineering, 34, 283–288.CrossRefGoogle ScholarPubMed
Picton, T. W., Alain, C., Otten, L., Ritter, W., & Achim, A. (2000a). Mismatch negativity: Different water in the same river. Audiology and Neuro-Otology, 5, 111–139.CrossRefGoogle Scholar
Picton, T. W., Bentin, S., Berg, P., Donchin, E., Hillyard, S. A., Johnson, R., Miller, G. A., Ritter, W., Ruchkin, D. S., Rugg, M. D., & Taylor, M. J. (2000b). Guidelines for using human event-related potentials to study cognition: Recording standards and publication criteria. Psychophysiology, 37, 127–152.CrossRefGoogle Scholar
Picton, T. W., John, M. S., Dimitrijevic, A., & Purcell, D. (2003). Human auditory steady-state responses. International Journal of Audiology, 42, 177–219.CrossRefGoogle ScholarPubMed
Picton, T. W., Roon, P., Armilio, M. L., Berg, P., Ille, N., & Scherg, M. (2000c). The correction of ocular artifacts: A topographic perspective. Clinical Neurophysiology, 111, 53–65.CrossRefGoogle Scholar
Pivik, R. T., Broughton, R. J., Coppola, R., Davidson, R. J., Fox, N., & Nuwer, M. R. (1993). Guidelines for the recording and quantitative analysis of electroencephalographic activity in research contexts. Psychophysiology, 30, 547–558.CrossRefGoogle ScholarPubMed
Porges, S. W., & Byrne, E. A. (1992). Research methods for measurement of heart rate and respiration. Biological Psychology, 34, 93–130.CrossRefGoogle ScholarPubMed
Porges, S. W. (1986). Respiratory sinus arrhythmia: Physiological basis, quantitative methods, and clinical implications. In Grossman, P., Janssen, K., & Vaitl, D. (Eds.), Cardiorespiratory and cardiosomatic psychophysiology (pp. 101–115). New York: Plenum.CrossRefGoogle Scholar
Pritchard, W. S. (1981). Psychophysiology of P300. Psychological Bulletin, 89, 506–540.CrossRefGoogle ScholarPubMed
Raine, A. (2002). Biosocial studies of antisocial and violent behavior in children and adults: A review. Journal of Abnormal Child Psychology, 30, 311–326.CrossRefGoogle ScholarPubMed
Richards, J. E. (2000). Localizing the development of covert attention in infants using scalp event-related-potentials. Developmental Psychology, 36, 91–108.CrossRefGoogle ScholarPubMed
Richards, J. E. (2004). Recovering dipole sources from scalp-recorded event-related-potentials using component analysis: Principal component analysis and independent component analysis. International Journal of Psychophysiology, 54, 201–220.CrossRefGoogle ScholarPubMed
Riniolo, T., & Porges, S. W. (1997). Inferential and descriptive influences on measures of respiratory sinus arrhythmia: Sampling rate, R-wave trigger accuracy, and variance estimates. Psychophysiology, 34, 613–621.CrossRefGoogle ScholarPubMed
Rogan, J. C., Keselman, H. J., & Mendoza, J. L. (1979). Analysis of repeated measurements. British Journal of Mathematical and Statistical Psychology, 32, 269–286.CrossRefGoogle Scholar
Saltzberg, B., Burton, W. D., Burch, N. R., Fletcher, J., & Michaels, R. (1986). Electrophysiological measures of regional neural interactive coupling. Linear and non-linear dependence relationships among multiple channel electroencephalographic recordings. International Journal of Biomedical Computation, 18, 77–87.CrossRefGoogle ScholarPubMed
Scerbo, A., Freedman, L. W., Raine, A., Dawson, M. E., & Venables, P. H. (1992). A major effect of recording site on measurement of electrodermal activity. Psychophysiology, 29, 241–246.CrossRefGoogle Scholar
Schachinger, H., Weinbacher, M., Kiss, A., Ritz, R., & Langewitz, W. (2001). Cardiovascular indices of peripheral and central sympathetic activation. Psychosomatic Medicine, 63, 788–796.CrossRefGoogle ScholarPubMed
Scherg, M., & Berg, P. (1991). Use of prior knowledge in brain electromagnetic source analysis. Brain Topography 4:143–150.CrossRefGoogle ScholarPubMed
Scherg, M., Ille, N., Bornfleth, H., & Berg, P. (2002). Advanced tools for digital EEG review: Virtual source montages, whole-head mapping, correlation, and phase analysis. Journal of Clinical Neurophysiology, 19, 91–112.CrossRefGoogle ScholarPubMed
Schmidt, L. A., & Fox, N. A. (1998). Fear-potentiated startle responses in temperamentally different human infants. Developmental Psychobiology, 32, 113–120.3.0.CO;2-S>CrossRefGoogle ScholarPubMed
Schmidt, L. A., Fox, N. A., & Long, J. M. (1998). Acoustic startle electromyographic (EMG) activity indexed from an electrooculargraphic (EOG) electrode placement: A methodological note. International Journal of Neuroscience, 93, 185–188.CrossRefGoogle Scholar
Segalowitz, S. J., & Davies, P. L. (2004). Charting the maturation of the frontal lobe: An electrophysiological strategy. Brain and Cognition, 55, 116–133.CrossRefGoogle ScholarPubMed
Segalowitz, S. J. & Graves, R. (1990). Suitability of the IBM PC/AT/PS2 keyboard, mouse and game port as response devices in reaction time paradigms. Behavior Research Methods, Instruments, & Computers, 22, 283–289.CrossRefGoogle Scholar
Sherwood, A., Allen, M. T., Fahrenberg, J., Kelsey, R. M., Lovallo, W. R., & Doornen, L. J. P. (1990). Methodological guidelines for impedance cardiography. Psychophysiology, 27, 1–23.Google ScholarPubMed
Simons, R. F., & Zelson, M. F. (1985). Engaging visual stimuli and blink modification. Psychophysiology, 22, 44–49.CrossRefGoogle ScholarPubMed
Sloan, R. P., Shapiro, P. A., Bagiella, E., Gorman, J. M., & Bigger, J. T. (1995). Temporal stability of heart period variability during a resting baseline and in response to psychological challenge. Psychophysiology, 32, 191–196.CrossRefGoogle ScholarPubMed
Sokolov, E. N. (1990). The orienting response and future directions of its development. Pavlovian Journal of Biological Science, 25, 142–150.Google ScholarPubMed
Somsen, R. J. & Beek, B. (1998). Ocular artifacts in children's EEG: Selection is better than correction. Biological Psychology, 48, 281–300.CrossRefGoogle ScholarPubMed
Sonin, A. A. (2001). The physical basis of dimensional analysis. Cambridge, MA: MIT.Google Scholar
Spangler, G. (1991). The emergence of adrenocortical circadian function in newborns and infants and its relationship to sleep, feeding, and maternal adrenocortical activity. Early Human Development, 25, 197–208.CrossRefGoogle ScholarPubMed
Squires, K. C., & Donchin, E. (1976). Beyond averaging: The use of discriminant functions to recognize event related potentials elicited by single auditory stimuli. Electroencephalography and Clinical Neurophysiology, 41, 449–459.CrossRefGoogle ScholarPubMed
Stern, J. A. (1964). Toward a definition of psychophysiology. Psychophysiology, 1, 90–91.CrossRefGoogle Scholar
Stern, J. A. (1968). Toward a developmental psychophysiology: My look into the crystal ball. Psychophysiology, 4, 403–420.CrossRefGoogle Scholar
Strauss, M. E. (2001). Demonstrating specific cognitive deficits: A psychometric perspective. Journal of Abnormal Psychology, 110, 6–14.CrossRefGoogle ScholarPubMed
Tassinary, L. G. & Cacioppo, J. T. (2000). The skeletomotor system: Surface electromyography. In Cacioppo, J. T., Tassinary, L. G., & Berntson, G. G. (Eds.), Handbook of psychophysiology (2nd ed., pp. 163–199). New York: Cambridge.Google Scholar
Taylor, M. J., McCarthy, G., Saliba, E., & Degiovanni, E. (1999). ERP evidence of developmental changes in processing of faces. Clinical Neurophysiology, 110, 910–915.CrossRefGoogle ScholarPubMed
Thomas, D. G., & Crow, C. D. (1994). Development of evoked electrical brain activity in infancy. In Dawson, G. & Fischer, K. W. (Eds.), Human behavior and the developing brain (pp. 207–231). New York: Guilford.Google Scholar
Tomarken, A. J. (1995). A psychometric perspective on psychophysiological measures. Psychological Assessment, 7, 387–395.CrossRefGoogle Scholar
Tomarken, A. J. (1999). Methodological issues in psychophysiological research. In Kendall, P. C., Butcher, J. N., & Holmbeck, G. N. (Eds.), Handbook of research methods in clinical psychology (2nd ed., pp. 251–275). New York: John Wiley & Sons.Google Scholar
Tomarken, A. J., Davidson, R. J., Wheeler, R. E., & Kinney, L. (1992). Psychometric properties of resting anterior EEG asymmetry: Temporal stability and internal consistency. Psychophysiology, 29, 576–592.CrossRefGoogle ScholarPubMed
Molen, M. W., & Ridderinkhof, K. R. (1998). Chronopsychophysiology of developmental changes in selective attention and processing speed: A selective review and re-analysis. Journal of Psychophysiology, 12, 223–235.Google Scholar
Vasey, M., & Thayer, J. F. (1987). The continuing problem of false positives in repeated measures ANOVA in psychophysiology: A multivariate solution. Psychophysiology, 24, 479–486.CrossRefGoogle ScholarPubMed
Wackermann, J., & Matousek, M. (1998). From the ‘EEG age’ to a rational scale of brain electric maturation. Electroencephalography and Clinical Neurophysiology, 107, 415–421.CrossRefGoogle ScholarPubMed
Wallstrom, G. L., Kass, R. E., Miller, A., Cohn, J. F., & Fox, N. A. (2004). Automatic correction of ocular artifacts in the EEG: A comparison of regression-based and component-based methods. International Journal of Psychophysiology, 53, 105–119.CrossRefGoogle ScholarPubMed
Walter, W. G., Cooper, R., Aldridge, V. J., McCallum, W. C., & Winter, A. L. (1964). Contingent negative variation: An electrical sign of sensorimotor association and expectancy in the human brain. Nature, 203, 380–384.CrossRefGoogle Scholar
Watamura, S. E., Donzella, B., Kertes, D. A., & Gunnar, M. R. (2004). Developmental changes in baseline cortisol activity in early childhood: Relations with napping and effortful control. Developmental Psychobiology, 45, 125–133.CrossRefGoogle ScholarPubMed
West, J. B. (2000). Respiratory physiology: The essentials. (6th ed.). Philadelphia: Lippincott Williams & Wilkins.Google Scholar
Wientjes, C. J. E. (1992). Respiration in psychophysiology: Methods and application. Biological Psychology, 34, 179–203.CrossRefGoogle Scholar
Wilhelm, F. H., Grossman, P., & Coyle, M. A. (2004). Improving estimation of cardiac vagal tone during spontaneous breathing using a paced breathing calibration. Biomedical Science Instrumentation, 40, 317–324.Google ScholarPubMed
Yao, D., Wang, L., Oostenveld, R., Nielsen, K. D., Arendt-Nielsen, L., & Chen, A. C. N. (2005). A comparative study of different references for EEG spectral mapping: The issue of neutral reference and the use of the infinity reference. Physiological Measurement, 26, 173–184.CrossRefGoogle ScholarPubMed
Yordanova, J., & Kolev, V. (1996). Developmental changes in the alpha response system. Electroencephalography and Clinical Neurophysiology, 99, 527–538.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×